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

Topological K-theory of quasi-BPS categories for Higgs bundles

Tudor Pădurariu and Yukinobu Toda
Abstract.

In a previous paper, we introduced quasi-BPS categories for moduli stacks of semistable Higgs bundles. Under a certain condition on the rank, Euler characteristic, and weight, the quasi-BPS categories (called BPS in this case) are non-commutative analogues of Hitchin integrable systems. We proposed a conjectural equivalence between BPS categories which swaps Euler characteristics and weights. The conjecture is inspired by the Dolbeault Geometric Langlands equivalence of Donagi–Pantev, by the Hausel–Thaddeus mirror symmetry, and by the χ\chi-independence phenomenon for BPS invariants of curves on Calabi-Yau threefolds.

In this paper, we show that the above conjecture holds at the level of topological K-theories. When the rank and the Euler characteristic are coprime, such an isomorphism was proved by Groechenig–Shen. Along the way, we show that the topological K-theory of BPS categories is isomorphic to the BPS cohomology of the moduli of semistable Higgs bundles.

1. Introduction

1.1. Hausel–Thaddeus mirror symmetry of Higgs bundles

Let CC be a smooth projective curve of genus gg, and let the group GG be either GL(r),SL(r)\mathrm{GL}(r),\mathrm{SL}(r) or PGL(r)\mathrm{PGL}(r). We denote by

π:MG(χ)B\displaystyle\pi\colon M_{G}(\chi)\to B

the moduli space of semistable GG-Higgs bundles with Euler characteristic χ\chi together with the Hitchin fibration with Hitchin base BB. In the case of G=GL(r)G=\mathrm{GL}(r), it consists of pairs

(F,θ),θ:FFΩC,\displaystyle(F,\theta),\ \theta\colon F\to F\otimes\Omega_{C},

where FF is a vector bundle on CC with (rank(F),χ(F))=(r,χ)(\operatorname{rank}(F),\chi(F))=(r,\chi) and (F,θ)(F,\theta) satisfies a stability condition.

In general, the stack MG(χ)M_{G}(\chi) is singular, however it is a smooth Deligne-Mumford stack if (r,χ)(r,\chi) are coprime. Suppose that both of χ\chi and ww are coprime with rr. In this case, Hausel–Thaddeus [HT03] proposed that the pair of smooth Deligne-Mumford stacks (together with some Brauer classes)

(MG(χ),MGL(w))\displaystyle(M_{G}(\chi),M_{G^{L}}(w))

is a mirror pair [SYZ01]. In particular, Hausel–Thaddeus proposed the equality of the stringy Hodge numbers of MG(χ)M_{G}(\chi) and MGL(w)M_{G^{L}}(w), which was proved by Groechenig–Wyss–Ziegler [GWZ20]. At the categorical level, one expects a derived equivalence [DP12] (called “the Dolbeault Langlands equivalence” [BZN18]):

(1.1) Db(MG(χ),αw)Db(MGL(w),βχ),\displaystyle D^{b}(M_{G}(\chi),\alpha^{-w})\stackrel{{\scriptstyle\sim}}{{\to}}D^{b}(M_{G^{L}}(w),\beta^{\chi}),

where α,β\alpha,\beta are some canonical Brauer classes, see for example [Hau22, Section 2.4]. Heuristically (following Donagi–Pantev [DP12]), the conjectural equivalence (1.1) may be regarded as a classical limit of the geometric de Rham Langlands correspondence [AG15]:

(1.2) Ind𝒩Db(ocSysG)D-mod(unGL).\displaystyle\mathrm{Ind}_{\mathcal{N}}D^{b}(\mathcal{L}\mathrm{ocSys}_{G})\simeq\text{D-mod}(\mathcal{B}\mathrm{un}_{G^{L}}).

Here ocSysG\mathcal{L}\mathrm{ocSys}_{G} is the moduli stack of GG-flat connections on CC and unGL\mathcal{B}\mathrm{un}_{G^{L}} is the moduli stack of GLG^{L}-bundles on CC. The equivalence (1.1) may be regarded as an extension over the full Hitchin base of the Fourier-Mukai equivalence between dual abelian schemes, which are constructed from the Picard schemes of the smooth spectral curves. The equivalence (1.1) should match Hecke and Wilson operators on the two sides [KW07, DP12, Hau22], but we do not discuss this aspect in the paper.

The existence of an equivalence (1.1) is an open problem in most of the cases. Besides the equality of Hodge stringy numbers conjectured in [HT03], it is known that versions of (1.1) hold in cohomology (by Groechenig–Wyss–Ziegler [GWZ20], see also [LW21]), for Hodge structures (by Maulik–Shen [MS21]), for Chow and Voevodsky motives (by Hoskins–Pepin Lehalleur [HLb]), and for relative Chow motives (by Maulik–Shen–Yin [MSY]). Groechenig–Shen considered a comparison in topological K-theory [GS], and showed an equivalence of (integral) topological K-theories spectra:

(1.3) Ktop(MG(χ),αw)Ktop(MGL(w),βχ).\displaystyle K^{\rm{top}}(M_{G}(\chi),\alpha^{-w})\stackrel{{\scriptstyle\sim}}{{\to}}K^{\rm{top}}(M_{G^{L}}(w),\beta^{\chi}).

The first main difficulty in proving the equivalence (1.1) is the construction of a candidate Fourier-Mukai kernel, which should be an extension of the Poincaré sheaf, which is initially defined only over the locus of smooth spectral curves. Such an extension was constructed by Arinkin [Ari13] over the elliptic locus (when the spectral curve is reduced and irreducible), by Melo–Rapagnetta–Viviani [MRV19a, MRV19b] over the locus of reduced spectral curves, and by Li [Li21] for rank two meromorphic Higgs bundles. However, this is not an impediment in proving the statements in [GWZ20, LW21, MS21, HLb, GS, MSY], which follow from a good understanding of the comparison of the two sides over the elliptic locus. For example, any extension over the full Hitchin base of Arinkin’s kernel induces the isomorphism (1.3).

The purpose of this paper is to give a version of the equivalence (1.3) for integers χ,w\chi,w which are not necessary coprime with rr, using the quasi-BPS categories studied in the previous paper [PTa].

1.2. Symmetry of quasi-BPS categories

So far in the literatures, the studies of mirror symmetry of Higgs bundles have been restricted to the case when (r,χ)(r,\chi) are coprime. In this case, the moduli space MG(χ)M_{G}(\chi) is a smooth Deligne-Mumford stack, and its derived category has several nice properties, for example it is smooth over \mathbb{C} and proper over the Hitchin base BB.

However, it is important to study the (derived) moduli stacks of semistable Higgs bundles G(χ)\mathcal{M}_{G}(\chi) for general χ\chi. First, all such moduli stacks are used in the definition of the categorical Hall algebra of the surface TotC(ΩC)\mathrm{Tot}_{C}(\Omega_{C}) [PS23], and needed to be studied in order to categorify theorems known for (Kontsevich–Soibelman [KS11]) cohomological Hall algebras [KK, DHSM].

Second, the stack ocSysG\mathcal{L}\mathrm{ocSys}_{G} degenerates to G(0)\mathcal{M}_{G}(0), see [Sim97, Proposition 4.1]. Thus, when studying the limit of the de Rham Langlands equivalence (1.2) (following Donagi–Pantev), the limit of the right hand side should be a category of ind-coherent sheaves on G(0)\mathcal{M}_{G}(0).

Third, for G=SL(r)\mathrm{G}=\mathrm{SL}(r), principal SL(r)\mathrm{SL}(r)-Higgs bundles have degree zero, thus their Euler characteristic is equal to r(1g)r(1-g), which is divisible by rr. Therefore, considering quasi-BPS categories for χ=r(1g)\chi=r(1-g) is essential in the categorical study of principal SL\mathrm{SL}-Higgs bundles.

In [PTa], we introduced some admissible subcategories, called quasi-BPS categories:

(1.4) 𝕋G(χ)wDb(G(χ))w\displaystyle\mathbb{T}_{G}(\chi)_{w}\subset D^{b}(\mathcal{M}_{G}(\chi))_{w}

for ww\in\mathbb{Z} corresponding to a weight with respect to the action of the center of GG. The construction of the categories (1.4) is part of the problem of categorifying BPS invariants on Calabi-Yau 3-folds [PTc, PTe, PTd], or more generally of categorifying the constructions and theorems from (numerical or cohomological) Donaldson-Thomas theory [Tod24]. Indeed, the quasi-BPS categories (1.4) have analogous properties to the BPS cohomology (defined by Kinjo–Koseki [KK] and Davison–Hennecart–Schlegel Mejia [DHSM]) of the local Calabi-Yau threefold X=TotC(ΩC)×𝔸1X=\mathrm{Tot}_{C}(\Omega_{C})\times\mathbb{A}^{1}_{\mathbb{C}}, see [PTa] for more details. In this paper, we make this relation precise by computing the topological K-theory of quasi-BPS categories in terms of BPS cohomology, see Proposition 4.1, Proposition 4.5, and Theorem 6.9.

If the vector (r,χ,w)(r,\chi,w) is primitive, i.e. gcd(r,χ,w)=1\gcd(r,\chi,w)=1, the category (1.4) is smooth over \mathbb{C} and proper over the Hitchin base BB. In this case, we regard it as a non-commutative analogue of the Hitchin system. Note that neither χ\chi nor ww may be coprime with rr, even if (r,χ,w)(r,\chi,w) is primitive. In [PTa], we conjectured the existence of an equivalence

(1.5) 𝕋G(w)χ𝕋GL(χ)w\displaystyle\mathbb{T}_{G}(w)_{-\chi}\stackrel{{\scriptstyle\sim}}{{\to}}\mathbb{T}_{G^{L}}(\chi)_{w}

extending the Donagi–Pantev equivalence over the locus of smooth spectral curves, which is nothing but the equivalence (1.1) if both χ\chi and ww are coprime with rr.

1.3. Main theorem

The purpose of this paper is to provide evidence towards the equivalence (1.5), namely to prove that extensions of the Poincaré sheaf induce an isomorphism of the (rational) topological K-theories of the two categories in (1.5). We thus obtain a generalization of the Groechenig–Shen theorem (1.3) beyond the coprime case. Note that (1.3) holds integrally, and that we also prove versions for integral topological K-theory for twisted Higgs bundles.

For a dg-category 𝒟\mathscr{D}, Blanc [Bla16] introduced its topological K-theory spectrum

K(𝒟)Sp.\displaystyle K(\mathscr{D})\in\mathrm{Sp}.

We denote by K(𝒟):=K(𝒟)HK(\mathscr{D})_{\mathbb{Q}}:=K(\mathscr{D})\wedge H\mathbb{Q} its rationalization.

We use the following notations

𝕋(r,χ)w:=𝕋GL(r)(χ)w,𝕋SL(r),w:=𝕋SL(r)(r(1g))w,𝕋PGL(r)(χ):=𝕋PGL(r)(χ)0.\displaystyle\mathbb{T}(r,\chi)_{w}:=\mathbb{T}_{\mathrm{GL}(r)}(\chi)_{w},\ \mathbb{T}_{\mathrm{SL}(r),w}:=\mathbb{T}_{\mathrm{SL}(r)}(r(1-g))_{w},\ \mathbb{T}_{\mathrm{PGL}(r)}(\chi):=\mathbb{T}_{\mathrm{PGL}(r)}(\chi)_{0}.

We also use the notation 𝕋(r,χ)wred\mathbb{T}(r,\chi)_{w}^{\rm{red}} for the reduced quasi-BPS category, which is a category obtained from the usual quasi-BPS categories by removing a redundant derived structure from G(χ)\mathcal{M}_{G}(\chi). The following is the main theorem in this paper:

Theorem 1.1.

(Theorem 6.11, Theorem 8.21)

(1) Suppose that the vector (r,χ,w)(r,\chi,w) is primitive. For G=GL(r)G=\mathrm{GL}(r), there is an equivalence

(1.6) Ktop(𝕋(r,w)χred)Ktop(𝕋(r,χ)wred).\displaystyle K^{\rm{top}}(\mathbb{T}(r,w)_{-\chi}^{\rm{red}})_{\mathbb{Q}}\stackrel{{\scriptstyle\sim}}{{\to}}K^{\rm{top}}(\mathbb{T}(r,\chi)_{w}^{\rm{red}})_{\mathbb{Q}}.

(2) Suppose that gcd(r,w)=1\gcd(r,w)=1. For (G,GL)=(PGL(r),SL(r))(G,G^{L})=(\mathrm{PGL}(r),\mathrm{SL}(r)), there is an equivalence

Ktop(𝕋PGL(r)(w))Ktop(𝕋SL(r),w).\displaystyle K^{\rm{top}}(\mathbb{T}_{\mathrm{PGL}(r)}(w))_{\mathbb{Q}}\stackrel{{\scriptstyle\sim}}{{\to}}K^{\rm{top}}(\mathbb{T}_{\mathrm{SL}(r),w})_{\mathbb{Q}}.

The main ingredient in the proof of the above theorem is a computation of the (rational) topological K-theory of BPS categories. For example, for G=GL(r)G=\mathrm{GL}(r), we show in Theorem 6.9 that

(1.7) 𝒦Mtop(𝕋(r,χ)wred)𝒫𝒮M[β±1],\displaystyle\mathcal{K}_{M}^{\rm{top}}(\mathbb{T}(r,\chi)_{w}^{\rm{red}})_{\mathbb{Q}}\cong\mathcal{BPS}_{M}[\beta^{\pm 1}],

where 𝒦Mtop()\mathcal{K}_{M}^{\rm{top}}(-) is the relative topological K-theory [Mou19] over MM, 𝒫𝒮M\mathcal{BPS}_{M} is the BPS sheaf defined by Kinjo–Koseki [KK] for M=MGL(r)(χ)M=M_{GL(r)}(\chi), and β\beta is of degree 22. The isomorphism (1.7) explains the use of the name of BPS categories. Note that 𝒫𝒮M\mathcal{BPS}_{M} is (a shift of) the constant sheaf M\mathbb{Q}_{M} if (r,χ)(r,\chi) are coprime, and that in general 𝒫𝒮M\mathcal{BPS}_{M} contains ICM\mathrm{IC}_{M} as a direct summand. The two sides on (1.6) are thus isomorphic because of (cohomological) χ\chi-independence [KK, Theorem 1.2]. We use the methods of [GS] to show that an extension of Arinkin’s kernel induces the isomorphism (1.6).

Note that the isomorphism (1.7) shows the weight independence of topological K-theory of BPS categories, which is a phenomenon we also discussed in [PTe, PTc]. The equivalence (1.5) thus interchanges the weight and χ\chi-independence phenomena for Higgs bundles.

1.4. The LL-twisted case

The result of Theorem 1.2 is deduced from analogous results for LL-twisted (i.e. meromorphic) Higgs bundles, where

LCL\to C

is a line bundle with degL>2g2\deg L>2g-2, using the method of vanishing cycles as in [KK, MS21]. We denote by GL(χ)\mathcal{M}^{L}_{G}(\chi) the moduli stack of LL-twisted semistable GG-Higgs bundles. Note that GL(χ)\mathcal{M}^{L}_{G}(\chi) is a smooth stack, whereas G(χ)\mathcal{M}_{G}(\chi) is, in general, singular and non-separated. In the case of G=GL(r)G=\mathrm{GL}(r), the moduli stack GL(χ)\mathcal{M}^{L}_{G}(\chi) consists of pairs

(F,θ),θ:FFL,\displaystyle(F,\theta),\ \theta\colon F\to F\otimes L,

where FF is a vector bundle and (F,θ)(F,\theta) satisfies a stability condition. We can similarly define the quasi-BPS category

𝕋GL(χ)wDb(L(χ))w.\displaystyle\mathbb{T}_{G}^{L}(\chi)_{w}\subset D^{b}(\mathcal{M}^{L}(\chi))_{w}.

We use the notation as in the previous subsection, e.g. 𝕋L(r,χ)w:=𝕋GL(r)L(χ)w\mathbb{T}^{L}(r,\chi)_{w}:=\mathbb{T}_{\mathrm{GL}(r)}^{L}(\chi)_{w}.

As before, we regard quasi-BPS categories as the categorical replacement of the BPS cohomology for the local Calabi-Yau threefold TotC(LL1ΩC)\mathrm{Tot}_{C}(L\oplus L^{-1}\otimes\Omega_{C}). Indeed, the derived category of coherent sheaves on L(χ)\mathcal{M}^{L}(\chi) has a semiorthogonal decomposition in Hall products of quasi-BPS categories, analogous to the decomposition of the cohomology of L(χ)\mathcal{M}^{L}(\chi) in terms of the BPS cohomology of ML:=MGL(r)L(χ)M^{L}:=M^{L}_{GL(r)}(\chi), which is isomorphic to its intersection cohomology [Mei]. In Propositions 4.1 and 4.5, we compute the (rational) topological K-theory of quasi-BPS categories in terms of BPS cohomology using the results and methods for quivers [PTe]. In particular, we show that, if (r,χ,w)(r,\chi,w) satisfies the BPS condition, then

(1.8) 𝒦MLtop(𝕋L(r,χ)w)ICML[dimML][β±1].\mathcal{K}_{M^{L}}^{\rm{top}}(\mathbb{T}^{L}(r,\chi)_{w})_{\mathbb{Q}}\cong\mathrm{IC}_{M^{L}}[-\dim M^{L}][\beta^{\pm 1}].

The vector (r,χ,w)(r,\chi,w) satisfies the BPS condition if the vector

(r,χ,w+1gsp)3\displaystyle(r,\chi,w+1-g^{\rm{sp}})\in\mathbb{Z}^{3}

is primitive, where gspg^{\rm{sp}} is the genus of the spectral curve, see the formula (2.9). Note that, if degL\deg L is even, the BPS condition is equivalent to the vector (r,χ,w)(r,\chi,w) being primitive. We say that (r,w)(r,w) satisfies the BPS condition if (r,0,w)(r,0,w) satisfies the BPS condition. One can also formulate, for LL-twisted Higgs bundles, a conjectural derived equivalence analogous to (1.1), (1.5), see [PTa, Conjecture 4.3]. We prove its version for topological K-theory:

Theorem 1.2.

(Theorem 5.10, Corollary 8.20) Let LCL\to C be a line bundle of degL>2g2\deg L>2g-2.

(1) Suppose that (r,χ,w)(r,\chi,w) satisfies the BPS condition. For G=GL(r)G=\mathrm{GL}(r), there is an equivalence

(1.9) Ktop(𝕋L(r,w+1gsp)χ+1gsp)Ktop(𝕋L(r,χ)w).\displaystyle K^{\rm{top}}(\mathbb{T}^{L}(r,w+1-g^{\rm{sp}})_{-\chi+1-g^{\rm{sp}}})\stackrel{{\scriptstyle\sim}}{{\to}}K^{\rm{top}}(\mathbb{T}^{L}(r,\chi)_{w}).

(2) Suppose that (r,w)(r,w) satisfies the BPS condition. For (G,GL)=(PGL(r),SL(r))(G,G^{L})=(\mathrm{PGL}(r),\mathrm{SL}(r)), there is an equivalence

Ktop(𝕋PGL(r)L(w))Ktop(𝕋SL(r),wL).\displaystyle K^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{PGL}(r)}(w))\stackrel{{\scriptstyle\sim}}{{\to}}K^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{SL}(r),w}).

As in the case L=ΩCL=\Omega_{C}, a first step in proving part (1) is to show that both sides in (1.9) have isomorphic rational topological K-theory. This follows from (1.8) and the cohomological χ\chi-independence for twisted Higgs bundles proved by Maulik–Shen [MS23]. Similarly, part (2) uses a cohomological SL/PGL-duality for twisted Higgs bundles [MS22].

Note that the equivalences of topological K-theories in Theorem 1.2 hold integrally. Indeed, we show that the topological K-groups in Theorem 1.2 are torsion-free. In Theorem 8.19, we give a slightly stronger statement of part (2) of Theorem 1.2 which also involves weight/Euler characteristics on left/hand hand sides.

1.5. Complements

As we mentioned in the beginning of Subsection 1.2, it is important to study all quasi-BPS categories. Thus it is natural to inquire whether a derived equivalence (1.5), or isomorphisms (1.6) or (1.9), may holds for all quasi-BPS categories. An analogous such comparison was discussed in [PTb, Section 1.3] for quasi-BPS categories of points on a surface, which includes points (i.e. rank zero Higgs sheaves) on TotC(ΩC)\mathrm{Tot}_{C}(\Omega_{C}). We mention a rational analogue of the isomorphism (1.9) in Subsection 4.2.

It is also interesting to pursue an integral version of Theorem 1.1.

The methods of this paper (and of [PTe]) may be used to compute the topological K-theory of quasi-BPS categories (or of noncommutative resolutions defined by Špenko–Van den Bergh [ŠdB17]) for other smooth (or quasi-smooth) symmetric stacks, for example for the moduli stack un(r,d)ss\mathcal{B}\mathrm{un}(r,d)^{\mathrm{ss}} of semistable vector bundles of rank rr and degree dd on a smooth projective curve CC. In Subsection 4.3, we briefly discuss the computation of topological K-theory of quasi-BPS categories of un(r,d)ss\mathcal{B}\mathrm{un}(r,d)^{\mathrm{ss}} in terms of the intersection cohomology of the good moduli space Bun(r,d)ss\mathrm{Bun}(r,d)^{\mathrm{ss}}. Note that in [PTa, Subsection 3.4] we explained that Db(un(r,d)ss)D^{b}(\mathcal{B}\mathrm{un}(r,d)^{\mathrm{ss}}) has a semiorthogonal decomposition in Hall products of quasi-BPS categories. Thus the results in loc.cit. and in Subsection 4.3 provide a K-theoretic (or categorical) version of a theorem of Mozgovoy–Reineke [MR, Theorem 1.3].

1.6. Acknowledgements

T. P. thanks MPIM Bonn and CNRS for their support during part of the preparation of this paper. This material is partially based upon work supported by the NSF under Grant No. DMS-1928930 and by the Alfred P. Sloan Foundation under grant G-2021-16778, while T. P. was in residence at SLMath in Berkeley during the Spring 2024 semester. T. P. thanks Kavli IPMU for their hospitality and excellent working conditions during a visit in May 2024.

This work started while Y. T. was visiting to Hausdorff Research Institute for Mathematics in Bonn on November 12-18, 2023. Y. T. thanks the hospitality of HIM Bonn during his visit. Y. T. is supported by World Premier International Research Center Initiative (WPI initiative), MEXT, Japan, and JSPS KAKENHI Grant Numbers JP19H01779, JP24H00180.

1.7. Notation and convention

In this paper, all the (derived) stacks are defined over \mathbb{C}. For a stack 𝒳\mathcal{X}, we use the notion of good moduli space from [Alp13]. It generalizes the notion of GIT quotient

R/GR//G,R/G\to R/\!\!/G,

where RR is an affine variety on which a reductive group GG acts.

For a (classical) stack 𝒳\mathcal{X}, denote by D(Sh(𝒳))D(\mathrm{Sh}_{\mathbb{Q}}(\mathcal{X})) the derived category of complexes of \mathbb{Q}-constructible sheaves [Ols07].

For a torus TT, a character χ\chi, and cocharacter λ\lambda, we denote by λ,χ\langle\lambda,\chi\rangle\in\mathbb{Z} its natural pairing. For a TT-representation VV, we denote by Vλ>0VV^{\lambda>0}\subset V the subspace spanned by TT-weights β\beta with λ,β>0\langle\lambda,\beta\rangle>0. We also write λ,Vλ>0:=λ,det(Vλ>0)\langle\lambda,V^{\lambda>0}\rangle:=\langle\lambda,\det(V^{\lambda>0})\rangle.

For a variety or stack 𝒳\mathcal{X}, we denote by Db(𝒳)D^{b}(\mathcal{X}) the bounded derived category of coherent sheaves, which is a pre-triangulated dg-category. We denote by Perf(𝒳)\mathrm{Perf}(\mathcal{X}) the category of perfect complexes and Dqcoh(𝒳)D_{\rm{qcoh}}(\mathcal{X}) the unbounded derived category of quasi-coherent sheaves. For pre-triangulated dg-subcategories 𝒞iDb(𝒳i)\mathcal{C}_{i}\subset D^{b}(\mathcal{X}_{i}) for i=1,2i=1,2, we denote by 𝒞1𝒞2\mathcal{C}_{1}\boxtimes\mathcal{C}_{2} the smallest pre-triangulated dg-subcategory of Db(𝒳1×𝒳2)D^{b}(\mathcal{X}_{1}\times\mathcal{X}_{2}) which contains objects E1E2E_{1}\boxtimes E_{2} for Ei𝒞iE_{i}\in\mathcal{C}_{i} and closed under direct summands.

For a dg-category 𝒟\mathcal{D} with Perf(B)\mathrm{Perf}(B)-module structure for a scheme BB, its semiorthogonal decomposition 𝒟=𝒞i|iI\mathcal{D}=\langle\mathcal{C}_{i}\,|\,i\in I\rangle is called Perf(B)\mathrm{Perf}(B)-linear if 𝒞iPerf(B)𝒞i\mathcal{C}_{i}\otimes\mathrm{Perf}(B)\subset\mathcal{C}_{i} for all iIi\in I.

For a smooth stack 𝒳\mathcal{X} and a regular function f:𝒳f\colon\mathcal{X}\to\mathbb{C}, we denote by MF(𝒳,f)\mathrm{MF}(\mathcal{X},f) the /2\mathbb{Z}/2-graded dg-category of matrix factorizations of ff. If there is a \mathbb{C}^{\ast}-action on 𝒳\mathcal{X} such that ff is of weight one, we also consider the dg-category of graded matrix factorizations MFgr(𝒳,f)\mathrm{MF}^{\rm{gr}}(\mathcal{X},f). We refer to [PT24, Section 2.6] for a review of (graded) matrix factorizations.

2. Quasi-BPS categories for GL-Higgs bundles

In this section, we first recall some basic properties of moduli stacks of semistable Higgs bundles for G=GL(r)G=\mathrm{GL}(r), such as its local description using quivers and the BNR spectral correspondence. We then recall the definition of quasi-BPS categories for Higgs bundles and their conjectural symmetry. We finally briefly discuss Joyce-Song pairs for Higgs bundles.

2.1. Twisted Higgs bundles

Let CC be a smooth projective curve of genus gg. Let LL be a line bundle on CC such that either

l:=degL>2g2 or L=ΩC.\displaystyle l:=\deg L>2g-2\mbox{ or }L=\Omega_{C}.

By definition, a LL-twisted Higgs bundle is a pair (F,θ)(F,\theta), where FF is a vector bundle on CC and θ\theta is a morphism

θ:FFL.\displaystyle\theta\colon F\to F\otimes L.

When L=ΩCL=\Omega_{C}, it is just called a Higgs bundle. The (semi)stable LL-twisted Higgs bundle is defined using the slope μ(F)=χ(F)/rank(F)\mu(F)=\chi(F)/\operatorname{rank}(F) in the usual way: a LL-twisted Higgs bundle (F,θ)(F,\theta) is (semi)stable if we have

μ(F)<()μ(F),\displaystyle\mu(F^{\prime})<(\leqslant)\mu(F),

for any sub-Higgs bundle (F,θ)(F,θ)(F^{\prime},\theta^{\prime})\subset(F,\theta) such that rank(F)<rank(F)\operatorname{rank}(F^{\prime})<\operatorname{rank}(F).

A LL-twisted Higgs bundle corresponds to a compactly supported pure one-dimensional coherent sheaf on the non-compact surface

p:S=TotC(L)C.\displaystyle p\colon S=\mathrm{Tot}_{C}(L)\to C.

The correspondence (called the spectral construction [BNR89]) is given as follows: for a given LL-twisted Higgs pair (F,θ)(F,\theta), the Higgs field θ\theta determines the p𝒪Sp_{\ast}\mathcal{O}_{S}-module structure on FF, which in turn gives a coherent sheaf on SS. Conversely, a pure one-dimensional compactly supported sheaf EE on SS pushes forward to a vector bundle FF with Higgs field θ\theta given by the p𝒪Sp_{\ast}\mathcal{O}_{S}-module structure on it.

2.2. Moduli stacks of Higgs bundles

We denote by

(2.1) L(r,χ)\displaystyle\mathcal{M}^{L}(r,\chi)

the derived moduli stack of semistable LL-twisted Higgs bundles (F,θ)(F,\theta) with

(rank(F),χ(F))=(r,χ).\displaystyle(\operatorname{rank}(F),\chi(F))=(r,\chi).

It is smooth (in particular classical) when degL>2g2\deg L>2g-2, and quasi-smooth when L=ΩCL=\Omega_{C}. We omit LL in the notation when L=ΩCL=\Omega_{C}, i.e. (r,χ):=ΩC(r,χ)\mathcal{M}(r,\chi):=\mathcal{M}^{\Omega_{C}}(r,\chi). We also denote by (,ϑ)(\mathcal{F},\vartheta) the universal Higgs bundle

(2.2) Coh(C×L(r,χ)),ϑ:L.\displaystyle\mathcal{F}\in\mathrm{Coh}(C\times\mathcal{M}^{L}(r,\chi)),\ \vartheta\colon\mathcal{F}\to\mathcal{F}\boxtimes L.

The stack (2.1) is equipped with the Hitchin map

h:L(r,χ)BL(r,χ):=i=1rH0(C,Li)\displaystyle h\colon\mathcal{M}^{L}(r,\chi)\to B^{L}(r,\chi):=\bigoplus_{i=1}^{r}H^{0}(C,L^{\otimes i})

sending (F,θ)(F,\theta) to tr(θi)\mathrm{tr}(\theta^{i}) for 1ir1\leqslant i\leqslant r. When r,χr,\chi are clear from the context, we write BL:=BL(r,χ)B^{L}:=B^{L}(r,\chi), B:=BΩCB:=B^{\Omega_{C}}. We have the factorization

h:L(r,χ)clπML(r,χ)BL,\displaystyle h\colon\mathcal{M}^{L}(r,\chi)^{\rm{cl}}\stackrel{{\scriptstyle\pi}}{{\to}}M^{L}(r,\chi)\to B^{L},

where the first morphism is the good moduli space morphism. A closed point yML(r,χ)y\in M^{L}(r,\chi) corresponds to a polystable Higgs bundle

(2.3) E=i=1kViEi,\displaystyle E=\bigoplus_{i=1}^{k}V_{i}\otimes E_{i},

where EiE_{i} is a stable LL-twisted Higgs bundle such that (rank(Ei),χ(Ei))=(ri,χi)(\operatorname{rank}(E_{i}),\chi(E_{i}))=(r_{i},\chi_{i}) satisfies χi/ri=χ/r\chi_{i}/r_{i}=\chi/r and ViV_{i} is a finite dimensional vector space. By abuse of notation, we also denote by yL(r,χ)y\in\mathcal{M}^{L}(r,\chi) the closed point represented by (2.3). It is the unique closed point in the fiber of L(r,χ)clML(r,χ)\mathcal{M}^{L}(r,\chi)^{\rm{cl}}\to M^{L}(r,\chi) at yy. We also have the Cartesian square

(2.8)

where ()st(-)^{\rm{st}} is the open locus of stable points, the horizontal arrows are open immersions, and the left vertical arrow is a good moduli space morphism which is a \mathbb{C}^{\ast}-gerbe.

Lemma 2.1.

([PTa, Lemma 2.3]) If l>0l>0, then ML=ML(r,χ)M^{L}=M^{L}(r,\chi) is Gorenstein with trivial dualizing sheaf ωML=𝒪ML\omega_{M^{L}}=\mathcal{O}_{M^{L}}, and the right vertical arrow in (2.8) is generically a \mathbb{C}^{\ast}-gerbe.

A point bBLb\in B^{L} corresponds to a support 𝒞bS\mathcal{C}_{b}\subset S of the sheaf on SS, called the spectral curve. We denote by gspg^{\rm{sp}} the arithmetic genus of the spectral curve, which is given by

(2.9) gsp:=1+(g1)rrl2+r2l2.\displaystyle g^{\rm{sp}}:=1+(g-1)r-\frac{rl}{2}+\frac{r^{2}l}{2}.

Let 𝒞BL\mathcal{C}\to B^{L} be the universal spectral curve, which is a closed subscheme of S×BLS\times B^{L}. By the spectral construction, the universal Higgs bundle corresponds to a universal sheaf

(2.10) Coh(𝒞×BLL(r,χ))\displaystyle\mathcal{E}\in\operatorname{Coh}(\mathcal{C}\times_{B^{L}}\mathcal{M}^{L}(r,\chi))

which is also regarded as a coherent sheaf on S×L(r,χ)S\times\mathcal{M}^{L}(r,\chi) by the closed immersion 𝒞×BLL(r,χ)S×L(r,χ)\mathcal{C}\times_{B^{L}}\mathcal{M}^{L}(r,\chi)\hookrightarrow S\times\mathcal{M}^{L}(r,\chi).

2.3. The local description of moduli stacks of Higgs bundles

The good moduli space

(2.11) π:L(r,χ)clML(r,χ)\displaystyle\pi\colon\mathcal{M}^{L}(r,\chi)^{\rm{cl}}\to M^{L}(r,\chi)

is, locally near a point yML(r,χ)y\in M^{L}(r,\chi), described in terms of the representations of the Ext-quiver of yy. In what follows, we write

(2.12) (r,χ)=d(r0,χ0),\displaystyle(r,\chi)=d(r_{0},\chi_{0}),

where d>0d\in\mathbb{Z}_{>0} and (r0,χ0)(r_{0},\chi_{0}) are coprime.

Let yML(r,χ)y\in M^{L}(r,\chi) be a closed point corresponding to the polystable object (2.3). The associated Ext-quiver QyQ_{y} consists of vertices {1,,k}\{1,\ldots,k\} with the number of arrows given by

(ij)=dimExtS1(Ei,Ej)={rirjl+δij,l>2g2,rirj(2g2)+2δij,L=ΩC.\displaystyle\sharp(i\to j)=\dim\operatorname{Ext}_{S}^{1}(E_{i},E_{j})=\begin{cases}r_{i}r_{j}l+\delta_{ij},&l>2g-2,\\ r_{i}r_{j}(2g-2)+2\delta_{ij},&L=\Omega_{C}.\end{cases}

Here, by the spectral construction, we regard EiE_{i} as a coherent sheaf on SS. The representation space of QyQ_{y}-representations of dimension vector 𝒅=(di)i=1k\bm{d}=(d_{i})_{i=1}^{k} for di=dimVid_{i}=\dim V_{i} is given by

RQy(𝒅):=(ij)QyHom(Vi,Vj)=ExtS1(E,E).\displaystyle R_{Q_{y}}(\bm{d}):=\bigoplus_{(i\to j)\in Q_{y}}\operatorname{Hom}(V_{i},V_{j})=\operatorname{Ext}_{S}^{1}(E,E).

Let G(𝒅)G(\bm{d}) be the algebraic group

G(𝒅):=i=1kGL(Vi)=Aut(E)\displaystyle G(\bm{d}):=\prod_{i=1}^{k}GL(V_{i})=\mathrm{Aut}(E)

which acts on RQy(𝒅)R_{Q_{y}}(\bm{d}) by the conjugation.

If l>2g2l>2g-2, by the Luna étale slice theorem, étale locally at yy, the map

π:L(r,χ)ML(r,χ)\pi\colon\mathcal{M}^{L}(r,\chi)\to M^{L}(r,\chi)

is isomorphic to

(2.13) 𝒳Qy(𝒅):=RQy(𝒅)/G(𝒅)XQy(𝒅):=RQy(𝒅)//G(𝒅).\displaystyle\mathcal{X}_{Q_{y}}(\bm{d}):=R_{Q_{y}}(\bm{d})/G(\bm{d})\to X_{Q_{y}}(\bm{d}):=R_{Q_{y}}(\bm{d})/\!\!/G(\bm{d}).

Note that the above quotient stack is the moduli stack of QyQ_{y}-representations of dimension 𝒅\bm{d}.

If L=ΩCL=\Omega_{C}, we can write RQy(𝒅)=R(𝒅)R(𝒅)R_{Q_{y}}(\bm{d})=R(\bm{d})\oplus R(\bm{d})^{\vee} for some G(𝒅)G(\bm{d})-representation R(𝒅)R(\bm{d}). Let μ\mu be the moment map

μ:RQy(𝒅)𝔤(𝒅),\displaystyle\mu\colon R_{Q_{y}}(\bm{d})\to\mathfrak{g}(\bm{d})^{\vee},

where 𝔤(𝒅)\mathfrak{g}(\bm{d}) is the Lie algebra of G(𝒅)G(\bm{d}). Then π:(r,χ)clM(r,χ)\pi\colon\mathcal{M}(r,\chi)^{\rm{cl}}\to M(r,\chi) is étale locally at yy isomorphic to ([Sac19], [Dav, Section 5], [HLa, Section 4.2]):

(2.14) 𝒫(d)cl:=μ1(0)cl/G(𝒅)P(d):=μ1(0)cl//G(𝒅),\displaystyle\mathscr{P}(d)^{\rm{cl}}:=\mu^{-1}(0)^{\rm{cl}}/G(\bm{d})\to P(d):=\mu^{-1}(0)^{\rm{cl}}/\!\!/G(\bm{d}),

see [PTc, Section 4.3] for more details.

The moduli space ML(r,χ)M^{L}(r,\chi) is stratified with strata indexed by the data (di,ri,χi)i=1k(d_{i},r_{i},\chi_{i})_{i=1}^{k} of the polystable object (2.3). The deepest stratum corresponds to k=1k=1 with (d1,r1,χ1)=(d,r0,χ0)(d_{1},r_{1},\chi_{1})=(d,r_{0},\chi_{0}), which consist of polystable objects VE0V\otimes E_{0} where dimV=d\dim V=d and E0E_{0} is stable with

(rank(E0),χ(E0))=(r0,d0).\displaystyle(\operatorname{rank}(E_{0}),\chi(E_{0}))=(r_{0},d_{0}).

The associated Ext-quiver at the deepest stratum has one vertex and (1+lr02)(1+lr_{0}^{2})-loops. We have the following lemma, also see [PTc, Lemma 4.3] for an analogous statement for K3 surfaces.

Lemma 2.2.

([PTa, Lemma 2.2]) For each closed point yML(r,χ)y\in M^{L}(r,\chi), and a closed point xML(r,χ)x\in M^{L}(r,\chi) which lies in the deepest stratum, there exists a closed point yML(r,χ)y^{\prime}\in M^{L}(r,\chi) which is sufficiently close to xx such that Qy=QyQ_{y}=Q_{y^{\prime}}.

2.4. Quasi-BPS categories

We consider the bounded derived category of coherent sheaves Db(L(r,χ))D^{b}(\mathcal{M}^{L}(r,\chi)). Note that there is an orthogonal decomposition

Db(L(r,χ))=wDb(L(r,χ))w,D^{b}(\mathcal{M}^{L}(r,\chi))=\bigoplus_{w\in\mathbb{Z}}D^{b}(\mathcal{M}^{L}(r,\chi))_{w},

where Db(L(r,χ))wD^{b}(\mathcal{M}^{L}(r,\chi))_{w} is the subcategory of Db(L(r,χ))D^{b}(\mathcal{M}^{L}(r,\chi)) of weight ww complexes with respect to the action of the scalar automorphisms \mathbb{C}^{*} at each point of L(r,χ)\mathcal{M}^{L}(r,\chi). In this subsection, we define the quasi-BPS categories, which are subcategories of Db(L(r,χ))wD^{b}(\mathcal{M}^{L}(r,\chi))_{w}.

We first construct a line bundle on L(r,χ)\mathcal{M}^{L}(r,\chi). We write (r,χ)=d(r0,χ0)(r,\chi)=d(r_{0},\chi_{0}) as in (2.12), and take (a,b)2(a,b)\in\mathbb{Z}^{2} such that

(2.15) aχ0+br0+(1g)ar0=1\displaystyle a\chi_{0}+br_{0}+(1-g)ar_{0}=1

which is possible as (r0,χ0)(r_{0},\chi_{0}) are coprime. Let uK(C)u\in K(C) be such that (rank(u),χ(u))=(a,b)(\operatorname{rank}(u),\chi(u))=(a,b). Define the following line bundle on L(r,χ)\mathcal{M}^{L}(r,\chi)

(2.16) δ:=det(Rp(u))Pic(L(r,χ)),\displaystyle\delta:=\det(Rp_{\mathcal{M}\ast}(u\boxtimes\mathcal{F}))\in\mathrm{Pic}(\mathcal{M}^{L}(r,\chi)),

where \mathcal{F} is the universal Higgs bundle (2.2) and pp_{\mathcal{M}} is the projection onto \mathcal{M}. It has diagonal \mathbb{C}^{\ast}-weight dd because of the condition (2.15) and the Riemann-Roch theorem.

Before defining quasi-BPS categories, we need to introduce some more notations. An object ADb(B)A\in D^{b}(B\mathbb{C}^{\ast}) decomposes into a direct sum

A=wAw,A=\bigoplus_{w\in\mathbb{Z}}A_{w},

where AwA_{w} is of \mathbb{C}^{\ast}-weight ww. Denote by wt(A)\mathrm{wt}(A) the set of ww\in\mathbb{Z} such that Aw0A_{w}\neq 0. In the case that AA is a line bundle on BB\mathbb{C}^{\ast}, then wt(A)\mathrm{wt}(A) consists of one element wt(A)\mathrm{wt}(A)\in\mathbb{Z}. We also write A>0:=w>0AwA^{>0}:=\bigoplus_{w>0}A_{w}.

Definition 2.3.

In the case of l>2g2l>2g-2, define the quasi-BPS category

(2.17) 𝕋L(r,χ)wDb(L(r,χ))w.\displaystyle\mathbb{T}^{L}(r,\chi)_{w}\subset D^{b}(\mathcal{M}^{L}(r,\chi))_{w}.

to be consisting of objects \mathcal{E} such that, for any map ν:B=L(r,χ)\nu\colon B\mathbb{C}^{\ast}\to\mathcal{M}=\mathcal{M}^{L}(r,\chi), we have

(2.18) wt(ν)[12wtdet((ν𝕃)>0),12wtdet((ν𝕃)>0)]+wdwt(νδ).\displaystyle\mathrm{wt}(\nu^{\ast}\mathcal{E})\subset\left[-\frac{1}{2}\mathrm{wt}\det((\nu^{\ast}\mathbb{L}_{\mathcal{M}})^{>0}),\frac{1}{2}\mathrm{wt}\det((\nu^{\ast}\mathbb{L}_{\mathcal{M}})^{>0})\right]+\frac{w}{d}\mathrm{wt}(\nu^{\ast}\delta).
Remark 2.4.

In the notation of (2.13), let

(2.19) 𝕋Qy(𝒳Qy(𝒅))wDb(𝒳Qy(𝒅))\displaystyle\mathbb{T}_{Q_{y}}(\mathcal{X}_{Q_{y}}(\bm{d}))_{w}\subset D^{b}(\mathcal{X}_{Q_{y}}(\bm{d}))

be generated by Γ(χ)𝒪𝒳Qy(𝒅)\Gamma(\chi)\otimes\mathcal{O}_{\mathcal{X}_{Q_{y}}(\bm{d})}, where Γ(χ)\Gamma(\chi) is the irreducible G(𝒅)G(\bm{d})-representation whose highest weight χ\chi is such that

χ+ρ12sum[0,β]+wdδy.\displaystyle\chi+\rho\in\frac{1}{2}\mathrm{sum}[0,\beta]+\frac{w}{d}\delta_{y}.

Here, M(𝒅)M(\bm{d}) is the weight lattice of the maximal torus T(𝒅)G(𝒅)T(\bm{d})\subset G(\bm{d}), sum[0,β]\mathrm{sum}[0,\beta] is the Minkowski sum of weights in RQy(𝒅)R_{Q_{y}}(\bm{d}), ρ\rho is the half the sum of positive roots, and δy:=δ|y\delta_{y}:=\delta|_{y}. The above category (2.19) is the quasi-BPS category of the quiver QyQ_{y} and gives an étale local model of 𝕋L(r,χ)w\mathbb{T}^{L}(r,\chi)_{w}. For more details, see [PTa, Lemma 3.5, Remark 3.7].

We next recall the definition of (reduced or not) quasi-BPS categories in the case of L=ΩCL=\Omega_{C}. Below we fix pCp\in C. There is a closed embedding

(2.20) j:(r,χ)ΩC(p)(r,χ)\displaystyle j\colon\mathcal{M}(r,\chi)\hookrightarrow\mathcal{M}^{\Omega_{C}(p)}(r,\chi)

sending (F,θ)(F,\theta) for θ:FFΩC\theta\colon F\to F\otimes\Omega_{C} to (F,θ)(F,\theta^{\prime}), where θ\theta^{\prime} is the composition

θ:FθFΩCFΩC(p).\displaystyle\theta^{\prime}\colon F\stackrel{{\scriptstyle\theta}}{{\to}}F\otimes\Omega_{C}\hookrightarrow F\otimes\Omega_{C}(p).

Given (F,θ)(F,\theta^{\prime}) in ΩC(p)(r,χ)\mathcal{M}^{\Omega_{C}(p)}(r,\chi), it comes from the image of (2.20) if and only if θ|p:F|pF|pΩC(p)|p\theta^{\prime}|_{p}\colon F|_{p}\to F|_{p}\otimes\Omega_{C}(p)|_{p} is zero. Globally, let (,ϑ)(\mathcal{F},\vartheta) be the universal Higgs bundle (2.2) for L=ΩC(p)L=\Omega_{C}(p), and set

p:=|p×ΩC(p)(r,χ)Coh(ΩC(p)(r,χ)).\displaystyle\mathcal{F}_{p}:=\mathcal{F}|_{p\times\mathcal{M}^{\Omega_{C}(p)}(r,\chi)}\in\mathrm{Coh}(\mathcal{M}^{\Omega_{C}(p)}(r,\chi)).

By fixing an isomorphism ΩC(p)|p\Omega_{C}(p)|_{p}\cong\mathbb{C}, the correspondence (F,θ)(F|p,θ|p)(F,\theta^{\prime})\mapsto(F|_{p},\theta^{\prime}|_{p}) gives a section ss of the vector bundle

(2.23)

Then we have an equivalence of derived stacks

(2.24) (r,χ)s1(0),\displaystyle\mathcal{M}(r,\chi)\stackrel{{\scriptstyle\sim}}{{\to}}s^{-1}(0),

where the right hand side is the derived zero locus of the section ss.

Definition 2.5.

Suppose that L=ΩCL=\Omega_{C}. We define the subcategory

(2.25) 𝕋(r,χ)wDb((r,χ))w\displaystyle\mathbb{T}(r,\chi)_{w}\subset D^{b}(\mathcal{M}(r,\chi))_{w}

to be consisting of objects Db((r,χ))\mathcal{E}\in D^{b}(\mathcal{M}(r,\chi)) such that, for all ν:B(r,χ)\nu\colon B\mathbb{C}^{\ast}\to\mathcal{M}(r,\chi), we have

(2.26) wt(νjj)[12wtdet(ν𝕃𝒱)>0,12wtdet(ν𝕃𝒱)>0]+wdwt(νδ).\displaystyle\mathrm{wt}(\nu^{\ast}j^{\ast}j_{\ast}\mathcal{E})\subset\left[-\frac{1}{2}\mathrm{wt}\det(\nu^{\ast}\mathbb{L}_{\mathcal{V}})^{>0},\frac{1}{2}\mathrm{wt}\det(\nu^{\ast}\mathbb{L}_{\mathcal{V}})^{>0}\right]+\frac{w}{d}\mathrm{wt}(\nu^{\ast}\delta).

Here, jj is the closed immersion (2.20).

The section ss is indeed a section s0s_{0} of the subbundle 𝒱0𝒱\mathcal{V}_{0}\subset\mathcal{V} consisting of traceless endomorphisms. The reduced stack is the derived zero locus of s0s_{0}

(2.27) (r,χ)red:=s01(0)(r,χ).\displaystyle\mathcal{M}(r,\chi)^{\rm{red}}:=s_{0}^{-1}(0)\subset\mathcal{M}(r,\chi).

Note that the classical truncations of (r,χ)red\mathcal{M}(r,\chi)^{\rm{red}} and (r,χ)\mathcal{M}(r,\chi) are the same. The reduced quasi-BPS category

(2.28) 𝕋(r,χ)wredDb((r,χ)red)\displaystyle\mathbb{T}(r,\chi)_{w}^{\rm{red}}\subset D^{b}(\mathcal{M}(r,\chi)^{\rm{red}})

is also similarly defined using the closed immersion (r,χ)redΩC(p)(r,χ)\mathcal{M}(r,\chi)^{\rm{red}}\hookrightarrow\mathcal{M}^{\Omega_{C}(p)}(r,\chi).

Remark 2.6.

The construction of the category (2.17) is local over ML=ML(r,χ)M^{L}=M^{L}(r,\chi). It follows that, for any open subset UMLU\subset M^{L}, there is an associated subcategory

𝕋L(r,χ)w|UDb(L(r,χ)×MLU),\displaystyle\mathbb{T}^{L}(r,\chi)_{w}|_{U}\subset D^{b}(\mathcal{M}^{L}(r,\chi)\times_{M^{L}}U),

see [PTa, Remark 3.11]. The same remark also applies to the reduced quasi-BPS categories.

We say that the tuple (r,χ,w)(r,\chi,w) satisfies the BPS condition if the tuple

(r,χ,w+1gsp)3\displaystyle(r,\chi,w+1-g^{\rm{sp}})\in\mathbb{Z}^{3}

is primitive, i.e. if gcd(r,χ,w+1gsp)=1\gcd(r,\chi,w+1-g^{\rm{sp}})=1. If (r,χ,w)(r,\chi,w) satisfies the BPS condition, we say that the category (2.25), (2.28) is a (reduced or not) BPS category. We now recall the basic properties of BPS categories.

Theorem 2.7.

([PTa, Theorem 1.2]) Suppose that the tuple (r,χ,w)(r,\chi,w) satisfies the BPS condition.

(1) If l>2g2l>2g-2, then 𝕋L(r,χ)w\mathbb{T}^{L}(r,\chi)_{w} is a smooth dg-category over \mathbb{C}, which is proper and Calabi-Yau over BLB^{L}.

(2) If L=ΩCL=\Omega_{C}, then 𝕋(r,χ)wred\mathbb{T}(r,\chi)^{\rm{red}}_{w} is a smooth dg-category over \mathbb{C}, which is proper and Calabi-Yau over BB.

2.5. The conjectural symmetry

We now recall the main conjecture about the symmetry of BPS categories for G=GL(r)G=\mathrm{GL}(r) proposed in [PTa].

We denote by (BL)smBL(B^{L})^{\rm{sm}}\subset B^{L} the open subset corresponding to smooth and irreducible spectral curves, and let sm\mathcal{E}^{\rm{sm}} be the restriction of \mathcal{E} to (BL)sm(B^{L})^{\rm{sm}}. We denote by

L(r,χ)sm:=L(r,χ)×BL(BL)sm,ML(r,χ)sm:=ML(r,χ)×BL(BL)sm.\displaystyle\mathcal{M}^{L}(r,\chi)^{\rm{sm}}:=\mathcal{M}^{L}(r,\chi)\times_{B^{L}}(B^{L})^{\rm{sm}},\ M^{L}(r,\chi)^{\rm{sm}}:=M^{L}(r,\chi)\times_{B^{L}}(B^{L})^{\rm{sm}}.

Note that L(r,χ)sm\mathcal{M}^{L}(r,\chi)^{\rm{sm}} is the relative Picard stack of the family of smooth spectral curves 𝒞sm(BL)sm\mathcal{C}^{\mathrm{sm}}\to(B^{L})^{\rm{sm}}. Let 𝒫sm\mathcal{P}^{\rm{sm}} be the Poincaré line bundle

𝒫smL(r,w+1gsp)sm×(BL)smL(r,χ)sm\displaystyle\mathcal{P}^{\rm{sm}}\to\mathcal{M}^{L}(r,w+1-g^{\rm{sp}})^{\rm{sm}}\times_{(B^{L})^{\rm{sm}}}\mathcal{M}^{L}(r,\chi)^{\rm{sm}}

defined by

(2.29) 𝒫sm:=detRp13(p12p23)\displaystyle\mathcal{P}^{\rm{sm}}:=\det Rp_{13\ast}(p_{12}^{\ast}\mathcal{E}^{\prime}\otimes p_{23}^{\ast}\mathcal{E}) detRp13(p12)1\displaystyle\otimes\det Rp_{13\ast}(p_{12}^{\ast}\mathcal{E}^{\prime})^{-1}
detRp13(p23)1detRp13𝒪,\displaystyle\otimes\det Rp_{13\ast}(p_{23}^{\ast}\mathcal{E})^{-1}\otimes\det Rp_{13\ast}\mathcal{O},

where pijp_{ij} are the projections from

L(r,w+1gsp)sm×(BL)sm×𝒞sm×(BL)sm×L(r,χ)sm\displaystyle\mathcal{M}^{L}(r,w+1-g^{\rm{sp}})^{\rm{sm}}\times_{(B^{L})^{\rm{sm}}}\times\mathcal{C}^{\rm{sm}}\times_{(B^{L})^{\rm{sm}}}\times\mathcal{M}^{L}(r,\chi)^{\rm{sm}}

onto the corresponding factors, and \mathcal{E}^{\prime} is the universal sheaf on the product 𝒞×BLL(r,w+1gsp)\mathcal{C}\times_{B^{L}}\mathcal{M}^{L}(r,w+1-g^{\rm{sp}}). It determines the Fourier-Mukai equivalence [Muk81, DP12]:

(2.30) Φ𝒫sm:Db(L(r,w+1gsp)sm)χ+1gspDb(L(r,χ)sm)w.\displaystyle\Phi_{\mathcal{P}^{\rm{sm}}}\colon D^{b}(\mathcal{M}^{L}(r,w+1-g^{\rm{sp}})^{\rm{sm}})_{-\chi+1-g^{\rm{sp}}}\stackrel{{\scriptstyle\sim}}{{\to}}D^{b}(\mathcal{M}^{L}(r,\chi)^{\rm{sm}})_{w}.

For l>2g2l>2g-2, we rewrite the equivalence above as the following equivalence:

(2.31) 𝕋L(r,w+1gsp)χ+1gsp|(BL)sm𝕋L(r,χ)w|(BL)sm.\displaystyle\mathbb{T}^{L}(r,w+1-g^{\rm{sp}})_{-\chi+1-g^{\rm{sp}}}|_{(B^{L})^{\rm{sm}}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathbb{T}^{L}(r,\chi)_{w}|_{(B^{L})^{\rm{sm}}}.

The following is the main conjecture in [PTa], which says that the above equivalence extends over the full Hitchin base BLB^{L}:

Conjecture 2.8.

([PTa, Conjecture 4.3]) Suppose that l>2g2l>2g-2 and that the tuple (r,χ,w)(r,\chi,w) satisfies the BPS condition. Then there is a BLB^{L}-linear equivalence

(2.32) 𝕋L(r,w+1gsp)χ+1gsp𝕋L(r,χ)w\displaystyle\mathbb{T}^{L}(r,w+1-g^{\rm{sp}})_{-\chi+1-g^{\rm{sp}}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathbb{T}^{L}(r,\chi)_{w}

which extends the equivalence (2.31), i.e. it commutes with the restriction functors to (BL)sm(B^{L})^{\rm{sm}}.

In the case of L=ΩCL=\Omega_{C}, we propose the following conjecture:

Conjecture 2.9.

([PTa, Conjecture 4.6]) Suppose that the tuple (r,χ,w)(r,\chi,w) is primitive. Then there is a BB-linear equivalence

(2.33) 𝕋(r,w)χred𝕋(r,χ)wred\displaystyle\mathbb{T}(r,w)_{-\chi}^{\rm{red}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathbb{T}(r,\chi)_{w}^{\rm{red}}

which extends the equivalence (2.31).

Note that, when ww and χ\chi are both coprime with rr and G=GL(r)G=\mathrm{GL}(r), the equivalence (2.33) recovers the (still conjectural) equivalence (1.1).

2.6. Semiorthogonal decompositions for moduli of Higgs bundles

We write (r,χ)=d(r0,χ0)(r,\chi)=d(r_{0},\chi_{0}) for d>0d\in\mathbb{Z}_{>0} and (r0,χ0)(r_{0},\chi_{0}) coprime. For a decomposition d=d1++dkd=d_{1}+\cdots+d_{k}, let ilL(d1,,dk)\mathcal{F}il^{L}(d_{1},\ldots,d_{k}) be the (derived) moduli stack of filtrations of LL-twisted Higgs bundles

(2.34) 0=E0E1Ek,\displaystyle 0=E_{0}\subset E_{1}\subset\cdots\subset E_{k},

where Ei/Ei1E_{i}/E_{i-1} has (rank(Ei/Ei1),χ(Ei/Ei1))=di(r0,χ0)(\operatorname{rank}(E_{i}/E_{i-1}),\chi(E_{i}/E_{i-1}))=d_{i}(r_{0},\chi_{0}). There are natural evaluation morphisms

×i=1kL(di(r0,χ0))qilL(d1,,dk)pL(r,χ),\displaystyle\times_{i=1}^{k}\mathcal{M}^{L}(d_{i}(r_{0},\chi_{0}))\stackrel{{\scriptstyle q}}{{\leftarrow}}\mathcal{F}il^{L}(d_{1},\ldots,d_{k})\stackrel{{\scriptstyle p}}{{\to}}\mathcal{M}^{L}(r,\chi),

where pp sends a filtration (2.34) to EkE_{k} and qq sends (2.34) to (Ei/Ei1)i=1k(E_{i}/E_{i-1})_{i=1}^{k}. The categorical Hall product is defined by (see [PS23]):

(2.35) :=RpLq:i=1kDb(L(di(r0,χ0)))Db(L(r,χ)).\displaystyle\ast:=Rp_{\ast}Lq^{\ast}\colon\boxtimes_{i=1}^{k}D^{b}(\mathcal{M}^{L}(d_{i}(r_{0},\chi_{0})))\to D^{b}(\mathcal{M}^{L}(r,\chi)).

We have the following semiorthogonal decomposition:

Theorem 2.10.

([PTa, Theorem 3.12]) For each ww\in\mathbb{Z}, there is a semiorthogonal decomposition

(2.36) Db(L(r,χ))w=i=1k𝕋L(dir0,diχ0)wi|v1d1<<vkdk,\displaystyle D^{b}(\mathcal{M}^{L}(r,\chi))_{w}=\left\langle\boxtimes_{i=1}^{k}\mathbb{T}^{L}(d_{i}r_{0},d_{i}\chi_{0})_{w_{i}}\,\Big{|}\,\frac{v_{1}}{d_{1}}<\cdots<\frac{v_{k}}{d_{k}}\right\rangle,

where the right hand side is after all partitions (d,w)=(d1,w1)++(dk,wk)(d,w)=(d_{1},w_{1})+\cdots+(d_{k},w_{k}) and where vi12v_{i}\in\frac{1}{2}\mathbb{Z} is given by

(2.37) vi=wilr022di(i>jdji<jdj).\displaystyle v_{i}=w_{i}-\frac{lr_{0}^{2}}{2}d_{i}\left(\sum_{i>j}d_{j}-\sum_{i<j}d_{j}\right).

The fully-faithful functor

i=1k𝕋L(dir0,diχ0)wiDb(L(r,χ))\displaystyle\boxtimes_{i=1}^{k}\mathbb{T}^{L}(d_{i}r_{0},d_{i}\chi_{0})_{w_{i}}\to D^{b}(\mathcal{M}^{L}(r,\chi))

is given by the restriction of the categorical Hall product (2.35).

2.7. Moduli spaces of Joyce-Song pairs

In this subsection, we mention a framed version of Theorem 2.10. We first discuss a version of Joyce-Song (JS) stable pairs [JS12] in the context of Higgs bundles. Fix an ample line bundle 𝒪C(1)\mathcal{O}_{C}(1) on CC of degree one and let m0m\gg 0.

Definition 2.11.

A tuple (F,θ,s)(F,\theta,s) is called a JS pair if (F,θ)(F,\theta) is a semistable LL-twisted Higgs bundle and sH0(F(m))s\in H^{0}(F(m)) is such that, for any surjection j:(F,θ)(F,θ)j\colon(F,\theta)\twoheadrightarrow(F^{\prime},\theta^{\prime}) of LL-twisted Higgs bundle with μ(F)=μ(F)\mu(F)=\mu(F^{\prime}), we have 0jsH0(F(m))0\neq j\circ s\in H^{0}(F^{\prime}(m)).

We denote by

(2.38) L=L(r,χ)JS\displaystyle\mathcal{M}^{L{\dagger}}=\mathcal{M}^{L}(r,\chi)^{\rm{JS}}

the moduli space of JS pairs (F,θ,s)(F,\theta,s) such that (rank(F),χ(F))=(r,χ)(\mathrm{rank}(F),\chi(F))=(r,\chi). It is a quasi-projective scheme with morphisms:

(2.39) ϕ:LLML,\displaystyle\phi\colon\mathcal{M}^{L{\dagger}}\to\mathcal{M}^{L}\to M^{L},

where L=L(r,χ)\mathcal{M}^{L}=\mathcal{M}^{L}(r,\chi) and ML=ML(r,χ)M^{L}=M^{L}(r,\chi), see [JS12, Section 12.1]. The first morphism is given by forgetting ss which is a smooth morphism, the second one is a good moduli space map, and the composition ϕ\phi is a projective morphism. In particular, L\mathcal{M}^{L{\dagger}} is smooth for l>2g2l>2g-2.

Let (r,χ)=d(r0,χ0)(r,\chi)=d(r_{0},\chi_{0}) with coprime (r0,χ0)(r_{0},\chi_{0}), and set

(2.40) N:=χ(E0(m))=mr0+χ0+r0(1g)\displaystyle N:=\chi(E_{0}(m))=mr_{0}+\chi_{0}+r_{0}(1-g)

where E0E_{0} is a vector bundle with (rank(E0),χ(E0))=(r0,χ0)(\mathrm{rank}(E_{0}),\chi(E_{0}))=(r_{0},\chi_{0}). Let yMLy\in M^{L} be the point corresponding to a polystable object (2.3), and let QyQ_{y} be its associated Ext-quiver. The morphism (2.39) is étale locally on MM at yy described as follows. Let QyQ_{y}^{{\dagger}} be the quiver framed quiver of the Ext-quiver QyQ_{y}, with vertices {0,1,,k}\{0,1,\ldots,k\} and with

(0i)=dimh0(Ei(m))=Nmi,\displaystyle\sharp(0\to i)=\dim h^{0}(E_{i}(m))=Nm_{i},

edges between the added vertex {0}\{0\} and a vertex ii of QyQ_{y}, and where mim_{i} is such that (rank(Ei),χ(Ei))=mi(r0,d0)(\mathrm{rank}(E_{i}),\chi(E_{i}))=m_{i}(r_{0},d_{0}). Let di=dimVid_{i}=\dim V_{i} and 𝒅=(di)i=1k\bm{d}=(d_{i})_{i=1}^{k}. The space of QyQ_{y}^{{\dagger}}-representations of dimension vector (1,𝒅)(1,\bm{d}) is given by

RQy(𝒅)=i=1kViNmiRQy(𝒅).\displaystyle R_{Q_{y}^{{\dagger}}}(\bm{d})=\bigoplus_{i=1}^{k}V_{i}^{\oplus Nm_{i}}\oplus R_{Q_{y}}(\bm{d}).

Then the morphism (2.39) is étale locally at yy isomorphic to

𝒳Qy(𝒅):=RQy(𝒅)ss/G(𝒅)RQy(𝒅)/G(𝒅)RQy(𝒅)//G(𝒅).\displaystyle\mathcal{X}_{Q_{y}}^{{\dagger}}(\bm{d}):=R_{Q_{y}^{{\dagger}}}(\bm{d})^{\rm{ss}}/G(\bm{d})\to R_{Q_{y}}(\bm{d})/G(\bm{d})\to R_{Q_{y}}(\bm{d})/\!\!/G(\bm{d}).

Here, the semistable locus is with respect to the character G(𝒅)G(\bm{d})\to\mathbb{C}^{\ast} given by (gi)i=1ki=1kdetgi(g_{i})_{i=1}^{k}\mapsto\prod_{i=1}^{k}\det g_{i}. The semistable locus consists of QyQ_{y}^{{\dagger}}-representations which are generated by the images of the edges from the vertex {0}\{0\}, see [Tod24, Lemma 6.1.9].

The following is a framed version of Theorem 2.10.

Theorem 2.12.

Suppose that l>2g2l>2g-2. There is a semiorthogonal decomposition

(2.41) Db(L(r,χ)JS)=i=1k𝕋L(dir0,diχ0)wi| 0v1d1<<vkdk<N,\displaystyle D^{b}(\mathcal{M}^{L}(r,\chi)^{\rm{JS}})=\left\langle\boxtimes_{i=1}^{k}\mathbb{T}^{L}(d_{i}r_{0},d_{i}\chi_{0})_{w_{i}}\,\Big{|}\,0\leqslant\frac{v_{1}}{d_{1}}<\cdots<\frac{v_{k}}{d_{k}}<N\right\rangle,

where the right hand side is after all partitions d=d1++dkd=d_{1}+\cdots+d_{k} and all (wi)i=1kk(w_{i})_{i=1}^{k}\in\mathbb{Z}^{k}, and where vi12v_{i}\in\frac{1}{2}\mathbb{Z} is given by

(2.42) vi=wilr022di(i>jdji<jdj).\displaystyle v_{i}=w_{i}-\frac{lr_{0}^{2}}{2}d_{i}\left(\sum_{i>j}d_{j}-\sum_{i<j}d_{j}\right).
Proof.

Let yMLy\in M^{L} lie in the deepest stratum, and let QyQ_{y}^{{\dagger}} be the framed quiver as above. There is a semiorthogonal decomposition, see [PTd, Theorem 4.18]:

Db(𝒳Qy(d))=i=1k𝕋Qy(di)wi| 0v1d1<<vkdk<N,\displaystyle D^{b}(\mathcal{X}_{Q_{y}^{{\dagger}}}(d))=\left\langle\boxtimes_{i=1}^{k}\mathbb{T}_{Q_{y}}(d_{i})_{w_{i}}\,\Big{|}\,0\leqslant\frac{v_{1}}{d_{1}}<\cdots<\frac{v_{k}}{d_{k}}<N\right\rangle,

where viv_{i} is given by (2.42). Similarly to the proof of Theorem 2.10 in [PTa, Theorem 3.12] (also see [PTc, Theorem 5.1]), we can reduce the proof of the semiorthogonal decomposition (2.41) to the above local statement at yy. ∎

3. Topological K-theory

In this section, we review the topological K-theory of dg-categories due to Blanc [Bla16] and its relative version due to Moulinos [Mou19]. We also prove some technical lemmas which will be used later.

3.1. Topological K-theory

We use the notion of spectrum, which is an object representing a generalized cohomology theory and plays a central role in stable homotopy theory. Basic references are [Ada74, May99]. A spectrum consists of a sequence of pointed spaces E={En}nE=\{E_{n}\}_{n\in\mathbb{N}} together with maps ΣEnEn+1\Sigma E_{n}\to E_{n+1} where Σ\Sigma is the suspension functor. There is a notion of homotopy groups of a spectrum denoted by π(E)\pi_{\bullet}(E). The \infty-category of spectra is constructed in [Lur, Section 1.4] and denoted by Sp\mathrm{Sp}. A map of spectra EFE\to F in Sp\mathrm{Sp} is an equivalence if it induces isomorphisms on homotopy groups.

For a topological space XX, we denote by Kitop(X)K^{\rm{top}}_{i}(X) its ii-th topological K-group. The K-groups Ktop(X)K^{\rm{top}}_{\bullet}(X) are obtained as homotopy groups of a spectrum: there is a spectrum KU={BU×,U,}KU=\{BU\times\mathbb{Z},U,\ldots\}, called the topological K-theory spectrum, such that we have (see [May99, Sections 22-24])

Ktop(X)=πKUX,KUX:=[ΣX,KU].\displaystyle K_{\bullet}^{\rm{top}}(X)=\pi_{\bullet}KU_{X},\ KU_{X}:=[\Sigma^{\infty}X,KU].

Here [,KU][-,KU] is the mapping spectrum with target KUKU and Σ()\Sigma^{\infty}(-) is the infinite suspension functor. We call Ktop(X)K^{\rm{top}}(X) the topological K-theory spectrum of XX.

For a dg-category 𝒟\mathscr{D}, Blanc [Bla16] defined the topological K-theory spectrum of 𝒟\mathscr{D}, denoted by

(3.1) Ktop(𝒟)Sp.\displaystyle K^{\mathrm{top}}(\mathscr{D})\in\mathrm{Sp}.

If 𝒟=Perf(X)\mathcal{D}=\mathrm{Perf}(X) for a separated \mathbb{C}-scheme XX of finite type, then Blanc’s topological K-theory spectrum is Ktop(𝒟)KUXK^{\rm{top}}(\mathcal{D})\simeq KU_{X}. For ii\in\mathbb{Z}, consider its ii-th rational homotopy group, which is a \mathbb{Q}-vector space:

Kitop(𝒟):=Kitop(𝒟),Kitop(𝒟):=πi(Ktop(𝒟)).\displaystyle K^{\mathrm{top}}_{i}(\mathscr{D})_{\mathbb{Q}}:=K^{\mathrm{top}}_{i}(\mathscr{D})\otimes_{\mathbb{Z}}\mathbb{Q},\ K^{\mathrm{top}}_{i}(\mathscr{D}):=\pi_{i}(K^{\mathrm{top}}(\mathscr{D})).

There are isomorphisms Kitop(𝒟)Ki+2top(𝒟)K^{\mathrm{top}}_{i}(\mathscr{D})\cong K^{\mathrm{top}}_{i+2}(\mathscr{D}) for every ii\in\mathbb{Z} obtained by multiplication with a Bott element, see [Bla16, Definition 1.6]. The topological K-theory spectrum sends exact triangles of dg-categories to exact triangles of spectra [Bla16, Theorem 1.1(c)].

3.2. Relative topological K-theory

In this subsection, we review the relative version of topological K-theory developed in [Mou19]. Below we use the notation in [GS, Section 2] for sheaves of spectra.

For a complex variety MM and an \infty-category with arbitrary small limits, we denote by Sh𝒞(Man)\mathrm{Sh}_{\mathcal{C}}(M^{\rm{an}}) the 𝒞\mathcal{C}-valued hypersheaves on ManM^{\rm{an}} as in [Lur, Section 1.3.1], where ManM^{\rm{an}} is the underlying complex analytic space. There is a rationalization functor, see [GS, Definition 2.6]:

Rat:ShSp(Man)ShD()(Man)=D(Sh(Man))\displaystyle\mathrm{Rat}\colon\mathrm{Sh}_{\mathrm{Sp}}(M^{\rm{an}})\to\mathrm{Sh}_{D(\mathbb{Q})}(M^{\rm{an}})=D(\mathrm{Sh}_{\mathbb{Q}}(M^{\rm{an}}))

given by H\mathcal{F}\mapsto\mathcal{F}\wedge H\mathbb{Q}. In the above, HH\mathbb{Q} is the Eilenberg-Maclane spectrum of \mathbb{Q} and the right hand side is the derived category of sheaves of \mathbb{Q}-vector spaces on ManM^{\rm{an}}. We write :=Rat(F)\mathcal{F}_{\mathbb{Q}}:=\mathrm{Rat}(F)_{\mathbb{Q}}. It satisfies π()=π()\pi_{\bullet}(\mathcal{F}_{\mathbb{Q}})=\pi_{\bullet}(\mathcal{F})\otimes\mathbb{Q}.

By [GS, Lemma 2.7], the sheaf of topological K-theory spectra KU¯M\underline{KU}_{M} on ManM^{\rm{an}} satisfies

(3.2) (KU¯M)M[β±1]=nM[2n],\displaystyle(\underline{KU}_{M})_{\mathbb{Q}}\cong\mathbb{Q}_{M}[\beta^{\pm 1}]=\bigoplus_{n\in\mathbb{Z}}\mathbb{Q}_{M}[2n],

where β\beta is of degree 22. The above isomorphism follows from the degeneration of the Atiyah–Hirzebruch spectral sequence.

Let Catperf(M)\mathrm{Cat}^{\rm{perf}}(M) be the \infty-category of Perf(M)\mathrm{Perf}(M)-linear stable \infty-categories. By [Mou19], there is a functor

𝒦Mtop:Catperf(M)ShSp(Man)\displaystyle\mathcal{K}_{M}^{\rm{top}}\colon\mathrm{Cat}^{\rm{perf}}(M)\to\mathrm{Sh}_{\mathrm{Sp}}(M^{\rm{an}})

such that, for a Brauer class α\alpha on MM, we have

(3.3) 𝒦Mtop(Perf(M,α))=KU¯Mα^.\displaystyle\mathcal{K}_{M}^{\rm{top}}(\mathrm{Perf}(M,\alpha))=\underline{KU}^{\hat{\alpha}}_{M}.

In the above, Perf(M,α)\mathrm{Perf}(M,\alpha) is the dg-category of α\alpha-twisted perfect complexes, α^H3(M,)\hat{\alpha}\in H^{3}(M,\mathbb{Z}) is the class associated with α\alpha by the natural map Br(M)H3(M,)\mathrm{Br}(M)\to H^{3}(M,\mathbb{Z}), and the right hand side is the sheaf of α^\hat{\alpha}-twisted topological K-theory spectrum. For M=SpecM=\operatorname{Spec}\mathbb{C}, it agrees with the spectrum (3.1): 𝒦Spectop(𝒟)=Ktop(𝒟)\mathcal{K}_{\mathrm{Spec}\mathbb{C}}^{\rm{top}}(\mathscr{D})=K^{\rm{top}}(\mathscr{D}).

3.3. Topological G-theory

For a complex variety MM, there is an embedding Perf(M)Db(M)\mathrm{Perf}(M)\subset D^{b}(M) which is not equivalence unless MM is smooth. A version of topological K-theory for Db(M)D^{b}(M) is called topological G-theory. It was introduced by Thomason in [Tho88], and we denote it by Gtop(M)G^{\rm{top}}_{\bullet}(M).

There is also its spectrum version KUM,cKU_{M,c}^{\vee}, called locally compact supported K-homology, see [HLP20, Lemma 2.6]. In [HLP20, Theorem 2.10], it is proved that there is an equivalence

(3.4) Ktop(Db(M))KUM,c.\displaystyle K^{\rm{top}}(D^{b}(M))\simeq KU_{M,c}^{\vee}.

We denote by KU¯M,c\underline{KU}_{M,c}^{\vee} the sheaf of locally compact supported K-homology. If MM is smooth, then we have KU¯M,cKU¯M\underline{KU}_{M,c}^{\vee}\cong\underline{KU}_{M}. The following is the sheaf version of the equivalence (3.4):

Lemma 3.1.

([PTe, Lemma 4.1]) For a quasi-projective scheme MM, there is a natural equivalence

(3.5) 𝒦Mtop(Db(M))KU¯M,c.\displaystyle\mathcal{K}_{M}^{\rm{top}}(D^{b}(M))\stackrel{{\scriptstyle\sim}}{{\to}}\underline{KU}_{M,c}^{\vee}.

In particular, there is an equivalence 𝒦Mtop(Db(M))ωM[β±1]\mathcal{K}_{M}^{\rm{top}}(D^{b}(M))_{\mathbb{Q}}\simeq\omega_{M}[\beta^{\pm 1}], where ωM=𝔻M\omega_{M}=\mathbb{D}\mathbb{Q}_{M} is the dualizing complex.

3.4. Push-forward of relative topological K-theories

We will use the following property of topological K-theory under proper push-forward.

Theorem 3.2.

([Mou19, Proposition 7.8], [GS, Theorem 2.12]) Let ϕ:MM\phi\colon M\to M^{\prime} be a proper morphism. Then for 𝒟Catperf(M)\mathscr{D}\in\mathrm{Cat}^{\rm{perf}}(M), we have

𝒦Mtop(ϕ𝒟)ϕ𝒦Mtop(𝒟).\displaystyle\mathcal{K}_{M^{\prime}}^{\rm{top}}(\phi_{\ast}\mathscr{D})\cong\phi_{\ast}\mathcal{K}_{M}^{\rm{top}}(\mathscr{D}).

Here, ϕ𝒟\phi_{\ast}\mathscr{D} is the category 𝒟\mathscr{D} with Perf(M)\mathrm{Perf}(M^{\prime})-linear structure induced by the pullback ϕ:Perf(M)Perf(M)\phi^{\ast}\colon\mathrm{Perf}(M^{\prime})\to\mathrm{Perf}(M).

Below we often write 𝒦Mtop(ϕ𝒟)\mathcal{K}_{M^{\prime}}^{\rm{top}}(\phi_{\ast}\mathscr{D}) as 𝒦Mtop(𝒟)\mathcal{K}_{M^{\prime}}^{\rm{top}}(\mathscr{D}) when ϕ\phi is clear from the context. The following is a version of Theorem 3.2 for open immersions.

Lemma 3.3.

([PTe, Lemma 4.3]) Let j:UMj\colon U\subset M be an open immersion. Then there is a natural equivalence

(3.6) 𝒦Mtop(jPerf(U))j𝒦Utop(Perf(U)).\displaystyle\mathcal{K}_{M}^{\rm{top}}(j_{\ast}\mathrm{Perf}(U))\stackrel{{\scriptstyle\sim}}{{\to}}j_{\ast}\mathcal{K}_{U}^{\rm{top}}(\mathrm{Perf}(U)).

3.5. Relative topological K-theories of semiorthogonal summands

We also need a version of Theorem 3.2 for global sections over non-proper schemes. We prove it for semiorthogonal summands of Perf(M)\mathrm{Perf}(M), Db(M)D^{b}(M), or of categories of matrix factorizations. We first state the following lemma:

Lemma 3.4.

([PTe, Lemma 4.4]) Let BB be a quasi-projective scheme. Then, for ϕ:BSpec\phi\colon B\to\operatorname{Spec}\mathbb{C} and 𝒟Cartperf(B)\mathscr{D}\in\mathrm{Cart}^{\rm{perf}}(B), there is a natural morphism

(3.7) η:Ktop(𝒟)ϕ𝒦Btop(𝒟).\displaystyle\eta\colon K^{\rm{top}}(\mathscr{D})\to\phi_{\ast}\mathcal{K}_{B}^{\rm{top}}(\mathscr{D}).

We will use the following results:

Lemma 3.5.

([PTe, Lemma 4.5]) Let 𝒟=𝒞1,𝒞2\mathscr{D}=\langle\mathcal{C}_{1},\mathcal{C}_{2}\rangle be a Perf(M)\mathrm{Perf}(M)-linear semiorthogonal decomposition. Then we have

(3.8) 𝒦Mtop(𝒟)=𝒦Mtop(𝒞1)𝒦Mtop(𝒞2).\displaystyle\mathcal{K}_{M}^{\rm{top}}(\mathscr{D})=\mathcal{K}_{M}^{\rm{top}}(\mathcal{C}_{1})\oplus\mathcal{K}_{M}^{\rm{top}}(\mathcal{C}_{2}).
Proposition 3.6.

([PTe, Proposition 4.6]) Let h:MBh\colon M\to B be a proper morphism. Let 𝒟\mathscr{D} be either Perf(M)\mathrm{Perf}(M), Db(M)D^{b}(M) or MF(M,f)\mathrm{MF}(M,f) for a non-zero function ff on MM, where we assume that MM is smooth in the last case. Let 𝒟=𝒞1,𝒞2\mathscr{D}=\langle\mathcal{C}_{1},\mathcal{C}_{2}\rangle be a Perf(B)\mathrm{Perf}(B)-linear semiorthogonal decomposition. Then the natural maps

ηi:Ktop(𝒞i)ϕ𝒦Btop(𝒞i)\displaystyle\eta_{i}\colon K^{\rm{top}}(\mathcal{C}_{i})\to\phi_{\ast}\mathcal{K}_{B}^{\rm{top}}(\mathcal{C}_{i})

in Lemma 3.4 are equivalences.

3.6. Pull-back of topological K-theory

Let π:M\pi\colon\mathcal{M}^{{\dagger}}\to M be a proper morphism of quasi-projective schemes, and let g:MΔg\colon M\to\Delta be a flat morphism such that the composition g:MΔg^{{\dagger}}\colon\mathcal{M}^{{\dagger}}\to M\to\Delta is also flat. For a closed point 0Δ0\in\Delta, let

M0:=g1(0),0:=(g)1(0),M_{0}:=g^{-1}(0),\mathcal{M}_{0}^{{\dagger}}:=(g^{{\dagger}})^{-1}(0),

so that we have the Cartesian square

Let Db()=𝒞,𝒞D^{b}(\mathcal{M}^{{\dagger}})=\langle\mathcal{C}^{\prime},\mathcal{C}\rangle be a Perf(M)\mathrm{Perf}(M)-linear semiorthogonal decomposition, which is strong in the sense that 𝒞\mathcal{C} is both left and right admissible in Db()D^{b}(\mathcal{M}^{{\dagger}}). Then there is an induced semiorthogonal decomposition, see [Kuz11, Theorem 5.6]:

(3.9) Db(0)=𝒞0,𝒞0.\displaystyle D^{b}(\mathcal{M}_{0}^{{\dagger}})=\langle\mathcal{C}_{0}^{\prime},\mathcal{C}_{0}\rangle.

In the next lemma, i1i^{-1} is the left adjoint to the functor i:D(Sh(M0))D(Sh(M))i_{*}\colon D(\mathrm{Sh}_{\mathbb{Q}}(M_{0}))\to D(\mathrm{Sh}_{\mathbb{Q}}(M)).

Lemma 3.7.

In the above setting, there is a natural isomorphism

i1𝒦Mtop(𝒞)𝒦M0top(𝒞0).\displaystyle i^{-1}\mathcal{K}_{M}^{\rm{top}}(\mathcal{C})_{\mathbb{Q}}\stackrel{{\scriptstyle\cong}}{{\to}}\mathcal{K}_{M_{0}}^{\rm{top}}(\mathcal{C}_{0})_{\mathbb{Q}}.
Proof.

The morphism ii^{{\dagger}} is quasi-smooth, so the pull-back (i)(i^{{\dagger}})^{\ast} gives a functor Db()Db(0)D^{b}(\mathcal{M}^{{\dagger}})\to D^{b}(\mathcal{M}_{0}^{{\dagger}}). The above functor restricts to a Perf(M)\mathrm{Perf}(M)-linear functor i:𝒞𝒞0i^{{\dagger}\ast}\colon\mathcal{C}\to\mathcal{C}_{0}, see [Kuz11, Theorem 5.6]. Therefore there is an induced morphism

𝒦Mtop(𝒞)𝒦Mtop(𝒞0)i𝒦M0top(𝒞0).\displaystyle\mathcal{K}_{M}^{\rm{top}}(\mathcal{C})_{\mathbb{Q}}\to\mathcal{K}_{M}^{\rm{top}}(\mathcal{C}_{0})_{\mathbb{Q}}\simeq i_{\ast}\mathcal{K}_{M_{0}}^{\rm{top}}(\mathcal{C}_{0})_{\mathbb{Q}}.

Here, the last equivalence follows from Theorem 3.2. By adjunction, we obtain a morphism

η𝒞:i1𝒦Mtop(𝒞)𝒦M0top(𝒞0).\displaystyle\eta_{\mathcal{C}}\colon i^{-1}\mathcal{K}_{M}^{\rm{top}}(\mathcal{C})_{\mathbb{Q}}\to\mathcal{K}_{M_{0}}^{\rm{top}}(\mathcal{C}_{0})_{\mathbb{Q}}.

The above construction applied for 𝒞=Db()\mathcal{C}=D^{b}(\mathcal{M}^{{\dagger}}) gives a morphism

(3.10) i1𝒦Mtop(Db())𝒦M0top(Db(0)).\displaystyle i^{-1}\mathcal{K}_{M}^{\rm{top}}(D^{b}(\mathcal{M}^{{\dagger}}))_{\mathbb{Q}}\to\mathcal{K}_{M_{0}}^{\rm{top}}(D^{b}(\mathcal{M}_{0}^{{\dagger}}))_{\mathbb{Q}}.

The above morphism is an equivalence. Indeed, we have

𝒦Mtop(Db())=πω[β±1],𝒦M0top(Db(0))=π0ω0[β±1]\displaystyle\mathcal{K}_{M}^{\rm{top}}(D^{b}(\mathcal{M}^{{\dagger}}))_{\mathbb{Q}}=\pi_{\ast}\omega_{\mathcal{M}^{{\dagger}}}[\beta^{\pm 1}],\ \mathcal{K}_{M_{0}}^{\rm{top}}(D^{b}(\mathcal{M}_{0}^{{\dagger}}))_{\mathbb{Q}}=\pi_{0\ast}\omega_{\mathcal{M}_{0}^{{\dagger}}}[\beta^{\pm 1}]

by Theorem 3.2 and Lemma 3.1, and then the equivalence of (3.10) follows from the proper base change theorem. From the Perf(M)\mathrm{Perf}(M)-linear semiorthogonal decompositions (3.9), the morphism (3.10) is identified with

η𝒞η𝒞:i1𝒦Mtop(𝒞)i1𝒦Mtop(𝒞)𝒦M0top(𝒞0)𝒦M0top(𝒞0).\displaystyle\eta_{\mathcal{C}}\oplus\eta_{\mathcal{C}^{\prime}}\colon i^{-1}\mathcal{K}_{M}^{\rm{top}}(\mathcal{C})_{\mathbb{Q}}\oplus i^{-1}\mathcal{K}_{M}^{\rm{top}}(\mathcal{C}^{\prime})_{\mathbb{Q}}\to\mathcal{K}_{M_{0}}^{\rm{top}}(\mathcal{C}_{0})_{\mathbb{Q}}\oplus\mathcal{K}_{M_{0}}^{\rm{top}}(\mathcal{C}_{0}^{\prime})_{\mathbb{Q}}.

Therefore η𝒞\eta_{\mathcal{C}} is an equivalence, as desired. ∎

4. Topological K-theory of BPS categories: the case G=GL and l>2g2l>2g-2

In this section, we prove part (1) of Theorem 1.2 after rationalization. The main ingredient is Proposition 4.1, where we compute the (relative) topological K-theory of BPS categories in terms of the intersection complex of the good moduli space. To show part (1) of Theorem 1.2, we use an extension or Arinkin’s sheaf over the full Hitchin base. We also discuss computations of topological K-theory of quasi-BPS categories beyond the BPS condition, and of Špenko–Van den Bergh noncommutative resolutions for the moduli of stable vector bundles on a curve.

4.1. The comparison with intersection complexes

Let L=L(r,χ)\mathcal{M}^{L}=\mathcal{M}^{L}(r,\chi) and ML=ML(r,χ)M^{L}=M^{L}(r,\chi) as in (2.39). Let

𝕋L=𝕋L(r,χ)wDb(L)\displaystyle\mathbb{T}^{L}=\mathbb{T}^{L}(r,\chi)_{w}\subset D^{b}(\mathcal{M}^{L})

be a quasi-BPS category. The above subcategory is closed under the action of Perf(ML)\mathrm{Perf}(M^{L}), and the semiorthogonal decomposition (2.36) is Perf(ML)\mathrm{Perf}(M^{L})-linear. In this subsection, we compute the topological K-theory of 𝕋L\mathbb{T}^{L} in terms of the BPS cohomology of MLM^{L}. Note that the BPS cohomology for MLM^{L} is the intersection cohomology IH(ML)\mathrm{IH}^{\star}(M^{L}) as the category of semistable LL-twisted Higgs bundles on CC has homological dimension one, see [Mei].

Proposition 4.1.

Suppose that l>2g2l>2g-2 and that the tuple (r,χ,w)(r,\chi,w) satisfies the BPS condition. There is an isomorphism in D(Sh(ML))D(\mathrm{Sh}_{\mathbb{Q}}(M^{L})):

𝒦MLtop(𝕋L)ICML[dimML][β±1].\displaystyle\mathcal{K}_{M^{L}}^{\rm{top}}(\mathbb{T}^{L})_{\mathbb{Q}}\cong\mathrm{IC}_{M^{L}}[-\dim M^{L}][\beta^{\pm 1}].
Proof.

Let ϕ\phi be the morphism (2.39). Since ϕ\phi is proper, there is the following commutative diagram, see [GS, Theorem 2.12]:

Then we have

ϕ𝒦Ltop(Db(L))𝒦MLtop(ϕDb(L)).\displaystyle\phi_{\ast}\mathcal{K}^{\rm{top}}_{\mathcal{M}^{L{\dagger}}}(D^{b}(\mathcal{M}^{L{\dagger}}))_{\mathbb{Q}}\cong\mathcal{K}_{M^{L}}^{\rm{top}}(\phi_{\ast}D^{b}(\mathcal{M}^{L{\dagger}}))_{\mathbb{Q}}.

By (3.2), (3.3), we have

𝒦Ltop(Db(L))(KU¯L)L[β±1].\displaystyle\mathcal{K}^{\rm{top}}_{\mathcal{M}^{L{\dagger}}}(D^{b}(\mathcal{M}^{L{\dagger}}))_{\mathbb{Q}}\cong(\underline{KU}_{\mathcal{M}^{L{\dagger}}})_{\mathbb{Q}}\cong\mathbb{Q}_{\mathcal{M}^{L{\dagger}}}[\beta^{\pm 1}].

It follows that we have

ϕL[β±1]𝒦MLtop(ϕDb(L)).\displaystyle\phi_{\ast}\mathbb{Q}_{\mathcal{M}^{L{\dagger}}}[\beta^{\pm 1}]\cong\mathcal{K}_{M^{L}}^{\rm{top}}(\phi_{\ast}D^{b}(\mathcal{M}^{L{\dagger}}))_{\mathbb{Q}}.

We may assume that 0w<d0\leqslant w<d. By Theorem 2.12, the subcategory 𝕋LDb(L)\mathbb{T}^{L}\subset D^{b}(\mathcal{M}^{L{\dagger}}) fits into a Perf(ML)\mathrm{Perf}(M^{L})-linear semiorthogonal decomposition. Therefore, by Lemma 3.5, 𝒦MLtop(𝕋L)\mathcal{K}_{M^{L}}^{\rm{top}}(\mathbb{T}^{L})_{\mathbb{Q}} is a direct summand of 𝒦MLtop(ϕDb(L))\mathcal{K}_{M^{L}}^{\rm{top}}(\phi_{\ast}D^{b}(\mathcal{M}^{L{\dagger}}))_{\mathbb{Q}}, hence a direct summand of ϕL[β±1]\phi_{\ast}\mathbb{Q}_{\mathcal{M}^{L{\dagger}}}[\beta^{\pm 1}]. By the BBDG decomposition theorem [BBD82], we have that:

(4.1) ϕL[dimL]pAp[p],\displaystyle\phi_{\ast}\mathbb{Q}_{\mathcal{M}^{L{\dagger}}}[\dim\mathcal{M}^{L{\dagger}}]\cong\bigoplus_{p\in\mathbb{Z}}A_{p}[-p],

where ApPerv(ML)A_{p}\in\mathrm{Perv}(M^{L}) is a semisimple perverse sheaf. The morphism ϕ\phi restricted to the stable locus (ML)stML(M^{L})^{\rm{st}}\subset M^{L} is a dN1\mathbb{P}^{dN-1}-bundle, and the decomposition (4.1) over (ML)st(M^{L})^{\rm{st}} is

ϕL|(ML)st(ML)st(ML)st[2](ML)st[2dN+2].\displaystyle\phi_{\ast}\mathbb{Q}_{\mathcal{M}^{L{\dagger}}}|_{(M^{L})^{\rm{st}}}\cong\mathbb{Q}_{(M^{L})^{\rm{st}}}\oplus\mathbb{Q}_{(M^{L})^{\rm{st}}}[-2]\oplus\cdots\oplus\mathbb{Q}_{(M^{L})^{\rm{st}}}[-2dN+2].

It follows that

(4.2) ϕL[β±1]|(ML)st(ML)st[β±1](ML)st[β±1],\displaystyle\phi_{\ast}\mathbb{Q}_{\mathcal{M}^{L{\dagger}}}[\beta^{\pm 1}]|_{(M^{L})^{\rm{st}}}\cong\mathbb{Q}_{(M^{L})^{\rm{st}}}[\beta^{\pm 1}]\oplus\cdots\oplus\mathbb{Q}_{(M^{L})^{\rm{st}}}[\beta^{\pm 1}],

where there are dNdN-direct sums in the right hand side. On the other hand, the semiorthogonal decomposition in Theorem 2.12 restricted to (ML)st(M^{L})^{\rm{st}} is

Db(ϕ1((ML)st))=𝕋L(r,χ)0|(ML)st,,𝕋L(r,χ)dN1|(ML)st.\displaystyle D^{b}\Big{(}\phi^{-1}((M^{L})^{\rm{st}})\Big{)}=\Big{\langle}\mathbb{T}^{L}(r,\chi)_{0}|_{(M^{L})^{\rm{st}}},\ldots,\mathbb{T}^{L}(r,\chi)_{dN-1}|_{(M^{L})^{\rm{st}}}\Big{\rangle}.

Therefore, we have

𝒦MLtop(Db(L))|(ML)sti=0dN1𝒦MLtop(𝕋L(r,χ)i)|(ML)st.\displaystyle\mathcal{K}_{M^{L}}^{\rm{top}}(D^{b}(\mathcal{M}^{L{\dagger}}))_{\mathbb{Q}}|_{(M^{L})^{\rm{st}}}\cong\bigoplus_{i=0}^{dN-1}\mathcal{K}_{M^{L}}^{\rm{top}}(\mathbb{T}^{L}(r,\chi)_{i})_{\mathbb{Q}}|_{(M^{L})^{\rm{st}}}.

Note that each 𝕋L(r,χ)i|(ML)st\mathbb{T}^{L}(r,\chi)_{i}|_{(M^{L})^{\rm{st}}} is étale locally on (ML)st(M^{L})^{\rm{st}} independent of ii up to equivalence, since a choice of ii corresponds to a power of a Brauer class of (ML)st(M^{L})^{\rm{st}}. By comparing with (4.2), we conclude that

𝒦MLtop(𝕋L)|(ML)st(ML)st[β±1].\displaystyle\mathcal{K}_{M^{L}}^{\rm{top}}(\mathbb{T}^{L})_{\mathbb{Q}}|_{(M^{L})^{\rm{st}}}\cong\mathbb{Q}_{(M^{L})^{\rm{st}}}[\beta^{\pm 1}].

As 𝒦MLtop(𝕋L)\mathcal{K}_{M^{L}}^{\rm{top}}(\mathbb{T}^{L})_{\mathbb{Q}} is a direct summand of (4.1), we can write

𝒦MLtop(𝕋L)ICML[dimML][β±1]P0[β±1]P1[1][β±1],\displaystyle\mathcal{K}_{M^{L}}^{\rm{top}}(\mathbb{T}^{L})_{\mathbb{Q}}\cong\mathrm{IC}_{M^{L}}[-\dim M^{L}][\beta^{\pm 1}]\oplus P_{0}[\beta^{\pm 1}]\oplus P_{1}[1][\beta^{\pm 1}],

where PiPerv(ML)P_{i}\in\mathrm{Perv}(M^{L}) for i=0,1i=0,1 are semisimple perverse sheaves. It is enough to show that P0=P1=0P_{0}=P_{1}=0.

We take yMLy\in M^{L} and let QyQ_{y} be the corresponding Ext-quiver with the associated dimension vector 𝒅\bm{d}, the moduli stack 𝒳Qy(𝒅)\mathcal{X}_{Q_{y}}(\bm{d}) and its good moduli space XQy(𝒅)X_{Q_{y}}(\bm{d}) as in (2.13). Let

𝕋Qy(𝒅)wDb(𝒳y(𝒅))\displaystyle\mathbb{T}_{Q_{y}}(\bm{d})_{w}\subset D^{b}(\mathcal{X}_{y}(\bm{d}))

be the quasi-BPS category for the quiver QyQ_{y}, see Remark 2.4. By [PTe, Theorem 8.26], there is an isomorphism

𝒦XQy(𝒅)(𝕋Qy(𝒅)w)ICXQy(𝒅)[dimXQy(𝒅)][β±1].\displaystyle\mathcal{K}_{X_{Q_{y}}(\bm{d})}(\mathbb{T}_{Q_{y}}(\bm{d})_{w})_{\mathbb{Q}}\cong\mathrm{IC}_{X_{Q_{y}}(\bm{d})}[-\dim X_{Q_{y}}(\bm{d})][\beta^{\pm 1}].

As 𝕋Qy(𝒅)w\mathbb{T}_{Q_{y}}(\bm{d})_{w} gives an étale local model for 𝕋L\mathbb{T}^{L}, we have Pi=0P_{i}=0 for i=0,1i=0,1. ∎

We remark that the method used in [PTe, Proof of Theorem 8.14 assuming Theorem 8.15] also applies to compute the topological K-theory of the quasi-BPS category 𝕋L\mathbb{T}^{L} for all tuples (r,χ,w)(r,\chi,w). We do not present full arguments as we will not use the following result in this paper. Further, we do not compute the topological K-theory of quasi-BPS categories for general tuples (r,χ,w)(r,\chi,w) in the case L=ΩCL=\Omega_{C}, as the proof of Theorem 6.9 uses the support lemma [PTc, Theorem 6.6], which holds for tuples satisfying the BPS condition, see the proof of Lemma 6.6.

To state the result, we first need to introduce some notation. Write (r,χ)=d(r0,χ0)(r,\chi)=d(r_{0},\chi_{0}) for (r0,χ0)(r_{0},\chi_{0}) coprime. We recall the set SwdS^{d}_{w} of partitions (di)i=1k(d_{i})_{i=1}^{k} of dd considered in [PTe, Section 8.1], and computed in [PTe, Proposition 8.5 and Lemma 8.6]. The set SwdS^{d}_{w} labels the summands of the topological K-theory of quasi-BPS categories for symmetric quivers, in this case for the quiver with one vertex and e=1+lr02e=1+lr_{0}^{2} loops. Note that this is the Ext-quiver for points in the deepest stratum on MLM^{L}, see Subsection 2.3.

Definition 4.2.

Let QQ be the quiver with one vertex and ee loops. Let d>0d>0 and ww\in\mathbb{Z}. The set SwdS^{d}_{w} consists of all partitions (di)i=1k(d_{i})_{i=1}^{k} of dd such that

wi:=e12di(j<idjj>idj)+wdidw_{i}:=\frac{e-1}{2}d_{i}\left(\sum_{j<i}d_{j}-\sum_{j>i}d_{j}\right)+\frac{wd_{i}}{d}\in\mathbb{Z}

for all 1ik1\leqslant i\leqslant k.

Note that, if ee is odd, the condition above is that each did_{i} is divisible by d/gcd(d,w)d/\gcd(d,w). Thus SwdS^{d}_{w} is in a natural bijection with the set of partitions of gcd(d,w)\gcd(d,w).

In the rest of this subsection, we assume e=1+lr02e=1+lr_{0}^{2}. We will refer to elements of SwdS^{d}_{w} either as partitions (di)i=1k(d_{i})_{i=1}^{k} or as tuples (di,wi)i=1k(d_{i},w_{i})_{i=1}^{k}.

We define the set S(r,χ,w)S(r,\chi,w) as follows. Let m:=gcd(r,χ,w+1gsp)m:=\gcd(r,\chi,w+1-g^{\mathrm{sp}}) and let (r~,χ~,w~):=1m(r,χ,w+1gsp)(\widetilde{r},\widetilde{\chi},\widetilde{w}):=\frac{1}{m}(r,\chi,w+1-g^{\mathrm{sp}}). Then S(r,χ,w)S(r,\chi,w) is the set of partitions (ri,χi,wi)i=1k(r_{i},\chi_{i},w_{i})_{i=1}^{k} of (r,χ,w+1gsp)(r,\chi,w+1-g^{\mathrm{sp}}) such that (ri,χi,wi)=mi(r~,χ~,w~)(r_{i},\chi_{i},w_{i})=m_{i}(\widetilde{r},\widetilde{\chi},\widetilde{w}) for some mi>0m_{i}\in\mathbb{Z}_{>0}. Thus there is a natural bijection between S(r,χ,w)S(r,\chi,w) and the set of partitions of gg. We note the following:

Proposition 4.3.

There is a natural bijection of sets

SwdS(r,χ,w),(di,wi)i=1k(dir0,diχ0,w~i)i=1kS^{d}_{w}\xrightarrow{\cong}S(r,\chi,w),\,\,(d_{i},w_{i})_{i=1}^{k}\mapsto(d_{i}r_{0},d_{i}\chi_{0},\widetilde{w}_{i})_{i=1}^{k}

such that r0|wiw~ir_{0}|w_{i}-\widetilde{w}_{i} for all 1ik1\leqslant i\leqslant k. In particular, if the vector (r,χ,w+1gsp)(r,\chi,w+1-g^{\mathrm{sp}}) is primitive, then Swd={(d)}S^{d}_{w}=\{(d)\}.

Proof.

By definition, the set SwdS^{d}_{w} contains partitions (di)i=1k(d_{i})_{i=1}^{k} such that

wi:=lr022di(ddi)+wdid.w_{i}:=\frac{lr_{0}^{2}}{2}d_{i}(d-d_{i})+\frac{wd_{i}}{d}\in\mathbb{Z}.

Alternatively, we need to have that

lr02(d1)di2+wdid, or wi:=(lr02(d1)d+2w)di2d.\frac{lr_{0}^{2}(d-1)d_{i}}{2}+\frac{wd_{i}}{d}\in\mathbb{Z},\text{ or }w^{\circ}_{i}:=\frac{(lr_{0}^{2}(d-1)d+2w)d_{i}}{2d}\in\mathbb{Z}.

Note that r0|wiwir_{0}|w_{i}-w^{\circ}_{i}. Further, we have that 2d|lr02(d1)dlr0d(r0d1)2d|lr_{0}^{2}(d-1)d-lr_{0}d(r_{0}d-1). Thus, SwdS^{d}_{w} consists of partitions (di)i=1k(d_{i})_{i=1}^{k} such that

w~i:=lr0d(r0d1)+2w2ddi.\widetilde{w}_{i}:=\frac{lr_{0}d(r_{0}d-1)+2w}{2d}d_{i}\in\mathbb{Z}.

Note that r0|wiw~ir_{0}|w_{i}-\widetilde{w}_{i} for all 1ik1\leqslant i\leqslant k. By a direct computation, we have that

m=gcd(d,w+1gsp)=gcd(d,w+lr0d(r0d1)2).m=\gcd(d,w+1-g^{\mathrm{sp}})=\gcd\left(d,w+\frac{lr_{0}d(r_{0}d-1)}{2}\right).

Then SwdS^{d}_{w} is in bijection with partitions of (d,w+lr0d(r0d1)2)\left(d,w+\frac{lr_{0}d(r_{0}d-1)}{2}\right) with all terms divisible by 1m(d,w+lr0d(r0d1)2)\frac{1}{m}\left(d,w+\frac{lr_{0}d(r_{0}d-1)}{2}\right). This last set of partitions is in natural bijection with S(r,χ,w)S(r,\chi,w). ∎

We define some direct sums of IC (alternatively, BPS) sheaves associated with partitions of dd. For a partition A=(di)i=1kA=(d_{i})_{i=1}^{k} of dd, its length is defined to be (A):=k\ell(A):=k. Assume the set {d1,,dk}={e1,,es}\{d_{1},\ldots,d_{k}\}=\{e_{1},\ldots,e_{s}\} has cardinality ss and that, for each 1is1\leqslant i\leqslant s, there are mim_{i} elements in {d1,,dk}\{d_{1},\ldots,d_{k}\} equal to eie_{i}. We define the following maps, given by the direct sums of polystable Higgs bundles:

i\displaystyle\oplus_{i} :ML(eir0,eiχ0)×miML(mieir0,mieiχ0),\displaystyle\colon M^{L}(e_{i}r_{0},e_{i}\chi_{0})^{\times m_{i}}\to M^{L}(m_{i}e_{i}r_{0},m_{i}e_{i}\chi_{0}),
\displaystyle\oplus^{\prime} :×i=1sML(mieir0,mieiχ0)ML.\displaystyle\colon\times_{i=1}^{s}M^{L}(m_{i}e_{i}r_{0},m_{i}e_{i}\chi_{0})\to M^{L}.

The above maps are finite maps. We define the following perverse sheaves:

Symmi(ICML(eir0,eiχ0))\displaystyle\mathrm{Sym}^{m_{i}}\big{(}\mathrm{IC}_{M^{L}(e_{i}r_{0},e_{i}\chi_{0})}\big{)} :=i,(ICML(eir0,eiχ0)mi)𝔖mi,\displaystyle:=\oplus_{i,\ast}\left(\mathrm{IC}_{M^{L}(e_{i}r_{0},e_{i}\chi_{0})}^{\boxtimes m_{i}}\right)^{\mathfrak{S}_{m_{i}}},
(4.3) ICA\displaystyle\mathrm{IC}_{A} :=(i=1sSymmi(ICML(eir0,eiχ0))).\displaystyle:=\oplus^{\prime}_{\ast}\left(\boxtimes_{i=1}^{s}\mathrm{Sym}^{m_{i}}\big{(}\mathrm{IC}_{M^{L}(e_{i}r_{0},e_{i}\chi_{0})}\big{)}\right).
Definition 4.4.

For a tuple (r,χ,w)(r,\chi,w) with (r,χ)=d(r0,χ0)(r,\chi)=d(r_{0},\chi_{0}) such that (r0,χ0)(r_{0},\chi_{0}) are coprime, define the following direct sum of symmetric products of BPS sheaves:

(4.4) ICMLw\displaystyle\mathrm{IC}^{w}_{M^{L}} :=ASwdICA[(A)]D(Sh(ML)).\displaystyle:=\bigoplus_{A\in S^{d}_{w}}\mathrm{IC}_{A}[-\ell(A)]\in D(\mathrm{Sh}_{\mathbb{Q}}(M^{L})).

Then, as in [PTe, Proof of Theorem 8.14 assuming Theorem 8.15], to which we refer the reader for full details, one shows that:

Proposition 4.5.

Suppose that l>2g2l>2g-2. There is an isomorphism in D(Sh(ML))D(\mathrm{Sh}_{\mathbb{Q}}(M^{L})):

(4.5) ICMLw[dimML][β±1]𝒦MLtop(𝕋L).\displaystyle\mathrm{IC}^{w}_{M^{L}}[-\dim M^{L}][\beta^{\pm 1}]\cong\mathcal{K}_{M^{L}}^{\rm{top}}(\mathbb{T}^{L})_{\mathbb{Q}}.
Remark 4.6.

We explain how one obtains, for ASwdA\in S^{d}_{w}, a map

(4.6) ICA[(A)dimML][β±1]B=:ICA𝒦MLtop(𝕋L),\displaystyle\mathrm{IC}_{A}[-\ell(A)-\dim M^{L}][\beta^{\pm 1}]\oplus B=:\mathrm{IC}^{\dagger}_{A}\to\mathcal{K}_{M^{L}}^{\rm{top}}(\mathbb{T}^{L})_{\mathbb{Q}},

which induces one of summands of (4.5). Here, BB is a direct sum of shifted perverse sheaves of support strictly contained in the support of ICA\mathrm{IC}_{A}.

Let A=(di,wi)i=1kSwdA=(d_{i},w_{i})_{i=1}^{k}\in S^{d}_{w}, and consider the corresponding tuple (ri,χi,w~i)i=1kS(r,χ,w)(r_{i},\chi_{i},\widetilde{w}_{i})_{i=1}^{k}\in S(r,\chi,w) from Proposition 4.3. There are natural equivalences

(4.7) 𝕋L(r,χ)w𝕋L(r,χ)w\displaystyle\mathbb{T}^{L}(r,\chi)_{w}\cong\mathbb{T}^{L}(r,\chi^{\prime})_{w^{\prime}}

if r|χχr|\chi^{\prime}-\chi and r|wwr|w^{\prime}-w. The map (4.6) is induced by applying topological K-theory to the following functor, which is the composition of the Hall product with equivalences (4.7), see [PTe, Proposition 8.2] and [PTe, Proof of Theorem 8.14 assuming Theorem 8.15]):

(4.8) i=1k𝕋L(ri,χi)w~ii=1k𝕋L(ri,χi)wi𝕋L(r,χ)w.\displaystyle\bigotimes_{i=1}^{k}\mathbb{T}^{L}(r_{i},\chi_{i})_{\widetilde{w}_{i}}\cong\bigotimes_{i=1}^{k}\mathbb{T}^{L}(r_{i},\chi_{i})_{w_{i}}\to\mathbb{T}^{L}(r,\chi)_{w}.

4.2. Equivalences of rational topological K-theories for l>2g2l>2g-2

For simplicity, we write

L=L(r,χ),L=L(r,w+1gsp).\displaystyle\mathcal{M}^{L}=\mathcal{M}^{L}(r,\chi),\ \mathcal{M}^{L^{\prime}}=\mathcal{M}^{L}(r,w+1-g^{\rm{sp}}).

Let (BL)ellBL(B^{L})^{\rm{ell}}\subset B^{L} be the elliptic locus, i.e. the locus corresponding to reduced and irreducible spectral curves. We set

(L)ell=L×BL(BL)ell,(L)ell=L×BL(BL)ell.\displaystyle(\mathcal{M}^{L})^{\rm{ell}}=\mathcal{M}^{L}\times_{B^{L}}(B^{L})^{\rm{ell}},\ (\mathcal{M}^{L^{\prime}})^{\rm{ell}}=\mathcal{M}^{L^{\prime}}\times_{B^{L}}(B^{L})^{\rm{ell}}.

Let 𝒫sm\mathcal{P}^{\rm{sm}} be the Poincare line bundle on (L)sm×(BL)sm(L)sm(\mathcal{M}^{L^{\prime}})^{\rm{sm}}\times_{(B^{L})^{\rm{sm}}}(\mathcal{M}^{L})^{\rm{sm}} as in (2.29). By [Ari13], it uniquely extends to a maximal Cohen-Macaulay sheaf

𝒫ellCoh((L)ell×(BL)ell(L)ell)\displaystyle\mathcal{P}^{\rm{ell}}\in\operatorname{Coh}((\mathcal{M}^{L^{\prime}})^{\rm{ell}}\times_{(B^{L})^{\rm{ell}}}(\mathcal{M}^{L})^{\rm{ell}})

which induces an equivalence

(4.9) Φ𝒫ell:Db((L)ell)χgsp+1Db((L)ell)w.\displaystyle\Phi_{\mathcal{P}^{\rm{ell}}}\colon D^{b}((\mathcal{M}^{L^{\prime}})^{\rm{ell}})_{-\chi-g^{\rm{sp}}+1}\stackrel{{\scriptstyle\sim}}{{\to}}D^{b}((\mathcal{M}^{L})^{\rm{ell}})_{w}.
Lemma 4.7.

There is an object

𝒫𝕋L(r,w+1gsp)χ+gsp1BL𝕋L(r,χ)wDb(L×BLL)\displaystyle\mathcal{P}\in\mathbb{T}^{L}(r,w+1-g^{\rm{sp}})_{\chi+g^{\rm{sp}}-1}\boxtimes_{B^{L}}\mathbb{T}^{L}(r,\chi)_{w}\subset D^{b}(\mathcal{M}^{L^{\prime}}\times_{B^{L}}\mathcal{M}^{L})

such that 𝒫|(BL)ell𝒫ell\mathcal{P}|_{(B^{L})^{\rm{ell}}}\cong\mathcal{P}^{\rm{ell}}.

Proof.

The restriction functor

(4.10) Db(L×BLL)Db((L)ell×(BL)ell(L)ell)\displaystyle D^{b}(\mathcal{M}^{L^{\prime}}\times_{B^{L}}\mathcal{M}^{L})\to D^{b}((\mathcal{M}^{L^{\prime}})^{\rm{ell}}\times_{(B^{L})^{\rm{ell}}}(\mathcal{M}^{L})^{\rm{ell}})

is essentially surjective. Let 𝒫Db(L×BLL)\mathcal{P}^{\prime}\in D^{b}(\mathcal{M}^{L^{\prime}}\times_{B^{L}}\mathcal{M}^{L}) be a lift of 𝒫ell\mathcal{P}^{\rm{ell}}. From the semiorthogonal decomposition in Theorem 2.10, the subcategory

(4.11) 𝕋L(r,w+1gsp)χ+gsp1BL𝕋L(r,χ)wDb(L×BLL)\displaystyle\mathbb{T}^{L}(r,w+1-g^{\rm{sp}})_{\chi+g^{\rm{sp}}-1}\boxtimes_{B^{L}}\mathbb{T}^{L}(r,\chi)_{w}\subset D^{b}(\mathcal{M}^{L^{\prime}}\times_{B^{L}}\mathcal{M}^{L})

is a part of a semiorthogonal decomposition by [Kuz11, Theorem 5.8]. Its semiorthogonal complements are generated by categorical Hall products, so they are sent to zero under the functor (4.10). Therefore by taking the projection of 𝒫\mathcal{P}^{\prime} to the subcategory (4.11), we obtain a desired 𝒫\mathcal{P}. ∎

For an object 𝒫\mathcal{P} as in Lemma 4.7, there is an induced functor

(4.12) Φ𝒫:𝕋L(r,w+1gsp)χ+1gsp𝕋L(r,χ)w.\displaystyle\Phi_{\mathcal{P}}\colon\mathbb{T}^{L}(r,w+1-g^{\rm{sp}})_{-\chi+1-g^{\rm{sp}}}\to\mathbb{T}^{L}(r,\chi)_{w}.

In general we cannot expect the above Fourier-Mukai functor to be an equivalence, since there is an ambiguity in the choice of 𝒫\mathcal{P}. However, following Groechenig–Shen [GS], the functor induces an isomorphism in topological K-theory:

Proposition 4.8.

Suppose that l>2g2l>2g-2. The functor Φ𝒫\Phi_{\mathcal{P}} in (4.12) induces a (rational) equivalence:

(4.13) Ktop(𝕋L(r,w+1gsp)χgsp+1)Ktop(𝕋L(r,χ)w).\displaystyle K^{\rm{top}}(\mathbb{T}^{L}(r,w+1-g^{\rm{sp}})_{-\chi-g^{\rm{sp}}+1})_{\mathbb{Q}}\stackrel{{\scriptstyle\sim}}{{\to}}K^{\rm{top}}(\mathbb{T}^{L}(r,\chi)_{w})_{\mathbb{Q}}.
Proof.

Note that we have the following diagram

Here π\pi, π\pi^{\prime} are good moduli space morphisms. We write 𝕋L=𝕋L(r,χ)w\mathbb{T}^{L}=\mathbb{T}^{L}(r,\chi)_{w} and 𝕋L=𝕋L(r,w+1gsp)χgsp+1\mathbb{T}^{L^{\prime}}=\mathbb{T}^{L}(r,w+1-g^{\rm{sp}})_{-\chi-g^{\rm{sp}}+1}. The functor Φ𝒫\Phi_{\mathcal{P}} is linear over Perf(BL)\mathrm{Perf}(B^{L}), so it induces a morphism in D(Sh(BL))D(\mathrm{Sh}_{\mathbb{Q}}(B^{L})):

(4.14) Φ𝒫K:𝒦BLtop(h𝕋L)𝒦BLtop(h𝕋L).\displaystyle\Phi_{\mathcal{P}\mathbb{Q}}^{K}\colon\mathcal{K}_{B^{L}}^{\rm{top}}(h^{\prime}_{\ast}\mathbb{T}^{L^{\prime}})_{\mathbb{Q}}\to\mathcal{K}_{B^{L}}^{\rm{top}}(h_{\ast}\mathbb{T}^{L})_{\mathbb{Q}}.

It is enough to show that (4.14) is an equivalence, as (4.13) is given by taking the global section of (4.14), see Proposition 3.6.

By Proposition 4.1, we have

𝒦BLtop(h𝕋L)h𝒦MLtop(𝕋L)hICML[dimML][β±1].\displaystyle\mathcal{K}_{B^{L}}^{\rm{top}}(h_{\ast}\mathbb{T}^{L})_{\mathbb{Q}}\cong h_{\ast}\mathcal{K}_{M^{L}}^{\rm{top}}(\mathbb{T}^{L})_{\mathbb{Q}}\cong h_{\ast}\mathrm{IC}_{M^{L}}[-\dim M^{L}][\beta^{\pm 1}].

Similarly, we have

𝒦BLtop(h𝕋L)hICML[dimML][β±1].\displaystyle\mathcal{K}_{B^{L}}^{\rm{top}}(h^{\prime}_{\ast}\mathbb{T}^{L^{\prime}})_{\mathbb{Q}}\cong h^{\prime}_{\ast}\mathrm{IC}_{M^{L^{\prime}}}[-\dim M^{L^{\prime}}][\beta^{\pm 1}].

On the other hand, as hh is proper, we may apply the BBDG decomposition theorem [BBD82] and write

hICML=iICZi(Li)[ki],\displaystyle h_{\ast}\mathrm{IC}_{M^{L}}=\bigoplus_{i}\mathrm{IC}_{Z_{i}}(L_{i})[-k_{i}],

where ZiBLZ_{i}\subset B^{L} is an irreducible closed subset and LiL_{i} is a local system on a dense open subset of ZiZ_{i}. By [MS23, Theorem 0.4], each generic point of ZiZ_{i} is contained in (BL)ell(B^{L})^{\rm{ell}}. We have the same support property for hICMLh^{\prime}_{\ast}\mathrm{IC}_{M^{L^{\prime}}}. Therefore it is enough to check that (4.14) is an equivalence on (BL)ellBL(B^{L})^{\rm{ell}}\subset B^{L}.

The restriction of (4.14) to (BL)ell(B^{L})^{\rm{ell}} is

Φ𝒫ellK:𝒦(BL)elltop(h𝕋L|(BL)ell)𝒦(BL)elltop(h𝕋L|(BL)ell).\displaystyle\Phi_{\mathcal{P}^{\rm{ell}}}^{K}\colon\mathcal{K}_{(B^{L})^{\rm{ell}}}^{\rm{top}}(h_{\ast}\mathbb{T}^{L^{\prime}}|_{(B^{L})^{\rm{ell}}})_{\mathbb{Q}}\to\mathcal{K}_{(B^{L})^{\rm{ell}}}^{\rm{top}}(h_{\ast}\mathbb{T}^{L}|_{(B^{L})^{\rm{ell}}})_{\mathbb{Q}}.

The above map is an equivalence since

𝕋L|(BL)ell=Db((L)ell)w,𝕋L|(BL)ell=Db((L)ell)χgsp+1\displaystyle\mathbb{T}^{L}|_{(B^{L})^{\rm{ell}}}=D^{b}((\mathcal{M}^{L})^{\rm{ell}})_{w},\ \mathbb{T}^{L^{\prime}}|_{(B^{L})^{\rm{ell}}}=D^{b}((\mathcal{M}^{L^{\prime}})^{\rm{ell}})_{-\chi-g^{\rm{sp}}+1}

and Φ𝒫ell\Phi_{\mathcal{P}^{\rm{ell}}} induces the equivalence (4.9). ∎

A similar statement also holds for more general quasi-BPS categories. Namely, for any tuple (r,χ,w)(r,\chi,w), there exists an object

𝒫𝕋L(r,w+1gsp)χ+gsp1BL𝕋L(r,χ)wDb(L×BLL)\mathcal{P}^{\prime}\in\mathbb{T}^{L}(r,w+1-g^{\rm{sp}})_{\chi+g^{\rm{sp}}-1}\boxtimes_{B^{L}}\mathbb{T}^{L}(r,\chi)_{w}\subset D^{b}(\mathcal{M}^{L^{\prime}}\times_{B^{L}}\mathcal{M}^{L})

which induces an equivalence of rational topological K-theory:

(4.15) Ktop(𝕋L(r,w+1gsp)χgsp+1)Ktop(𝕋L(r,χ)w).\displaystyle K^{\rm{top}}(\mathbb{T}^{L}(r,w+1-g^{\rm{sp}})_{-\chi-g^{\rm{sp}}+1})_{\mathbb{Q}}\stackrel{{\scriptstyle\sim}}{{\to}}K^{\rm{top}}(\mathbb{T}^{L}(r,\chi)_{w})_{\mathbb{Q}}.

We explain the construction of such an object. The proof that it induces the isomorphism (4.15) is analogous to the proof of Proposition 4.8 using the explicit form of the supports in the BBDG decomposition theorem.

First, using Proposition 4.3 one shows that there is a bijection of sets

S(r,χ,w)S(r,w+1gsp,χ+1gsp),A=(ri,χi,wi)i=1kA=(ri,wi,χi)i=1kS(r,\chi,w)\cong S(r,w+1-g^{\mathrm{sp}},-\chi+1-g^{\mathrm{sp}}),\,A=(r_{i},\chi_{i},w_{i})_{i=1}^{k}\mapsto A^{\prime}=(r_{i},w^{\prime}_{i},-\chi^{\prime}_{i})_{i=1}^{k}

such that ri|χiχir_{i}|\chi^{\prime}_{i}-\chi_{i} and ri|wiwir_{i}|w^{\prime}_{i}-w_{i} for every 1ik1\leqslant i\leqslant k. Indeed, there is a bijection

S(r,χ,w)S(r,w+1gsp,χ1+gsp),(ri,χi,wi)i=1k(ri,wi,χi)i=1k.S(r,\chi,w)\cong S(r,w+1-g^{\mathrm{sp}},-\chi-1+g^{\mathrm{sp}}),\,(r_{i},\chi_{i},w_{i})_{i=1}^{k}\mapsto(r_{i},w_{i},-\chi_{i})_{i=1}^{k}.

We have r|2(1gsp)r|2(1-g^{\mathrm{sp}}), so there is a bijection

S(r,w+1gsp,χ1+gsp)S(r,w+1gsp,χ+1gsp)S(r,w+1-g^{\mathrm{sp}},-\chi-1+g^{\mathrm{sp}})\cong S(r,w+1-g^{\mathrm{sp}},-\chi+1-g^{\mathrm{sp}})

which sends (ri,wi,χi)i=1k(ri,wi,χi)i=1k(r_{i},w_{i},-\chi_{i})_{i=1}^{k}\mapsto(r_{i},w_{i},-\chi^{\prime}_{i})_{i=1}^{k} such that ri|χiχir_{i}|\chi^{\prime}_{i}-\chi_{i} for all 1ik1\leqslant i\leqslant k. The functors (4.8) for AA and AA^{\prime} induce the summands ICA\mathrm{IC}^{\dagger}_{A} and ICA\mathrm{IC}^{\dagger}_{A^{\prime}} (see (4.6)) for 𝒦topML(𝕋)\mathcal{K}^{\mathrm{top}}_{M^{L}}(\mathbb{T})_{\mathbb{Q}} and 𝒦topML(𝕋)\mathcal{K}^{\mathrm{top}}_{M^{L^{\prime}}}(\mathbb{T}^{\prime})_{\mathbb{Q}}, respectively. For 1ik1\leqslant i\leqslant k, consider the Poincaré line bundle (tensored with equivalences 4.7):

𝒫iellDb(L(ri,χ)ell)wiDb(L(ri,wi)ell)χi.\mathcal{P}_{i}^{\mathrm{ell}}\in D^{b}\left(\mathcal{M}^{L}(r_{i},\chi)^{\mathrm{ell}}\right)_{w_{i}}\otimes D^{b}\left(\mathcal{M}^{L}(r_{i},w^{\prime}_{i})^{\mathrm{ell}}\right)_{-\chi^{\prime}_{i}}.

Then 𝒫A\mathcal{P}_{A} induces a map:

Φ𝒫A:hICAhICA\Phi_{\mathcal{P}_{A}}\colon h_{*}\mathrm{IC}_{A}^{\dagger}\to h_{*}\mathrm{IC}_{A^{\prime}}^{\dagger}

which is an isomorphism onto the summands of largest support, see the argument in the proof of Proposition 4.8:

hICA[β±1]hICA[β±1].h_{*}\mathrm{IC}_{A}[\beta^{\pm 1}]\xrightarrow{\sim}h_{*}\mathrm{IC}_{A^{\prime}}[\beta^{\pm 1}].

Using the above observation and the χ\chi-independence phenomenon from [MS23], we may choose eAe_{A}\in\mathbb{N} inductively on (A)\ell(A) such that the kernel

𝒫:=ASdw𝒫AeA\mathcal{P}^{\prime}:=\bigoplus_{A\in S^{d}_{w}}\mathcal{P}_{A}^{\oplus e_{A}}

induces the isomorphism (4.15).

4.3. Topological K-theory and the moduli of semistable vector bundles on a curve

Let CC be a smooth projective curve of genus g1g\geqslant 1. Consider the moduli stack of slope semistable vector bundles (r,χ):=un(r,χ)ss\mathcal{B}(r,\chi):=\mathcal{B}\mathrm{un}(r,\chi)^{\mathrm{ss}} of rank rr and degree χ\chi on the curve CC with good moduli space:

(r,χ)B(r,χ).\mathcal{B}(r,\chi)\to\mathrm{B}(r,\chi).

Recall from [PTa, Subsection 3.4] the categories

𝔹(r,χ)wDb((r,χ))w\mathbb{B}(r,\chi)_{w}\subset D^{b}(\mathcal{B}(r,\chi))_{w}

which are defined similarly to Definition 2.3, see also [ŠdB17, Păd]. These categories are twisted non-commutative resolutions of singularities of B(r,χ)\mathrm{B}(r,\chi). There is a semiorthogonal decomposition of Db((r,χ))wD^{b}\left(\mathcal{B}(r,\chi)\right)_{w} in terms of Hall products of such categories analogous to the decomposition from Theorem 2.36, see [PTa, Theorem 3.17].

In this subsection, we mention a computation of the (rational) topological K-theory of 𝔹(r,χ)w\mathbb{B}(r,\chi)_{w}. This computation is not used later in the paper, but it may be of independent interest, and it complements the discussion in [PTa, Subsection 3.4]. Recall from loc.cit. that the good moduli space B(r,χ)\mathrm{B}(r,\chi) has a stratification as in Subsection 2.3. Write (r,χ)=d(r0,χ0)(r,\chi)=d(r_{0},\chi_{0}) for (r0,χ0)(r_{0},\chi_{0}) coprime. The deepest stratum corresponds to vector bundles VE0V\otimes E_{0}, where E0E_{0} is a vector bundle of rank r0r_{0} and Euler characteristic χ0\chi_{0} and VV is a vector space of dimension dd. The Ext-quiver corresponding to such a point has one vertex and r02(g1)+1r_{0}^{2}(g-1)+1 loops. Consider the set of partitions SdwS^{d}_{w} for the quiver with one vertex and e=r02(g1)+1e=r_{0}^{2}(g-1)+1 loops, see Definition 4.2. For ASdwA\in S^{d}_{w}, define ICA\mathrm{IC}_{A} as in (4.1), and then define ICwBun(r,χ)ss\mathrm{IC}^{w}_{\mathrm{Bun}(r,\chi)^{\mathrm{ss}}} as in (4.4). The following is proved as [PTe, Proof of Theorem 8.14 assuming Theorem 8.15]:

Proposition 4.9.

There is an isomorphism in D(Sh(B(r,χ)))D(\mathrm{Sh}_{\mathbb{Q}}(\mathrm{B}(r,\chi))):

ICwB(r,χ)[dimB(r,χ)][β±1]𝒦B(r,χ)top(𝔹(r,χ)w).\mathrm{IC}^{w}_{\mathrm{B}(r,\chi)}[-\dim\mathrm{B}(r,\chi)][\beta^{\pm 1}]\cong\mathcal{K}_{\mathrm{B}(r,\chi)}^{\rm{top}}(\mathbb{B}(r,\chi)_{w})_{\mathbb{Q}}.

5. Torsion freeness of topological K-theories of quasi-BPS categories for l>2g2l>2g-2

We continue the discussion from Section 4, in particular we continue to assume that G=GL(r)G=\mathrm{GL}(r) and l>2g2l>2g-2. In this section, we prove, following Groechenig–Shen [GS], that there is an equivalence (4.13) without rationalization. This claim follows from the torsion freeness of Ktop(𝕋L(r,χ)w)K^{\rm{top}}(\mathbb{T}^{L}(r,\chi)_{w}). Following the argument in [GS], we use localization with respect to the \mathbb{C}^{\ast}-action on Higgs bundles which scales the Higgs field, where the fixed part corresponds to moduli stacks of chains on the curve.

5.1. The abelian category of chains

The moduli stack L(r,χ)\mathcal{M}^{L}(r,\chi) admits a \mathbb{C}^{\ast}-action given by t(F,θ)=(F,tθ)t\cdot(F,\theta)=(F,t\theta) for tt\in\mathbb{C}^{\ast}. A \mathbb{C}^{\ast}-fixed LL-twisted Higgs bundle corresponds to a chain, see [HT03, Lemma 9.2]:

(5.1) 0ϕ11ϕkk.\displaystyle\mathcal{E}_{0}\stackrel{{\scriptstyle\phi_{1}}}{{\to}}\mathcal{E}_{1}\to\cdots\stackrel{{\scriptstyle\phi_{k}}}{{\to}}\mathcal{E}_{k}.

Here, each i\mathcal{E}_{i} is a vector bundle on CC and ϕi:i1i\phi_{i}\colon\mathcal{E}_{i-1}\to\mathcal{E}_{i} is a morphism of coherent sheaves, such that (r,χ)={(ri,χi)}0ik(r_{\bullet},\chi_{\bullet})=\{(r_{i},\chi_{i})\}_{0\leqslant i\leqslant k} for (ri,χi)=(rank(i),χ(i))(r_{i},\chi_{i})=(\mathrm{rank}(\mathcal{E}_{i}),\chi(\mathcal{E}_{i})) satisfies

(5.2) (r0,χ0)++(rk,χk)=(r,χ+lk(k+1)/2).\displaystyle(r_{0},\chi_{0})+\cdots+(r_{k},\chi_{k})=(r,\chi+lk(k+1)/2).

Given a chain (5.1), there is a corresponding LL-twisted Higgs bundle:

(5.3) =i=0kiLi\displaystyle\mathcal{E}=\bigoplus_{i=0}^{k}\mathcal{E}_{i}\otimes L^{-i}

with Higgs field ϕ:L\phi\colon\mathcal{E}\to\mathcal{E}\otimes L naturally induced by ϕi\phi_{i}.

We denote by 𝒜k\mathcal{A}_{k} the abelian category of chains (5.1) such that each i\mathcal{E}_{i} is a coherent sheaf on CC.

Lemma 5.1.

For ,𝒜k\mathcal{E}_{\bullet},\mathcal{E}_{\bullet}^{\prime}\in\mathcal{A}_{k}, we have Ext3(,)=0\operatorname{Ext}^{\geqslant 3}(\mathcal{E}_{\bullet},\mathcal{E}_{\bullet}^{\prime})=0. Moreover, by setting (rank(i),χ(i))=(ri,χi)(\mathrm{rank}(\mathcal{E}_{i}),\chi(\mathcal{E}_{i}))=(r_{i},\chi_{i}), (rank(i),χ(i))=(ri,χi)(\mathrm{rank}(\mathcal{E}_{i}^{\prime}),\chi(\mathcal{E}_{i}^{\prime}))=(r_{i}^{\prime},\chi_{i}^{\prime}), and

χ((r,χ),(r,χ)):=i=0k1(ri(χiχi+1)χi(riri+1)),\displaystyle\chi((r_{\bullet},\chi_{\bullet}),(r_{\bullet}^{\prime},\chi_{\bullet}^{\prime})):=\sum_{i=0}^{k-1}\left(r_{i}(\chi_{i}^{\prime}-\chi_{i+1}^{\prime})-\chi_{i}(r_{i}^{\prime}-r_{i+1}^{\prime})\right),

we have

(5.4) i0(1)idimExti(,)=χ((r,χ),(r,χ)).\displaystyle\sum_{i\geqslant 0}(-1)^{i}\dim\operatorname{Ext}^{i}(\mathcal{E}_{\bullet},\mathcal{E}_{\bullet}^{\prime})=\chi((r_{\bullet},\chi_{\bullet}),(r_{\bullet}^{\prime},\chi_{\bullet}^{\prime})).
Proof.

The lemma follows from the fact that RHom(,)R\operatorname{Hom}(\mathcal{E}_{\bullet},\mathcal{E}_{\bullet}^{\prime}) is computed by the hypercohomology of the following complex, see [GPHS14, Proposition 4.4]:

iom(i,i)iom(i,i+1).\displaystyle\bigoplus_{i}\mathcal{H}om(\mathcal{E}_{i},\mathcal{E}_{i}^{\prime})\to\bigoplus_{i}\mathcal{H}om(\mathcal{E}_{i},\mathcal{E}_{i+1}^{\prime}).

For α=(αi)0ik\alpha_{\bullet}=(\alpha_{i})_{0\leqslant i\leqslant k} with αi\alpha_{i}\in\mathbb{R}, define the slope of a chain (5.1) to be

μα():=i=0k(χi+αiri)i=0kri.\displaystyle\mu_{\alpha_{\bullet}}(\mathcal{E}_{\bullet}):=\frac{\sum_{i=0}^{k}(\chi_{i}+\alpha_{i}r_{i})}{\sum_{i=0}^{k}r_{i}}.

There is a notion of μα\mu_{\alpha_{\bullet}}-stability on 𝒜k\mathcal{A}_{k}: a chain (5.1) is μα\mu_{\alpha_{\bullet}}-(semi)stable if we have

μα()<()μα()\displaystyle\mu_{\alpha_{\bullet}}(\mathcal{E}_{\bullet}^{\prime})<(\leqslant)\mu_{\alpha_{\bullet}}(\mathcal{E}_{\bullet})

for any non-zero subobject \mathcal{E}_{\bullet}^{\prime}\subsetneq\mathcal{E}_{\bullet} in 𝒜k\mathcal{A}_{k}. The μα\mu_{\alpha_{\bullet}}-stability corresponds to the stability of the LL-twisted Higgs bundle (5.3) when αi=il\alpha_{i}=-il, as in this case μα()\mu_{\alpha}(\mathcal{E}_{\bullet}) is the usual slope χ/r\chi/r of (5.3).

Lemma 5.2.

Suppose that αiαi+1>2g2\alpha_{i}-\alpha_{i+1}>2g-2. Then, for any μα\mu_{\alpha_{\bullet}}-semistable \mathcal{E}_{\bullet}, \mathcal{E}_{\bullet}^{\prime} with μα()μα()\mu_{\alpha_{\bullet}}(\mathcal{E}_{\bullet})\geqslant\mu_{\alpha_{\bullet}}(\mathcal{E}_{\bullet}^{\prime}), we have Ext2(,)=0\operatorname{Ext}^{\geqslant 2}(\mathcal{E}_{\bullet},\mathcal{E}_{\bullet}^{\prime})=0.

Proof.

The lemma follows from the argument of [GPHS14, Lemma 4.6]. ∎

5.2. Moduli stacks of chains

Let 𝒞(r,χ)α\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}} be the moduli stack of μα\mu_{\alpha_{\bullet}}-semistable chains (5.1) such that (rank(i),χ(i))=(ri,χi)(\mathrm{rank}(\mathcal{E}_{i}),\chi(\mathcal{E}_{i}))=(r_{i},\chi_{i}). Denote by

(5.5) 𝒞(r,χ)α𝒞(r,χ)α\displaystyle\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}\circ}\subset\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}}

the open substack such that ϕi0\phi_{i}\neq 0 when ri1ri0r_{i-1}r_{i}\neq 0. By [HT03, Lemma 9.2], the \mathbb{C}^{\ast}-fixed stack of L(r,χ)\mathcal{M}^{L}(r,\chi) is given by

(5.6) (r0,χ0)++(rk,χk)=(r,χ+lk(k+1)/2)𝒞(r,χ)α, for αi=il,0ik.\displaystyle\coprod_{(r_{0},\chi_{0})+\cdots+(r_{k},\chi_{k})=(r,\chi+lk(k+1)/2)}\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}\circ},\text{ for }\alpha_{i}=-il,0\leqslant i\leqslant k.

For Coh(C)\mathcal{F}\in\operatorname{Coh}(C), we use the same symbol 𝒜k\mathcal{F}\in\mathcal{A}_{k} to denote the constant chain

ididid.\displaystyle\mathcal{F}\stackrel{{\scriptstyle\operatorname{id}}}{{\to}}\mathcal{F}\stackrel{{\scriptstyle\operatorname{id}}}{{\to}}\cdots\stackrel{{\scriptstyle\operatorname{id}}}{{\to}}\mathcal{F}.
Definition 5.3.

For m0m\gg 0, a pair

(5.7) (,s),𝒜k,s:𝒪C(m)\displaystyle(\mathcal{E}_{\bullet},s),\ \mathcal{E}_{\bullet}\in\mathcal{A}_{k},\ s\colon\mathcal{O}_{C}(-m)\to\mathcal{E}_{\bullet}

is called a JS (Joyce–Song) α\alpha_{\bullet}-stable pair if \mathcal{E}_{\bullet} is μα\mu_{\alpha_{\bullet}}-semistable and, for any surjection j:j\colon\mathcal{E}_{\bullet}\twoheadrightarrow\mathcal{E}_{\bullet}^{\prime} in 𝒜k\mathcal{A}_{k} with μα()=μα()\mu_{\alpha_{\bullet}}(\mathcal{E}_{\bullet}^{\prime})=\mu_{\alpha_{\bullet}}(\mathcal{E}_{\bullet}), we have js0j\circ s\neq 0.

Let 𝒜k\mathcal{A}_{k}^{{\dagger}} be the abelian category of pairs

W𝒪C(m),\displaystyle W\otimes\mathcal{O}_{C}(-m)\to\mathcal{E}_{\bullet},

where WW is a finite dimensional vector space and 𝒜k\mathcal{E}_{\bullet}\in\mathcal{A}_{k}. Note that 𝒜k𝒜k\mathcal{A}_{k}\subset\mathcal{A}_{k}^{{\dagger}} is an abelian subcategory by regarding \mathcal{E}_{\bullet} as a pair (0)(0\to\mathcal{E}_{\bullet}). We denote by 𝒞(r,χ)α\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}{\dagger}} the moduli space of JS α\alpha_{\bullet}-stable pairs (5.7) such that (rank(i),χ(i))=(ri,χi)(\mathrm{rank}(\mathcal{E}_{i}),\chi(\mathcal{E}_{i}))=(r_{i},\chi_{i}). The natural projection

𝒞(r,χ)α𝒞(r,χ)α\displaystyle\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}{\dagger}}\to\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}}

is smooth with image contained in the open substack (5.5) by the JS stability.

Lemma 5.4.

Suppose that αiαi+12g2\alpha_{i}-\alpha_{i+1}\geqslant 2g-2. Then, for any μα\mu_{\alpha_{\bullet}}-semistable 𝒜k\mathcal{E}_{\bullet}\in\mathcal{A}_{k} and any JS α\alpha_{\bullet}-stable pair I=(𝒪C(m))I=(\mathcal{O}_{C}(-m)\to\mathcal{E}_{\bullet}^{\prime}) in 𝒜k\mathcal{A}_{k}^{{\dagger}} with μα()=μα()\mu_{\alpha_{\bullet}}(\mathcal{E}_{\bullet})=\mu_{\alpha_{\bullet}}(\mathcal{E}_{\bullet}^{\prime}), we have

Ext2(,I)=0.\operatorname{Ext}^{\geqslant 2}(\mathcal{E}_{\bullet},I)=0.

Moreover, we have

(5.8) dimHom(,I)dimExt1(,I)=χ((r,χ),(r,χ))\displaystyle\dim\operatorname{Hom}(\mathcal{E}_{\bullet},I)-\dim\operatorname{Ext}^{1}(\mathcal{E}_{\bullet},I)=\chi((r_{\bullet},\chi_{\bullet}),(r_{\bullet}^{\prime},\chi_{\bullet}^{\prime}))

where (rank(i),χ(i))=(ri,χi)(\mathrm{rank}(\mathcal{E}_{i}),\chi(\mathcal{E}_{i}))=(r_{i},\chi_{i}), (rank(i),χ(i))=(ri,χi)(\mathrm{rank}(\mathcal{E}_{i}^{\prime}),\chi(\mathcal{E}_{i}^{\prime}))=(r_{i}^{\prime},\chi_{i}^{\prime}).

Proof.

The complex RHom(,I)R\operatorname{Hom}(\mathcal{E}_{\bullet},I) is given by the cone of the map

RHom(,𝒪C(m))RHom(,).\displaystyle R\operatorname{Hom}(\mathcal{E}_{\bullet},\mathcal{O}_{C}(-m))\to R\operatorname{Hom}(\mathcal{E}_{\bullet},\mathcal{E}_{\bullet}^{\prime}).

Therefore, by Lemma 5.1, we get Ext3(,I)=0\operatorname{Ext}^{\geqslant 3}(\mathcal{E}_{\bullet},I)=0 and we obtain the exact sequence

(5.9) Ext2(,𝒪C(m))Ext2(,)Ext2(,I)0.\displaystyle\operatorname{Ext}^{2}(\mathcal{E}_{\bullet},\mathcal{O}_{C}(-m))\to\operatorname{Ext}^{2}(\mathcal{E}_{\bullet},\mathcal{E}_{\bullet}^{\prime})\to\operatorname{Ext}^{2}(\mathcal{E}_{\bullet},I)\to 0.

The vanishing Ext2(,I)=0\operatorname{Ext}^{2}(\mathcal{E}_{\bullet},I)=0 follows when αiαi+1>2g2\alpha_{i}-\alpha_{i+1}>2g-2 by Lemma 5.2. Suppose that αiαi+1=2g2\alpha_{i}-\alpha_{i+1}=2g-2. The dual of the first map in (5.9) is given by (see [GPHS14, Lemma 4.5]):

(5.10) Hom𝒜k+1(,+1ΩC)Hom𝒜k+1(𝒪C(m),+1ΩC).\displaystyle\operatorname{Hom}_{\mathcal{A}_{k+1}}(\mathcal{E}_{\bullet}^{\prime},\mathcal{E}_{\bullet+1}\otimes\Omega_{C})\to\operatorname{Hom}_{\mathcal{A}_{k+1}}(\mathcal{O}_{C}(-m),\mathcal{E}_{\bullet+1}\otimes\Omega_{C}).

Here \mathcal{E}_{\bullet}^{\prime} and +1\mathcal{E}_{\bullet+1} are regarded as objects in 𝒜k+1\mathcal{A}_{k+1} by

=(0k0),+1=(00k).\displaystyle\mathcal{E}_{\bullet}^{\prime}=(\mathcal{E}_{0}^{\prime}\to\cdots\to\mathcal{E}_{k}^{\prime}\to 0),\ \mathcal{E}_{\bullet+1}=(0\to\mathcal{E}_{0}\to\cdots\to\mathcal{E}_{k}).

The object +1ΩC\mathcal{E}_{\bullet+1}\otimes\Omega_{C} is μα\mu_{\alpha_{\bullet}}-semistable whose slope is the same as μα()\mu_{\alpha_{\bullet}}(\mathcal{E}_{\bullet}^{\prime}). Therefore the map (5.10) is injective by the definition of JS stability of II, hence Ext2(,I)=0\operatorname{Ext}^{2}(\mathcal{E}_{\bullet},I)=0. The formula (5.8) follows from the above vanishing of Ext2(,I)\operatorname{Ext}^{\geqslant 2}(\mathcal{E}_{\bullet},I) together with the equality (5.4). ∎

Let L\mathcal{M}^{L{\dagger}} be the moduli space of JS stable Higgs bundles as in (2.38). The \mathbb{C}^{\ast}-action on LL-twisted Higgs bundles naturally lifts to the action on L\mathcal{M}^{L{\dagger}} by t(F,θ,s)=(F,tθ,s)t\cdot(F,\theta,s)=(F,t\theta,s) for tt\in\mathbb{C}^{\ast}. Similarly to (5.6), the \mathbb{C}^{\ast}-fixed locus of L\mathcal{M}^{L{\dagger}} is

(5.11) (L)=(r0,χ0)++(rk,χk)=(r,χ+lk(k+1)/2)𝒞(r,χ)α,αi=il.\displaystyle(\mathcal{M}^{L{\dagger}})^{\mathbb{C}^{\ast}}=\coprod_{(r_{0},\chi_{0})+\cdots+(r_{k},\chi_{k})=(r,\chi+lk(k+1)/2)}\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}{\dagger}},\ \alpha_{i}=-il.

Note that, for l>2g2l>2g-2, each 𝒞(r,χ)α\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}{\dagger}} is a smooth projective variety since it is a component of a \mathbb{C}^{\ast}-fixed locus of the smooth quasi-projective variety L\mathcal{M}^{L{\dagger}} and it is supported on the fiber at 0B0\in B.

5.3. The Grothendieck ring of stacks

We will study the class of 𝒞(r,χ)α,\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet},{\dagger}} in the Grothendieck ring of varieties to show the torsion freeness of topological K-theory. In this subsection and in the next subsection, we discuss some terminology and lemmas about Grothendieck rings of varieties, stacks, and motivic Hall algebras. These are basic tools in the wall-crossing arguments of Donaldson-Thomas theory, and we refer the reader to [Bri12] for an introduction.

For an Artin stack 𝒮\mathcal{S} over \mathbb{C}, we denote by K(St/𝒮)K(\mathrm{St}/\mathcal{S}) the Grothendieck ring of stacks over 𝒮\mathcal{S}. Its underlying \mathbb{Q}-vector space is generated by symbols

[ρ:𝒳𝒮]\displaystyle[\rho\colon\mathcal{X}\to\mathcal{S}]

where 𝒳\mathcal{X} is an Artin stack of finite type over \mathbb{C} with affine geometric stabilizers, and these symbols satisfy certain motivic relations, see [Bri12, Definition 3.10] for its precise definition. We write K(St):=K(St/Spec)K(\mathrm{St}):=K(\mathrm{St}/\operatorname{Spec}\mathbb{C}).

Let K(Var)K(St)K(\mathrm{Var})\subset K(\mathrm{St}) be the subspace spanned by the class of varieties. We also denote by 𝕃K(Var)\mathbb{L}\in K(\mathrm{Var}) the class of the affine line 𝔸1\mathbb{A}^{1}. It is proved in [Bri12, Lemma 3.9] that we have the identity

(5.12) K(St)=K(Var)[1𝕃,1𝕃1,1𝕃n++𝕃+1|n1].\displaystyle K(\mathrm{St})=K(\mathrm{Var})\left[\frac{1}{\mathbb{L}},\frac{1}{\mathbb{L}-1},\frac{1}{\mathbb{L}^{n}+\cdots+\mathbb{L}+1}\,\Big{|}\,n\geqslant 1\right].

Let K^(Var)\widehat{K}(\mathrm{Var}) be the dimensional completion of K(Var)[𝕃1]K(\mathrm{Var})[\mathbb{L}^{-1}]. By expanding the denominators in the right hand side of (5.12) in terms of 𝕃1\mathbb{L}^{-1}, we obtain a map

K(St)K^(Var).\displaystyle K(\mathrm{St})\to\widehat{K}(\mathrm{Var}).

Let KK be a field. For a smooth projective variety XX over \mathbb{C}, set

PK(X,t):=idimK(Hi(X,K))ti[t].\displaystyle P_{K}(X,t):=\sum_{i}\dim_{K}(H^{i}(X,K))t^{i}\in\mathbb{Z}[t].

There exists an extension of PK(X,t)P_{K}(X,t) for any complex algebraic variety YY which satisfies the relation

PK(Y,t)=PK(YZ,t)+PK(Z,t)\displaystyle P_{K}(Y,t)=P_{K}(Y\setminus Z,t)+P_{K}(Z,t)

for any closed subvariety ZYZ\subset Y. The correspondence YPK(Y,t)Y\mapsto P_{K}(Y,t) induces the map, see [GS, Section 6]:

PK:K^(Var)[t][[t1]].\displaystyle P_{K}\colon\widehat{K}(\mathrm{Var})\to\mathbb{Q}[t][[t^{-1}]].

5.4. The motivic Hall algebra

Recall the abelian category 𝒜k\mathcal{A}_{k} from Subsection 5.1. Let 𝒪bj(𝒜k)\mathcal{O}bj(\mathcal{A}_{k}) be the moduli stack of objects in 𝒜k\mathcal{A}_{k}. We set

H(𝒜k):=K(St/𝒪bj(𝒜k)).\displaystyle H(\mathcal{A}_{k}):=K(\mathrm{St}/\mathcal{O}bj(\mathcal{A}_{k})).

There is an associative algebra structure on H(𝒜k)H(\mathcal{A}_{k}), called the motivic Hall algebra, defined as follows. Let x(𝒜k)\mathcal{E}x(\mathcal{A}_{k}) be the moduli stack of short exact sequences in 𝒜k\mathcal{A}_{k}

(5.13) 01320.\displaystyle 0\to\mathcal{E}_{1\bullet}\to\mathcal{E}_{3\bullet}\to\mathcal{E}_{2\bullet}\to 0.

There are evaluation morphisms

pi:x(𝒪bj(𝒜k))𝒪bj(𝒜k)\displaystyle p_{i}\colon\mathcal{E}x(\mathcal{O}bj(\mathcal{A}_{k}))\to\mathcal{O}bj(\mathcal{A}_{k})

sending (5.13) to i\mathcal{E}_{i\bullet}. The \ast-product on H(𝒜k)H(\mathcal{A}_{k}) is given by

[𝒳1ρ1𝒪bj(𝒜k)][𝒳2ρ2𝒪bj(𝒜k)]=[𝒳3ρ3𝒪bj(𝒜k)]\displaystyle[\mathcal{X}_{1}\stackrel{{\scriptstyle\rho_{1}}}{{\to}}\mathcal{O}bj(\mathcal{A}_{k})]\ast[\mathcal{X}_{2}\stackrel{{\scriptstyle\rho_{2}}}{{\to}}\mathcal{O}bj(\mathcal{A}_{k})]=[\mathcal{X}_{3}\stackrel{{\scriptstyle\rho_{3}}}{{\to}}\mathcal{O}bj(\mathcal{A}_{k})]

where (𝒳3,ρ3=p3η)(\mathcal{X}_{3},\rho_{3}=p_{3}\circ\eta) is given by the following diagram

where the left square is Cartesian.

Let 𝒪bj(𝒜k)\mathcal{O}bj(\mathcal{A}_{k}^{{\dagger}}) be the moduli stack of objects in 𝒜k\mathcal{A}_{k}^{{\dagger}}. Similarly to above, the \mathbb{Q}-vector space

H(𝒜k):=H(St/𝒪bj(𝒜k))\displaystyle H(\mathcal{A}_{k}^{{\dagger}}):=H(\mathrm{St}/\mathcal{O}bj(\mathcal{A}_{k}^{{\dagger}}))

admits a \ast-algebra structure given by the stack of short exact sequences in 𝒜k\mathcal{A}_{k}^{{\dagger}}. Since 𝒜k\mathcal{A}_{k} is an abelian subcategory of 𝒜k\mathcal{A}_{k}^{{\dagger}} by (0)\mathcal{E}_{\bullet}\mapsto(0\to\mathcal{E}_{\bullet}), there is an injective algebra homomorphism H(𝒜k)H(𝒜k)H(\mathcal{A}_{k})\to H(\mathcal{A}_{k}^{{\dagger}}). In particular, there are right and left actions of H(𝒜k)H(\mathcal{A}_{k}) on H(𝒜k)H(\mathcal{A}_{k}^{{\dagger}}).

Recall the moduli stack 𝒞(r,χ)α\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}} of semistable chains and the moduli space 𝒞(r,χ)α\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}{\dagger}} of JS stable chains. We set

δ(r,χ)α:=[𝒞(r,χ)α𝒪bj(𝒜k)]H(𝒜k),\displaystyle\delta_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}}:=[\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}}\to\mathcal{O}bj(\mathcal{A}_{k})]\in H(\mathcal{A}_{k}),
δ(r,χ)α:=[𝒞(r,χ)α𝒪bj(𝒜k)]H(𝒜k).\displaystyle\delta_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}{\dagger}}:=[\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}{\dagger}}\to\mathcal{O}bj(\mathcal{A}_{k}^{{\dagger}})]\in H(\mathcal{A}_{k}^{{\dagger}}).

We denote by 𝒞(r,χ)α\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}\sharp} the moduli stack of pairs (𝒪C(m))(\mathcal{O}_{C}(-m)\to\mathcal{E}_{\bullet}) such that \mathcal{E}_{\bullet} is μα\mu_{\alpha_{\bullet}}-semistable with (rank(i),χ(i))=(ri,χi)(\mathrm{rank}(\mathcal{E}_{i}),\chi(\mathcal{E}_{i}))=(r_{i},\chi_{i}) without imposing the JS stability. We also set

δ(r,χ)α:=[𝒞(r,χ)α𝒪bj(𝒜k)]H(𝒜k).\displaystyle\delta_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}\sharp}:=[\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}\sharp}\to\mathcal{O}bj(\mathcal{A}_{k}^{{\dagger}})]\in H(\mathcal{A}_{k}^{{\dagger}}).
Lemma 5.5.

The following identity holds in H(𝒜k)H(\mathcal{A}_{k}^{{\dagger}}):

(5.14) δ(r,χ)α=(r1,χ1)+(r2,χ2)=(r,χ)μα(ri,χi)=μα(r,χ)δ(r1,χ1)αδ(r2,χ2)α.\displaystyle\delta_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}\sharp}=\sum_{\begin{subarray}{c}(r_{\bullet}^{1},\chi_{\bullet}^{1})+(r_{\bullet}^{2},\chi_{\bullet}^{2})=(r_{\bullet},\chi_{\bullet})\\ \mu_{\alpha_{\bullet}}(r_{\bullet}^{i},\chi_{\bullet}^{i})=\mu_{\alpha_{\bullet}}(r_{\bullet},\chi_{\bullet})\end{subarray}}\delta_{(r_{\bullet}^{1},\chi_{\bullet}^{1})}^{\alpha_{\bullet}{\dagger}}\ast\delta_{(r_{\bullet}^{2},\chi_{\bullet}^{2})}^{\alpha_{\bullet}}.
Proof.

For any pair (𝒪C(m))(\mathcal{O}_{C}(-m)\to\mathcal{E}_{\bullet}), there is an exact sequence in 𝒜k\mathcal{A}_{k}^{{\dagger}}, unique up to isomorphism

0(𝒪C(m)1)(𝒪C(m))(02)0.\displaystyle 0\to(\mathcal{O}_{C}(-m)\to\mathcal{E}_{1\bullet})\to(\mathcal{O}_{C}(-m)\to\mathcal{E}_{\bullet})\to(0\to\mathcal{E}_{2\bullet})\to 0.

Here (𝒪C(m)1)(\mathcal{O}_{C}(-m)\to\mathcal{E}_{1\bullet}) is a JS stable pair and 2\mathcal{E}_{2\bullet} is μα\mu_{\alpha_{\bullet}}-semistable such that μα(i)=μα()\mu_{\alpha_{\bullet}}(\mathcal{E}_{i\bullet})=\mu_{\alpha_{\bullet}}(\mathcal{E}_{\bullet}). The above exact sequence is nothing but the Harder-Narasimhan filtration with respect to the JS stability. Then the lemma follows by describing the above Harder-Narasimhan filtration in terms of motivic Hall algebras, see [JS12, Formula (3.11)]. ∎

5.5. Proof of torsion freeness

In this subsection, we prove the torsion freeness of the topological K-theory of quasi-BPS categories for LL-twisted Higgs bundles using the technique of wall-crossing in Donaldson-Thomas theory [JS12, KS]. We use it to show that the equivalence in Proposition 4.8 holds integrally.

We denote by ΛK(Var)\Lambda\subset K(\mathrm{Var}) the \mathbb{Q}-subspace spanned by the classes of the products of Symi(C)\mathrm{Sym}^{i}(C) for ii\in\mathbb{Z} and 𝕃K(Var)\mathbb{L}\in K(\mathrm{Var}). Let Λ^\widehat{\Lambda} be the dimensional completion of Λ[𝕃1]\Lambda[\mathbb{L}^{-1}], and we use the same symbol Λ^\widehat{\Lambda} to denote its image in K^(Var)\widehat{K}(\mathrm{Var}).

Lemma 5.6.

For a smooth projective variety YY, suppose that its class [Y]K^(Var)[Y]\in\widehat{K}(\mathrm{Var}) lies in Λ^\widehat{\Lambda}. Then H(Y,)H^{\ast}(Y,\mathbb{Z}) is torsion free.

Proof.

The argument is the same as in [GS, Theorem 6.1]. It is enough to show

P(Y,t)=P𝔽p(Y,t)P_{\mathbb{Q}}(Y,t)=P_{\mathbb{F}_{p}}(Y,t)

for any prime pp. As [Y]Λ^[Y]\in\widehat{\Lambda}, it is enough to check this for Y=Symi(C)Y=\mathrm{Sym}^{i}(C), where it is known that H(Symi(C),)H^{\ast}(\mathrm{Sym}^{i}(C),\mathbb{Z}) is torsion-free, see [Mac62, Equation (12.3)]. ∎

Proposition 5.7.

Suppose that αiαi+12g2\alpha_{i}-\alpha_{i+1}\geqslant 2g-2. Then the class of the variety 𝒞(r,χ)α\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}{\dagger}} lies in Λ^\widehat{\Lambda}.

Proof.

We denote by Π\Pi the composition

Π:H(𝒜k)K(St)K^(Var),\displaystyle\Pi\colon H(\mathcal{A}_{k}^{{\dagger}})\to K(\mathrm{St})\to\widehat{K}(\mathrm{Var}),

where the first map forgets the map to 𝒪bj(𝒜k)\mathcal{O}bj(\mathcal{A}_{k}^{{\dagger}}). The stack 𝒞(r,χ)α\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}\sharp} is a vector bundle over 𝒞(r,χ)α\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}} with fiber 𝔸Hom(𝒪C(m),)\mathbb{A}^{\operatorname{Hom}(\mathcal{O}_{C}(-m),\mathcal{E}_{\bullet})}. For m0m\gg 0, we have

dimHom(𝒪C(m),)\displaystyle\dim\operatorname{Hom}(\mathcal{O}_{C}(-m),\mathcal{E}_{\bullet}) =dimHom(𝒪C(m),0)\displaystyle=\dim\operatorname{Hom}(\mathcal{O}_{C}(-m),\mathcal{E}_{0})
=mr0+χ0+(1g)r0.\displaystyle=mr_{0}+\chi_{0}+(1-g)r_{0}.

Therefore the image of Π\Pi of the left hand side of (5.14) is 𝕃mr0+χ0+(1g)r0[𝒞(r,χ)α]\mathbb{L}^{mr_{0}+\chi_{0}+(1-g)r_{0}}[\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}}].

On the other hand, by Lemma 5.4, the stack representing δ(r1,χ1)αδ(r2,χ2)α\delta_{(r_{\bullet}^{1},\chi_{\bullet}^{1})}^{\alpha_{\bullet}{\dagger}}\ast\delta_{(r_{\bullet}^{2},\chi_{\bullet}^{2})}^{\alpha_{\bullet}} in the right hand side of (5.14) is the vector bundle stack over 𝒞(r1,χ1)α×𝒞(r2,χ2)α\mathcal{C}_{(r^{1}_{\bullet},\chi^{1}_{\bullet})}^{\alpha_{\bullet}{\dagger}}\times\mathcal{C}_{(r^{2}_{\bullet},\chi^{2}_{\bullet})}^{\alpha_{\bullet}} with fiber of the form 𝔸a/𝔸b\mathbb{A}^{a}/\mathbb{A}^{b} such that ab=χ((r2,χ2),(r1,χ1))a-b=-\chi((r^{2}_{\bullet},\chi^{2}_{\bullet}),(r^{1}_{\bullet},\chi^{1}_{\bullet})). Therefore its class is 𝕃χ((r2,χ2),(r1,χ1))[𝒞(r1,χ1)α×𝒞(r2,χ2)α]\mathbb{L}^{-\chi((r^{2}_{\bullet},\chi^{2}_{\bullet}),(r^{1}_{\bullet},\chi^{1}_{\bullet}))}[\mathcal{C}_{(r^{1}_{\bullet},\chi^{1}_{\bullet})}^{\alpha_{\bullet}{\dagger}}\times\mathcal{C}_{(r^{2}_{\bullet},\chi^{2}_{\bullet})}^{\alpha_{\bullet}}]. By applying Π\Pi to (5.14), we obtain the following identity in K^(Var)\widehat{K}(\mathrm{Var}):

(5.15) 𝕃mr0+χ0+(1g)r0[𝒞(r,χ)α]=\displaystyle\mathbb{L}^{mr_{0}+\chi_{0}+(1-g)r_{0}}[\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}}]=
(r1,χ1)+(r2,χ2)=(r,χ)μα(ri,χi)=μα(r,χ)𝕃χ((r2,χ2),(r1,χ1))[𝒞(r1,χ1)α×𝒞(r2,χ2)α].\displaystyle\sum_{\begin{subarray}{c}(r_{\bullet}^{1},\chi_{\bullet}^{1})+(r_{\bullet}^{2},\chi_{\bullet}^{2})=(r_{\bullet},\chi_{\bullet})\\ \mu_{\alpha_{\bullet}}(r_{\bullet}^{i},\chi_{\bullet}^{i})=\mu_{\alpha_{\bullet}}(r_{\bullet},\chi_{\bullet})\end{subarray}}\mathbb{L}^{-\chi((r^{2}_{\bullet},\chi^{2}_{\bullet}),(r^{1}_{\bullet},\chi^{1}_{\bullet}))}[\mathcal{C}_{(r^{1}_{\bullet},\chi^{1}_{\bullet})}^{\alpha_{\bullet}{\dagger}}\times\mathcal{C}_{(r^{2}_{\bullet},\chi^{2}_{\bullet})}^{\alpha_{\bullet}}].

The class [𝒞(r,χ)α][\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}}] lies in Λ^\widehat{\Lambda}, see [GPH13, Theorem B]. Therefore from (5.15) and using induction on r0++rkr_{0}+\cdots+r_{k}, we conclude that [𝒞(r,χ)α][\mathcal{C}_{(r_{\bullet},\chi_{\bullet})}^{\alpha_{\bullet}{\dagger}}] also lies in Λ^\widehat{\Lambda}. ∎

Corollary 5.8.

For l>2g2l>2g-2, the singular cohomology H((L),)H^{\ast}((\mathcal{M}^{L{\dagger}})^{\mathbb{C}^{\ast}},\mathbb{Z}) is torsion free.

Proof.

The corollary follows from the decomposition (5.11), Lemma 5.6, and Proposition 5.7. ∎

Consider a tuple (r,χ,w)(r,\chi,w) and let 𝕋L=𝕋L(r,χ)w\mathbb{T}^{L}=\mathbb{T}^{L}(r,\chi)_{w} be the corresponding quasi-BPS category. We have the following torsion freeness of its topological K-theory:

Proposition 5.9.

Suppose that l>2g2l>2g-2. Then the topological K-group Ktop(𝕋L)K_{\ast}^{\rm{top}}(\mathbb{T}^{L}) is torsion free.

Proof.

By [HL21, Theorem A.4], the variety L\mathcal{M}^{L{\dagger}} decomposes into the direct sum of the components of its \mathbb{C}^{\ast}-fixed loci up to Tate twist in Voevodsky’s triangulated category of mixed motives with integer coefficient. Therefore, from Corollary 5.8 and applying the Betti realization [Lec08], the singular cohomology H(L,)H^{\ast}(\mathcal{M}^{L{\dagger}},\mathbb{Z}) is also torsion free. Then Ktop(L)K_{\ast}^{\rm{top}}(\mathcal{M}^{L{\dagger}}) is torsion free, see [GS, Proposition 6.6]. Since 𝕋L\mathbb{T}^{L} is a semiorthogonal summand of Db(L)D^{b}(\mathcal{M}^{L{\dagger}}), we have that Ktop(𝕋L)K_{\ast}^{\rm{top}}(\mathbb{T}^{L}) is a direct summand of Ktop(L)K_{\ast}^{\rm{top}}(\mathcal{M}^{L{\dagger}}), hence it is torsion-free. ∎

Theorem 5.10.

Suppose that l>2g2l>2g-2. The functor Φ𝒫\Phi_{\mathcal{P}} in (4.12) induces an equivalence of topological K-theory spectra

(5.16) Ktop(𝕋L(r,w+1gsp)χgsp+1)Ktop(𝕋L(r,χ)w).\displaystyle K^{\rm{top}}(\mathbb{T}^{L}(r,w+1-g^{\rm{sp}})_{-\chi-g^{\rm{sp}}+1})\stackrel{{\scriptstyle\sim}}{{\to}}K^{\rm{top}}(\mathbb{T}^{L}(r,\chi)_{w}).
Proof.

By Proposition 4.8 and Proposition 5.9, the result follows from the argument of [GS, Theorem 3.10], which we explain below. We use the same notation in the proof of Proposition 4.8. Let 𝒫\mathcal{P}^{\prime} be an object

𝒫𝕋L(r,χ)wBL𝕋L(r,w+1gsp)χgsp+1\displaystyle\mathcal{P}^{\prime}\in\mathbb{T}^{L}(r,\chi)_{-w}\boxtimes_{B^{L}}\mathbb{T}^{L}(r,w+1-g^{\rm{sp}})_{-\chi-g^{\rm{sp}}+1}

which restricts to the kernel object of the inverse of the equivalence (4.9). Note that the above 𝒫\mathcal{P}^{\prime} exists by the argument of Lemma 4.7. Let

Φ𝒫:𝕋L𝕋L\displaystyle\Phi_{\mathcal{P}^{\prime}}\colon\mathbb{T}^{L}\to\mathbb{T}^{L^{\prime}}

be the induced functor. We have the following commutative diagram

The vertical arrows are injective by Proposition 5.9. The bottom arrows are given by taking the global sections of direct sum of shifts of perverse sheaves on BB, see the morphism (4.14). Therefore, from [GS, Lemma 3.3], the composition of bottom arrows of the above diagram is a unipotent endomorphism. Then the composition of top arrows Φ𝒫KΦ𝒫K\Phi_{\mathcal{P}^{\prime}}^{K}\circ\Phi_{\mathcal{P}}^{K} is a unipotent map of free abelian groups, which implies that Φ𝒫K\Phi_{\mathcal{P}}^{K} is injective. Applying the same argument to Φ𝒫KΦ𝒫K\Phi_{\mathcal{P}}^{K}\circ\Phi_{\mathcal{P}^{\prime}}^{K}, we obtain the surjectivity of Φ𝒫K\Phi_{\mathcal{P}}^{K}. Therefore Φ𝒫K\Phi_{\mathcal{P}}^{K} is an isomorphism. ∎

6. Topological K-theory of BPS categories: the case G=GL and L=ΩCL=\Omega_{C}

In this section, we prove part (1) of Theorem 1.1. We consider topological K-theory of (reduced) quasi-BPS categories for the usual Hitchin moduli spaces, i.e. for L=ΩCL=\Omega_{C}, and prove the expected symmetry for rational topological K-theories.

6.1. Quasi-BPS categories in the case of L=ΩCL=\Omega_{C}

As before, we write red=(r,χ)red\mathcal{M}^{\rm{red}}=\mathcal{M}(r,\chi)^{\rm{red}}, M=M(r,χ)M=M(r,\chi), ΩC(p)=ΩC(p)(r,χ)\mathcal{M}^{\Omega_{C}(p)}=\mathcal{M}^{\Omega_{C}(p)}(r,\chi), etc. Recall the reduced quasi-BPS category

𝕋red:=𝕋(r,χ)wredDb(red).\displaystyle\mathbb{T}^{\rm{red}}:=\mathbb{T}(r,\chi)_{w}^{\rm{red}}\subset D^{b}(\mathcal{M}^{\rm{red}}).

Note that, for a fixed pCp\in C, we have the following diagram, see Subsection 2.4:

(6.5)

Here, each vertical arrow is a good moduli space morphism, the function ff is given by

f(x,v)=tr(s0(x)v),xΩC(p),v𝒱0|x,\displaystyle f(x,v)=\mathrm{tr}(s_{0}(x)\circ v),\ x\in\mathcal{M}^{\Omega_{C}(p)},v\in\mathcal{V}_{0}|_{x},

where s0s_{0} is the section of 𝒱0\mathcal{V}_{0} as in (2.27). The critical locus Crit(f)\mathrm{Crit}(f) is isomorphic to the classical truncation of the (1)(-1)-shifted cotangent of red\mathcal{M}^{\rm{red}}, see [Tod24, Chapter 2]. Recall that 𝒱\mathcal{V} consists of

(6.6) (F,θ,u),θ:FFΩC(p),uEnd(F|p),\displaystyle(F,\theta,u),\ \theta\colon F\to F\otimes\Omega_{C}(p),\ u\in\mathrm{End}(F|_{p}),

where (F,θ)(F,\theta) is a semistable ΩC(p)\Omega_{C}(p)-Higgs bundle, and the subbundle 𝒱0𝒱\mathcal{V}_{0}\subset\mathcal{V} corresponds to (6.6) such that tr(u)=0\mathrm{tr}(u)=0. Let \mathbb{C}^{\ast} acts on fibers of 𝒱0ΩC(p)\mathcal{V}_{0}\to\mathcal{M}^{\Omega_{C}(p)} and

𝕋Db(𝒱0),𝕋Db(𝒱0)\displaystyle\mathbb{T}_{\mathbb{C}^{\ast}}^{\prime}\subset D^{b}_{\mathbb{C}^{\ast}}(\mathcal{V}_{0}),\ \mathbb{T}^{\prime}\subset D^{b}(\mathcal{V}_{0})

be the subcategory consisting of objects 𝒫\mathcal{P} such that, for all ν:B𝒱0\nu\colon B\mathbb{C}^{\ast}\to\mathcal{V}_{0}, the set of weights ν𝒫\nu^{\ast}\mathcal{P} satisfies the weight condition as in (2.26), i.e.

(6.7) wt(ν𝒫)[12wtdet(ν𝕃𝒱)>0,12wtdet(ν𝕃𝒱)>0]+wdwt(νδ).\displaystyle\mathrm{wt}(\nu^{\ast}\mathcal{P})\subset\left[-\frac{1}{2}\mathrm{wt}\det(\nu^{\ast}\mathbb{L}_{\mathcal{V}})^{>0},\frac{1}{2}\mathrm{wt}\det(\nu^{\ast}\mathbb{L}_{\mathcal{V}})^{>0}\right]+\frac{w}{d}\mathrm{wt}(\nu^{\ast}\delta).

By [PTd, Lemma 2.6, Corollary 3.15], there is a Koszul equivalence

(6.8) 𝕋redMFgr(𝕋,f).\displaystyle\mathbb{T}^{\rm{red}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathrm{MF}^{\rm{gr}}(\mathbb{T}^{\prime}_{\mathbb{C}^{\ast}},f).

We now define the JS stable pair version of the vector bundle 𝒱\mathcal{V}. For m0m\gg 0, define 𝒱\mathcal{V}^{{\dagger}} to be consisting of tuples

(6.9) (F,θ,u,s),θ:FFΩC(p),uEnd(F|p),s:𝒪C(m)F,\displaystyle(F,\theta,u,s),\ \theta\colon F\to F\otimes\Omega_{C}(p),\ u\in\mathrm{End}(F|_{p}),\ s\colon\mathcal{O}_{C}(-m)\to F,

where (F,θ,u)(F,\theta,u) is as in (6.6), and the tuple (6.9) satisfies the JS stability: for any (F,θ,u)(F^{\prime},\theta^{\prime},u^{\prime}) as in (6.6) with μ(F)=μ(F)\mu(F)=\mu(F^{\prime}) and a surjection η:(F,θ)(F,θ)\eta\colon(F,\theta)\twoheadrightarrow(F^{\prime},\theta^{\prime}) of ΩC(p)\Omega_{C}(p)-Higgs bundles which fits into a commutative diagram

we have ηs0\eta\circ s\neq 0. We also define 𝒱0𝒱\mathcal{V}_{0}^{{\dagger}}\subset\mathcal{V}^{{\dagger}} by the condition tr(u)=0\mathrm{tr}(u)=0.

Lemma 6.1.

The stacks 𝒱\mathcal{V}^{{\dagger}}, 𝒱0\mathcal{V}_{0}^{{\dagger}} are smooth algebraic spaces such that the compositions

(6.10) α:𝒱πJS𝒱π𝒱N,α:𝒱0πJS𝒱0π𝒱N0\displaystyle\alpha\colon\mathcal{V}^{{\dagger}}\stackrel{{\scriptstyle\pi_{\rm{JS}}}}{{\to}}\mathcal{V}\stackrel{{\scriptstyle\pi_{\mathcal{V}}}}{{\to}}N,\ \alpha\colon\mathcal{V}_{0}^{{\dagger}}\stackrel{{\scriptstyle\pi_{\rm{JS}}}}{{\to}}\mathcal{V}_{0}\stackrel{{\scriptstyle\pi_{\mathcal{V}}}}{{\to}}N_{0}

are proper morphisms. Here, the morphism πJS\pi_{\mathrm{JS}} sends (F,θ,u,s)(F,\theta,u,s) to (F,θ,u)(F,\theta,u).

Proof.

We prove the lemma only for 𝒱\mathcal{V}^{{\dagger}}. It is enough to prove the claim étale locally at any point in MΩC(p)M^{\Omega_{C}(p)}. Let yMΩC(p)y\in M^{\Omega_{C}(p)} be a closed point. By Lemma 2.2, we may assume that yy lies in the deepest stratum, corresponding to VE0V\otimes E_{0} where dimV=d\dim V=d and E0E_{0} is a stable ΩC(p)\Omega_{C}(p)-Higgs bundle with (rank(E0),χ(E0))=(r0,χ0)(\mathrm{rank}(E_{0}),\chi(E_{0}))=(r_{0},\chi_{0}). Using the étale local description of ΩC(p)MΩC(p)\mathcal{M}^{\Omega_{C}(p)}\to M^{\Omega_{C}(p)} as in Subsection 2.3, (also see the proof of [PTa, Proposition 3.23]), the composition

𝒱𝒱N\displaystyle\mathcal{V}^{{\dagger}}\to\mathcal{V}\to N

is étale locally on MΩC(p)M^{\Omega_{C}(p)} at yy isomorphic to

(6.11) (𝔤𝔩(V)(1+2r02g)Vχ(E0(m)))ss/GL(V)𝔤𝔩(V)(1+2r02g)//GL(V),\displaystyle(\mathfrak{gl}(V)^{\oplus(1+2r_{0}^{2}g)}\oplus V^{\otimes\chi(E_{0}(m))})^{\rm{ss}}/GL(V)\to\mathfrak{gl}(V)^{\oplus(1+2r_{0}^{2}g)}/\!\!/GL(V),

where the semistable locus is with respect to the determinant character of GL(V)GL(V). The left hand side is the moduli space of stable representations of a quiver with vertices {0,1}\{0,1\}, (1+2r02g)(1+2r_{0}^{2}g)-loops at 11, χ(E0(m))\chi(E_{0}(m))-arrows from 0 to 11, and with dimension vector (1,d)(1,d). The stable representations correspond to those generated by the images from the maps from 0 to 11, see [Tod24, Lemma 6.1.9]. The source of the map (6.11) is smooth because it consists of stable representations. By [HdlPn02, Theorem 4.1], the map (6.11) is also projective since the map

(𝔤𝔩(V)(1+2r02g)Vχ(E0(m)))/GL(V)𝔤𝔩(V)(1+2r02g)//GL(V)\displaystyle(\mathfrak{gl}(V)^{\oplus(1+2r_{0}^{2}g)}\oplus V^{\otimes\chi(E_{0}(m))})/GL(V)\to\mathfrak{gl}(V)^{\oplus(1+2r_{0}^{2}g)}/\!\!/GL(V)

is the good moduli space morphism. We therefore obtain the desired conclusion. ∎

Lemma 6.2.

The compositions

𝕋Db(𝒱0)πJSDb(𝒱0),𝕋Db(𝒱0)πJSDb(𝒱0)\displaystyle\mathbb{T}^{\prime}_{\mathbb{C}^{\ast}}\subset D_{\mathbb{C}^{\ast}}^{b}(\mathcal{V}_{0})\stackrel{{\scriptstyle\pi_{\rm{JS}}^{\ast}}}{{\to}}D_{\mathbb{C}^{\ast}}^{b}(\mathcal{V}_{0}^{{\dagger}}),\ \mathbb{T}^{\prime}\subset D^{b}(\mathcal{V}_{0})\stackrel{{\scriptstyle\pi_{\rm{JS}}^{\ast}}}{{\to}}D^{b}(\mathcal{V}_{0}^{{\dagger}})

are fully-faithful and admit right adjoints.

Proof.

Using the étale local description (6.11) in terms of the Ext-quiver, an argument similar to Theorem 2.12 applies. See also the proof of [PTe, Proposition 3.7]. ∎

6.2. Topological K-theory of /2\mathbb{Z}/2-graded BPS categories

Recall the equivalence (6.8). We introduce the /2\mathbb{Z}/2-graded version of quasi-BPS category by replacing the right hand side in (6.8) with the /2\mathbb{Z}/2-graded dg-category of matrix factorizations

𝕋red,/2:=MF(𝕋,f)MF(𝒱0,f).\displaystyle\mathbb{T}^{\rm{red},\mathbb{Z}/2}:=\mathrm{MF}(\mathbb{T}^{\prime},f)\subset\mathrm{MF}(\mathcal{V}_{0},f).

Let gg be the function on 𝒱0\mathcal{V}_{0}^{{\dagger}} defined by the following commutative diagram

By Lemma 6.2 and using [PT24, Proposition 2.5], the following composition functor is fully-faithful with right adjoint

(6.12) 𝕋red,/2MF(𝒱0,f)πJSMF(𝒱0,g).\displaystyle\mathbb{T}^{\rm{red},\mathbb{Z}/2}\subset\mathrm{MF}(\mathcal{V}_{0},f)\stackrel{{\scriptstyle\pi_{\rm{JS}}^{\ast}}}{{\to}}\mathrm{MF}(\mathcal{V}_{0}^{{\dagger}},g).
Lemma 6.3.

There is an equivalence

(6.13) 𝒦𝒱0top(MF(𝒱0,g))ϕginv[β±1].\displaystyle\mathcal{K}_{\mathcal{V}_{0}^{{\dagger}}}^{\rm{top}}(\mathrm{MF}(\mathcal{V}_{0}^{{\dagger}},g))_{\mathbb{Q}}\simeq\phi_{g}^{\rm{inv}}[\beta^{\pm 1}].

In the above, ϕginv\phi_{g}^{\rm{inv}} is the monodromy invariant vanishing cycle of ϕg:=ϕg(𝒱0)\phi_{g}:=\phi_{g}(\mathbb{Q}_{\mathcal{V}_{0}^{{\dagger}}}):

ϕginv:=Cone(ϕg1Tϕg)=ϕg[γ],\displaystyle\phi_{g}^{\rm{inv}}:=\mathrm{Cone}(\phi_{g}\stackrel{{\scriptstyle 1-T}}{{\to}}\phi_{g})=\phi_{g}\otimes\mathbb{Q}[\gamma],

where [γ]=γ\mathbb{Q}[\gamma]=\mathbb{Q}\oplus\mathbb{Q}\gamma with degγ=1\deg\gamma=1 and TT is the monodromy operator, which in this case vanishes T=0T=0.

Proof.

The equivalence (6.13) is proved in [PTe, Lemma 6.5]. The monodromy operator TT vanishes because the \mathbb{C}^{\ast}-action on the fibers of 𝒱0ΩC(p)\mathcal{V}_{0}\to\mathcal{M}^{\Omega_{C}(p)} lifts to an action on 𝒱0\mathcal{V}_{0}^{{\dagger}}, and gg is of weight one with respect to the above \mathbb{C}^{\ast}-action. ∎

Remark 6.4.

A reason of considering /2\mathbb{Z}/2-periodic version is that the graded category MFgr(𝒱0,g)\mathrm{MF}^{\rm{gr}}(\mathcal{V}_{0}^{{\dagger}},g) is not linear over Perf(𝒱0)\mathrm{Perf}(\mathcal{V}_{0}^{{\dagger}}), rather it is linear over Perfgr(𝒱0)\mathrm{Perf}^{\rm{gr}}(\mathcal{V}_{0}^{{\dagger}}). Because of a lack of reference of relative topological K-theory linear over categories of graded perfect complexes, we use the /2\mathbb{Z}/2-version which is linear over Perf(𝒱0)\mathrm{Perf}(\mathcal{V}_{0}^{{\dagger}}).

For a variety MM, we say an object PD(Sh(M))P\in D(\mathrm{Sh}_{\mathbb{Q}}(M)) is perverse-split if PP is isomorphic to a direct sum iAi[i]\oplus_{i\in\mathbb{Z}}A_{i}[-i] with AiA_{i} is a perverse sheaf on MM.

Lemma 6.5.

The object

(6.14) 𝒦N0top(α𝕋red,/2)D(Sh(N0))\displaystyle\mathcal{K}_{N_{0}}^{\rm{top}}(\alpha_{\ast}\mathbb{T}^{\rm{red},\mathbb{Z}/2})_{\mathbb{Q}}\in D(\mathrm{Sh}_{\mathbb{Q}}(N_{0}))

is perverse-split. In the above, α\alpha is the morphism (6.10). Moreover, the object (6.14) contains ϕfN0(ICN0)[β±1,γ]\phi_{f_{N_{0}}}(\mathrm{IC}_{N_{0}})[\beta^{\pm 1},\gamma] as a direct summand.

Proof.

Since α\alpha is projective, by Theorem 3.2 and Lemma 6.3 we have

(6.15) 𝒦N0top(αMF(𝒱0,g))α𝒦𝒱0top(MF(𝒱0,g))αϕginv[β±1].\displaystyle\mathcal{K}_{N_{0}}^{\rm{top}}(\alpha_{\ast}\mathrm{MF}(\mathcal{V}_{0}^{{\dagger}},g))_{\mathbb{Q}}\cong\alpha_{\ast}\mathcal{K}_{\mathcal{V}_{0}^{{\dagger}}}^{\rm{top}}(\mathrm{MF}(\mathcal{V}_{0}^{{\dagger}},g))_{\mathbb{Q}}\cong\alpha_{\ast}\phi_{g}^{\rm{inv}}[\beta^{\pm 1}].

We have

αϕginvϕfN0(α𝒱0)[γ]\displaystyle\alpha_{\ast}\phi_{g}^{\rm{inv}}\cong\phi_{f_{N_{0}}}(\alpha_{\ast}\mathbb{Q}_{\mathcal{V}_{0}^{{\dagger}}})\otimes_{\mathbb{Q}}\mathbb{Q}[\gamma]

which is perverse-split by the BBDG decomposition theorem and the fact that the vanishing cycle functor ϕfN0\phi_{f_{N_{0}}} preserves the perverse t-structure. Therefore (6.15) is perverse-split. Since (6.12) is a part of a semiorthogonal decomposition, the object (6.14) is a direct summand of (6.15). Therefore (6.14) is perverse-split. The second statement follows as in the proof of Proposition 4.1, noting that 𝒱0N0\mathcal{V}_{0}^{{\dagger}}\to N_{0} is a projective bundle over a dense open smooth subset N0stN0N_{0}^{\rm{st}}\subset N_{0} (e.g. we can take N0stN_{0}^{\rm{st}} to be the preimage of the stable part in MΩC(p)M^{\Omega_{C}(p)} under the morphism N0MΩC(p)N_{0}\to M^{\Omega_{C}(p)} in (6.5)), and the category 𝕋red,/2\mathbb{T}^{\rm{red},\mathbb{Z}/2} restricted to N0stN_{0}^{\rm{st}} is equivalent to the category of matrix factorization of fN0:Nst0f_{N_{0}}\colon N^{\rm{st}}_{0}\to\mathbb{C}, possibly twisted by some Brauer class. ∎

Let 0:ΩC(p)𝒱00\colon\mathcal{M}^{\Omega_{C}(p)}\to\mathcal{V}_{0} be the zero section. It induces the morphism on good moduli spaces

0¯:MΩC(p)N0\displaystyle\overline{0}\colon M^{\Omega_{C}(p)}\to N_{0}

which is a section of the projection N0MΩC(p)N_{0}\to M^{\Omega_{C}(p)}. In particular, 0¯\overline{0} is a closed immersion.

Lemma 6.6.

Assume that the tuple (r,χ,w)(r,\chi,w) is primitive. Then there is a closed subscheme ZN0Z\subset N_{0} whose support is the image of 0¯\overline{0} such that the Perf(N0)\mathrm{Perf}(N_{0})-linear structure on 𝕋red,/2\mathbb{T}^{\rm{red},\mathbb{Z}/2} descends to the Perf(Z)\mathrm{Perf}(Z)-linear structure via the restriction functor Perf(N0)Perf(Z)\mathrm{Perf}(N_{0})\to\mathrm{Perf}(Z).

Proof.

As in [PTc, Theorem 6.4] (the result in loc.cit. is stated for moduli of sheaves on K3 surfaces, but it applies ad litteram to the local Calabi-Yau surface TotC(ΩC)\mathrm{Tot}_{C}(\Omega_{C})), the graded version 𝕋red\mathbb{T}^{\rm{red}} admits a strong generator \mathcal{E}. Indeed, the dg-category 𝕋red\mathbb{T}^{\rm{red}} is smooth, so \mathcal{E} is constructed by taking the direct sum of second factors of objects in (𝕋red)op𝕋red(\mathbb{T}^{\rm{red}})^{\rm{op}}\boxtimes\mathbb{T}^{\rm{red}} which generates the diagonal (𝕋red)op𝕋red(\mathbb{T}^{\rm{red}})^{\rm{op}}\boxtimes\mathbb{T}^{\rm{red}}-module. There is a natural morphism in Db(N0)D^{b}(N_{0}):

(6.16) 𝒪N0π𝒱Rom(,).\displaystyle\mathcal{O}_{N_{0}}\to\pi_{\mathcal{V}\ast}R\mathcal{H}om(\mathcal{E},\mathcal{E}).

By categorical support lemma [PTc, Theorem 6.6] (stated for K3 surfaces, but the argument applies ad litteram to Higgs bundles as well), [PTd, Lemma 5.4], any object in 𝕋red\mathbb{T}^{\rm{red}} is supported over 𝒱0×N0Im(0¯)\mathcal{V}_{0}\times_{N_{0}}\mathrm{Im}(\overline{0}), where 𝒱0N0\mathcal{V}_{0}\to N_{0} is the good moduli space morphism. Therefore the right hand side in (6.16) is supported on the image of 0¯\overline{0}. It follows that there is a closed subscheme i:ZN0i\colon Z\hookrightarrow N_{0} with support the image of 0¯\overline{0} such that the right hand side in (6.16) lies in the image of i:Db(Z)Db(N0)i_{\ast}\colon D^{b}(Z)\to D^{b}(N_{0}). Then the morphism (6.16) factors through 𝒪N0𝒪Z\mathcal{O}_{N_{0}}\twoheadrightarrow\mathcal{O}_{Z}.

Let /2𝕋red,/2\mathcal{E}^{\mathbb{Z}/2}\in\mathbb{T}^{\rm{red},\mathbb{Z}/2} be the object by forgetting the grading of \mathcal{E}. By the above argument, there is a morphism:

(6.17) 𝒪Zπ𝒱Rom(/2,/2)=:N0,\displaystyle\mathcal{O}_{Z}\to\pi_{\mathcal{V}\ast}R\mathcal{H}om(\mathcal{E}^{\mathbb{Z}/2},\mathcal{E}^{\mathbb{Z}/2})=:\mathscr{B}_{N_{0}},

which is a morphism of sheaves of /2\mathbb{Z}/2-graded dg-algebras over N0N_{0}. By the above construction of \mathcal{E}, the object /2\mathcal{E}^{\mathbb{Z}/2} is also a strong generator of 𝕋red,/2\mathbb{T}^{\rm{red},\mathbb{Z}/2}. Therefore, we have the fully-faithful functor

(6.18) 𝕋red,/2D(N0),()π𝒱Rom(/2,),\displaystyle\mathbb{T}^{\rm{red},\mathbb{Z}/2}\hookrightarrow D(\mathscr{B}_{N_{0}}),(-)\mapsto\pi_{\mathcal{V}\ast}R\mathcal{H}om(\mathcal{E}^{\mathbb{Z}/2},-),

where the right hand side is the derived category of /2\mathbb{Z}/2-graded dg-modules over N0\mathscr{B}_{N_{0}}. The above functor is Perf(N0)\mathrm{Perf}(N_{0})-linear, and the Perf(N0)\mathrm{Perf}(N_{0})-linear structure on D(N0)D(\mathscr{B}_{N_{0}}) descends to the Perf(Z)\mathrm{Perf}(Z)-linear structure on it where the action of Perf(Z)\mathrm{Perf}(Z) is given by the tensor product over 𝒪Z\mathcal{O}_{Z} through the morphism (6.17). Since the image of the pull-back functor Perf(N0)Perf(Z)\mathrm{Perf}(N_{0})\to\mathrm{Perf}(Z) generates Perf(Z)\mathrm{Perf}(Z), the above Perf(Z)\mathrm{Perf}(Z)-linear structure on D(N0)D(\mathscr{B}_{N_{0}}) restricts to the Perf(Z)\mathrm{Perf}(Z)-linear structure on 𝕋red,/2\mathbb{T}^{\rm{red},\mathbb{Z}/2} under the embedding (6.18), i.e. 𝕋red,/2Perf(Z)𝕋red,/2\mathbb{T}^{\rm{red},\mathbb{Z}/2}\otimes\mathrm{Perf}(Z)\subset\mathbb{T}^{\rm{red},\mathbb{Z}/2}. ∎

6.3. BPS sheaves

Let XX be the non-compact Calabi-Yau 3-fold:

X=TotC(𝒪CΩC)pC,\displaystyle X=\mathrm{Tot}_{C}(\mathcal{O}_{C}\oplus\Omega_{C})\stackrel{{\scriptstyle p}}{{\to}}C,

where pp is the natural projection. Let X\mathcal{M}_{X} be the moduli stack of compactly supported coherent sheaves EE on XX with rank(pE)=r\mathrm{rank}(p_{\ast}E)=r and χ(pE)=χ\chi(p_{\ast}E)=\chi. Note that X\mathcal{M}_{X} consists of tuples

(6.19) (F,θ,ι),\displaystyle(F,\theta,\iota),

where (F,θ)(F,\theta) is a Higgs bundle on CC and ι\iota is an endomorphism of (F,θ)(F,\theta). Consider the closed subscheme

XredX\displaystyle\mathcal{M}_{X}^{\rm{red}}\subset\mathcal{M}_{X}

consisting of tuples (6.19) such that the trace of ι\iota is zero. Then there are isomorphisms

Ω[1]clX,Ωred[1]clXred,\displaystyle\Omega_{\mathcal{M}}[-1]^{\rm{cl}}\cong\mathcal{M}_{X},\ \Omega_{\mathcal{M}^{\rm{red}}}[-1]^{\rm{cl}}\cong\mathcal{M}_{X}^{\rm{red}},

where Ω[1]\Omega_{\mathcal{M}}[-1] is the (1)(-1)-shifted cotangent of \mathcal{M}, see [Tod24] for the above isomorphisms. Consider the following diagram

such that Xred=Crit(f)\mathcal{M}_{X}^{\rm{red}}=\mathrm{Crit}(f), where recall that =(r,χ)\mathcal{M}=\mathcal{M}(r,\chi) and M=M(r,χ)M=M(r,\chi). Let

ϕPerv(Xred)\displaystyle\phi_{\mathcal{M}}\in\mathrm{Perv}(\mathcal{M}_{X}^{\rm{red}})

be the DT perverse sheaf [BBBBJ15] on Xred\mathcal{M}_{X}^{\rm{red}} with respect to the orientation data canonically defined as a (1)(-1)-shifted cotangent, see [Tod24, Section 3.3.3]. The orientation data determined by the embedding Xred𝒱0\mathcal{M}_{X}^{\rm{red}}\hookrightarrow\mathcal{V}_{0} matches with the above one, so ϕ\phi_{\mathcal{M}} is isomorphic to ϕf(IC𝒱0)\phi_{f}(\mathrm{IC}_{\mathcal{V}_{0}}), see [KK, Proposition 2.4].

The BPS sheaf on MXredM_{X}^{\rm{red}} is defined by

𝒫𝒮MXred:=p1(πXϕ)Perv(MXred),\displaystyle\mathcal{BPS}_{M_{X}^{\rm{red}}}:={}^{p}\mathcal{H}^{1}(\pi_{X\ast}\phi_{\mathcal{M}})\in\mathrm{Perv}(M_{X}^{\rm{red}}),

where p1(){}^{p}\mathcal{H}^{1}(-) is the first cohomology with respect to the perverse t-structure. By the support lemma [KK, Proposition 5.1], the sheaf 𝒫𝒮MXred\mathcal{BPS}_{M_{X}^{\rm{red}}} is supported on the image of the map 0¯M:MMXred\overline{0}_{M}\colon M\to M_{X}^{\rm{red}} induced by the zero section 0:redXred0\colon\mathcal{M}^{\rm{red}}\to\mathcal{M}_{X}^{\rm{red}}. The BPS sheaf on MM is given by, see [KK, Proposition 5.12]:

𝒫𝒮M:=p¯𝒫𝒮MXredPerv(M).\displaystyle\mathcal{BPS}_{M}:=\overline{p}_{\ast}\mathcal{BPS}_{M_{X}^{\rm{red}}}\in\mathrm{Perv}(M).
Remark 6.7.

In [KK, Proposition 5.11], the support lemma is given for the unreduced moduli space MXM_{X}. The reduced version removes the extra 𝔸1\mathbb{A}^{1}-factor in loc. cit.

The following lemma is proved similarly to [KK, Proposition 3.10].

Lemma 6.8.

There is an isomorphism

𝒫𝒮MXredϕfN0(ICN0).\displaystyle\mathcal{BPS}_{M_{X}^{\rm{red}}}\cong\phi_{f_{N_{0}}}(\mathrm{IC}_{N_{0}}).

Let (r,χ,w)(r,\chi,w) satisfy the BPS condition and 𝕋red=𝕋(r,χ)wred\mathbb{T}^{\rm{red}}=\mathbb{T}(r,\chi)_{w}^{\rm{red}} be the reduced BPS category. We have the following relation of its topological K-theory with BPS sheaves:

Theorem 6.9.

There is an isomorphism

(6.20) 𝒦Mtop(𝕋red)𝒫𝒮M[β±1].\displaystyle\mathcal{K}_{M}^{\rm{top}}(\mathbb{T}^{\rm{red}})_{\mathbb{Q}}\cong\mathcal{BPS}_{M}[\beta^{\pm 1}].
Proof.

Let i:ZN0i\colon Z\hookrightarrow N_{0} be a closed subscheme as in Lemma 6.6. By Theorem 3.2, we have

(6.21) 𝒦N0top(i𝕋red,/2)i𝒦Ztop(𝕋red,/2)\displaystyle\mathcal{K}_{N_{0}}^{\rm{top}}(i_{\ast}\mathbb{T}^{\rm{red},\mathbb{Z}/2})_{\mathbb{Q}}\cong i_{\ast}\mathcal{K}_{Z}^{\rm{top}}(\mathbb{T}^{\rm{red},\mathbb{Z}/2})_{\mathbb{Q}}

which is perverse-split by Lemma 6.5. Since the composition ZN0MZ\hookrightarrow N_{0}\to M is proper (indeed, a homeomorphism), we conclude that 𝒦Mtop(𝕋red,/2)\mathcal{K}_{M}^{\rm{top}}(\mathbb{T}^{\rm{red},\mathbb{Z}/2})_{\mathbb{Q}} is also perverse-split. We have the following relation

(6.22) 𝒦Mtop(𝕋red,/2)𝒦Mtop(𝕋red)[γ]\displaystyle\mathcal{K}_{M}^{\rm{top}}(\mathbb{T}^{\rm{red},\mathbb{Z}/2})_{\mathbb{Q}}\cong\mathcal{K}_{M}^{\rm{top}}(\mathbb{T}^{\rm{red}})_{\mathbb{Q}}\otimes\mathbb{Q}[\gamma]

see [PTe, Proposition 7.5] (the statement in loc. cit. is for the absolute version, but the argument for the relative version is the same). It follows that the left hand side of (6.20) is perverse-split. Moreover by the second statement of Lemma 6.5 and Lemma 6.8, it contains 𝒫𝒮M[β±1]\mathcal{BPS}_{M}[\beta^{\pm 1}] as a direct summand.

It follows that we can write

𝒦Mtop(𝕋red)=𝒫𝒮M[β±1]P0[β±1]P1[1][β±1]\displaystyle\mathcal{K}_{M}^{\rm{top}}(\mathbb{T}^{\rm{red}})_{\mathbb{Q}}=\mathcal{BPS}_{M}[\beta^{\pm 1}]\oplus P_{0}[\beta^{\pm 1}]\oplus P_{1}[1][\beta^{\pm 1}]

for some PiPerv(M)P_{i}\in\mathrm{Perv}(M). It is enough to show that P0=P1=0P_{0}=P_{1}=0.

The argument of the vanishing Pi=0P_{i}=0 is the same as in the last part of the proof of Proposition 4.1. Namely for a closed point yMy\in M, let 𝒫Qy(𝒅)\mathscr{P}_{Q_{y}}(\bm{d}) and PQy(𝒅)P_{Q_{y}}(\bm{d}) be as in (2.14). Then from [PTe, Theorem 8.26, Theorem 10.6], we see that

𝒦PQy(𝒅)(𝕋Qy(𝒅)wred)𝒫𝒮PQy(𝒅)[β±1]\displaystyle\mathcal{K}_{P_{Q_{y}}(\bm{d})}(\mathbb{T}_{Q_{y}}(\bm{d})_{w}^{\rm{red}})\cong\mathcal{BPS}_{P_{Q_{y}}(\bm{d})}[\beta^{\pm 1}]

where 𝕋Qy(𝒅)wredDb(𝒫(𝒅)red)w\mathbb{T}_{Q_{y}}(\bm{d})_{w}^{\rm{red}}\subset D^{b}(\mathscr{P}(\bm{d})^{\rm{red}})_{w} is the preprojective reduced quasi-BPS category, see [PTd, Definition 2.14]. Therefore Pi=0P_{i}=0 étale locally at each yMy\in M, hence Pi=0P_{i}=0. ∎

Lemma 6.10.

The natural map

(6.23) η:Ktop(𝕋red)Γ(𝒦topB(𝕋red))\displaystyle\eta\colon K^{\rm{top}}(\mathbb{T}^{\rm{red}})_{\mathbb{Q}}\to\Gamma(\mathcal{K}^{\rm{top}}_{B}(\mathbb{T}^{\rm{red}})_{\mathbb{Q}})

given in Lemma 3.4 is an isomorphism.

Proof.

By Lemma 6.6 and by Theorem 3.2 for proper morphisms

N0ZMB,\displaystyle N_{0}\hookleftarrow Z\to M\to B,

we have that

Γ(𝒦N0top(𝕋red,/2))=Γ(𝒦Btop(𝕋red,/2)).\displaystyle\Gamma(\mathcal{K}_{N_{0}}^{\rm{top}}(\mathbb{T}^{\rm{red},\mathbb{Z}/2})_{\mathbb{Q}})=\Gamma(\mathcal{K}_{B}^{\rm{top}}(\mathbb{T}^{\rm{red},\mathbb{Z}/2})_{\mathbb{Q}}).

By applying Proposition 3.6 for the proper morphism 𝒱0N0\mathcal{V}_{0}^{{\dagger}}\to N_{0} and using the fact that 𝕋red,/2\mathbb{T}^{\rm{red},\mathbb{Z}/2} is a semiorthogonal summand of MF(𝒱0,g)\mathrm{MF}(\mathcal{V}_{0}^{{\dagger}},g), see (6.12), the natural map

Ktop(𝕋red,/2)Γ(𝒦topN0(𝕋red,/2))\displaystyle K^{\rm{top}}(\mathbb{T}^{\rm{red},\mathbb{Z}/2})_{\mathbb{Q}}\to\Gamma(\mathcal{K}^{\rm{top}}_{N_{0}}(\mathbb{T}^{\rm{red},\mathbb{Z}/2})_{\mathbb{Q}})

is an isomorphism. Noting (6.22) and also its absolute version in [PTe, Proposition 7.5], it follows that

η1[γ]:Ktop(𝕋red)[γ]Γ(KBtop(𝕋red))[γ]\displaystyle\eta\otimes 1_{\mathbb{Q}[\gamma]}\colon K^{\rm{top}}(\mathbb{T}^{\rm{red}})_{\mathbb{Q}}\otimes\mathbb{Q}[\gamma]\to\Gamma(K_{B}^{\rm{top}}(\mathbb{T}^{\rm{red}})_{\mathbb{Q}})\otimes\mathbb{Q}[\gamma]

is an isomorphism, where η\eta is the morphism (6.23). Therefore η\eta is an isomorphism. ∎

6.4. Duality of rational topological K-theories

In this subsection, we prove part (1) of Theorem 1.1.

Theorem 6.11.

There is an equivalence

(6.24) Ktop(𝕋(r,w)redχ)Ktop(𝕋(r,χ)wred).\displaystyle K^{\rm{top}}(\mathbb{T}(r,w)^{\rm{red}}_{-\chi})_{\mathbb{Q}}\stackrel{{\scriptstyle\sim}}{{\to}}K^{\rm{top}}(\mathbb{T}(r,\chi)_{w}^{\rm{red}})_{\mathbb{Q}}.
Proof.

Let LL be a line bundle on CC such that degL0\deg L\gg 0 is even and admits a surjection

(6.25) 𝒪CΩCL.\displaystyle\mathcal{O}_{C}\oplus\Omega_{C}\twoheadrightarrow L.

We write L=L(r,χ)\mathcal{M}^{L}=\mathcal{M}^{L}(r,\chi), L=L(r,w+1gsp)\mathcal{M}^{\prime L}=\mathcal{M}^{L}(r,w+1-g^{\rm{sp}}) etc. By [KM24, Theorem 5.6, Proposition 5.7], [KK, Section 3.3], there is a commutative diagram

(6.32)

such that Xred=Crit(w¯)\mathcal{M}_{X}^{\rm{red}}=\mathrm{Crit}(\overline{w}), and the embedding XredL\mathcal{M}_{X}^{\rm{red}}\hookrightarrow\mathcal{M}^{L} is induced by (6.25). Here hX:XredBXh_{X}\colon\mathcal{M}_{X}^{\rm{red}}\to B_{X} is a Hitchin-type map, where BXB_{X} is given by

BX=i=1rH0(C,Symi(𝒪CΩC)).\displaystyle B_{X}=\bigoplus_{i=1}^{r}H^{0}(C,\mathrm{Sym}^{i}(\mathcal{O}_{C}\oplus\Omega_{C})).

The map hXh_{X} sends a compactly supported coherent sheaf on XX to its support, see [KK, Section 2.4]. By [KK, Proposition 3.10], there is an isomorphism

𝒫𝒮Mp¯ϕw¯(ICML).\displaystyle\mathcal{BPS}_{M}\cong\overline{p}_{\ast}\phi_{\overline{w}}(\mathrm{IC}_{M^{L}}).

Therefore using Proposition 4.1, Theorem 6.9, and Theorem 3.2, there are isomorphisms (cf. Remark 6.12):

(6.33) pBϕwB(𝒦BLtop(hL𝕋L))\displaystyle p_{B\ast}\phi_{w_{B}}(\mathcal{K}_{B^{L}}^{\rm{top}}(h^{L}_{\ast}\mathbb{T}^{L}))_{\mathbb{Q}} pBϕwB(hLICML[dimML][β±1])\displaystyle\cong p_{B\ast}\phi_{w_{B}}(h^{L}_{\ast}\mathrm{IC}_{M^{L}}[-\dim M^{L}][\beta^{\pm 1}])
pBhXϕw¯(ICML)[β±1][dimML]\displaystyle\cong p_{B\ast}h_{X\ast}\phi_{\overline{w}}(\mathrm{IC}_{M^{L}})[\beta^{\pm 1}][-\dim M^{L}]
hp¯ϕw¯(ICML)[β±1][dimML]\displaystyle\cong h_{\ast}\overline{p}_{\ast}\phi_{\overline{w}}(\mathrm{IC}_{M^{L}})[\beta^{\pm 1}][-\dim M^{L}]
h𝒫𝒮M[β±1][dimML]\displaystyle\cong h_{\ast}\mathcal{BPS}_{M}[\beta^{\pm 1}][-\dim M^{L}]
h𝒦Mtop(𝕋red)[dimML]\displaystyle\cong h_{\ast}\mathcal{K}_{M}^{\rm{top}}(\mathbb{T}^{\rm{red}})_{\mathbb{Q}}[-\dim M^{L}]
𝒦Btop(𝕋red)[dimML].\displaystyle\cong\mathcal{K}_{B}^{\rm{top}}(\mathbb{T}^{\rm{red}})_{\mathbb{Q}}[-\dim M^{L}].

Applying pBϕwBp_{B\ast}\phi_{w_{B}} to the isomorphism (4.14), we obtain the isomorphism

𝒦Btop(𝕋(r,w+1gsp)redχgsp+1)𝒦Btop(𝕋(r,χ)wred).\displaystyle\mathcal{K}_{B}^{\rm{top}}(\mathbb{T}(r,w+1-g^{\rm{sp}})^{\rm{red}}_{-\chi-g^{\rm{sp}}+1})_{\mathbb{Q}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathcal{K}_{B}^{\rm{top}}(\mathbb{T}(r,\chi)_{w}^{\rm{red}})_{\mathbb{Q}}.

The isomorphism (6.24) follows by taking the global section of the above isomorphism (see Lemma 6.10) and the following equivalence from [PTa, Lemma 3.8]

𝕋(r,w+1gsp)χgsp+1red𝕋(r,w)χred.\displaystyle\mathbb{T}(r,w+1-g^{\rm{sp}})_{-\chi-g^{\rm{sp}}+1}^{\rm{red}}\simeq\mathbb{T}(r,w)_{-\chi}^{\rm{red}}.

Remark 6.12.

By [KK, Remark 2.5], the map BXBLB_{X}\to B^{L} restricted to Im(hX)\mathrm{Im}(h_{X}) is injective. The notation pBϕwB()p_{B\ast}\phi_{w_{B}}(-) means that ϕwB()\phi_{w_{B}}(-) lies in the image from Im(hX)\mathrm{Im}(h_{X}), then apply pBp_{B\ast} by regarding ϕwB()\phi_{w_{B}}(-) as a sheaf on BXB_{X}. The same notation will also appear in the later sections.

7. Review of quasi-BPS categories for SL/PGL Higgs bundles

In this section, we recall the SL/PGL Higgs moduli spaces and the definition of their quasi-BPS categories from [PTa].

7.1. SL-Higgs moduli spaces

Let CC be a smooth projective curve of genus gg, and let LL be a line bundle on CC such that l=degL>2g2l=\deg L>2g-2 or L=ΩCL=\Omega_{C}. For each decomposition r=r1++rkr=r_{1}+\cdots+r_{k}, we set

SL(r):=Ker(i=1kGL(ri)det),(gi)1ikdeti=1kdetgi.\displaystyle\mathrm{SL}(r_{\bullet}):=\operatorname{Ker}\left(\prod_{i=1}^{k}\operatorname{GL}(r_{i})\stackrel{{\scriptstyle\det}}{{\to}}\mathbb{C}^{\ast}\right),\ (g_{i})_{1\leqslant i\leqslant k}\stackrel{{\scriptstyle\det}}{{\mapsto}}\prod_{i=1}^{k}\det g_{i}.

For a tuple of integers χ=(χ1,,χk)\chi_{\bullet}=(\chi_{1},\ldots,\chi_{k}), the moduli stack of (L,χ)(L,\chi_{\bullet})-twisted SL(r)\mathrm{SL}(r_{\bullet})-Higgs bundle is given as follows. Consider the closed substack

L(r,χ)tr=0i=1kL(ri,χi)\displaystyle\mathcal{M}^{L}(r_{\bullet},\chi_{\bullet})^{\rm{tr}=0}\subset\prod_{i=1}^{k}\mathcal{M}^{L}(r_{i},\chi_{i})

given by the derived zero locus of

tr:i=1kL(ri,χi)H0(L)der,{(Fi,θi)}1iki=1ktr(θi).\displaystyle\mathrm{tr}\colon\prod_{i=1}^{k}\mathcal{M}^{L}(r_{i},\chi_{i})\to H^{0}(L)^{\rm{der}},\ \{(F_{i},\theta_{i})\}_{1\leqslant i\leqslant k}\to\sum_{i=1}^{k}\mathrm{tr}(\theta_{i}).

Here H0(L)derH^{0}(L)^{\rm{der}} is the derived space of global sections:

H0(L)der=SpecSym(RΓ(L))={H0(L),l>2g2H0(ωC)×Spec[ϵ],L=ΩC,\displaystyle H^{0}(L)^{\rm{der}}=\operatorname{Spec}\mathrm{Sym}(R\Gamma(L)^{\vee})=\begin{cases}H^{0}(L),&l>2g-2\\ H^{0}(\omega_{C})\times\operatorname{Spec}\mathbb{C}[\epsilon],&L=\Omega_{C},\end{cases}

where degϵ=1\deg\epsilon=-1. Consider the map

det:L(r,χ)tr=0𝒫ic(C),{(Fi,θi)}1iki=1kdetFi.\displaystyle\det\colon\mathcal{M}^{L}(r_{\bullet},\chi_{\bullet})^{\rm{tr}=0}\to\mathcal{P}ic(C),\ \{(F_{i},\theta_{i})\}_{1\leqslant i\leqslant k}\mapsto\otimes_{i=1}^{k}\det F_{i}.

Here 𝒫ic(C)\mathcal{P}ic(C) is the Picard stack of line bundles on CC, which is a trivial \mathbb{C}^{\ast}-gerbe

𝒫ic(C)=Pic(C)/.\displaystyle\mathcal{P}ic(C)=\mathrm{Pic}(C)/\mathbb{C}^{\ast}.

We fix a line bundle AA on CC of degree χ+r(g1)\chi+r(g-1). The (L,χ)(L,\chi_{\bullet})-twisted SL(r)\mathrm{SL}(r_{\bullet})-Higgs moduli stack is given by the Cartesian square

where the bottom arrow corresponds to AA. The stack SL(r)L(χ)\mathcal{M}_{\mathrm{SL}(r_{\bullet})}^{L}(\chi_{\bullet}) is smooth for l>2g2l>2g-2 and it is quasi-smooth for L=ΩCL=\Omega_{C}. In particular, for k=1k=1, we obtain the (L,χ)(L,\chi)-twisted SL(r)\mathrm{SL}(r)-Higgs moduli stack LSL(r)(χ)\mathcal{M}^{L}_{\mathrm{SL}(r)}(\chi).

Note that the center of SL(r)\mathrm{SL}(r_{\bullet}) is

Z(SL(r))=Ker(()k),(ti)1iki=1ktiri,\displaystyle Z(\mathrm{SL}(r_{\bullet}))=\operatorname{Ker}\left((\mathbb{C}^{\ast})^{k}\to\mathbb{C}^{\ast}\right),(t_{i})_{1\leqslant i\leqslant k}\mapsto\prod_{i=1}^{k}t_{i}^{r_{i}},

and we have

Hom(Z(SL(r)),)=k/(r1,,rk).\displaystyle\operatorname{Hom}(Z(\mathrm{SL}(r_{\bullet})),\mathbb{C}^{\ast})=\mathbb{Z}^{\oplus k}/(r_{1},\ldots,r_{k})\mathbb{Z}.

There is a corresponding orthogonal decomposition

Db(SL(r)L(χ))=wk/(r1,,rk)Db(SL(r)L(χ))w\displaystyle D^{b}(\mathcal{M}_{\mathrm{SL}(r_{\bullet})}^{L}(\chi_{\bullet}))=\bigoplus_{w_{\bullet}\in\mathbb{Z}^{\oplus k}/(r_{1},\ldots,r_{k})\mathbb{Z}}D^{b}(\mathcal{M}_{\mathrm{SL}(r_{\bullet})}^{L}(\chi_{\bullet}))_{w_{\bullet}}

where for w=0w=0 each summand corresponds to the weight (w1,,wk)(w_{1},\ldots,w_{k})-component with respect to the action of Z(SL(r))Z(\mathrm{SL}(r_{\bullet})).

7.2. PGL-Higgs moduli spaces

For each decomposition r=r1++rkr=r_{1}+\cdots+r_{k}, we set

PGL(r):=(i=1kGL(ri))/.\displaystyle\mathrm{PGL}(r_{\bullet}):=\left(\prod_{i=1}^{k}\operatorname{GL}(r_{i})\right)/\mathbb{C}^{\ast}.

There is a natural action of 𝒫ic(C)\mathcal{P}ic(C) on the disjoint union of L(r,χ)tr=0\mathcal{M}^{L}(r_{\bullet},\chi_{\bullet})^{\rm{tr}=0} for χk\chi_{\bullet}\in\mathbb{Z}^{k} via the tensor product. The moduli stack of LL-twisted PGL(r)\mathrm{PGL}(r_{\bullet})-Higgs bundles is given by

LPGL(r)\displaystyle\mathcal{M}^{L}_{\mathrm{PGL}(r_{\bullet})} :=(χkL(r,χ)tr=0)/𝒫ic(C)\displaystyle:=\left(\coprod_{\chi_{\bullet}\in\mathbb{Z}^{k}}\mathcal{M}^{L}(r_{\bullet},\chi_{\bullet})^{\rm{tr}=0}\right)/\mathcal{P}ic(C)
=χk/(r1,,rk)LPGL(r)(χ),\displaystyle=\coprod_{\chi_{\bullet}\in\mathbb{Z}^{\oplus k}/(r_{1},\ldots,r_{k})\mathbb{Z}}\mathcal{M}^{L}_{\mathrm{PGL}(r_{\bullet})}(\chi_{\bullet}),

where each component is given by

(7.1) LPGL(r)(χ)=L(r,χ)tr=0/𝒫ic0(C).\displaystyle\mathcal{M}^{L}_{\mathrm{PGL}(r_{\bullet})}(\chi_{\bullet})=\mathcal{M}^{L}(r_{\bullet},\chi_{\bullet})^{\rm{tr}=0}/\mathcal{P}ic^{0}(C).

In particular, for k=1k=1, we obtain the LL-twisted PGL(r)\mathrm{PGL}(r)-Higgs moduli stack PGL(r)L(χ)\mathcal{M}_{\mathrm{PGL}(r)}^{L}(\chi) for χ/r\chi\in\mathbb{Z}/r\mathbb{Z}.

By taking the quotient by Pic0(C)\mathrm{Pic}^{0}(C) in (7.1) instead of the quotient by 𝒫ic0(C)\mathcal{P}ic^{0}(C), we obtain the \mathbb{C}^{\ast}-gerbe

~LPGL(r)(χ)LPGL(r)(χ).\displaystyle\widetilde{\mathcal{M}}^{L}_{\mathrm{PGL}(r_{\bullet})}(\chi_{\bullet})\to\mathcal{M}^{L}_{\mathrm{PGL}(r_{\bullet})}(\chi_{\bullet}).

There is an orthogonal decomposition of the derived category into subcategories of fixed \mathbb{C}^{\ast}-weight:

(7.2) Db(~LPGL(r)(χ))=wDb(LPGL(r)(χ))w\displaystyle D^{b}(\widetilde{\mathcal{M}}^{L}_{\mathrm{PGL}(r_{\bullet})}(\chi_{\bullet}))=\bigoplus_{w\in\mathbb{Z}}D^{b}(\mathcal{M}^{L}_{\mathrm{PGL}(r_{\bullet})}(\chi_{\bullet}))_{w}

Note that the w=0w=0 component is equivalent to Db(LPGL(r)(χ))D^{b}(\mathcal{M}^{L}_{\mathrm{PGL}(r_{\bullet})}(\chi_{\bullet})).

The center of PGL(r)\mathrm{PGL}(r_{\bullet}) is

Z(PGL(r))=()k/\displaystyle Z(\mathrm{PGL}(r_{\bullet}))=(\mathbb{C}^{\ast})^{k}/\mathbb{C}^{\ast}

and we have

Hom(Z(PGL(r)),)=Ker(k),(w1,,wk)i=1kwi.\displaystyle\operatorname{Hom}(Z(\mathrm{PGL}(r_{\bullet})),\mathbb{C}^{\ast})=\operatorname{Ker}\left(\mathbb{Z}^{\oplus k}\to\mathbb{Z}\right),\ (w_{1},\ldots,w_{k})\mapsto\sum_{i=1}^{k}w_{i}.

There is a corresponding orthogonal decomposition:

Db(LPGL(r)(χ))w=w1++wk=wDb(LPGL(r)(χ))w\displaystyle D^{b}(\mathcal{M}^{L}_{\mathrm{PGL}(r_{\bullet})}(\chi_{\bullet}))_{w}=\bigoplus_{w_{1}+\cdots+w_{k}=w}D^{b}(\mathcal{M}^{L}_{\mathrm{PGL}(r_{\bullet})}(\chi_{\bullet}))_{w_{\bullet}}

where each summand corresponds to the weight (w1,,wk)(w_{1},\ldots,w_{k})-component with respect to the action of Z(PGL(r))Z(\mathrm{PGL}(r_{\bullet})).

7.3. Semiorthogonal decompositions of SL/PGL-Higgs moduli stacks

We recall the SL/PGL versions of the semiorthogonal decomposition from Theorem 2.10:

Theorem 7.1.

([PTa, Theorem 7.2]) For each rkr_{\bullet}\in\mathbb{Z}^{k}, χk\chi_{\bullet}\in\mathbb{Z}^{k} and wk/(r1,,rk)w_{\bullet}\in\mathbb{Z}^{k}/(r_{1},\ldots,r_{k})\mathbb{Z}, there is a subcategory

𝕋LSL(r)(χ)wDb(LSL(r)(χ))w\displaystyle\mathbb{T}^{L}_{\mathrm{SL}(r_{\bullet})}(\chi_{\bullet})_{w_{\bullet}}\subset D^{b}(\mathcal{M}^{L}_{\mathrm{SL}(r_{\bullet})}(\chi_{\bullet}))_{w_{\bullet}}

such that there is a semiorthogonal decomposition

(7.3) Db(LSL(r)(χ))w=𝕋LSL(r)(χ)w|v1r1<<vkrk.\displaystyle D^{b}(\mathcal{M}^{L}_{\mathrm{SL}(r)}(\chi))_{w}=\left\langle\mathbb{T}^{L}_{\mathrm{SL}(r_{\bullet})}(\chi_{\bullet})_{w_{\bullet}}\,\Big{|}\,\frac{v_{1}}{r_{1}}<\cdots<\frac{v_{k}}{r_{k}}\right\rangle.

The right hand side is after all partitions r=r1++rkr=r_{1}+\cdots+r_{k}, χ=χ1++χk\chi=\chi_{1}+\cdots+\chi_{k} such that χi/ri=χ/r\chi_{i}/r_{i}=\chi/r, and wk/(r1,,rk)w_{\bullet}\in\mathbb{Z}^{k}/(r_{1},\ldots,r_{k})\mathbb{Z} with w1++wr=ww_{1}+\cdots+w_{r}=w in /r\mathbb{Z}/r\mathbb{Z}, and where vk/(r1,,rk)v_{\bullet}\in\mathbb{Q}^{\oplus k}/(r_{1},\ldots,r_{k})\mathbb{Q} is determined by

wi=vi+l2ri(i>jrji<jrj).\displaystyle w_{i}=v_{i}+\frac{l}{2}r_{i}\left(\sum_{i>j}r_{j}-\sum_{i<j}r_{j}\right).

The fully-faithful functor

𝕋LSL(r)(χ)wDb(LSL(r)(χ))w\displaystyle\mathbb{T}^{L}_{\mathrm{SL}(r_{\bullet})}(\chi_{\bullet})_{w_{\bullet}}\to D^{b}(\mathcal{M}^{L}_{\mathrm{SL}(r)}(\chi))_{w}

is induced by the categorical Hall product.

Theorem 7.2.

([PTa, Theorem 7.3]) For each rkr_{\bullet}\in\mathbb{Z}^{k}, χk/(r1,,rk)\chi_{\bullet}\in\mathbb{Z}^{k}/(r_{1},\ldots,r_{k})\mathbb{Z} and wkw_{\bullet}\in\mathbb{Z}^{k}, there is a subcategory

𝕋LPGL(r)(χ)wDb(LPGL(r)(χ))w\displaystyle\mathbb{T}^{L}_{\mathrm{PGL}(r_{\bullet})}(\chi_{\bullet})_{w_{\bullet}}\subset D^{b}(\mathcal{M}^{L}_{\mathrm{PGL}(r_{\bullet})}(\chi_{\bullet}))_{w_{\bullet}}

such that there is a semiorthogonal decomposition

(7.4) Db(LPGL(r)(χ))w=𝕋LPGL(r)(χ)w|v1r1<<vkrk.\displaystyle D^{b}(\mathcal{M}^{L}_{\mathrm{PGL}(r)}(\chi))_{w}=\left\langle\mathbb{T}^{L}_{\mathrm{PGL}(r_{\bullet})}(\chi_{\bullet})_{w_{\bullet}}\,\Big{|}\,\frac{v_{1}}{r_{1}}<\cdots<\frac{v_{k}}{r_{k}}\right\rangle.

The right hand side is after all partitions r=r1++rkr=r_{1}+\cdots+r_{k}, χ=χ1++χk\chi=\chi_{1}+\cdots+\chi_{k} such that χ1/r1==χk/rk\chi_{1}/r_{1}=\cdots=\chi_{k}/r_{k}and wkw_{\bullet}\in\mathbb{Z}^{\oplus k} with w1++wk=ww_{1}+\cdots+w_{k}=w, and vi12v_{i}\in\frac{1}{2}\mathbb{Z} is determined by

wi=vi+l2ri(i>jrji<jrj).\displaystyle w_{i}=v_{i}+\frac{l}{2}r_{i}\left(\sum_{i>j}r_{j}-\sum_{i<j}r_{j}\right).

The fully-faithful functor

𝕋LPGL(r)(χ)wDb(LPGL(r)(χ))w\displaystyle\mathbb{T}^{L}_{\mathrm{PGL}(r_{\bullet})}(\chi_{\bullet})_{w_{\bullet}}\to D^{b}(\mathcal{M}^{L}_{\mathrm{PGL}(r)}(\chi))_{w}

is induced by the categorical Hall product.

In particular by setting k=1k=1 in Theorem 7.1 and Theorem 7.2, we obtain the subcategories

𝕋SL(r)L(χ)wDb(SL(r)L(χ))w,𝕋PGL(r)L(χ)wDb(PGL(r)L(χ))w\displaystyle\mathbb{T}_{\mathrm{SL}(r)}^{L}(\chi)_{w}\subset D^{b}(\mathcal{M}_{\mathrm{SL}(r)}^{L}(\chi))_{w},\ \mathbb{T}_{\mathrm{PGL}(r)}^{L}(\chi)_{w}\subset D^{b}(\mathcal{M}_{\mathrm{PGL}(r)}^{L}(\chi))_{w}

which are the quasi-BPS categories for SL/PGL Higgs bundles. Again we omit LL from the notation if L=ΩCL=\Omega_{C}.

7.4. The SL/PGL symmetry conjecture

The Hitchin maps give the diagram

(7.9)

where B2LB_{\geqslant 2}^{L} is defined by

B2L:=i=2rH0(C,Li).\displaystyle B_{\geqslant 2}^{L}:=\bigoplus_{i=2}^{r}H^{0}(C,L^{\otimes i}).

The maps in (7.9) have relative dimension gspgg^{\rm{sp}}-g. The following is the SL/PGL version of Conjecture 2.8:

Conjecture 7.3.

([PTa, Conjecture 7.5]) Suppose that the tuple (r,χ,w)(r,\chi,w) satisfies the BPS condition. Then there is an equivalence

𝕋LPGL(r)(w+1gsp)χ+1gsp𝕋LSL(r)(χ)w.\displaystyle\mathbb{T}^{L}_{\mathrm{PGL}(r)}(w+1-g^{\rm{sp}})_{-\chi+1-g^{\rm{sp}}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathbb{T}^{L}_{\mathrm{SL}(r)}(\chi)_{w}.

8. The SL/PGL-duality of topological K-theories

In this section, we provide evidence towards Conjecture 7.3, namely we prove part (2) of Theorem 1.1 and part (2) of Theorem 1.2.

8.1. Topological K-theory of SL-moduli spaces

In this subsection, we compute the topological K-theory of the quasi-BPS categories for SL-moduli spaces using Joyce–Song pairs. The argument is similar to the one used to prove Proposition 4.1 and Theorem 6.9.

Recall the moduli stack of SL-Higgs bundles together with the good moduli space

LSL(r)(χ)MLSL(r)(χ)\displaystyle\mathcal{M}^{L}_{\mathrm{SL}(r)}(\chi)\to M^{L}_{\mathrm{SL}(r)}(\chi)

and the Hitchin map

(8.1) h0:MLSL(r)(χ)BL2.\displaystyle h_{0}\colon M^{L}_{\mathrm{SL}(r)}(\chi)\to B^{L}_{\geqslant 2}.

Note that MLSL(r)(χ)M^{L}_{\mathrm{SL}(r)}(\chi) is the fiber of the smooth morphism

(8.2) ML(r,χ)H0(L)×Pic(C),(F,θ)(tr(θ),detF)\displaystyle M^{L}(r,\chi)\to H^{0}(L)\times\mathrm{Pic}(C),\ (F,\theta)\mapsto(\mathrm{tr}(\theta),\det F)

at (0,A)(0,A), where AA is a fixed line bundle of degree χ+r(g1)\chi+r(g-1). We define the space of Joyce–Song pairs LSL(r)(χ)JS\mathcal{M}^{L}_{\mathrm{SL}(r)}(\chi)^{\rm{JS}} to be the fiber of the composition

L(r,χ)JSML(r,χ)H0(L)×Pic(C)\displaystyle\mathcal{M}^{L}(r,\chi)^{\rm{JS}}\to M^{L}(r,\chi)\to H^{0}(L)\times\mathrm{Pic}(C)

at (0,A)(0,A). For simplicity, we write ML=ML(r,χ)M^{L}=M^{L}(r,\chi), L=L(r,χ)JS\mathcal{M}^{L{\dagger}}=\mathcal{M}^{L}(r,\chi)^{\rm{JS}}, MSLL=MLSL(r)(χ)M_{\mathrm{SL}}^{L}=M^{L}_{\mathrm{SL}(r)}(\chi) and SLL=LSL(r)(χ)JS\mathcal{M}_{\mathrm{SL}}^{L{\dagger}}=\mathcal{M}^{L}_{\mathrm{SL}(r)}(\chi)^{\rm{JS}}. We first discuss the case of l>2g2l>2g-2.

Proposition 8.1.

Assume that l>2g2l>2g-2 and (r,χ,w)(r,\chi,w) satisfies the BPS condition. Then the complex of sheaves

𝒦BL2top(𝕋LSL(r)(χ)w)D(Sh(BL2))\displaystyle\mathcal{K}_{B^{L}_{\geqslant 2}}^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{SL}(r)}(\chi)_{w})_{\mathbb{Q}}\in D(\mathrm{Sh}_{\mathbb{Q}}(B^{L}_{\geqslant 2}))

is of the form (PQ[1])[β±1](P\oplus Q[1])[\beta^{\pm 1}] for semisimple perverse sheaves P,QP,Q whose generic supports are contained in BL2(BL)ellB^{L}_{\geqslant 2}\cap(B^{L})^{\rm{ell}}.

Proof.

We use the notation from Lemma 3.7. By setting 𝒞=𝕋L(r,χ)wDb(L)\mathcal{C}=\mathbb{T}^{L}(r,\chi)_{w}\subset D^{b}(\mathcal{M}^{L{\dagger}}), its base change 𝒞SLDb(SLL)\mathcal{C}_{\mathrm{SL}}\subset D^{b}(\mathcal{M}_{\mathrm{SL}}^{L{\dagger}}) as in Lemma 3.7 is equivalent to 𝕋SL(r)L(χ)w\mathbb{T}_{\mathrm{SL}(r)}^{L}(\chi)_{w}, which follows by taking the base-change of the semiorthogonal decomposition in Theorem 2.12. By Lemma 3.7, we have the equivalence

𝒦MSLLtop(𝕋SL(r)L(χ)w)i1𝒦MLtop(𝕋L(r,χ)w).\displaystyle\mathcal{K}_{M_{\mathrm{SL}}^{L}}^{\rm{top}}(\mathbb{T}_{\mathrm{SL}(r)}^{L}(\chi)_{w})_{\mathbb{Q}}\simeq i^{-1}\mathcal{K}_{M^{L}}^{\rm{top}}(\mathbb{T}^{L}(r,\chi)_{w})_{\mathbb{Q}}.

Note that we have

𝒦MLtop(𝕋L(r,χ)w)ICML[dimML][β±1]\displaystyle\mathcal{K}_{M^{L}}^{\rm{top}}(\mathbb{T}^{L}(r,\chi)_{w})_{\mathbb{Q}}\cong\mathrm{IC}_{M^{L}}[-\dim M^{L}][\beta^{\pm 1}]

by Proposition 4.1. Since the map (8.2) is smooth, we have

i1ICML[dimML]ICMSLL[dimMSLL].\displaystyle i^{-1}\mathrm{IC}_{M^{L}}[-\dim M^{L}]\cong\mathrm{IC}_{M_{\mathrm{SL}}^{L}}[-\dim M_{\mathrm{SL}}^{L}].

Therefore we have

𝒦MSLLtop(𝕋SL(r)L(χ)w)ICMSLL[dimMSLL][β±1].\displaystyle\mathcal{K}_{M_{\mathrm{SL}}^{L}}^{\rm{top}}(\mathbb{T}_{\mathrm{SL}(r)}^{L}(\chi)_{w})_{\mathbb{Q}}\cong\mathrm{IC}_{M_{\mathrm{SL}}^{L}}[-\dim M_{\mathrm{SL}}^{L}][\beta^{\pm 1}].

Using Theorem 3.2, we conclude that there is an isomorphism

𝒦BL2top(𝕋LSL(r)(χ)w)h0ICMSLL[dimMSLL][β±1]\displaystyle\mathcal{K}_{B^{L}_{\geqslant 2}}^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{SL}(r)}(\chi)_{w})_{\mathbb{Q}}\cong h_{0\ast}\mathrm{IC}_{M_{\mathrm{SL}}^{L}}[-\dim M_{\mathrm{SL}}^{L}][\beta^{\pm 1}]

where h0h_{0} is the Hitchin map (8.1). It is proved in [MS22, Theorem 0.1] that h0ICMSLLh_{0\ast}\mathrm{IC}_{M_{\mathrm{SL}}^{L}} is of the form iAi[i]\oplus_{i}A_{i}[-i] for semisimple perverse sheaves AiA_{i} with generic supports contained in BL2(BL)ellB^{L}_{\geqslant 2}\cap(B^{L})^{\rm{ell}}. Therefore, we obtain the desired conclusion. ∎

Lemma 8.2.

Suppose that l>2g2l>2g-2. Then the topological K-groups Ktop(𝕋LSL(r)(χ)w)K_{\ast}^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{SL}(r)}(\chi)_{w}) are torsion free.

Proof.

We use the notation in the proof of Proposition 8.1. Since 𝕋LSL(r)(χ)w\mathbb{T}^{L}_{\mathrm{SL}(r)}(\chi)_{w} is a semiorthogonal summand of Db(SLL)D^{b}(\mathcal{M}_{\mathrm{SL}}^{L{\dagger}}), it is enough to prove that Ktop(SLL)K_{\ast}^{\rm{top}}(\mathcal{M}_{\mathrm{SL}}^{L{\dagger}}) is torsion free. Note that the \mathbb{C}^{\ast}-action on L\mathcal{M}^{L{\dagger}} scaling the Higgs field restricts to the action of SLL\mathcal{M}_{\mathrm{SL}}^{L{\dagger}}. As in the proof of Proposition 5.9, it is enough to prove that H((SLL),)H^{\ast}\big{(}(\mathcal{M}_{\mathrm{SL}}^{L{\dagger}})^{\mathbb{C}^{\ast}},\mathbb{Z}\big{)} is torsion free. Since SLL\mathcal{M}_{\mathrm{SL}}^{L{\dagger}} is independent of a choice of APic(C)A\in\mathrm{Pic}(C) of degree χ+r(g1)\chi+r(g-1), we may assume that A=𝒪C((χ+r(g1))c0)A=\mathcal{O}_{C}((\chi+r(g-1))c_{0}) for a fixed point c0Cc_{0}\in C.

Recall that, by Proposition 5.7, the class of (L)(\mathcal{M}^{L{\dagger}})^{\mathbb{C}^{\ast}} in K^(Var)\widehat{K}(\mathrm{Var}) is a linear combination of classes of products of 𝕃\mathbb{L} and Symi(C)\mathrm{Sym}^{i}(C) for ii\in\mathbb{Z}. We have the following pull-back square:

We then apply the argument of [GS, Proposition 6.4] by replacing ˇ\check{\mathcal{M}} in loc. cit. with (SLL)(\mathcal{M}_{\mathrm{SL}}^{L{\dagger}})^{\mathbb{C}^{\ast}} to conclude that the class (SLL)×Pic0(C)(\mathcal{M}_{\mathrm{SL}}^{L{\dagger}})^{\mathbb{C}^{\ast}}\times\mathrm{Pic}^{0}(C) in K^(Var)\widehat{K}(\mathrm{Var}) is a linear combination of the classes of the form iSymli(C)~×𝕃m\widetilde{\prod_{i}\mathrm{Sym}^{l_{i}}(C)}\times\mathbb{L}^{m}, where iSymli(C)~\widetilde{\prod_{i}\mathrm{Sym}^{l_{i}}(C)} is defined to be the pull-back diagram:

Here μ(k)\mu(k) is the map

μ(k):Syml(C)Pic0(C),(c1,,cl)𝒪C(c1++cllc0)k.\displaystyle\mu(k)\colon\mathrm{Sym}^{l}(C)\to\mathrm{Pic}^{0}(C),\ (c_{1},\ldots,c_{l})\mapsto\mathcal{O}_{C}(c_{1}+\cdots+c_{l}-lc_{0})^{\otimes k}.

It is proved in [GS, Lemma 6.8] that H(iSymli(C)~,)H^{\ast}\big{(}\widetilde{\prod_{i}\mathrm{Sym}^{l_{i}}(C)},\mathbb{Z}\big{)} is torsion free. Therefore as in the proof of Lemma 5.6, we conclude that H((SLL)×Pic0(C),)H^{\ast}\big{(}(\mathcal{M}_{\mathrm{SL}}^{L{\dagger}})^{\mathbb{C}^{\ast}}\times\mathrm{Pic}^{0}(C),\mathbb{Z}\big{)} is torsion free, hence H((SLL),)H^{\ast}\big{(}(\mathcal{M}_{\mathrm{SL}}^{L{\dagger}})^{\mathbb{C}^{\ast}},\mathbb{Z}\big{)} is also torsion free. ∎

We next consider the case of L=ΩCL=\Omega_{C}. The computation is analogous to the one from Theorem 6.9. For simplicity, we write M=MΩC(r,χ)M=M^{\Omega_{C}}(r,\chi), MSL=MSLΩC(r,χ)M_{\mathrm{SL}}=M_{\mathrm{SL}}^{\Omega_{C}}(r,\chi), etc. Let i:MSLMi\colon M_{\mathrm{SL}}\hookrightarrow M be the closed immersion, and set

𝒫𝒮MSL:=i1𝒫𝒮M[2g].\displaystyle\mathcal{BPS}_{M_{\mathrm{SL}}}:=i^{-1}\mathcal{BPS}_{M}[-2g].
Lemma 8.3.

There is an isomorphism

𝒦MSLtop(𝕋SL(r)(χ)w)𝒫𝒮MSL[β±1].\displaystyle\mathcal{K}_{M_{\mathrm{SL}}}^{\rm{top}}(\mathbb{T}_{\mathrm{SL}(r)}(\chi)_{w})_{\mathbb{Q}}\cong\mathcal{BPS}_{M_{\mathrm{SL}}}[\beta^{\pm 1}].
Proof.

The lemma follows from the definition of the sheaf 𝒫𝒮MSL\mathcal{BPS}_{M_{\mathrm{SL}}}, and from Theorem 6.9 and Lemma 3.7. ∎

Let 𝒪CΩCL\mathcal{O}_{C}\oplus\Omega_{C}\twoheadrightarrow L be a surjection as in the proof of Theorem 6.11. In the notation of the diagram (6.32), we have the commutative diagram

Here, the bottom arrow is the inclusion induced by ΩC𝒪CΩCL\Omega_{C}\hookrightarrow\mathcal{O}_{C}\oplus\Omega_{C}\twoheadrightarrow L. By taking the (classical) fiber at (0,A)H0(L)×Pic(C)(0,A)\in H^{0}(L)\times\mathrm{Pic}(C), we obtain the commutative diagram

(8.7)

where MX,SLM_{X,\mathrm{SL}} is the fiber of the morphism MXredH0(L)×Pic(C)M_{X}^{\rm{red}}\to H^{0}(L)\times\mathrm{Pic}(C) at (0,A)(0,A). The function w¯\overline{w} is given in the diagram (6.32), which by [KM24, Proposition 5.7] can be described as, for (F,θ)(F,\theta) a LL-twisted Higgs bundle:

(8.8) w¯(F,θ)=α,tr(θ2)\displaystyle\overline{w}(F,\theta)=\langle\alpha,\mathrm{tr}(\theta^{2})\rangle

for some αH0(L2)\alpha\in H^{0}(L^{\otimes 2})^{\vee}. The element α\alpha corresponds to the extension class

0L𝒪CΩCL0\displaystyle 0\to L^{\prime}\to\mathcal{O}_{C}\oplus\Omega_{C}\to L\to 0

under the isomorphism

ExtC1(L,L)=ExtC1(L,L1ΩC)H0(L2).\displaystyle\operatorname{Ext}_{C}^{1}(L,L^{\prime})=\operatorname{Ext}_{C}^{1}(L,L^{-1}\otimes\Omega_{C})\cong H^{0}(L^{\otimes 2})^{\vee}.
Lemma 8.4.

There is an isomorphism

𝒫𝒮MSLpSLϕωSL(ICMLSL).\displaystyle\mathcal{BPS}_{M_{\mathrm{SL}}}\cong p_{\mathrm{SL}\ast}\phi_{\omega_{\mathrm{SL}}}(\mathrm{IC}_{M^{L}_{\mathrm{SL}}}).
Proof.

Recall that there is an isomorphism, see [KK, Proposition 3.10]:

𝒫𝒮Mp¯ϕw¯(ICML).\displaystyle\mathcal{BPS}_{M}\cong\overline{p}_{\ast}\phi_{\overline{w}}(\mathrm{IC}_{M^{L}}).

It follows that

𝒫𝒮MSLi1p¯ϕw¯(ICML)[2g]pSLi1ϕw¯(ICML)[2g].\displaystyle\mathcal{BPS}_{M_{\mathrm{SL}}}\cong i^{-1}\overline{p}_{\ast}\phi_{\overline{w}}(\mathrm{IC}_{M^{L}})[-2g]\cong p_{\mathrm{SL}\ast}i^{-1}\phi_{\overline{w}}(\mathrm{IC}_{M^{L}})[-2g].

It is enough to show that

(8.9) i1ϕw¯(ICML)ϕwSL(ICMSL)[2g].\displaystyle i^{-1}\phi_{\overline{w}}(\mathrm{IC}_{M^{L}})\cong\phi_{w_{\mathrm{SL}}}(\mathrm{IC}_{M_{\mathrm{SL}}})[2g].

We have the following commutative diagram

Here jj is given by j(x)=(x,0,A)j(x)=(x,0,A), the right square is Cartesian, and the map γ\gamma is given by

γ((F,θ),η,)=(F,θ+1Fη).\displaystyle\gamma((F,\theta),\eta,\mathcal{L})=\mathcal{L}\otimes(F,\theta+1_{F}\otimes\eta).

From the above diagram, we have

i1ϕw¯(ICML)\displaystyle i^{-1}\phi_{\overline{w}}(\mathrm{IC}_{M^{L}}) j1γ1ϕw¯(ICML)\displaystyle\cong j^{-1}\gamma^{-1}\phi_{\overline{w}}(\mathrm{IC}_{M^{L}})
j1ϕγw¯(γ1ICML)\displaystyle\cong j^{-1}\phi_{\gamma^{\ast}\overline{w}}(\gamma^{-1}\mathrm{IC}_{M^{L}})
j1ϕγw¯(ICMSLL[h0(L)+g]).\displaystyle\cong j^{-1}\phi_{\gamma^{\ast}\overline{w}}(\mathrm{IC}_{M_{\mathrm{SL}}^{L}}\boxtimes\mathbb{Q}[h^{0}(L)+g]).

By the above formula for γ\gamma and w¯\overline{w}, we have

γw¯((F,θ),η,)=α,tr(θ+1Fη)2=ωSLrα,η2.\displaystyle\gamma^{\ast}\overline{w}((F,\theta),\eta,\mathcal{L})=\langle\alpha,\mathrm{tr}(\theta+1_{F}\otimes\eta)^{2}\rangle=\omega_{\mathrm{SL}}\boxplus r\langle\alpha,\eta^{2}\rangle.

Let q(η)=α,η2q(\eta)=\langle\alpha,\eta^{2}\rangle, which is a quadratic function on H0(L)H^{0}(L). From the identity Xred=Crit(w)\mathcal{M}_{X}^{\rm{red}}=\mathrm{Crit}(w) in the rank one case, we see that Crit(q)=H0(ΩC)H0(L)\mathrm{Crit}(q)=H^{0}(\Omega_{C})\subset H^{0}(L). By using the Thom–Sebastiani theorem, we have

j1ϕγw¯(ICMSLL[h0(L)+g])\displaystyle j^{-1}\phi_{\gamma^{\ast}\overline{w}}(\mathrm{IC}_{M_{\mathrm{SL}}^{L}}\boxtimes\mathbb{Q}[h^{0}(L)+g]) j1(ϕwSL(ICMSLL)ϕq([h0(L)+g]))\displaystyle\cong j^{-1}(\phi_{w_{\mathrm{SL}}}(\mathrm{IC}_{\mathrm{M}_{\mathrm{SL}}^{L}})\boxtimes\phi_{q}(\mathbb{Q}[h^{0}(L)+g]))
j1(ϕwSL(ICMSLL[2g])\displaystyle\cong j^{-1}(\phi_{w_{\mathrm{SL}}}(\mathrm{IC}_{M_{\mathrm{SL}}^{L}}\boxtimes\mathbb{Q}[2g])
ϕwSL(ICMSLL[2g]).\displaystyle\cong\phi_{w_{\mathrm{SL}}}(\mathrm{IC}_{M_{\mathrm{SL}}^{L}}[2g]).

Therefore the isomorphism (8.9) holds. ∎

We have the following commutative diagram, where the vertical arrows are Hitchin maps and the top arrows are as in the diagram (8.7):

Proposition 8.5.

There is an isomorphism

𝒦B2top(𝕋SL(r)(χ)w)pBϕwB(𝒦BL2top(𝕋LSL(r)(χ)w)).\displaystyle\mathcal{K}_{B_{\geqslant 2}}^{\rm{top}}(\mathbb{T}_{\mathrm{SL}(r)}(\chi)_{w})_{\mathbb{Q}}\cong p_{B\ast}\phi_{w_{B}}(\mathcal{K}_{B^{L}_{\geqslant 2}}^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{SL}(r)}(\chi)_{w})_{\mathbb{Q}}).
Proof.

The conclusion follows using the same argument as in (6.33) and Lemma 8.3 and Lemma 8.4. ∎

8.2. Parabolic framing of PGL-moduli spaces

In this and in the next subsections, we compute the topological K-theory of quasi-BPS categories for the PGL-moduli spaces. The computation is more difficult than in the SL case. A first issue is that we cannot use the straightforward generalization of the Joyce–Song stable pairs, as we explain below.

Recall the PGL-Higgs moduli space

PGL(r)L(χ)=L(r,χ)tr=0/𝒫ic0(C).\displaystyle\mathcal{M}_{\mathrm{PGL}(r)}^{L}(\chi)=\mathcal{M}^{L}(r,\chi)^{\rm{tr}=0}/\mathcal{P}ic^{0}(C).

In this subsection, we construct a framed version of the PGL-Higgs moduli space. A subtlety here is that there is no natural action of 𝒫ic0(C)\mathcal{P}ic^{0}(C) on the moduli of stable Joyce–Song pairs L(r,χ)JS\mathcal{M}^{L}(r,\chi)^{\rm{JS}}, so that we cannot take its quotient. Instead, we use parabolic framing to rigidify automorphisms and to construct a space with an action of Pic0(C)\mathrm{Pic}^{0}(C). A similar idea also appeared in [Tod14, Tod].

We fix pCp\in C, and define L(r,χ)par\mathcal{M}^{L}(r,\chi)^{\rm{par}} to be the moduli stack of tuples

(8.10) (F,θ,ξ),θ:FFL,ξF|p{0}\displaystyle(F,\theta,\xi),\theta\colon F\to F\otimes L,\ \xi\in F|_{p}\setminus\{0\}

such that (F,θ)(F,\theta) is a semistable LL-twisted Higgs bundle, and that for any surjection η:(F,θ)(F,θ)\eta\colon(F,\theta)\twoheadrightarrow(F^{\prime},\theta^{\prime}) with μ(F)=μ(F)\mu(F)=\mu(F^{\prime}), we have η|p(ξ)0\eta|_{p}(\xi)\neq 0.

Remark 8.6.

Let S=TotC(L)S=\mathrm{Tot}_{C}(L) and let ECoh(S)E\in\operatorname{Coh}(S) correspond to (F,θ)(F,\theta) by the spectral construction. Then giving ξ\xi is equivalent to giving a morphism

ξ:𝒪Fp[1]E,\displaystyle\xi\colon\mathcal{O}_{F_{p}}[-1]\to E,

where FpSF_{p}\subset S is the fiber of SCS\to C at pp. The pair (E,ξ)(E,\xi) is nothing but the parabolic stable pair considered in [Tod14].

Similarly to [Tod14, Theorem 2.10], the moduli stack L(r,χ)par\mathcal{M}^{L}(r,\chi)^{\rm{par}} is a quasi-projective scheme such that the composition

L(r,χ)parL(r,χ)ML(r,χ)\displaystyle\mathcal{M}^{L}(r,\chi)^{\rm{par}}\to\mathcal{M}^{L}(r,\chi)\to M^{L}(r,\chi)

is projective. Here, the first morphism is forgetting ξ\xi, which is a smooth morphism. When l>2g2l>2g-2, the moduli space L(r,χ)par\mathcal{M}^{L}(r,\chi)^{\rm{par}} is also smooth since L(r,χ)\mathcal{M}^{L}(r,\chi) is a smooth stack.

There is a natural action of Pic0(C)\mathrm{Pic}^{0}(C) on L(r,χ)par\mathcal{M}^{L}(r,\chi)^{\rm{par}} as follows. We identify Pic0(C)\mathrm{Pic}^{0}(C) with the moduli stack of pairs

(,ι),ι:|p,\displaystyle(\mathcal{L},\iota),\ \iota\colon\mathcal{L}|_{p}\stackrel{{\scriptstyle\cong}}{{\to}}\mathbb{C},

where \mathcal{L} is a line bundle on CC of degree zero. Note that the isomorphism ι\iota rigidifies the automorphisms of the line bundle \mathcal{L}. Then the action is given by

(8.11) (,ι)(F,θ,ξ)=(F,θ1,(1Fι)(ξ)).\displaystyle(\mathcal{L},\iota)\circ(F,\theta,\xi)=(F\otimes\mathcal{L},\theta\otimes 1_{\mathcal{L}},(1_{F}\otimes\iota)(\xi)).

We denote by

(8.12) L(r,χ)tr=0,parL(r,χ)par\displaystyle\mathcal{M}^{L}(r,\chi)^{\rm{tr}=0,\rm{par}}\subset\mathcal{M}^{L}(r,\chi)^{\rm{par}}

the closed subscheme consisting of tuples (8.10) satisfying tr(θ)=0\mathrm{tr}(\theta)=0. The above action (8.11) of Pic0(C)\mathrm{Pic}^{0}(C) restricts to the action on the closed subscheme (8.12). The parabolic framed PGL-moduli space is defined to be the quotient stack

LPGL(r)(χ)par:=L(r,χ)tr=0,par/Pic0(C).\displaystyle\mathcal{M}^{L}_{\mathrm{PGL}(r)}(\chi)^{\rm{par}}:=\mathcal{M}^{L}(r,\chi)^{\rm{tr}=0,\rm{par}}/\mathrm{Pic}^{0}(C).

The above stack is a Deligne-Mumford stack which is smooth for l>2g2l>2g-2. Alternatively, let SL(r)L(χ)par\mathcal{M}_{\mathrm{SL}(r)}^{L}(\chi)^{\rm{par}} be the fiber of the morphism

L(r,χ)tr=0Pic(C),(F,θ,ξ)detF\displaystyle\mathcal{M}^{L}(r,\chi)^{\rm{tr=0}}\to\mathrm{Pic}(C),\ (F,\theta,\xi)\mapsto\det F

at a fixed APic(C)A\in\mathrm{Pic}(C) of degree χ+r(g1)\chi+r(g-1). Then we have

(8.13) PGL(r)L(χ)par=LSL(r)(χ)par/Γ[r],\displaystyle\mathcal{M}_{\mathrm{PGL}(r)}^{L}(\chi)^{\rm{par}}=\mathcal{M}^{L}_{\mathrm{SL}(r)}(\chi)^{\rm{par}}/\Gamma[r],

where Γ[r]Pic0(C)\Gamma[r]\subset\mathrm{Pic}^{0}(C) is the subgroup of rr-torsion elements. We note the following analogue of Theorem 2.12:

Proposition 8.7.

Assume that l>2g2l>2g-2. There is a semiorthogonal decomposition

(8.14) Db(LPGL(r)(χ)par)=𝕋LPGL(r)(χ)w| 0v1r1<<vkrk<1.\displaystyle D^{b}(\mathcal{M}^{L}_{\mathrm{PGL}(r)}(\chi)^{\rm{par}})=\left\langle\mathbb{T}^{L}_{\mathrm{PGL}(r_{\bullet})}(\chi_{\bullet})_{w_{\bullet}}\,\Big{|}\,0\leqslant\frac{v_{1}}{r_{1}}<\cdots<\frac{v_{k}}{r_{k}}<1\right\rangle.

The right hand side is after all partitions r=r1++rkr=r_{1}+\cdots+r_{k}, χ=χ1++χk\chi=\chi_{1}+\cdots+\chi_{k} for rkr_{\bullet}\in\mathbb{Z}^{\oplus k}, χk/(r1,,rk)\chi_{\bullet}\in\mathbb{Z}^{\oplus k}/(r_{1},\ldots,r_{k})\mathbb{Z} such that (χ1/r1,,χk/rk)=(χ/r,,χ/r)(\chi_{1}/r_{1},\ldots,\chi_{k}/r_{k})=(\chi/r,\ldots,\chi/r) in k/\mathbb{Q}^{\oplus k}/\mathbb{Q}, and wkw_{\bullet}\in\mathbb{Z}^{\oplus k}. Each vi12v_{i}\in\frac{1}{2}\mathbb{Z} is determined by

vi=wil2ri(i>jrji<jrj).\displaystyle v_{i}=w_{i}-\frac{l}{2}r_{i}\left(\sum_{i>j}r_{j}-\sum_{i<j}r_{j}\right).
Proof.

Similarly to the proof of Theorem 2.12, there is a semiorthogonal decomposition

(8.15) Db(L(r,χ)par)=i=1k𝕋L(ri,χi)wi| 0v1r1<<vkrk<1.\displaystyle D^{b}(\mathcal{M}^{L}(r,\chi)^{\rm{par}})=\left\langle\boxtimes_{i=1}^{k}\mathbb{T}^{L}(r_{i},\chi_{i})_{w_{i}}\,\Big{|}\,0\leqslant\frac{v_{1}}{r_{1}}<\cdots<\frac{v_{k}}{r_{k}}<1\right\rangle.

By restricting it to the trace free part, and taking the quotient by Pic0(C)\mathrm{Pic}^{0}(C) as in the proof of [PTa, Theorem 7.2], we obtain the semiorthogonal decomposition (8.14). ∎

Remark 8.8.

If (r,χ)(r,\chi) are coprime, then L(r,χ)parML(r,χ)\mathcal{M}^{L}(r,\chi)^{\rm{par}}\to M^{L}(r,\chi) is a r1\mathbb{P}^{r-1}-bundle, and the semiorthogonal decomposition (8.15) is

Db(L(r,χ)par)=𝕋L(r,χ)w| 0wr1.\displaystyle D^{b}(\mathcal{M}^{L}(r,\chi)^{\rm{par}})=\left\langle\mathbb{T}^{L}(r,\chi)_{w}\,\Big{|}\,0\leqslant w\leqslant r-1\right\rangle.

Then the semiorthogonal decomposition (8.7) is

Db(LPGL(r)(χ)par)=𝕋LPGL(r)(χ)w| 0wr1.\displaystyle D^{b}(\mathcal{M}^{L}_{\mathrm{PGL}(r)}(\chi)^{\rm{par}})=\left\langle\mathbb{T}^{L}_{\mathrm{PGL}(r)}(\chi)_{w}\,\Big{|}\,0\leqslant w\leqslant r-1\right\rangle.

The above semiorthogonal decomposition also holds for L=ΩCL=\Omega_{C}, since M(r,χ)M(r,\chi) is smooth if (r,χ)(r,\chi) are coprime and M(r,χ)parM(r,χ)M(r,\chi)^{\rm{par}}\to M(r,\chi) is a r1\mathbb{P}^{r-1}-bundle as well.

8.3. Fixed loci of Γ[r]\Gamma[r]-actions

The computation of the topological K-theory of quasi-BPS categories for PGL has contributions from the Γ[r]\Gamma[r]-fixed loci, and uses the topological K-theory for moduli of Higgs bundles on certain étale covers of CC. This argument is standard when studying Hausel–Thaddeus mirror symmetry, see for example the arguments in [GS, MS22, MS21].

For γΓ[r]\gamma\in\Gamma[r], let

L(r,χ)γparL(r,χ)par\displaystyle\mathcal{M}^{L}(r,\chi)_{\gamma}^{\rm{par}}\subset\mathcal{M}^{L}(r,\chi)^{\rm{par}}

be the γ\gamma-fixed subscheme. We describe the above fixed locus in terms of parabolic-framed moduli space on an étale cover of CC.

Let mm be the order of γ\gamma, so we have r=mrr=mr^{\prime} for a positive integer rr^{\prime}. If γ\gamma corresponds to (γ,ι)(\mathcal{L}_{\gamma},\iota), then

(8.16) η:γm𝒪C,η|p=ιm.\displaystyle\eta\colon\mathcal{L}_{\gamma}^{\otimes m}\stackrel{{\scriptstyle\cong}}{{\to}}\mathcal{O}_{C},\ \eta|_{p}=\iota^{\otimes m}.

The isomorphism η\eta in (8.16) determines a finite étale cover of degree mm, with Galois group Gγ=/mG_{\gamma}=\mathbb{Z}/m:

πγ:C~C.\displaystyle\pi_{\gamma}\colon\widetilde{C}\to C.

The curve C~\widetilde{C} is given by

C~=Spec(𝒪Cγ1γm+1).\displaystyle\widetilde{C}=\operatorname{Spec}(\mathcal{O}_{C}\oplus\mathcal{L}_{\gamma}^{-1}\oplus\cdots\oplus\mathcal{L}_{\gamma}^{-m+1}).

The isomorphism ι:γ|p\iota\colon\mathcal{L}_{\gamma}|_{p}\stackrel{{\scriptstyle\cong}}{{\to}}\mathbb{C} determines a πγ𝒪C~\pi_{\gamma\ast}\mathcal{O}_{\widetilde{C}}-module structure on 𝒪p\mathcal{O}_{p}, which determines a lift of pp to a point p~C~\widetilde{p}\in\widetilde{C}. Let L~=πγL\widetilde{L}=\pi_{\gamma}^{\ast}L and let L~(r,χ)par\mathcal{M}^{\widetilde{L}}(r^{\prime},\chi)^{\rm{par}} be the parabolic-framed moduli stack for (C~,L~)(\widetilde{C},\widetilde{L}), with parabolic framing at p~\widetilde{p}.

Lemma 8.9.

There is an isomorphism

L~(r,χ)parL(r,χ)γpar\displaystyle\mathcal{M}^{\widetilde{L}}(r^{\prime},\chi)^{\rm{par}}\stackrel{{\scriptstyle\cong}}{{\to}}\mathcal{M}^{L}(r,\chi)_{\gamma}^{\rm{par}}

by sending (F,θ,ξ)(F,\theta,\xi) to (πγF,πγθ,πγξ)(\pi_{\gamma\ast}F,\pi_{\gamma\ast}\theta,\pi_{\gamma\ast}\xi), where πγξ\pi_{\gamma\ast}\xi is given by the composition

ξF|pgGγF|g(p)=(πγF)|p.\displaystyle\mathbb{C}\stackrel{{\scriptstyle\xi}}{{\hookrightarrow}}F|_{p}\hookrightarrow\bigoplus_{g\in G_{\gamma}}F|_{g(p)}=(\pi_{\gamma\ast}F)|_{p}.

Here, the second arrow is the inclusion into the direct summand.

Proof.

The argument of [Tod, Proposition 4.3] applies. ∎

Let

LSL(r)(χ)γparLSL(r)(χ)par\displaystyle\mathcal{M}^{L}_{\mathrm{SL}(r)}(\chi)_{\gamma}^{\rm{par}}\subset\mathcal{M}^{L}_{\mathrm{SL}(r)}(\chi)^{\rm{par}}

be the γ\gamma-fixed subscheme. We also denote by

L~SL(r)/πγ(χ)parL~(r,χ)par\displaystyle\mathcal{M}^{\widetilde{L}}_{\mathrm{SL}(r^{\prime})/\pi_{\gamma}}(\chi)^{\rm{par}}\subset\mathcal{M}^{\widetilde{L}}(r^{\prime},\chi)^{\rm{par}}

be the closed subscheme determined by tr(πγθ)=0\mathrm{tr}(\pi_{\gamma\ast}\theta)=0 and det(πγF)=A\det(\pi_{\gamma\ast}F)=A. By Lemma 8.9, we obtain the following:

Lemma 8.10.

There is an isomorphism

L~SL(r)/πγ(χ)parLSL(r)(χ)γpar\displaystyle\mathcal{M}^{\widetilde{L}}_{\mathrm{SL}(r^{\prime})/\pi_{\gamma}}(\chi)^{\rm{par}}\stackrel{{\scriptstyle\cong}}{{\to}}\mathcal{M}^{L}_{\mathrm{SL}(r)}(\chi)_{\gamma}^{\rm{par}}

by sending (F,θ,ξ)(F,\theta,\xi) to (πγF,πγθ,πγξ)(\pi_{\gamma\ast}F,\pi_{\gamma\ast}\theta,\pi_{\gamma\ast}\xi).

8.4. Topological K-theory of PGL-moduli spaces

In this subsection, we prove Proposition 8.15 about the supports of the relative topological K-theory of quasi-BPS categories for PGL-moduli spaces when l>2g2l>2g-2. For simplicity, we set L=SL(r)L(χ)par\mathcal{M}^{L\sharp}=\mathcal{M}_{\mathrm{SL}(r)}^{L}(\chi)^{\rm{par}} and let h:LBL2h^{\sharp}\colon\mathcal{M}^{L\sharp}\to B^{L}_{\geqslant 2} be the Hitchin map.

Lemma 8.11.

Suppose that l>2g2l>2g-2. There is a natural isomorphism

(8.17) 𝒦BL2top(Db(L/Γ[r]))γΓ[r]𝒦BL2top(Db(Lγ))Γ[r].\displaystyle\mathcal{K}_{B^{L}_{\geqslant 2}}^{\rm{top}}(D^{b}(\mathcal{M}^{L\sharp}/\Gamma[r]))_{\mathbb{C}}\stackrel{{\scriptstyle\cong}}{{\to}}\bigoplus_{\gamma\in\Gamma[r]}\mathcal{K}_{B^{L}_{\geqslant 2}}^{\rm{top}}(D^{b}(\mathcal{M}^{L\sharp}_{\gamma}))_{\mathbb{C}}^{\Gamma[r]}.
Proof.

For a complex analytic space MM with an action of a finite group GG with a quotient π:MM//G\pi\colon M\to M/\!\!/G, and a morphism h:M//GBh\colon M/\!\!/G\to B, let KU(M,G)/BKU_{(M,G)/B} be the sheaf of spectra which assigns to each open subset UBU\subset B the GG-equivariant topological K-theory spectra of π1h1(U)\pi^{-1}h^{-1}(U) (cf. [GS, Appendix]). By [GS, Proposition 2.25], there is an equivalence

𝒦L//Γ[r]top(Db(L/Γ[r]))KU(L,Γ[r])/(L//Γ[r]).\displaystyle\mathcal{K}_{\mathcal{M}^{L\sharp}/\!\!/\Gamma[r]}^{\rm{top}}(D^{b}(\mathcal{M}^{L\sharp}/\Gamma[r]))\stackrel{{\scriptstyle\sim}}{{\to}}KU_{(\mathcal{M}^{L\sharp},\Gamma[r])/(\mathcal{M}^{L\sharp}/\!\!/\Gamma[r])}.

By pushing forward via hh^{\sharp}, we obtain an equivalence

𝒦BL2top(Db(L/Γ[r]))KU(L,Γ[r])/BL2.\displaystyle\mathcal{K}_{B^{L}_{\geqslant 2}}^{\rm{top}}(D^{b}(\mathcal{M}^{L\sharp}/\Gamma[r]))\stackrel{{\scriptstyle\sim}}{{\to}}KU_{(\mathcal{M}^{L\sharp},\Gamma[r])/B^{L}_{\geqslant 2}}.

By a theorem of Atiyah-Segal [AS89], there is an equivalence over \mathbb{C}:

(KU(L,Γ[r])/BL2)γΓ[r](KUγL/BL2)Γ[r].\displaystyle(KU_{(\mathcal{M}^{L\sharp},\Gamma[r])/B^{L}_{\geqslant 2}})_{\mathbb{C}}\stackrel{{\scriptstyle\sim}}{{\to}}\bigoplus_{\gamma\in\Gamma[r]}(KU_{\mathcal{M}_{\gamma}^{L\sharp}/B^{L}_{\geqslant 2}})_{\mathbb{C}}^{\Gamma[r]}.

Then the lemma follows from the equivalence

𝒦BL2top(Db(γL))KUγL/BL2.\displaystyle\mathcal{K}_{B^{L}_{\geqslant 2}}^{\rm{top}}(D^{b}(\mathcal{M}_{\gamma}^{L\sharp}))\stackrel{{\scriptstyle\sim}}{{\to}}KU_{\mathcal{M}_{\gamma}^{L\sharp}/B^{L}_{\geqslant 2}}.

We have the following commutative diagram

(8.26)

In the above, L~SL(r)/πγ(χ)\mathcal{M}^{\widetilde{L}}_{\mathrm{SL}(r^{\prime})/\pi_{\gamma}}(\chi) is the closed substack of L~(r,χ)\mathcal{M}^{\widetilde{L}}(r^{\prime},\chi) given by

(tr(πγθ),det(πγF))=(0,A),(\mathrm{tr}(\pi_{\gamma\ast}\theta),\det(\pi_{\gamma\ast}F))=(0,A),

MSL(r)L(χ)γM_{\mathrm{SL}(r)}^{L}(\chi)_{\gamma} is the γ\gamma-fixed closed subscheme of MSL(r)L(χ)M_{\mathrm{SL}(r)}^{L}(\chi), and the middle vertical arrows are good moduli space morphisms. The closed subscheme BLγBL2B^{L}_{\gamma}\subset B^{L}_{\geqslant 2} is the image of hh restricted to MSL(r)L(χ)γM_{\mathrm{SL}(r)}^{L}(\chi)_{\gamma}, and BL(πγ)B^{L}(\pi_{\gamma}) is the Hitchin base for ML~SL(r)/πγ(χ)M^{\widetilde{L}}_{\mathrm{SL}(r^{\prime})/\pi_{\gamma}}(\chi). The map qγ:BL(πγ)BLγq_{\gamma}\colon B^{L}(\pi_{\gamma})\to B^{L}_{\gamma} is a quotient map with respect to the GπγG_{\pi_{\gamma}}-action, see [MS21, Section 1.5].

Remark 8.12.

By [MS21, Remark 1.6], there is a dense open subset BL(πγ)BL(πγ)B^{L}(\pi_{\gamma})^{\ast}\subset B^{L}(\pi_{\gamma}) upon which the GπγG_{\pi_{\gamma}}-action is free. It follows that the map qγq_{\gamma} sends the generic point of BL(πγ)B^{L}(\pi_{\gamma}) to a point in the elliptic locus in BB.

Using the commutative diagram (8.26), Lemma 8.10, and Lemma 8.11, we obtain the following:

Corollary 8.13.

Suppose that l>2g2l>2g-2. There is a natural isomorphism

(8.27) 𝒦BL2top(Db(LPGL(r)(χ)par))γΓ[r]iγ𝒦BLγtop(Db(L~SL(r)/πγ(χ)par))Γ[r].\displaystyle\mathcal{K}_{B^{L}_{\geqslant 2}}^{\rm{top}}(D^{b}(\mathcal{M}^{L}_{\mathrm{PGL}(r)}(\chi)^{\rm{par}}))_{\mathbb{C}}\stackrel{{\scriptstyle\cong}}{{\to}}\bigoplus_{\gamma\in\Gamma[r]}i_{\gamma\ast}\mathcal{K}_{B^{L}_{\gamma}}^{\rm{top}}(D^{b}(\mathcal{M}^{\widetilde{L}}_{\mathrm{SL}(r^{\prime})/\pi_{\gamma}}(\chi)^{\rm{par}}))_{\mathbb{C}}^{\Gamma[r]}.
Proposition 8.14.

Assume that l>2g2l>2g-2 and that (r,χ)(r^{\prime},\chi) are coprime. Then the object

𝒦BLγtop(Db(L~SL(r)/πγ(χ))par)Γ[r]D(Sh(BLγ))\displaystyle\mathcal{K}_{B^{L}_{\gamma}}^{\rm{top}}(D^{b}(\mathcal{M}^{\widetilde{L}}_{\mathrm{SL}(r^{\prime})/\pi_{\gamma}}(\chi))^{\rm{par}})_{\mathbb{Q}}^{\Gamma[r]}\in D(\mathrm{Sh}_{\mathbb{Q}}(B^{L}_{\gamma}))

is of the form (PQ[1])[β±1](P\oplus Q[1])[\beta^{\pm 1}] for semisimple perverse sheaves P,QP,Q with full support BLγB^{L}_{\gamma}.

Proof.

For simplicity, we write M~L=ML~SL(r)/πγ(χ)\widetilde{M}^{L}=M^{\widetilde{L}}_{\mathrm{SL}(r^{\prime})/\pi_{\gamma}}(\chi). When (r,χ)(r^{\prime},\chi) are coprime, the morphism

L~SL(r)/πγ(χ)parM~L\displaystyle\mathcal{M}^{\widetilde{L}}_{\mathrm{SL}(r^{\prime})/\pi_{\gamma}}(\chi)^{\rm{par}}\to\widetilde{M}^{L}

is an étale locally trivial r1\mathbb{P}^{r^{\prime}-1}-bundle. It follows that there is a semiorthogonal decomposition, see Remark 8.8:

Db(L~SL(r)/πγ(χ)par)=Db(M~L,αi)| 0ir1\displaystyle D^{b}(\mathcal{M}^{\widetilde{L}}_{\mathrm{SL}(r^{\prime})/\pi_{\gamma}}(\chi)^{\rm{par}})=\left\langle D^{b}(\widetilde{M}^{L},\alpha^{i})\,\Big{|}\,0\leqslant i\leqslant r^{\prime}-1\right\rangle

for a Brauer class α\alpha on M~L\widetilde{M}^{L}. It is enough to show that the object

(8.28) 𝒦BγLtop(Db(M~L,αi))Γ[r]D(Sh(BγL))\displaystyle\mathcal{K}_{B_{\gamma}^{L}}^{\rm{top}}(D^{b}(\widetilde{M}^{L},\alpha^{i}))_{\mathbb{Q}}^{\Gamma[r]}\in D(\mathrm{Sh}_{\mathbb{Q}}(B_{\gamma}^{L}))

is of the form (PQ[1])[β±1](P\oplus Q[1])[\beta^{\pm 1}] for semisimple perverse sheaves P,QP,Q with full support BγLB_{\gamma}^{L}.

Since the class α^H3(M,)\widehat{\alpha}\in H^{3}(M,\mathbb{Z}) associated with α\alpha is a torsion class, it does not affect the topological K-theory after the rationalization. Therefore we have

𝒦M~Ltop(Db(M~L,αi))M~L[β±1].\displaystyle\mathcal{K}_{\widetilde{M}^{L}}^{\rm{top}}(D^{b}(\widetilde{M}^{L},\alpha^{i}))_{\mathbb{Q}}\cong\mathbb{Q}_{\widetilde{M}^{L}}[\beta^{\pm 1}].

Then we have

𝒦BLγtop(Db(M~L,αi))hγICM~L[d][β±1],\displaystyle\mathcal{K}_{B^{L}_{\gamma}}^{\rm{top}}(D^{b}(\widetilde{M}^{L},\alpha^{i}))_{\mathbb{Q}}\cong h_{\gamma\ast}\mathrm{IC}_{\widetilde{M}^{L}}[-d][\beta^{\pm 1}],

where d=dimM~Ld=\dim\widetilde{M}^{L} and hγh_{\gamma} is the morphism in the diagram (8.26). It is proved in [MS21, Theorem 2.3] that the Γ[r]\Gamma[r]-fixed part of h~ICM~L\widetilde{h}_{\ast}\mathrm{IC}_{\widetilde{M}^{L}} is of the form iAi[i]\oplus_{i}A_{i}[-i] where AiA_{i} is a semisimple perverse sheaf on BL(πγ)B^{L}(\pi_{\gamma}) with full support BL(πγ)B^{L}(\pi_{\gamma}). Since qπq_{\pi} is a quotient map and hγ=qγh~h_{\gamma}=q_{\gamma}\circ\widetilde{h}, it follows that hγICM~Lh_{\gamma\ast}\mathrm{IC}_{\widetilde{M}^{L}} is of the form iAi[i]\oplus_{i}A_{i}^{\prime}[-i] where AiA_{i}^{\prime} is a semisimple perverse sheaf on BLγB^{L}_{\gamma} with full support BLγB^{L}_{\gamma}. Therefore we obtain the desired conclusion. ∎

Proposition 8.15.

Suppose that l>2g2l>2g-2 and that the tuple (r,χ,w)(r,\chi,w) satisfies the BPS condition. Moreover, assume that either (r,χ)(r,\chi) are coprime or rr is a prime number. Then the object

(8.29) 𝒦BL2top(𝕋LPGL(r)(χ)w)D(Sh(BL2))\displaystyle\mathcal{K}_{B^{L}_{\geqslant 2}}^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{PGL}(r)}(\chi)_{w})_{\mathbb{Q}}\in D(\mathrm{Sh}_{\mathbb{Q}}(B^{L}_{\geqslant 2}))

is of the form (PQ[1])[β±1](P\oplus Q[1])[\beta^{\pm 1}] for semisimple perverse sheaves P,QP,Q on BL2B^{L}_{\geqslant 2} whose generic supports are contained in BL2(BL)ellB^{L}_{\geqslant 2}\cap(B^{L})^{\rm{ell}}.

Proof.

It is enough to prove the proposition after taking the tensor product with \mathbb{C}. By Proposition 8.7 and Corollary 8.13, the object (8.29) over \mathbb{C} is a direct summand of the following direct sum

(8.30) γΓ[π]iγ𝒦BLγtop(Db(L~SL(r)/πγ(χ)par))Γ[r]\displaystyle\bigoplus_{\gamma\in\Gamma[\pi]}i_{\gamma\ast}\mathcal{K}_{B^{L}_{\gamma}}^{\rm{top}}(D^{b}(\mathcal{M}^{\widetilde{L}}_{\mathrm{SL}(r^{\prime})/\pi_{\gamma}}(\chi)^{\rm{par}}))_{\mathbb{C}}^{\Gamma[r]}

The above object is of the form γ(PγQγ[1])[β±1]\oplus_{\gamma}(P_{\gamma}\oplus Q_{\gamma}[1])[\beta^{\pm 1}] for semisimple perverse sheaves Pγ,QγP_{\gamma},Q_{\gamma} on B2LB_{\geqslant 2}^{L}. It follows that (8.29) is of the form (PQ[1])[β±1](P\oplus Q[1])[\beta^{\pm 1}] for semisimple perverse sheaves P,QP,Q. It is enough to show that their generic supports are contained in (BL)ell(B^{L})^{\rm{ell}}.

By the assumption that (r,χ)(r,\chi) are coprime or rr is prime, a summand in (8.30) corresponding to γ1\gamma\neq 1 satisfies that gcd(r,χ)=1\gcd(r^{\prime},\chi)=1. Therefore, by Proposition 8.14, the perverse sheaves Pγ,QγP_{\gamma},Q_{\gamma} for γ1\gamma\neq 1 have full support BLγB^{L}_{\gamma}. In particular, their generic support is contained in (BL)ell(B^{L})^{\rm{ell}}, see Remark 8.12. As for the γ=1\gamma=1 summand, by Proposition 8.1, the object

𝒦BL2top(𝕋SL(r)L(χ)w)D(Sh(BL2))\displaystyle\mathcal{K}_{B^{L}_{\geqslant 2}}^{\rm{top}}(\mathbb{T}_{\mathrm{SL}(r)}^{L}(\chi)_{w})_{\mathbb{Q}}\in D(\mathrm{Sh}_{\mathbb{Q}}(B^{L}_{\geqslant 2}))

is of the form (PQ[1])[β±1](P^{\prime}\oplus Q^{\prime}[1])[\beta^{\pm 1}] for semisimple perverse sheaves PP^{\prime}, QQ^{\prime} whose generic supports are contained in (BL)ell(B^{L})^{\rm{ell}}. The isomorphism (8.27) sends 𝒦BL2top(𝕋LPGL(r)(χ)w)\mathcal{K}_{B^{L}_{\geqslant 2}}^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{PGL}(r)}(\chi)_{w})_{\mathbb{C}} to

𝒦BL2top(𝕋SL(r)L(χ)w)Γ[r]γ1iγ𝒦BLγtop(Db(L~SL(r)/πγ(χ)par))Γ[r].\displaystyle\mathcal{K}_{B^{L}_{\geqslant 2}}^{\rm{top}}(\mathbb{T}_{\mathrm{SL}(r)}^{L}(\chi)_{w})_{\mathbb{C}}^{\Gamma[r]}\oplus\bigoplus_{\gamma\neq 1}i_{\gamma\ast}\mathcal{K}_{B^{L}_{\gamma}}^{\rm{top}}(D^{b}(\mathcal{M}^{\widetilde{L}}_{\mathrm{SL}(r^{\prime})/\pi_{\gamma}}(\chi)^{\rm{par}}))_{\mathbb{C}}^{\Gamma[r]}.

Therefore P,QP,Q also have generic supports contained in (BL)ell(B^{L})^{\rm{ell}}. ∎

8.5. Torsion-freeness for PGL-moduli spaces

In this subsection, we discuss the torsion-freeness of topological K-theories of quasi-BPS categories for PGL\mathrm{PGL}-Higgs moduli spaces when l>2g2l>2g-2.

Proposition 8.16.

For l>2g2l>2g-2, the topological KK-group Ktop(𝕋PGL(r)L(χ)w)K_{\ast}^{\rm{top}}(\mathbb{T}_{\mathrm{PGL}(r)}^{L}(\chi)_{w}) is torsion free.

Proof.

Since 𝕋PGL(r)L(χ)w\mathbb{T}_{\mathrm{PGL}(r)}^{L}(\chi)_{w} is a semiorthogonal summand of Db(PGL(r)L(χ)par)D^{b}(\mathcal{M}_{\mathrm{PGL}(r)}^{L}(\chi)^{\rm{par}}) by Proposition 8.7, it is enough to show that

KU(PGL(r)L(χ)par)=π(KUΓ[r](SL(r)L(χ)par))\displaystyle KU_{\ast}(\mathcal{M}_{\mathrm{PGL}(r)}^{L}(\chi)^{\rm{par}})=\pi_{\ast}(KU_{\Gamma[r]}(\mathcal{M}_{\mathrm{SL}(r)}^{L}(\chi)^{\rm{par}}))

is torsion free, where the right hand side is the Γ[r]\Gamma[r]-equivariant topological K-theory. The proof of this claim is the same as in the unframed case [GS, Theorem 6.11]. Below we give its outline.

Let J=Pic0(C)J=\mathrm{Pic}^{0}(C). There is a JJ-torsor:

(8.31) L(r,χ)tr=0,parLPGL(r)(χ)par.\displaystyle\mathcal{M}^{L}(r,\chi)^{\rm{tr}=0,\mathrm{par}}\to\mathcal{M}^{L}_{\mathrm{PGL}(r)}(\chi)^{\rm{par}}.

By [GS, Lemma 6.12], there is a corresponding \mathbb{C}^{\ast}-gerbe β\beta on LPGL(r)(χ)par×J\mathcal{M}^{L}_{\mathrm{PGL}(r)}(\chi)^{\rm{par}}\times J and a derived equivalence

Db(L(r,χ)tr=0,par)Db(LPGL(r)(χ)par×J,β).\displaystyle D^{b}(\mathcal{M}^{L}(r,\chi)^{\rm{tr}=0,\mathrm{par}})\stackrel{{\scriptstyle\sim}}{{\to}}D^{b}(\mathcal{M}^{L}_{\mathrm{PGL}(r)}(\chi)^{\rm{par}}\times J,\beta).

The gerbe β\beta splits after pull-back via [r]:JJ[r]\colon J\to J. Indeed, the gerbe [r]β[r]^{\ast}\beta corresponds to the JJ-torsor given by push-forward of the torsor (8.31) by [r]:JJ[r]\colon J\to J, which is a trivial JJ-torsor by the isomorphism

(π,det):L(r,χ)tr=0,par/Γ[r]PGL(r)L(χ)par×J,\displaystyle(\pi,\det):\mathcal{M}^{L}(r,\chi)^{\rm{tr}=0,\mathrm{par}}/\Gamma[r]\stackrel{{\scriptstyle\cong}}{{\to}}\mathcal{M}_{\mathrm{PGL}(r)}^{L}(\chi)^{\rm{par}}\times J,

where π\pi is the natural projection.

For a prime number pp, write r=parr=p^{a}r^{\prime} with where gcd(p,r)=1\gcd(p,r^{\prime})=1. Let β\beta^{\prime} be the pull-back of β\beta by [r]:JJ[r^{\prime}]\colon J\to J. By [GS, Lemma 6.12], there is a derived equivalence

Db(L(r,χ)tr=0,par/Γ[r])Db(LPGL(r)(χ)par×J,β)\displaystyle D^{b}(\mathcal{M}^{L}(r,\chi)^{\rm{tr}=0,\mathrm{par}}/\Gamma[r^{\prime}])\stackrel{{\scriptstyle\sim}}{{\to}}D^{b}(\mathcal{M}^{L}_{\mathrm{PGL}(r)}(\chi)^{\rm{par}}\times J,\beta^{\prime})

which induces an equivalence of spectra

(8.32) KUΓ[r](L(r,χ)tr=0,par)KUβ^(LPGL(r)(χ)par×J).\displaystyle KU_{\Gamma[r^{\prime}]}(\mathcal{M}^{L}(r,\chi)^{\rm{tr}=0,\mathrm{par}})\stackrel{{\scriptstyle\sim}}{{\to}}KU^{\hat{\beta}^{\prime}}(\mathcal{M}^{L}_{\mathrm{PGL}(r)}(\chi)^{\rm{par}}\times J).

Assume by contradiction that the pp-torsion part of KU(PGL(r)L(χ)par)KU_{\ast}(\mathcal{M}_{\mathrm{PGL}(r)}^{L}(\chi)^{\rm{par}}) is non-zero. Let pr2\mathrm{pr}_{2} be the projection

pr2:LPGL(r)(χ)par×JJ\displaystyle\mathrm{pr}_{2}\colon\mathcal{M}^{L}_{\mathrm{PGL}(r)}(\chi)^{\rm{par}}\times J\to J

and consider the following locally constant sheaf of spectra on JJ

=pr2KU¯β^(LPGL(r)(χ)par×J)\displaystyle\mathcal{F}=\mathrm{pr}_{2\ast}\underline{KU}^{\hat{\beta}^{\prime}}(\mathcal{M}^{L}_{\mathrm{PGL}(r)}(\chi)^{\mathrm{par}}\times J)

with fiber KU(PGL(r)L(χ)par)KU(\mathcal{M}_{\mathrm{PGL}(r)}^{L}(\chi)^{\rm{par}}). The pull-back of \mathcal{F} by [pa]:JJ[p^{a}]\colon J\to J is a constant sheaf, as [pa]β=[r]β[p^{a}]^{\ast}\beta^{\prime}=[r]^{\ast}\beta is a trivial gerbe. Therefore, by [GS, Lemma 6.18], the homotopy groups of Γ()\Gamma(\mathcal{F}) have non-zero pp-torsions. It follows that the homotopy groups of (8.32) have non-zero pp-torsions. On the other hand, by the Atiyah-Segal completion theorem, there is an isomorphism

(8.33) π^(KUΓ[r](L(r,χ)tr=0,par))π(KU(L(r,χ)tr=0,par)hΓ[r]).\displaystyle\widehat{\pi_{\ast}}(KU_{\Gamma[r^{\prime}]}(\mathcal{M}^{L}(r,\chi)^{\rm{tr}=0,\mathrm{par}}))\cong\pi_{\ast}(KU(\mathcal{M}^{L}(r,\chi)^{\rm{tr}=0,\mathrm{par}})^{h\Gamma[r^{\prime}]}).

Here, the left hand side is the the completion with respect to the augmentation ideal of [Γ[r]]\mathbb{Z}[\Gamma[r^{\prime}]], and hΓh\Gamma^{\prime} in the right hand side means the homotopy invariant with respect to the Γ[r]\Gamma[r^{\prime}]-action. By Lemma 8.17 below and using that the order of Γ[r]\Gamma[r^{\prime}] is coprime with pp, we obtain that the right hand side of (8.33) does not have non-zero pp-torsion, which is a contradiction. ∎

We have used the following lemma:

Lemma 8.17.

For l>2g2l>2g-2, the topological KK-group KU(L(r,χ)tr=0,par)KU_{\ast}(\mathcal{M}^{L}(r,\chi)^{\rm{tr}=0,\mathrm{par}}) is torsion free.

Proof.

It is enough to show that KU(L(r,χ)par)KU_{\ast}(\mathcal{M}^{L}(r,\chi)^{\mathrm{par}}) is torsion free, because of the isomorphism

L(r,χ)tr=0,par×H0(L)L(r,χ)par.\displaystyle\mathcal{M}^{L}(r,\chi)^{\mathrm{tr}=0,\mathrm{par}}\times H^{0}(L)\stackrel{{\scriptstyle\cong}}{{\to}}\mathcal{M}^{L}(r,\chi)^{\mathrm{par}}.

By the semiorthogonal decomposition (8.15), we have the direct sum decomposition of KU(L(r,χ)par)KU_{\ast}(\mathcal{M}^{L}(r,\chi)^{\mathrm{par}}) into the direct sum of topological K-groups of the products of quasi-BPS categories. These direct summands are part of the direct summands of Ktop(L)K^{\rm{top}}(\mathcal{M}^{L{\dagger}}) for L=L(r,χ)JS\mathcal{M}^{L{\dagger}}=\mathcal{M}^{L}(r,\chi)^{\rm{JS}} by the semiorthogonal decomposition in Theorem 2.12. Therefore the desired torsion-freeness follows from the torsion-freeness of KU(L)KU_{\ast}(\mathcal{M}^{L{\dagger}}), which is discussed in the proof of Proposition 5.9. ∎

8.6. Topological K-theory for PGL-moduli spaces and L=ΩCL=\Omega_{C}

We next consider PGL(r)(χ):=PGL(r)ΩC(χ)\mathcal{M}_{\mathrm{PGL}(r)}(\chi):=\mathcal{M}_{\mathrm{PGL}(r)}^{\Omega_{C}}(\chi). Assume that (r,χ)(r,\chi) are coprime. In this case, both of SL(r)(χ)\mathcal{M}_{\mathrm{SL}(r)}(\chi) and PGL(r)(χ)\mathcal{M}_{\mathrm{PGL}(r)}(\chi) are smooth Deligne-Mumford stacks. In particular, we have

ΩSL(r)(χ)[1]=SL(r)(χ).\displaystyle\Omega_{\mathcal{M}_{\mathrm{SL}(r)}(\chi)}[-1]=\mathcal{M}_{\mathrm{SL}(r)}(\chi).

Consider a surjection 𝒪CΩCL\mathcal{O}_{C}\oplus\Omega_{C}\twoheadrightarrow L as in the proof of Theorem 6.11, and for simplicity write PGLL=PGL(r)L(χ)\mathcal{M}_{\mathrm{PGL}}^{L}=\mathcal{M}_{\mathrm{PGL}(r)}^{L}(\chi), PGLL=PGL(r)L(χ)par\mathcal{M}_{\mathrm{PGL}}^{L\sharp}=\mathcal{M}_{\mathrm{PGL}(r)}^{L}(\chi)^{\rm{par}}, etc. Consider the following diagram

Here, the horizontal inclusions are induced by the embedding ΩCL\Omega_{C}\hookrightarrow L, and the function wBw_{B} is given as in the diagram (6.32).

Proposition 8.18.

Suppose that (r,χ)(r,\chi) are coprime. There is an isomorphism

(8.34) pBϕwB(𝒦BL2top(𝕋LPGL(r)(χ)w))𝒦B2top(𝕋PGL(r)(χ)w).\displaystyle p_{B\ast}\phi_{w_{B}}(\mathcal{K}_{B^{L}_{\geqslant 2}}^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{PGL}(r)}(\chi)_{w})_{\mathbb{Q}})\cong\mathcal{K}_{B_{\geqslant 2}}^{\rm{top}}(\mathbb{T}_{\mathrm{PGL}(r)}(\chi)_{w})_{\mathbb{Q}}.
Proof.

It is enough to prove the proposition after tensoring \otimes_{\mathbb{Q}}\mathbb{C}. Indeed, both sides over \mathbb{C} are of the form (PQ[1])[β±1](P_{\mathbb{C}}\oplus Q_{\mathbb{C}}[1])[\beta^{\pm 1}] for semisimple perverse sheaves PP, QQ with \mathbb{C}-coefficients. Then both sides in (8.34) are of the form (PQ[1])[β±1](P\oplus Q[1])[\beta^{\pm 1}] for semisimple perverse sheaves with \mathbb{Q}-coefficients. If two of such objects are isomorphic over \mathbb{C}, they are also isomorphic over \mathbb{Q}.

We first prove that

(8.35) pBϕwB(𝒦BL2top(Db(PGLL)))𝒦B2top(Db(PGL)).\displaystyle p_{B\ast}\phi_{w_{B}}(\mathcal{K}_{B^{L}_{\geqslant 2}}^{\rm{top}}(D^{b}(\mathcal{M}_{\mathrm{PGL}}^{L\sharp})))_{\mathbb{C}}\cong\mathcal{K}_{B_{\geqslant 2}}^{\rm{top}}(D^{b}(\mathcal{M}_{\mathrm{PGL}}^{\sharp}))_{\mathbb{C}}.

By the semiorthogonal decomposition (8.15), Lemma 8.11, and Lemma 8.10, there are decompositions

𝒦BL2top(Db(PGLL))\displaystyle\mathcal{K}_{B^{L}_{\geqslant 2}}^{\rm{top}}(D^{b}(\mathcal{M}_{\mathrm{PGL}}^{L\sharp}))_{\mathbb{C}} =γΓ[r](0wr1jγL𝒦BL(πγ)top(𝕋L~SL(r)/πγ(χ)w))Γ[r],\displaystyle=\bigoplus_{\gamma\in\Gamma[r]}\left(\bigoplus_{0\leqslant w\leqslant r^{\prime}-1}j_{\gamma\ast}^{L}\mathcal{K}_{B^{L}(\pi_{\gamma})}^{\rm{top}}(\mathbb{T}^{\widetilde{L}}_{\mathrm{SL}(r^{\prime})/\pi_{\gamma}}(\chi)_{w})_{\mathbb{C}}\right)^{\Gamma[r]},
𝒦B2top(Db(PGL))\displaystyle\mathcal{K}_{B_{\geqslant 2}}^{\rm{top}}(D^{b}(\mathcal{M}_{\mathrm{PGL}}^{\sharp}))_{\mathbb{C}} =γΓ[r](0wr1jγ𝒦B(πγ)top(𝕋SL(r)/πγ(χ)w))Γ[r].\displaystyle=\bigoplus_{\gamma\in\Gamma[r]}\left(\bigoplus_{0\leqslant w\leqslant r^{\prime}-1}j_{\gamma\ast}\mathcal{K}_{B(\pi_{\gamma})}^{\rm{top}}(\mathbb{T}_{\mathrm{SL}(r^{\prime})/\pi_{\gamma}}(\chi)_{w})_{\mathbb{C}}\right)^{\Gamma[r]}.

Here, we have that r=r/mr^{\prime}=r/m, where mm is the order of γ\gamma, πγ:C~C\pi_{\gamma}\colon\widetilde{C}\to C is the Galois cover associated with γ\gamma, L~=πγL\widetilde{L}=\pi_{\gamma}^{\ast}L, and we have used the notation of the diagram

The subcategory

𝕋L~SL(r)/πγ(χ)wDb(L~SL(r)/πγ(χ))w\displaystyle\mathbb{T}^{\widetilde{L}}_{\mathrm{SL}(r^{\prime})/\pi_{\gamma}}(\chi)_{w}\subset D^{b}(\mathcal{M}^{\widetilde{L}}_{\mathrm{SL}(r^{\prime})/\pi_{\gamma}}(\chi))_{w}

is defined similarly to 𝕋SL(r)L(χ)w\mathbb{T}_{\mathrm{SL}(r)}^{L}(\chi)_{w}, which is indeed equivalent to the right hand side as (r,χ)(r^{\prime},\chi) are coprime. By the formula (8.8), the function

L~SL(r)(χ)parBL(πγ)jγLBL2wB\displaystyle\mathcal{M}^{\widetilde{L}}_{\mathrm{SL}(r^{\prime})}(\chi)^{\rm{par}}\to B^{L}(\pi_{\gamma})\stackrel{{\scriptstyle j_{\gamma}^{L}}}{{\to}}B^{L}_{\geqslant 2}\stackrel{{\scriptstyle w_{B}}}{{\to}}\mathbb{C}

is given by the same formula as (8.8), i.e.

(F,θ,ξ)πγα,tr(θ2).\displaystyle(F,\theta,\xi)\mapsto\langle\pi_{\gamma}^{\ast}\alpha,\mathrm{tr}(\theta^{2})\rangle.

Then the isomorphism (8.35) follows from isomorphisms

pBϕw(jLγ𝒦BL(πγ)top(𝕋L~SL(r)/πγ(χ)w))jγ𝒦B(πγ)top(𝕋SL(r)/πγ(χ)w),p_{B\ast}\phi_{w}\left(j^{L}_{\gamma*}\mathcal{K}_{B^{L}(\pi_{\gamma})}^{\rm{top}}(\mathbb{T}^{\widetilde{L}}_{\mathrm{SL}(r^{\prime})/\pi_{\gamma}}(\chi)_{w})\right)_{\mathbb{C}}\cong j_{\gamma*}\mathcal{K}_{B(\pi_{\gamma})}^{\rm{top}}(\mathbb{T}_{\mathrm{SL}(r^{\prime})/\pi_{\gamma}}(\chi)_{w})_{\mathbb{C}},

which are proved completely analogously to Proposition 8.5.

By Proposition 8.7 (see also Remark 8.8), the isomorphism (8.35) is

0wr1pBϕwB(𝒦BL2top(𝕋LPGL(r)(χ)))0wr1𝒦B2top(𝕋PGL(r)(χ)w).\displaystyle\bigoplus_{0\leqslant w\leqslant r-1}p_{B\ast}\phi_{w_{B}}(\mathcal{K}_{B^{L}_{\geqslant 2}}^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{PGL}(r)}(\chi))_{\mathbb{C}})\cong\bigoplus_{0\leqslant w\leqslant r-1}\mathcal{K}_{B_{\geqslant 2}}^{\rm{top}}(\mathbb{T}_{\mathrm{PGL}(r)}(\chi)_{w})_{\mathbb{C}}.

It is straightforward to check that the above isomorphism preserves the direct summands. We thus obtain the isomorphism (8.34). ∎

8.7. The SL/PGL-duality of topological K-theories

For simplicity, we write

SLL=SL(r)L(χ),LPGL=PGL(r)L(w+1gsp).\displaystyle\mathcal{M}_{\mathrm{SL}}^{L}=\mathcal{M}_{\mathrm{SL}(r)}^{L}(\chi),\ \mathcal{M}^{L^{\prime}}_{\mathrm{PGL}}=\mathcal{M}_{\mathrm{PGL}(r)}^{L}(w+1-g^{\rm{sp}}).

We denote their pull-backs to the elliptic locus BL2(BL)ellB^{L}_{\geqslant 2}\cap(B^{L})^{\rm{ell}} by (SLL)ell(\mathcal{M}_{\mathrm{SL}}^{L})^{\rm{ell}}, (PGLL)ell(\mathcal{M}_{\mathrm{PGL}}^{L^{\prime}})^{\rm{ell}} respectively. By [GS, Section 4], the Poincare sheaf 𝒫ell\mathcal{P}^{\rm{ell}} on (L)ell×BL2(L)ell(\mathcal{M}^{L^{\prime}})^{\rm{ell}}\times_{B^{L}_{\geqslant 2}}(\mathcal{M}^{L})^{\rm{ell}} induces the maximal Cohen-Macaulay sheaf

𝒫¯ellCoh((PGLL)ell×BL2(SLL)ell)\displaystyle\overline{\mathcal{P}}^{\rm{ell}}\in\operatorname{Coh}((\mathcal{M}_{\mathrm{PGL}}^{L^{\prime}})^{\rm{ell}}\times_{B^{L}_{\geqslant 2}}(\mathcal{M}_{\mathrm{SL}}^{L})^{\rm{ell}})

and that there is a derived equivalence

(8.36) Φ𝒫¯ell:Db((PGLL)ell)χgsp+1Db((LSL)ell)w.\displaystyle\Phi_{\overline{\mathcal{P}}^{\rm{ell}}}\colon D^{b}((\mathcal{M}_{\mathrm{PGL}}^{L^{\prime}})^{\rm{ell}})_{-\chi-g^{\rm{sp}}+1}\stackrel{{\scriptstyle\sim}}{{\to}}D^{b}((\mathcal{M}^{L}_{\mathrm{SL}})^{\rm{ell}})_{w}.

As in Lemma 4.7, there is an extension of 𝒫¯ell\overline{\mathcal{P}}^{\rm{ell}}:

P¯𝕋LPGL(r)(w+1gsp)χ+gsp1BL2𝕋LSL(r)(χ)w\displaystyle\overline{P}\in\mathbb{T}^{L}_{\mathrm{PGL}(r)}(w+1-g^{\rm{sp}})_{\chi+g^{\rm{sp}}-1}\boxtimes_{B^{L}_{\geqslant 2}}\mathbb{T}^{L}_{\mathrm{SL}(r)}(\chi)_{w}

with the induced Fourier-Mukai functor:

Φ𝒫¯:𝕋LPGL(r)(w+1gsp)χgsp+1𝕋LSL(r)(χ)w.\displaystyle\Phi_{\overline{\mathcal{P}}}\colon\mathbb{T}^{L}_{\mathrm{PGL}(r)}(w+1-g^{\rm{sp}})_{-\chi-g^{\rm{sp}}+1}\to\mathbb{T}^{L}_{\mathrm{SL}(r)}(\chi)_{w}.

The above functor induces a map of spectra

(8.37) Φ𝒫¯:Ktop(𝕋LPGL(r)(w+1gsp)χgsp+1)Ktop(𝕋LSL(r)(χ)w).\displaystyle\Phi_{\overline{\mathcal{P}}}\colon K^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{PGL}(r)}(w+1-g^{\rm{sp}})_{-\chi-g^{\rm{sp}}+1})\to K^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{SL}(r)}(\chi)_{w}).
Theorem 8.19.

Suppose that l>2g2l>2g-2 and that the tuple (r,χ,w)(r,\chi,w) satisfies the BPS condition. Furthermore, assume that (r,w+1gsp)(r,w+1-g^{\rm{sp}}) are coprime or that rr is a prime number. Then the functor Φ𝒫¯\Phi_{\overline{\mathcal{P}}} induces an equivalence of spectra

(8.38) Φ𝒫¯:Ktop(𝕋LPGL(r)(w+1gsp)χgsp+1)Ktop(𝕋LSL(r)(χ)w).\displaystyle\Phi_{\overline{\mathcal{P}}}\colon K^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{PGL}(r)}(w+1-g^{\rm{sp}})_{-\chi-g^{\rm{sp}}+1})\stackrel{{\scriptstyle\sim}}{{\to}}K^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{SL}(r)}(\chi)_{w}).
Proof.

As in the proof of Proposition 4.8, we first show that Φ𝒫¯\Phi_{\overline{\mathcal{P}}} induces an isomorphism

(8.39) Φ𝒫¯:𝒦BL2top(𝕋LPGL(r)(w+1gsp)χgsp+1)𝒦BL2top(𝕋LSL(r)(χ)w).\displaystyle\Phi_{\overline{\mathcal{P}}}\colon\mathcal{K}_{B^{L}_{\geqslant 2}}^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{PGL}(r)}(w+1-g^{\rm{sp}})_{-\chi-g^{\rm{sp}}+1})_{\mathbb{Q}}\stackrel{{\scriptstyle\sim}}{{\to}}\mathcal{K}_{B^{L}_{\geqslant 2}}^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{SL}(r)}(\chi)_{w})_{\mathbb{Q}}.

By Proposition 8.1 and Proposition 8.15, both sides are of the form (PQ[1])[β±1](P\oplus Q[1])[\beta^{\pm 1}] for semisimple perverse sheaves P,QP,Q with generic supports contained in BL2(BL)ellB^{L}_{\geqslant 2}\cap(B^{L})^{\rm{ell}}. Therefore, it is enough to show the equivalence (8.39) over the elliptic locus, which follows from the equivalence (8.36). Therefore (8.39) is an equivalence. Taking the global section and using Proposition 3.6, we obtain the isomorphism

Φ𝒫¯:Ktop(𝕋LPGL(r)(w+1gsp)χgsp+1)Ktop(𝕋LSL(r)(χ)w)\displaystyle\Phi_{\overline{\mathcal{P}}}\colon K^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{PGL}(r)}(w+1-g^{\rm{sp}})_{-\chi-g^{\rm{sp}}+1})_{\mathbb{Q}}\stackrel{{\scriptstyle\sim}}{{\to}}K^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{SL}(r)}(\chi)_{w})_{\mathbb{Q}}

The homotopy groups of both sides in (8.37) are torsion free by Lemma 8.2 and Proposition 8.16, and the map (8.37) is an isomorphism after rationalization by the above argument. Therefore, as in the proof of Theorem 5.10, we have the equivalence (8.38). ∎

We now prove part (2) of Theorem 1.1 and part (2) of Theorem 1.2. Let SL(r)L\mathcal{M}_{\mathrm{SL}(r)}^{L} be the moduli stack of LL-twisted SL(r)\mathrm{SL}(r)-principal Higgs bundles, namely:

SL(r)L:=SL(r)L(χ=r(1g)).\displaystyle\mathcal{M}_{\mathrm{SL}(r)}^{L}:=\mathcal{M}_{\mathrm{SL}(r)}^{L}(\chi=r(1-g)).

The quasi-BPS category for principal SL\mathrm{SL}-Higgs bundles is

𝕋LSL(r),w:=𝕋LSL(r)(χ=r(1g))wDb(SL(r)L)w.\displaystyle\mathbb{T}^{L}_{\mathrm{SL}(r),w}:=\mathbb{T}^{L}_{\mathrm{SL}(r)}(\chi=r(1-g))_{w}\subset D^{b}(\mathcal{M}_{\mathrm{SL}(r)}^{L})_{w}.

We also set

𝕋LPGL(r)(χ):={𝕋PGL(r)L(χ)r/2,l is odd and r is even𝕋PGL(r)L(χ)0,otherwise.\displaystyle\mathbb{T}^{L}_{\mathrm{PGL}(r)}(\chi):=\begin{cases}\mathbb{T}_{\mathrm{PGL}(r)}^{L}(\chi)_{r/2},&l\mbox{ is odd and }r\mbox{ is even}\\ \mathbb{T}_{\mathrm{PGL}(r)}^{L}(\chi)_{0},&\mbox{otherwise. }\end{cases}

We say that (r,w)(r,w) satisfies the BPS condition if (r,r(1g),w)(r,r(1-g),w) satisfies the BPS condition, or equivalently if (r,w+1gsp)(r,w+1-g^{\rm{sp}}) are coprime. This is also equivalent to either one of the following conditions

  1. (1)

    ll is even and (r,w)(r,w) are coprime,

  2. (2)

    ll is odd and (r,w)(r,w) are coprime with r2(mod4)r\not\equiv 2\pmod{4},

  3. (3)

    ll is odd and (r,w)(r,w) has divisibility 22 with r/2r/2 odd.

Corollary 8.20.

Suppose that l>2g2l>2g-2 and (r,w)(r,w) satisfies the BPS condition. Then there is an equivalence of spectra

(8.40) Ktop(𝕋LPGL(r)(w+1gsp))Ktop(𝕋LSL(r),w).\displaystyle K^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{PGL}(r)}(w+1-g^{\rm{sp}}))\stackrel{{\scriptstyle\sim}}{{\to}}K^{\rm{top}}(\mathbb{T}^{L}_{\mathrm{SL}(r),w}).
Proof.

For χ=r(1g)\chi=r(1-g), we have

χgsp+1{r/2(modr),l is odd and r is even0(modr),otherwise.\displaystyle-\chi-g^{\rm{sp}}+1\equiv\begin{cases}r/2\pmod{r},&l\mbox{ is odd and }r\mbox{ is even}\\ 0\pmod{r},&\mbox{otherwise. }\end{cases}

It follows that we have

𝕋PGL(r)L(w+1gsp)r(1g)gsp+1𝕋PGL(r)L(w+1gsp).\displaystyle\mathbb{T}_{\mathrm{PGL}(r)}^{L}(w+1-g^{\rm{sp}})_{-r(1-g)-g^{\rm{sp}}+1}\simeq\mathbb{T}_{\mathrm{PGL}(r)}^{L}(w+1-g^{\rm{sp}}).

Therefore we obtain the desired equivalence (8.40) from Theorem 8.19. ∎

Recall that we denoted by 𝕋PGL(r)(w):=𝕋PGL(r)L=ΩC(w)\mathbb{T}_{\mathrm{PGL}(r)}(w):=\mathbb{T}_{\mathrm{PGL}(r)}^{L=\Omega_{C}}(w),  𝕋SL(r),w:=𝕋SL(r),wL=ΩC\mathbb{T}_{\mathrm{SL}(r),w}:=\mathbb{T}_{\mathrm{SL}(r),w}^{L=\Omega_{C}}. We finally obtain the following result:

Theorem 8.21.

Suppose that (r,w)(r,w) are coprime. Then there is an equivalence

Ktop(𝕋PGL(r)(w))Ktop(𝕋SL(r),w).\displaystyle K^{\rm{top}}(\mathbb{T}_{\mathrm{PGL}(r)}(w))_{\mathbb{Q}}\stackrel{{\scriptstyle\sim}}{{\to}}K^{\rm{top}}(\mathbb{T}_{\mathrm{SL}(r),w})_{\mathbb{Q}}.
Proof.

Note that 1gsp1-g^{\rm{sp}} is divisible by rr in the case of L=ΩCL=\Omega_{C}, so

𝕋PGL(r)(w+1gsp)𝕋PGL(r)(w).\mathbb{T}_{\mathrm{PGL}(r)}(w+1-g^{\rm{sp}})\cong\mathbb{T}_{\mathrm{PGL}(r)}(w).

The claim follows from applying pBϕwBp_{B\ast}\phi_{w_{B}} to the equivalence (8.38), and using Proposition 8.5 and Proposition 8.18. ∎

References

  • [Ada74] J. F. Adams, Stable homotopy and generalised homology, Chicago Lectures in Mathematics, University of Chicago Press, Chicago, Ill.-London, 1974.
  • [AG15] D. Arinkin and D. Gaitsgory, Singular support of coherent sheaves and the geometric Langlands conjecture, Selecta Math. (N.S.) 21 (2015), no. 1, 1–199.
  • [Alp13] J. Alper, Good moduli spaces for Artin stacks, Ann. Inst. Fourier (Grenoble) 63 (2013), no. 6, 2349–2402.
  • [Ari13] D. Arinkin, Autoduality of compactified Jacobians for curves with plane singularities, J. Algebraic Geom. 22 (2013), no. 2, 363–388.
  • [AS89] M. Atiyah and G. Segal, On equivariant Euler characteristics, J. Geom. Phys. 6 (1989), no. 4, 671–677.
  • [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), 1287–1359.
  • [BBD82] A. Beilinson, J. Bernstein, and P. Deligne, Faisceaux pervers, Analysis and topology on singular spaces I, Asterisque 100 (1982), 5–171.
  • [Bla16] A. Blanc, Topological K-theory of complex noncommutative spaces, Compos. Math. 152 (2016), no. 3, 489–555.
  • [BNR89] A. Beauville, M.S. Narasimhan, and S. Ramanan, Spectral curves and the generalised theta divisor, J. Reine Angew. Math.  398 (1989), 169–179.
  • [Bri12] T. Bridgeland, An introduction to motivic Hall algebras, Adv. Math.  229 (2012), 102–138.
  • [BZN18] D. Ben-Zvi and D. Nadler, Betti geometric Langlands, Algebraic geometry: Salt Lake City 2015, Proc. Sympos. Pure Math., vol. 97.2, Amer. Math. Soc., Providence, RI, 2018, pp. 3–41.
  • [Dav] B. Davison, The integrality conjecture and the cohomology of preprojective stacks, arXiv:1602.02110.
  • [DHSM] B. Davison, L. Hennecart, and S. Schlegel Mejia, BPS Lie algebras for totally negative 2-Calabi-Yau categories and nonabelian Hodge theory for stacks, arXiv:2212.07668.
  • [DP12] R. Donagi and T. Pantev, Langlands duality for Hitchin systems, Invent. Math. 189 (2012), no. 3, 653–735.
  • [GPH13] O. García-Prada and J. Heinloth, The yy-genus of the moduli space of PGLn{\rm PGL}_{n}-Higgs bundles on a curve (for degree coprime to nn), Duke Math. J. 162 (2013), no. 14, 2731–2749.
  • [GPHS14] O. García-Prada, J. Heinloth, and A. Schmitt, On the motives of moduli of chains and Higgs bundles, J. Eur. Math. Soc. (JEMS) 16 (2014), no. 12, 2617–2668.
  • [GS] M. Groechenig and S. Shen, Complex K-theory of moduli spaces of Higgs bundles, arXiv:2212.10695.
  • [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.
  • [Hau22] T. Hausel, Enhanced mirror symmetry for Langlands dual Hitchin systems, ICM—International Congress of Mathematicians. Vol. 3. Sections 1–4, EMS Press, Berlin, 2022, pp. 2228–2249.
  • [HdlPn02] L. Hille and J. A. de la Peña, Stable representations of quivers, J. Pure Appl. Algebra 172 (2002), no. 2-3, 205–224. MR 1906875
  • [HLa] D. Halpern-Leistner, Derived Θ\Theta-stratifications and the DD-equivalence conjecture, arXiv:2010.01127.
  • [HLb] V. Hoskins and S. Pepin Lehalleur, Motivic mirror symmetry for Higgs bundles, arXiv:2205.15393.
  • [HL21] by same author, On the Voevodsky motive of the moduli space of Higgs bundles on a curve, Selecta Math. (N.S.) 27 (2021), no. 1, Paper No. 11, 37.
  • [HLP20] D. Halpern-Leistner and D. Pomerleano, Equivariant Hodge theory and noncommutative geometry, Geom. Topol. 24 (2020), no. 5, 2361–2433.
  • [HT03] T. Hausel and M. Thaddeus, Mirror symmetry, Langlands duality, and the Hitchin system, Invent. Math. 153 (2003), no. 1, 197–229.
  • [JS12] D. Joyce and Y. Song, A theory of generalized Donaldson-Thomas invariants, Mem. Amer. Math. Soc. 217 (2012), no. 1020, iv+199.
  • [KK] T. Kinjo and N. Koseki, Cohomological χ\chi-independence for Higgs bundles and Gopakumar-Vafa invariants, arXiv:2112.10053, to appear in JEMS.
  • [KM24] T. Kinjo and N. Masuda, Global critical chart for local Calabi-Yau threefolds, Int. Math. Res. Not. IMRN (2024), no. 5, 4062–4093.
  • [KS] M. Kontsevich and Y. Soibelman, Stability structures, motivic Donaldson-Thomas invariants and cluster transformations, arXiv:0811.2435.
  • [KS11] by same author, Cohomological Hall algebra, exponential Hodge structures and motivic Donaldson-Thomas invariants, Commun. Number Theory Phys. 5 (2011), no. 2, 231–352.
  • [Kuz11] A. Kuznetsov, Base change for semiorthogonal decompositions, Compos. Math. 147 (2011), no. 3, 852–876.
  • [KW07] A. N. Kapustin and E. Witten, Electric-Magnetic Duality and the Geometric Langlands Program, Communications in Number Theory and Physics 1 (2007), no. 1, 1–236.
  • [Lec08] F. Lecomte, Réalisation de Betti des motifs de Voevodsky, C. R. Math. Acad. Sci. Paris 346 (2008), no. 19-20, 1083–1086.
  • [Li21] M. Li, Construction of Poincaré sheaf on stack of rank 2 Higgs bundles, Adv. Math. 376 (2021), Paper No. 107437, 55.
  • [Lur] J. Lurie, Spectral algebraic geometry, https://www.math.ias.edu/ lurie/papers/SAG-rootfile.pdf.
  • [LW21] F. Loeser and D. Wyss, Motivic integration on the Hitchin fibration, Algebraic Geometry 8 (2021), 196–230.
  • [Mac62] I. G. Macdonald, Symmetric products of an algebraic curve, Topology 1 (1962), 319–343.
  • [May99] J. P. May, A concise course in algebraic topology, Chicago Lectures in Mathematics, University of Chicago Press, Chicago, IL, 1999.
  • [Mei] S. Meinhardt, Donaldson-Thomas invariants vs. intersection cohomology for categories of homological dimension one, arXiv:1512.03343.
  • [Mou19] T. Moulinos, Derived Azumaya algebras and twisted KK-theory, Adv. Math. 351 (2019), 761–803.
  • [MR] S. Mozgovoy and M. Reineke, Intersection cohomology of moduli spaces of vector bundles over curves, arXiv:1512.04076.
  • [MRV19a] M. Melo, A. Rapagnetta, and F. Viviani, Fourier-Mukai and autoduality for compactified Jacobians. I, J. Reine Angew. Math. 755 (2019), 1–65.
  • [MRV19b] by same author, Fourier-Mukai and autoduality for compactified Jacobians. II, Geometry and Topology 23 (2019), 2335–2395.
  • [MS21] D. Maulik and J. Shen, Endoscopic decompositions and the Hausel-Thaddeus conjecture, Forum Math. Pi 9 (2021), Paper No. e8, 49.
  • [MS22] by same author, On the intersection cohomology of the moduli of SLn{\rm SL}_{n}-Higgs bundles on a curve, J. Topol. 15 (2022), no. 3, 1034–1057.
  • [MS23] by same author, Cohomological χ\chi-independence for moduli of one-dimensional sheaves and moduli of Higgs bundles, Geom. Topol. 27 (2023), no. 4, 1539–1586.
  • [MSY] D. Maulik, J. Shen, and Q. Yin, Algebraic cycles and Hitchin systems, arXiv:2407.05177.
  • [Muk81] S. Mukai, Duality between D(X){D}({X}) and D(X^){D}(\hat{X}) with its application to picard sheaves, Nagoya Math. J.  81 (1981), 101–116.
  • [Ols07] M. Olsson, Sheaves on Artin stacks, J. Reine Angew. Math. 603 (2007), 55–112.
  • [Păd] T. Pădurariu, Noncommutative resolutions and intersection cohomology for quotient singularities, arXiv:2103.06215.
  • [PS23] M. Porta and F. Sala, Two dimensional categorified Hall algebras, J. Eur. Math. Soc. 25 (2023), no. 3, 1113–1205.
  • [PTa] T. Pădurariu and Y. Toda, Quasi-BPS categories for Higgs bundles, arXiv:2408.02168.
  • [PTb] T. Pădurariu and Y. Toda, Categorical DT/PT correspondence for local surfaces, arXiv:2211.12182.
  • [PTc] by same author, Quasi-BPS categories for K3 surfaces, arXiv:2309.08437.
  • [PTd] by same author, Quasi-BPS categories for symmetric quivers with potential, arXiv:2309.08425.
  • [PTe] by same author, Topological K-theory of quasi-BPS categories of symmetric quivers with potential, arXiv:2309.08432.
  • [PT24] by same author, Categorical and K-theoretic Donaldson-Thomas theory of 3\mathbb{C}^{3} (part I), Duke Math. J.  173 (2024), 1973–2038.
  • [Sac19] G. Saccà, Relative compactified Jacobians of linear systems on Enriques surfaces, Trans. AMS.  371 (2019), 7791–7843.
  • [ŠdB17] Š. Špenko and M. Van den Bergh, Non-commutative resolutions of quotient singularities for reductive groups, Invent. Math. 210 (2017), no. 1, 3–67.
  • [Sim97] C. Simpson, The Hodge filtration on nonabelian cohomology, Algebraic geometry—Santa Cruz 1995, Proc. Sympos. Pure Math., vol. 62, Part 2, Amer. Math. Soc., Providence, RI, 1997, pp. 217–281.
  • [SYZ01] A. Strominger, S. T. Yau, and E. Zaslow, Mirror symmetry is TT-duality, Winter School on Mirror Symmetry, Vector Bundles and Lagrangian Submanifolds (Cambridge, MA, 1999), AMS/IP Stud. Adv. Math., vol. 23, Amer. Math. Soc., Providence, RI, 2001, pp. 333–347.
  • [Tho88] R. W. Thomason, Equivariant algebraic vs. topological KK-homology Atiyah-Segal-style, Duke Mathematical Journal 56 (1988), no. 3, 589 – 636.
  • [Tod] Y. Toda, Multiple cover formula of generalized DT invariants II: Jacobian localizations, arXiv:1108.4993.
  • [Tod14] by same author, Multiple cover formula of generalized DT invariants I: parabolic stable pairs, Adv. Math.  257 (2014), 476–526.
  • [Tod24] by same author, Categorical Donaldson-Thomas theory for local surfaces, Lecture Notes in Mathematics 2350 (2024), 1–309.

Tudor Pădurariu: Sorbonne Université and Université Paris Cité, CNRS, IMJ-PRG, F-75005 Paris, France.
E-mail address:
padurariu@imj-prg.fr

Yukinobu Toda: Kavli Institute for the Physics and Mathematics of the Universe (WPI), University of Tokyo, 5-1-5 Kashiwanoha, Kashiwa, 277-8583, Japan.
E-mail address:
yukinobu.toda@ipmu.jp