1. Introduction
1.1 Definitions of Segre and Verlinde invariants
Let
$Y$
be a smooth quasi-projective variety. For a torsion-free sheaf
$E$
, the Quot scheme
$\textrm {Quot}_Y(E,n)$
parameterizes quotients

such that

When
$Y={\mathbb{C}}^2$
is a toric surface or
$Y={\mathbb{C}}^4$
is a toric Calabi–Yau 4-fold, we shall define equivariant Segre and Verlinde invariants on these Quot schemes and find relations between them. In the non-equivariant case when
$Y$
is a smooth projective curve or surface, or a Calabi–Yau 4-fold, such relations are called Segre–Verlinde correspondences and are studied in [Reference Arbesfeld, Johnson, Lim, Oprea and PandharipandeAJL+21, Reference Behrend and FantechiBoj21b, Reference BojkoBoj21a, Reference Göttsche and MellitGM22, Reference Göttsche and KoolGK22, Reference Johnson, Oprea and PandharipandeJOP21, Reference Marian, Oprea and PandharipandeMOP21].
For a vector bundle
$V$
on
$Y$
, the Segre and Verlinde invariants associated to it are constructed using the tautological bundle

on
$\textrm {Quot}_Y(E,n)$
. Here

are projections, and
$\mathcal{F}$
is the universal quotient sheaf. This definition extends to the Grothendieck groups and associates to each
$\alpha \in K^0(Y)$
the tautological class
$\alpha ^{[n]}\in K^0(\textrm {Quot}_Y(E,n))$
.
1.1.1 Equivariant invariants on Hilbert schemes of surfaces.
Let
$S$
be a toric quasi-projective surface with an action of
${\mathsf{T}}=({\mathbb{C}}^*)^2$
. Before introducing the equivariant invariants, we give a brief summary of equivariant cohomology and equivariant integration with a further background in § 2.1. Given a
$\mathsf{T}$
-representation
$V$
, we define its equivariant characteristic classes by considering the associated bundle

and taking its characteristic classes in
$H^*(B{\mathsf{T}})=H^*_{\mathsf{T}}({\rm pt})$
. Denote by
$c^{\mathsf{T}},s^{\mathsf{T}},e_{\mathsf{T}},\textrm {ch}_{\mathsf{T}},{\textrm {td}}_{\mathsf{T}}$
the equivariant Chern class, Segre class, Euler class, Chern character, and Todd class, respectively. From here forward, we will always use equivariant classes for toric varieties, and we will omit the torus
$\mathsf{T}$
from the notations when it is clear from the context. In our equivariant setting, the integral
$\int _{Y}$
refers to the equivariant push-forward defined in (4), and
$\chi$
denotes the equivariant Euler characteristic of a K-theory class. By Euler characteristic, we always mean the holomorphic Euler characteristic.
Consider
$E={\mathcal{O}}_S$
, so
$\textrm {Quot}_S({\mathcal{O}}_S,n)$
is the smooth Hilbert scheme
$\textrm {Hilb}^n(S)$
parameterizing ideal sheaves of 0-dimensional subschemes of
$S$
of length
$n$
. The action of
$\mathsf{T}$
on
$S$
lifts to
$\textrm {Hilb}^n(S)$
, giving an equivariant structure to
${\alpha }^{[n]}$
for any equivariant K-theory class
${\alpha }\in K_{\mathsf{T}}(S)$
.
In the context of Donaldson invariants, the Segre and Chern series first appeared in [Reference TyurinTyu94], and are, respectively, given by

We use notation which resembles that of [Reference Göttsche and MellitGM22] as we regularly rely on the ideas presented there. Note that
$c(-E)=1/c(E)$
for a vector bundle
$E$
which, together with [Reference FultonFul13, § 3.2], implies that
$c(\alpha ^{[n]})=s(-\alpha ^{[n]})$
.
The Verlinde series was originally defined for moduli spaces of bundles on curves in the context of conformal field theory [Reference VerlindeVer88], and has since been extended to moduli spaces of torsion-free sheaves on surfaces in [Reference Göttsche and KoolGK22]. For Hilbert schemes on surfaces, the Verlinde series was considered in[Reference Ellingsrud, Göttsche and LehnEGL99, § 5]. To be consistent with the invariants above, we adopt the notation

1.1.2 Virtual invariants on Quot schemes.
In general, Quot schemes are not smooth, in which case we do not have a deformation invariant fundamental class to integrate against. One way to resolve this issue is to work with a virtual fundamental class
$[\textrm {Quot}_Y(E,n)]^{{\rm vir}}$
.
For a smooth projective surface
$S$
and a vector bundle
$E$
of rank
$N$
, a perfect obstruction theory for
$\textrm {Quot}_Y(E,n)$
of virtual dimension
$nN$
was constructed in [Reference Marian, Oprea and PandharipandeMOP15, Lemma 1] for trivial bundles and was generalized to arbitrary bundles in [Reference StarkSta24, § 2.3]. We remark that the more general case of coherent sheaves is covered in [Reference LimLim20, § 1.3] but will not be pursued here. To this perfect obstruction theory, one may associate a virtual fundamental class
$[\textrm {Quot}_S(E,n)]^{{\rm vir}}$
(following [Reference Behrend and FantechiBF98, Reference Li and TianLT96]), and a virtual structure sheaf
${\mathcal{O}}^{{\rm vir}}$
(first proposed in [Reference Behrend and FantechiBF98, Remark 5.4]). The virtual invariants are defined in a similar way to what has gone before, with the usual fundamental class replaced by the virtual class
$[\textrm {Quot}_S(E,n)]^{{\rm vir}}$
and the Euler characteristic replaced by the virtual Euler characteristic
$\chi ^{{\rm vir}}(-):=\chi (-\otimes {\mathcal{O}}^{{\rm vir}})$
, which are both invariant under deformations of the complex structure of
$S$
.
Unlike surfaces and Fano 3-folds, the usual obstruction theory for a Quot scheme of a Calabi–Yau 4-fold
$X$
is not perfect, so the previous method does not induce a virtual fundamental class. However, using the method of [Reference RicolfiRic20] and tools from [Reference Oh and ThomasOT20], a virtual class
$[\textrm {Quot}_X(E,n)]^{{\rm vir}}_{o(\mathcal{L})}\in H_{2nN}(\textrm {Quot}_X(E,n),{\mathbb{Z}})$
was constructed in [Reference BojkoBoj21a, § 2.1] when
$X$
is a strict Calabi–Yau 4-fold and
$E$
a simple rigid locally free sheaf. Here,
$\mathcal{L}$
denotes the determinant line bundle
$\det {\bf R}\mathscr{H}om_{q}(\mathcal{I},\mathcal{I})$
on
$\textrm {Quot}_X(E,n)$
, where
$\mathcal{I}$
is the universal subsheaf. As indicated by the subscript, this class is dependent on some choice of orientation
$o(\mathcal{L})$
, that is, a choice of a square root of the isomorphism

induced by Serre duality. There is also no canonical virtual structure sheaf, and instead we have the ‘twisted’ virtual structure sheaf
$\hat {{\mathcal{O}}}^{{\rm vir}}$
defined in [Reference Oh and ThomasOT20, (84), § 8]. The motivation for this notation comes from its relation to the twisted virtual structure sheaf
$\hat {{\mathcal{O}}}^{{\rm vir}}_{\textrm {NO}}$
of Nekrasov and Okounkov [Reference Nekrasov and OkounkovNO14]. Without going into details,
$\hat {{\mathcal{O}}}^{{\rm vir}}_{\textrm {NO}}$
is defined from the usual virtual structure sheaf
${\mathcal{O}}^{{\rm vir}}$
on a moduli space with obstruction theory by twisting with
${\rm det}^{\frac {1}{2}}\big (K^{{\rm vir}}\big )$
, where
$K^{{\rm vir}}$
is the virtual cotangent bundle. Square roots of determinant line bundles appear also in the construction of
$\hat {{\mathcal{O}}}^{{\rm vir}}$
, but this time they are used to make the result independent of the choices made. Another distinguishing feature is that

is not an integer class, precisely because of the square root determinant line bundles. To obtain integer invariants, the first author, in [Reference Behrend and FantechiBoj21b, § 5.3] and [Reference BojkoBoj21a, § 1.4], introduced the untwisted virtual structure sheaf

The notation
$(-)^{\frac {1}{2}}$
denotes the unique square root of a line bundle in
${\mathbb{Z}}[2^{-1}]$
-valued K-theory (as established in [Reference Oh and ThomasOT20, Remark 5.2]). In § 5.2, we motivate this definition further by showing, in the case
$X={\mathbb{C}}^4$
, that
$\mathsf{E}^{-\frac {1}{2}}$
naturally appears in the construction of
$\hat {{\mathcal{O}}}^{{\rm vir}}$
as the only term that is not an integer class. As a consequence, we prove the following.
Proposition 1.1 (Proposition 5.3). The untwisted virtual structure sheaf is integral:

Before defining equivariant virtual invariants on
$Y={\mathbb{C}}^d$
,
$d=2,4$
, we describe the torus action used on
$\textrm {Quot}_Y(E,n)$
. Let

act on
$Y$
naturally, where we quotient by the subgroup
$\langle \sim \rangle =\langle t_1t_2t_3t_4\rangle$
when
$d=4$
. Let
$ {\mathsf{T}}_1=({\mathbb{C}}^*)^N=\{(y_1,\ldots ,y_N):y_i\neq 0\}$
and
$E=\oplus _{i=1}^N{\mathcal{O}}_Y\langle y_i\rangle$
be the
${\mathsf{T}}_1$
-equivariant bundle of rank
$N$
with weights
$y_1,\ldots , y_N$
. This induces a
${\mathsf{T}}_0\times {\mathsf{T}}_1$
-action on
$\textrm {Quot}_Y(E,n)$
by acting on the middle term of the sequence

Let
$\alpha \in K_{{\mathsf{T}}_0}(Y)$
, then we can write

where
$v_1,\ldots ,v_{r+s}$
are its
${\mathsf{T}}_0$
-weights. However, instead of thinking of the
$v_i$
as
${\mathsf{T}}_0$
-weights, we would like to view them as generic parameters. Therefore we introduce an additional torus
${\mathsf{T}}_2=({\mathbb{C}}^*)^{r+s}$
acting on
${\mathbb{C}}^r\times {\mathbb{C}}^s$
. Set
${\mathsf{T}}:={\mathsf{T}}_0\times {\mathsf{T}}_1\times {\mathsf{T}}_2$
and write

where we quotient by the ideals
$(\sim )=(t_1t_2t_3t_4-1)$
and
$(\sim )=(\lambda _1+\lambda _2+\lambda _3+\lambda _4)$
, respectively, when
$d=4$
. Here
$\lambda _i,m_i,w_i$
are, respectively, the equivariant first Chern classes of the weights
$t_i,y_i,v_i$
.
In the surface case, the equivariant virtual invariants are defined using the virtual equivariant localization of [Reference Graber and PandharipandeGP97, Reference Ciocan-Fontanine and KapranovCFK09] in an analogous way to the non-virtual case. For toric Calabi–Yau 4-folds, the equivariant Donaldson invariants were first introduced by Cao and Leung in [Reference Cao and LeungCL14, § 8]; the K-theoretic equivariant invariants were predicted by Nekrasov [Reference NekrasovNek20] and Cao, Kool and Monavari [Reference Cao, Kool and MonavariCKM22]. They were formalized by Oh and Thomas [Reference Oh and ThomasOT20, § 7], using the twisted virtual structure sheaf
$\hat {{\mathcal{O}}}^{{\rm vir}}$
and their virtual equivariant localization. The equivariant virtual Segre, Chern and Verlinde series for
$\alpha$
on
$\textrm {Quot}_Y(E,n)$
are, respectively,

Here the symbols
$\int _{[\textrm {Quot}_Y(E,n)]^{{\rm vir}}},\chi ^{{\rm vir}}$
are defined using virtual versions of (4). See Definition 3.2 for their precise definitions.
Remark 1.2. It is important to note that the above definitions, when
$Y={\mathbb{C}}^4$
, are dependent on a choice of orientation
$o(\mathcal{L})$
. This induces signs at each fixed point
$Z\in \textrm {Quot}_X(E,n)^{\mathsf{T}}$
, which are suppressed in the notation. We denote the sign at
$Z$
to be
$(-1)^{o(\mathcal{L})|_{Z}}$
and call
$o(\mathcal{L})$
a choice of signs.
When
$N=1$
, the weight on
$E={\mathcal{O}}_Y\langle y_1\rangle$
is not necessary as this extra action does not affect the fixed locus, so we sometimes ignore it by setting
$y_1=1$
. By definition, the coefficients of the Chern and Segre series are rational functions in the cohomological parameters
$\lambda _1,\ldots ,\lambda _d,m_1,\ldots ,m_N,w_1,\ldots ,w_{r+s}$
. For the Verlinde series, they are rational functions in the K-theoretic parameters
$t_1,\ldots ,t_d,y_1\ldots ,y_N,v_1\ldots ,v_{r+s}$
. Using the identification of Remark 2.4, they can be viewed as functions in the cohomological parameters as well. Define

to be the part with total degree
$i$
in those variables for
$\ast \in \{{\mathcal{S}},{\mathcal{C}},{\mathcal{V}}\}$
. More precisely, by restricting a multi-variable function to a certain degree, we mean the following.
Definition 1.3. Let
$f(z_1,\ldots ,z_k)$
be a function in the ring of fractions of
${\mathbb{C}}[\![z_1,\ldots ,z_k]\!]$
. Consider the formal Laurent series expansion of
$f(bz_1,\ldots ,bz_k)$
in the variable
$b$
:

For
$i\in {\mathbb{Z}}$
, the part of
$f$
with total degree
$i$
is

1.1.3 Reduced invariants on Quot schemes of surfaces.
When
$S$
is a
$K$
-trivial surface, the obstruction on
$\textrm {Quot}_S(E,n)$
contains a trivial summand making
$e(T^{{\rm vir}})$
vanish and, as a result, the invariants all vanish. We may instead consider the reduced classes and invariants from Gromov–Witten theory and stable pair theory [Reference Kool and ThomasKT14], which have been employed to study the enumerative geometry of Hilbert schemes in, for instance, [Reference Gholampour, Sheshmani and YauGSY17]. In the case of
$\textrm {Quot}_S(E,n)$
, a reduced perfect obstruction theory can be obtained by removing one copy of
${\mathcal{O}}_{\textrm {Quot}_S(E,n)}$
from the usual obstruction for Quot schemes. The equivariant analogue of a
$K$
-trivial surface would be
$S={\mathbb{C}}^2$
with the action of the 1-dimensional torus
${\mathsf{T}}_0=\{(t_1,t_2):t_1t_2=1\}$
. Let
$T^{{\rm red}}$
be the virtual tangent bundle obtained from the reduced obstruction theory. For
$E=\oplus _{i=1}^N{\mathcal{O}}_S\langle y_i\rangle$
and
${\alpha }=[\oplus _{i=1}^r{\mathcal{O}}_S\langle v_i\rangle ]-[\oplus _{i=r+1}^s{\mathcal{O}}_S\langle v_i\rangle ]$
, we define the reduced Segre, Chern, and Verlinde series to be, respectively,

Again, the symbol
$\int _{[\textrm {Quot}_S(E,n)]^{{\rm red}}}$
is notational, and the precise definition is given in § 3.4.
1.2 Summary of results for surfaces
1.2.1 Computation of Chern series.
Consider the case
$S={\mathbb{C}}^2$
with the
${\mathsf{T}}=({\mathbb{C}}^*)^2$
-action. Using the tools from [Reference Göttsche and MellitGM22], we are able to obtain the virtual Chern series of line bundles for Hilbert schemes, as we show in § 4.1. Note that in the non-equivariant setting, this can be retrieved from [Reference Oprea and PandharipandeOP22, Corollary 15].
Corollary 1.4.
Let
$S={\mathbb{C}}^2$
, and let
$L={\mathcal{O}}_S\langle v_1\rangle$
be a
$\mathsf{T}$
-equivariant line bundle over
$S$
. We have

By extracting the part with the lowest total degree in
$\lambda _1,\lambda _2,w_1$
, we obtain the following 2-dimensional analogue to the Donaldson–Thomas partition function for
${\mathbb{C}}^3$
[Reference Maulik, Nekrasov, Okounkov and PandharipandeMNOP06b, Theorem 1], or Cao and Kool’s formulation of Nekrasov’s conjecture for
${\mathbb{C}}^4$
[Reference Cao and KoolCK17, Appendix B].
Corollary 1.5 (Corollary 4.2). For
$S={\mathbb{C}}^2$
, the following equality holds:

Remark 1.6. Suppose
$F(q)$
is a power series. If
$Y$
is compact and
$\sigma \in H^{2\dim Y}(Y;{\mathbb{Q}})$
, then
$\int _Y \sigma \in {\mathbb{Q}}$
, and we make the identification
$F(q)^\sigma :=F(q)^{\int _Y \sigma }$
. In the non-compact case
$Y={\mathbb{C}}^2$
or
$Y={\mathbb{C}}^4$
, working with the equivariant push-forward
$\int _Y:H_{\mathsf{T}}^*(Y)\rightarrow H_{\mathsf{T}}^*({\rm pt})_{{\rm loc}}$
, we will use a different notation. Suppose
${\gamma }\in H^*_{\mathsf{T}}({\rm pt})_{{\rm loc}}$
and
$F(0)=1$
. The expression
$F(q)^{{\gamma }}$
is to be interpreted as

Since
$F(0)=1$
, we have
$[q^0]\log F(q)=0$
, so the right-hand side does indeed belong to the ring
$H^*_{\mathsf{T}}({\rm pt})_{{\rm loc}}[\![q]\!]$
. We will be dealing exclusively with expressions of the form
$F(q)^{\int _Y\sigma }$
for
$\sigma \in H_{\mathsf{T}}^*(Y)$
.
1.2.2 Universal series expressions.
A common approach to find closed formulas for the Segre and Verlinde series is to compute their universal series. In the non-virtual case, for Hilbert schemes of projective surfaces, the main result of [Reference Ellingsrud, Göttsche and LehnEGL99] shows that the Segre and Verlinde invariants admit the following universal series expressions:


see [Reference Ellingsrud, Göttsche and LehnEGL99, Theorem 5.3], and [Reference LehnLeh99, § 4.3]. Here, the products in the exponents refer to intersection products. The series
$A_i(x),B_i(x)$
are universal, in the sense that they only depend on
$\alpha$
through its rank and are independent of the surface. Explicit formulas for these series were conjectured and computed in [Reference LehnLeh99, Reference Marian, Oprea and PandharipandeMOP17, Reference Marian, Oprea and PandharipandeMOP21, Reference Ellingsrud, Göttsche and LehnEGL99, Reference Göttsche and MellitGM22]. The Segre–Verlinde correspondence in this case concerns the relations between
$A_i$
and
$B_i$
. It was first proposed by Johnson and Marian, Oprea and Pandharipande in relation to the study of Le Potier’s strange duality [Reference Marian, Oprea and PandharipandeMOP17, Reference JohnsonJoh18], and was recently proved by Göttsche and Mellit [Reference Göttsche and MellitGM22].
For virtual invariants on Quot schemes of a smooth projective surface
$S$
and torsion-free sheaf
$E$
, the universal series expressions are given by [Reference BojkoBoj21a, Theorem 1.2]:

The explicit formulas are computed in [Reference Arbesfeld, Johnson, Lim, Oprea and PandharipandeAJL+21, Theorem 17] for
$A_1,A_2,B_1,B_2$
and in [Reference BojkoBoj21a, Theorem 1.2] for
$A_3,B_3$
.
Unlike the compact case where the invariants are simply numbers, the equivariant invariants can contain terms of various degrees in
$H_{\mathsf{T}}^*({\rm pt})_{{\rm loc}}$
. This is reflected in the following theorem, where the virtual equivariant Segre and Verlinde invariants are written as infinite products of series labeled by partitions. The notation for partitions is given in § 2.2. For a partition
$\mu$
and a K-theory class
$\alpha$
, we write

Theorem 1.7 (Theorem 3.5). Let
$S={\mathbb{C}}^2$
. For any
$r\in {\mathbb{Z}}$
,
$N\gt 0$
, there exist universal power series
$A_{\mu ,\nu ,\xi }(q),B_{\mu ,\nu ,\xi }(q)$
, dependent on
$N$
and
$r$
, such that for
$E=\oplus _{i=1}^{N}{\mathcal{O}}_S\langle y_i\rangle$
and
${\alpha }\in K_{\mathsf{T}}(S)$
of rank
$r$
, the equivariant virtual Segre and Verlinde series on
$\textrm {Quot}_S(E,n)$
can be written as the following infinite products:

The series in the above expressions are universal, in the sense that they depend on the input
$\alpha$
only by its rank
$r$
. Sometimes, for clarity, we will add superscripts
$N,r$
to indicate the ranks of
$E$
and
$\alpha$
. On the right-hand side of the above expressions, the exponents
$\int _{S}c_\mu ({\alpha }) c_\nu (S) c_\xi (E)c_1(S)$
are computed by (4). They are thus homogeneous rational functions in

of degree
$|\mu |+|\nu |+|\xi |-1$
. Degree 0 terms occur when one of
$\mu ,\nu ,\xi$
is the partition
$(1)$
and the rest are the empty partition
$(0)$
. The argument of § 3.2 shows that the series with degree 0 exponents are necessarily equal to the series from the projective case, that is

The universal series expressions of the reduced invariants take a much simpler form; as opposed to having series exponentiated to some powers of cohomology classes, we have the following additive expressions.
Theorem 1.8 (Theorem 3.11). When
$S={\mathbb{C}}^2$
, the equivariant reduced Segre and Verlinde series for
$E=\oplus _{i=1}^N{\mathcal{O}}_S\langle y_i\rangle$
and
${\alpha }\in K_{\mathsf{T}}(S)$
are

where
$A_{\mu ,\nu ,\xi }, B_{\mu ,\nu ,\xi }$
and
$C_{\mu ,\nu ,\xi }$
are the same series as in Theorem 1.7
.
1.2.3 Virtual Segre–Verlinde correspondence.
When
$Y$
is compact, the virtual Segre–Verlinde correspondence has been proved, for compact surfaces and Calabi–Yau 4-folds [Reference BojkoBoj21a, Theorem 1.6] and for torsion-free sheaves
$E$
, to be

As a corollary of this result, Theorem 1.7 and the relations (3), we first prove the following ‘weak’ equivariant Segre–Verlinde correspondence.
Corollary 1.9 (Corollary 3.7). In the setting of Theorem 1.7 , we have the following correspondence

whenever one of
$\mu ,\nu ,\xi$
is
$(1)$
and the other two are
$(0)$
. In particular, the degree 0 part satisfies

for some terms
$f_n\in H_{\mathsf{T}}^{2n-2}({\rm pt})$
dependent on
$\alpha$
through its rank and Chern classes.
This is weak, in the sense that only the series whose powers are degree 0 satisfy the usual correspondence, while computations for small values of
$n$
show that the terms
$f_n$
can be non-zero, so the naive correspondence does not hold in all degrees. One might instead ask whether there are any other relations between the series in all cohomological degrees, generalizing the above corollary.
We are able to give a complete answer relating
$B_{\mu ,\nu ,\xi }$
and
$C_{\mu ,\nu ,\xi }$
once the sizes of the partitions
$|\mu |,|\xi |$
are fixed and
$\nu = (0)$
. This is represented in Figure 1, but before stating our result, some further notation is needed. Given any partition
$\mu$
, integer
$a\in {\mathbb{Z}}$
and
$n\gt 0$
, the binomial coefficients for
$\mu$
are

and the downward factorial of
$a$
by
$n$
is



Figure 1. Relating Segre and Verlinde invariants by the common function
$\varphi$
.
Theorem 1.10 (Theorem 4.3). For rank
$r\geqslant 0$
and integers
$k_1,k_2\geqslant 0$
with
$k:=k_1+k_2$
, the universal series of Theorem 1.7 satisfy

Furthermore, we have

which can be compared to the identities above by replacing
$r$
with
$-r$
.
In particular, setting
$k=1$
, the first two equalities of this theorem give


This is consistent with the Segre–Verlinde correspondence in degree 0 from Corollary 1.9. Going one degree higher by setting
$k=2$
, we get the following correspondence.
Corollary 1.11 (Corollary 4.4). For rank
$r\geqslant 0$
, the universal series of Theorem 1.7 satisfy the following correspondences

It is natural to expect that applying Lagrange inversion to the expressions in Theorem 1.10 will simplify the formulae. In § 4.3, we use this to phrase the higher degree Segre and Verlinde series in terms of applying differential operators of comparable form to a single function
${\varphi }(t)=\log (1+t)$
with common variable changes except for the usual sign
$q\leadsto (-1)^Nq$
.
Theorem 1.12 (Theorem 4.6). Let
${\varphi }(t)=\log (1+t)$
and
$\psi (t)=Nt^{-1}+r(1+t)^{-1}$
. Define the differential operator

Furthermore, use the notation

for
$k\geqslant 0$
where
$D_\psi ^{-1}(-)$
denotes integrating
$\psi \cdot (-)$
assuming a constant term 0. In the setting of Theorem 1.10
, we have the following relations

where
$H_i$
are the Newton–Puiseux solutions to
$H_i^N=q(1+H_i)^{-r}$
.
If the rank
$r$
of
$\alpha$
is negative, then the above identities no longer hold. For such
$\alpha$
, we may instead consider the Chern series and
$C_{\mu ,(0),\xi }^{N,r}$
. Because of the previously discussed identity
$c(\alpha ) = s(-\alpha )$
, this is equivalent to working with
$A_{\mu ,(0),\xi }^{N,-r}(q)$
, which satisfies the above theorem for
$-r\gt 0$
.
Setting
$k=1$
allows us to retrieve the degree 0 parts for both the Segre and Verlinde invariants using the function
$\varphi$
, and hence the Segre–Verlinde correspondence of Corollary 1.9. When
$k=2$
, we get a relation in the degree 1 part which is equivalent to Corollary 1.11. In general, the Segre and Verlinde relations are obtained by applying
$D_{\mathcal{S}}^{(k)}$
and
$D_{\mathcal{V}}^{(k)}$
to
$\varphi$
. Figure 1 summarizes these observations.
Remark 1.13. The previous results only deal with the
$\nu =(0)$
case. When
$\nu =(1)$
, techniques similar to those of § 4.2 can be applied to the term [Reference BojkoBoj21a, (4.24)], if an explicit expression for it is found. For
$|\nu |\gt 1$
, the same techniques no longer work because of the limitations of Lemma 4.1.
A symmetry for virtual Segre series is also proved in [Reference BojkoBoj21a, Theorem 1.7] for the compact case, and states

for torsion-free sheaves
$E$
and
$V$
of rank
$N$
and
$r$
respectively. As in Corollary 1.9, we have the following weak version of this symmetry in the equivariant case.
Corollary 1.14.
In the setting of Theorem 1.7
, for
${\alpha }=V=\oplus _{i=1}^{r}{\mathcal{O}}_S\langle v_i\rangle$
, we have the following symmetry

whenever one of
$\mu ,\nu ,\xi$
is
$(1)$
and the other two are
$(0)$
. In degree 0, we have

for some terms
$g_n\in H_{\mathsf{T}}^{2n-2}({\rm pt})$
.
Motivated by our study of the higher degree Segre series in § 4.4 which, as we see after using Theorem 1.10, are symmetric under interchanging
$\mu$
and
$\xi$
, the partitions keeping track of Chern classes of
$\alpha$
and
$E$
, respectively, we conjecture the following strong Segre symmetry.
Conjecture 1.15. Let
$r,N\gt 0$
,
$S={\mathbb{C}}^2$
. For
$E=\oplus _{i=1}^N{\mathcal{O}}_S\langle y_i\rangle ,V=\oplus _{i=1}^r{\mathcal{O}}_S\langle v_i\rangle$
, we have the following symmetry

Using a program, we checked that this condition holds in the cases summarized in (28).
1.2.4 Reduced Segre–Verlinde correspondence.
For the reduced invariants on
$S={\mathbb{C}}^2$
, let us denote by
${\mathcal{S}}_i^{{\rm red}}$
and
${\mathcal{V}}_i^{{\rm red}}$
the degree
$i$
parts of the reduced Segre and Verlinde series, respectively, in the sense of Definition 1.3. In this setting, the fact that the series from Theorem 1.8 are the same as those from Theorem 1.7 allows us to obtain relations from the non-reduced case. For example, the Segre–Verlinde correspondence of Corollary 1.9 and symmetry of Corollary 1.14 give the following reduced Segre–Verlinde correspondence and symmetry in degree
$-1$
.
Corollary 1.16.
We have the following correspondence in degree
$-1$
:

When
${\alpha }=V$
is an equivariant vector bundle, we have the following symmetry:

Let
$\alpha \in K_{\mathsf{T}}(S)$
with rank
$r\gt 0$
. Write
$c_1\lambda =c_1({\alpha })$
and
$c_2\lambda^2=c_2({\alpha })$
for some
$c_1,c_2\in {\mathbb{Q}}$
. Then we can use the Theorem 1.8 to show that, for some series
$A_2(q),A_1(q),A_0(q),B_1(q),B_0(q)$
, dependent on
$r$
and
$N$
,

These degree 0 terms should correspond to the ones for compact K3-surfaces and their reduced invariants. In particular, one can use the results obtained in Theorems 1.10 and 1.12 and the ones below to predict the Segre and Verlinde series in the compact case and describe their correspondences.
Corollary 1.11 provides relations between
$A_1,A_2$
and
$B_1$
, and we have the following formula for
$B_1(q)$
obtained from Theorem 1.10 by setting
$k_1=2,k_2=0$
.
Corollary 1.17.
The series
$B_1(q)$
is explicitly given by

for
$n,r\gt 0$
.
We give some conjectural formulas for
$A_0,B_0$
in terms of
$A_1,B_1$
when
$N=1$
, which we have checked for
$n\leqslant 20,r\lt 5$
.
Conjecture 1.18. When
$N=1$
and
$r\in {\mathbb{Z}}$
, we have

When
$N=1$
,
$r\lt -1$
and
$n\gt 1$
, we have

1.3 Correspondence for 4-folds and other observations
1.3.1 Segre–Verlinde correspondence.
Since all toric Calabi–Yau 4-folds are non-compact, we do not know how the invariants in the non-compact case relate to the ones in the compact case. We have seen for the surface case that the exponents on the universal series have a factor of
$c_1(S)$
. Considering the series given in [Reference Behrend and FantechiBoj21b, Proposition 4.13] and [Reference BojkoBoj21a, (3.38)], together with § 5.4, we see that this term should be replaced by
$c_3(X)$
in the 4-fold case. Motivated by Corollaries 1.9 and 1.14, we conjecture a weak Segre–Verlinde correspondence and symmetry for
$X={\mathbb{C}}^4$
.
Conjecture 1.19. Let
$X={\mathbb{C}}^4$
,
$E=\oplus _{i=1}^{N}{\mathcal{O}}_X\langle y_i\rangle$
,
$V=\oplus _{i=1}^{r}{\mathcal{O}}_X\langle v_i\rangle$
, and
${\alpha }\in K_{\mathsf{T}}(X)$
, then for some choice of signs
$o(\mathcal{L})$
, we have the following symmetry and correspondence

for some terms
$F_n\in H_{\mathsf{T}}^{4n-6}({\rm pt})$
dependent on
$\alpha$
through its rank and Chern classes.
We checked both the correspondence and the symmetry using a computer program, as summarized in (40) and (41).
1.3.2 Nekrasov’s conjectures.
In [Reference Cao and KoolCK17, Appendix B], Cao and Kool gave a mathematical formulation of Nekrasov’s conjecture [Reference NekrasovNek20, § 5]. We generalize this to Quot schemes of
${\mathbb{C}}^2$
and
${\mathbb{C}}^4$
as follows.
Conjecture 1.20. Let
$Y={\mathbb{C}}^d$
,
$E=\oplus _{i=1}^{r+1}{\mathcal{O}}_Y\langle y_i\rangle$
, and
$V=\oplus _{i=1}^r{\mathcal{O}}_Y\langle v_i\rangle$
. When
$d=2$
, or
$d=4$
with some choice of signs
$o(\mathcal{L})$
, we have

Note that when
$N=1$
and
$Y={\mathbb{C}}^2$
, this is exactly Corollary 1.5. When
$Y={\mathbb{C}}^4$
, we shall show that this conjecture is a consequence of Nekrasov and Piazzalunga’s conjecture [Reference Nekrasov and PiazzalungaNP19, § 2.5] using a Quot scheme version of Cao, Kool and Monavari’s cohomological limit [Reference Cao, Kool and MonavariCKM22, Appendix A] in Proposition 5.9. For
$Y={\mathbb{C}}^2$
, we check it for
$N=2,3,4,5$
up to and including
$n=7,4,2,2$
, respectively.
Since
$[\textrm {Quot}_Y(E,n)]^{{\rm vir}}$
has virtual dimension
$nN$
, we have
$C^N_Y(V;q)=1$
when
$N\gt r$
in the compact case, simply for degree reason, which means the corresponding Verlinde series is also trivial because of the Segre–Verlinde correspondence. In the non-compact case, the above conjectures suggest that they may contain negative-degree terms. However, with a computer calculation for
$Y={\mathbb{C}}^2$
,
$N=2,3,4,5$
and all possible
$r$
, up to and including
$n=8,5,2,2$
, we see a complete vanishing when
$N\gt r+1$
for Chern numbers, and when
$r\lt N$
for rank
$-r$
Verlinde numbers.
Conjecture 1.21. Let
$Y={\mathbb{C}}^d$
,
$N\gt 1$
, and
$E=\oplus _{i=1}^{N}{\mathcal{O}}_Y\langle y_i\rangle$
. When
$d=2$
, or
$d=4$
with some choice of signs, we have, for
$r=0,1,\ldots ,N-2$
and
$V=\oplus _{i=1}^r{\mathcal{O}}_Y\langle v_i\rangle$
,

Furthermore, for
$r=1,\ldots ,N-1$
, we have

In Proposition 5.9, we also show that the Chern series part of this conjecture for
$d=4$
is a consequence of Nekrasov and Piazzalunga’s Conjecture 5.7.
2. Preliminaries
2.1 Equivariant cohomology and K-theory
Given a topological group
$G$
acting on a topological space
$M$
, the equivariant cohomology
$H_G^*(M)$
is defined to be
$H^*(EG\times M/G)$
, where
$EG\rightarrow BG$
is the universal principle
$G$
-bundle on the classifying space
$BG$
. The map
$M\rightarrow {\rm pt}$
induces a ring homomorphism
$H^*_G({\rm pt})\rightarrow H^*_G(M)$
, making
$H^*_G(M)$
a module over
$H^*_G({\rm pt})$
for any
$M$
, and we can view
$H^*_G({\rm pt})$
as a ‘coefficient ring’.
Definition 2.1. Given a
$G$
-representation
$V$
, viewed as a vector bundle
$V\rightarrow \{{\rm pt}\}$
, we define its equivariant characteristic classes by taking the associated bundle

and taking its characteristic classes in
$H^*(BG)=H^*_G({\rm pt})$
. We denote by
$c^G_i,e_G,\textrm {ch}_G,{\textrm {td}}_G$
the equivariant versions of the
$i$
-th Chern class, the Euler class, the Chern character, and the Todd class, respectively.
Example 2.2. For the action of a
$d$
-dimensional torus
${\mathsf{T}}=({\mathbb{C}}^*)^d=\{(t_1,\ldots ,t_d):t_i\neq 0\}$
, the coefficient ring is

and
$\lambda _1,\ldots ,\lambda _d$
are exactly the equivariant first Chern classes of 1-dimensional
$\mathsf{T}$
-representations with weights
$t_1,\ldots ,t_d$
, respectively. In general, from [Reference EdidinEdi97, § 3.2],

For the
$(d-1)$
-dimensional subtorus
${\mathsf{T}}^{\prime}=\{(t_1,\ldots ,t_d)\in {\mathbb{C}}^d: t_1\ldots t_d=1\}\subseteq {\mathsf{T}}$
, the inclusion induces the following isomorphism to the quotient ring:

We can construct the equivariant K-group
$K_G(M)$
from the
$G$
-equivariant vector bundles. When
$M={\mathbb{C}}^d$
, the vector bundles over
$M$
are trivial, but they may carry non-trivial
$G$
-actions. Therefore, the equivariant bundles on
${\mathbb{C}}^d$
correspond to finite-dimensional
$G$
-representations. The equivariant characteristic classes of vector bundles can then be extended to the K-theory classes. For example, the Euler class of
$\alpha =[V]-[W]\in K_G({\rm pt})$
is
$e^G({\alpha })=e^G(V)/e^G(W)$
, which lives in the ring of fractions
$H_G^*({\rm pt})_{{\rm loc}}$
; the Chern character is
$\textrm {ch}_G({\alpha })=\textrm {ch}_G(V)-\textrm {ch}_G(W)$
, which lives in
$\prod _{i=0}^\infty H^i_G({\rm pt})$
.
Example 2.3. When
$Y={\mathbb{C}}^d$
with the natural action by
${\mathsf{T}}=({\mathbb{C}}^*)^d$
or
${\mathsf{T}}^{\prime}\cong ({\mathbb{C}}^*)^{d-1}$
, we have the following character rings


where, for any weight
$w=(w_1,\ldots ,w_d)$
, the line bundle
${\mathcal{O}}_Y\langle t^w\rangle :=\mathcal{O}_Y\otimes t^w$
simply corresponds to its character
$t^w=t_1^{w_1}\cdots t_d^{w_d}$
.
Remark 2.4. We will occasionally identify Chern characters, which are power series in cohomology, with elements in K-theory by

This allows us to consider certain classes in cohomology as elements of
$K_{\mathsf{T}}({\rm pt})$
. For example, for the line bundle
$L={\mathcal{O}}_Y\langle t_1^{w_1}\cdots t_d^{w_d}\rangle$
, we write

The reason we consider equivariant cohomology is for equivariant integration. The integration formula of [Reference Edidin and GrahamEG95, Corollary 1] via equivariant localization states that, on a smooth complete variety
$Y$
with the action of a torus
$\mathsf{T}$
, for
$\lambda$
an equivariant cohomological class, we have

where the sum goes through the components
$F$
of the fixed locus,
$N_FY$
denotes the normal bundle,
$\pi$
denotes projection to a point, and
$i$
denotes the inclusion map. The right-hand side of this formula can be used to define equivariant integration in general; for
$Y$
an arbitrary smooth variety with finitely many fixed (reduced) points, the equivariant push-forward of
$\pi _Y$
is

Example 2.5. Again let
$Y={\mathbb{C}}^d$
with the natural
${\mathsf{T}}=({\mathbb{C}}^*)^d$
-action. The only
$\mathsf{T}$
-fixed point of
$Y$
is the origin. At the origin, the character for the tangent space is
$T_{0}Y=t_1+t_2+\cdots +t_d\in K_T({\rm pt})$
, so
$e_{\mathsf{T}}(T_{0}Y)=\lambda _1\cdots \lambda _d$
. Substituting into (4), we have

2.2 Partitions and solid partitions
A partition
$\mu$
is a finite sequence
$(\mu _1,\mu _2,\ldots ,\mu _\ell )$
of non-increasing positive integers. The size
$|\mu |$
is the sum of the
$\mu _i$
’s and we call
$\ell =\ell (\mu )$
the length of the partition
$\mu $
. The empty sequence
$(0)$
is the empty partition with size
$|(0)|=0$
. Each partition
$\mu$
corresponds to a Young diagram which consists of pairs of non-negative integers
$(i,j)\in {\mathbb{Z}}_{\geqslant 0}^2$
as follows:

A pair
$\square =(i,j)$
in the above set is called a box in
$\mu$
, which we denote
$\square \in \mu$
. The conjugate partition
$\mu ^t$
is defined to be the partition whose boxes are
$\{(j,i):(i,j)\in \mu \}$
. Denote by
$c(\square ),r(\square ),a(\square ),l(\square )$
the column index, row index, arm length and leg length of
$\square =(i,j)\in \mu$
, defined explicitly as follows:

When
$i,j\gt 0$
, a necessary condition for box
$(i,j)$
to be in
$\mu$
is that both
$(i-1,j)$
and
$(i,j-1)$
are in
$\mu$
. When
$i=0$
(respectively
$j=0$
), we only need
$(i,j-1)\in \mu$
(respectively
$(i-1,j)\in \mu$
).
A solid partition
$\pi$
is a finite sequence
$(\pi _{ijk})_{i,j,k\geqslant 1}$
of positive integers such that

The size of
$|\pi |$
is the sum of the integers
$\pi _{ijk}$
. As a 4-dimensional analogue to a partition, a solid partition can also be viewed as a collection of boxes

As for partitions, we have

For a positive integer
$N$
, an
$N$
-colored partition of size
$n$
is an
$N$
-tuple of partitions
$\mu =(\mu ^{(1)},\ldots ,\mu ^{(N)})$
such that
$|\mu |:=\sum |\mu ^{(i)}|=n$
. Figure 2 illustrates how the partitions are colored based on their index. Similarly, an
$N$
-colored solid partition is an
$N$
-tuple of solid partitions.

Figure 2. A
$3$
-colored partition
$\mu =(\mu ^{(1)},\mu ^{(2)},\mu ^{(3)})$
of size
$|\mu |=19$
where
$\mu ^{(1)}=(5,3,1)$
,
$\mu ^{(2)}=(4,1)$
,
$\mu ^{(3)}=(3,2)$
correspond to different colors.
2.3 Admissible functions and universal series
We consider the notion of admissibility in the sense of [Reference MellitMel18], which will be an important condition in finding universal series for equivariant invariants.
Definition 2.6. Let
$F(Q_1,Q_2\ldots ;q_1,\ldots ,q_d)\in {\mathbb{Q}}(q_1,\ldots ,q_d)[\![Q_1,Q_2,\ldots ]\!]$
be a series in finitely many variables
$Q_1,Q_2,\ldots$
with constant term equal to 1. Then, using the plethystic exponential
$\textrm {Exp}$
, we can write

such that
$L$
is a power series in the variables
$Q_1,Q_2,\ldots$
whose coefficients are rational functions in
$q_1,\ldots ,q_d$
. The series
$F$
is called admissible with respect to the variables
$q_1,\ldots , q_d$
if the coefficients of
$L$
are polynomials in
$q_1,\ldots ,q_d$
.
Suppose
$F(Q;m_1,\ldots ,m_N;w_1,\ldots ,w_r;q_1,\ldots ,q_d)\in {\mathbb{Q}}(q_1,\ldots ,q_d)[\![Q;m_1,\ldots ,m_N;w_1,\ldots ,$
$w_r]\!]$
is admissible with respect to
$q_1,\ldots ,q_d$
with constant term 1. We have the following Laurent expansion:

Since
$F$
is admissible, by the definition of plethystic exponential,

is regular in a neighbourhood of
$q_1=\ldots =q_d=0$
as a power series in
$q_1,\ldots ,q_d$
, meaning we have a lower bound
$k_1,\ldots ,k_d\geqslant -1$
for the above summation.
Furthermore, suppose
$F$
is symmetric in
$w_1,\ldots ,w_r$
and symmetric in
$m_1,\ldots ,m_N$
, then we can expand in the following elementary symmetric polynomial basis:

for some series
$H_{\mu ,\xi ,\vec {k}}$
.
Let
$Y={\mathbb{C}}^d$
and

with the natural actions on
$Y,E={\mathbb{C}}^N\otimes {\mathcal{O}}_Y,V={\mathbb{C}}^r\otimes {\mathcal{O}}_Y$
, respectively. We write
${\mathsf{T}}={\mathsf{T}}_0\times {\mathsf{T}}_1\times {\mathsf{T}}_2$
. Say the equivariant cohomology ring of
$\mathsf{T}$
is
$H_{\mathsf{T}}^*({\rm pt})={\mathbb{C}}[\lambda _1,\ldots ,\lambda _d;m_1,\ldots ,m_N;w_1,\ldots ,w_r]$
. Then
$V$
as a
$\mathsf{T}$
-equivariant bundle has equivariant Chern roots
$w_1,\ldots ,w_r$
, and
$E$
has Chern roots
$m_1,\ldots ,m_N$
, so
$e_i(m_1,\ldots ,m_N)=c_i^{\mathsf{T}}(E),e_i(w_1,\ldots ,w_r)=c_i^{\mathsf{T}}(V)$
. Therefore

For
$\vec k=(k_1,\ldots ,k_d)$
where
$k_1,\ldots ,k_d\geqslant -1$
, there exist polynomials
$E_{\vec k}$
such that

Now suppose
$F$
is symmetric in the variables
$q_1,\ldots ,q_d$
, so
$H_{\mu ,k}=H_{\mu ,\tau (k)}$
for any permutation
$\tau$
. Hence

Note the equivariant weights of the tangent space
$T_0Y$
are exactly
$\lambda _1,\ldots ,\lambda _d$
, so
$e_i(\lambda _1,\ldots ,\lambda _d)=c_i^{\mathsf{T}}(Y)$
. By Example 2.5, we have

Redistributing the terms, we get

for some series
$H_{\mu ,\nu ,\xi }$
. We exponentiate both sides, and we obtain the following universal series expression for
$F$
.
Proposition 2.7.
Let
$F(Q;\vec {m};\vec {w};\vec {q})\in {\mathbb{Q}}(q_1,\ldots ,q_d)[\![Q;m_1,\ldots ,m_N;w_1,\ldots ,w_r]\!]$
be admissible with respect to the variables
$q_1,\ldots ,q_d$
. Suppose
$F$
is symmetric in
$w_1,\ldots ,w_r$
, in
$m_1,\ldots ,m_N$
, and in
$q_1,\ldots ,q_d$
. Then there exist power series
$G_{\mu ,\nu ,\xi }(Q)$
labeled by partitions
$\mu ,\nu ,\xi$
, such that

3. Segre and Verlinde invariants on
${\mathbb{C}}^2$
3.1 Virtual equivariant invariants on Hilbert schemes
Before defining virtual invariants, we recall the notion of a perfect obstruction theory in the sense of [Reference Behrend and FantechiBF98, Definition 4.4]. For our purposes, we use the following simplified version.
Definition 3.1. Let
$X$
be a scheme over
$\mathbb{C}$
. An obstruction theory is a complex of vector bundles

for some
$a\in {\mathbb{Z}}$
, together with a morphism in the derived category
$D({\rm QCoh}(X))$
to the cotangent complex

such that
$h^0({\varphi })$
is an isomorphism and
$h^{-1}({\varphi })$
is surjective. It is a (2-term) perfect obstruction theory if
$E^i=0$
for
$i\neq 0,-1$
. The virtual tangent bundle
$T^{{\rm vir}}=E_\bullet =(E^\bullet )^*$
is the class of the dual complex of a given obstruction theory.
Let
$S$
be a surface and
$E$
a vector bundle. It is known that
$\textrm {Quot}_S(E,n)$
admits an obstruction theory given by the dual complex of
${\bf R}\mathscr{H}om_{p}(\mathcal{I},\mathcal{F})$
, where
$\mathcal{I},\mathcal{F}$
are, respectively, the universal subsheaf and the quotient sheaf. When
$S$
is a projective surface, this obstruction theory is perfect and of virtual dimension
$nN$
. An explicit construction of this perfect obstruction theory can be found in [Reference StarkSta24, § 2.3]. Using this, we can define a virtual fundamental class
$[\textrm {Quot}_S(E,n)]^{{\rm vir}}$
via the methods from [Reference Behrend and FantechiBF98, Reference Li and TianLT96], as well as a virtual structure sheaf
${\mathcal{O}}^{{\rm vir}}$
using [Reference Ciocan-Fontanine and KapranovCFK09]. Applying the same argument for
$S ={\mathbb{C}}^2$
gives us an equivariant perfect obstruction theory. We note that, since the quotients
$E\twoheadrightarrow F$
are of rank
$0$
and Euler characteristic
$n$
,
$F$
is compactly supported, and the
$\textrm {Ext}$
-groups
$\textrm {Ext}^i(E,F)$
are finite-dimensional vector spaces. Thus the steps involving Serre duality still work.
Let
$S={\mathbb{C}}^2$
and
$E=\oplus _{i=1}^N{\mathcal{O}}_S\langle y_i\rangle$
. Recall from the introduction the following tori:

Set
${\mathsf{T}}={\mathsf{T}}_0\times {\mathsf{T}}_1\times {\mathsf{T}}_2$
, with

where
$\lambda _i,m_i,w_i$
are, respectively, the equivariant first Chern classes of the 1-dimensional
$\mathsf{T}$
-representations with weights
$t_i,y_i,v_i$
(see Example 2.2). This gives us a
$\mathsf{T}$
-action on
$\textrm {Quot}_S(E,n)$
. For a
$\mathsf{T}$
-equivariant bundle
$V$
, the tautological bundle
$V^{[n]}$
is also
$\mathsf{T}$
-equivariant. The
${\mathsf{T}}_1$
-fixed quotients of
$\textrm {Quot}_S(E,n)$
decompose into the form

Thus the
${\mathsf{T}}_1$
-fixed locus can be identified as

The
${\mathsf{T}}_0$
-fixed locus of these Hilbert schemes consists of collections of finitely many reduced points, labeled by partitions. Consequently, the fixed locus
$\textrm {Quot}_S(E,n)^{\mathsf{T}}$
consists of finitely many reduced points of the form

labeled by
$N$
-colored partitions
$\mu =\left (\mu ^{(1)},\ldots ,\mu ^{(N)}\right )$
.
For a vector bundle
$V$
over
$Y$
, define

which extends to a homomorphism
$\Lambda _z:(K^0(Y),+)\rightarrow (K^0(Y)[\![z]\!],\cdot )$
. We define the following equivariant invariants using virtual equivariant localization [Reference Graber and PandharipandeGP97] and K-theoretic virtual equivariant localization [Reference Ciocan-Fontanine and KapranovCFK09, Theorem 5.3.1]. Since we are interested in comparing the Segre and Verlinde series, we convert the Verlinde invariants into cohomological invariants using the virtual Hirzebruch–Riemann–Roch formula [Reference Ravi and SreedharRS21, Corollary 1.2].
Definition 3.2. Let
$S={\mathbb{C}}^2$
and

The equivariant virtual Segre, Chern, and Verlinde series on Quot schemes are, respectively,

We shall describe how to calculate these invariants, and refer to [Reference Fasola, Monavari and RicolfiFMR21, § 5.1] and [Reference LimLim21, § 3.3] for the following argument. On each
${\mathsf{T}}_1$
-fixed locus

the universal subsheaf and universal quotient sheaf of
$D$
are
$\bigoplus _{i=1}^N I_{\mathcal{Z}_i}\langle y_i\rangle$
and
$\bigoplus _{j=1}^N {\mathcal{O}}_{\mathcal{Z}_j}\langle y_i\rangle$
where
$\mathcal{Z}_i$
is the universal subscheme of
$\textrm {Hilb}^{n_i}(S)$
. The virtual tangent bundle over
$D$
is then

where
$p:D\times X\rightarrow D$
is the projection. Further restricting to each
$Z_\mu =([Z_1],[Z_2],\ldots ,[Z_N])\in \textrm {Quot}_S(E,n)^{\mathsf{T}}$
gives the virtual tangent bundle at
$Z_{\mu }$
as follows:

To give an explicit formula for
$T^{{\rm vir}}$
, we consider a
${\mathsf{T}}_0$
-equivariant free resolution of
$I_{Z_i}$
. We refer to [Reference EisenbudEis95, p. 439] for the following Taylor resolution. Say
$I_{Z_i}$
is generated by monomials
$m_1,\ldots ,m_s$
. For each
$k=0,\ldots ,s$
, let
$F_k$
be the free
${\mathbb{C}}[x_1,\ldots , x_n]$
-module with basis
$\{e_I\}$
, indexed by subsets
$I\subseteq \{1,\ldots ,s\}$
of size
$k$
. Set

For
$k=1,\ldots , s$
, define the differential
$d_k:F_k\rightarrow F_{k-1}$
by

for each subset
$I=\{i_1,\ldots , i_k\}$
such that
$i_1\lt \cdots \lt i_k$
. Giving each
$e_I$
the weight of
$m_I$
, we obtain the
$T_0$
-equivariant free resolution

where

for some
$d_{kI}\in {\mathbb{Z}}^2$
. Define

Note that the character of
${\mathcal{O}}_S={\mathbb{C}}[x_1,x_2]$
is
$\sum _{i,j\geqslant 0}t_1^{-i}t_2^{-j}=1/(1-t_1^{-1})(1-t_2^{-1})$
, so the character of
${\mathcal{O}}_{Z_i}={\mathcal{O}}_S/I_{Z_i}$
is

Therefore, the character of
$T^{{\rm vir}}_{Z_{\mu }}$
in
$K_{{\mathsf{T}}}({\rm pt})$
can be expressed as

where
$\overline {(\cdot )}$
denotes the involution
$t_i\mapsto t_i^{-1}$
.
For the
$\mathsf{T}$
-equivariant bundle
$V=\oplus _{i=1}^r{\mathcal{O}}_S\langle v_i\rangle$
, the fiber of
$V^{[n]}$
over
${Z_{\mu }}=(Z_1,\ldots Z_N)$
is the
$rn$
-dimensional representation

We could also replace the terms
$\sum _{\square \in \mu ^{(j)}}t_1^{-c(\square )}t_2^{-r(\square )}$
by
$Q_j$
, but we elect to use the above expansion to match with one of our main references [Reference Göttsche and MellitGM22].
Substituting the above calculations into the definition, we obtain the following expressions for the Chern and Verlinde series of vector bundles

3.2 Relation to projective toric surfaces
We consider what the equivariant invariants will be for a smooth projective toric surface
$S^{\prime}$
, and compare them with the case
$S={\mathbb{C}}^2$
. More details on this reduction can be found in [Reference Göttsche and MellitGM22, § 3.2]; see also [Reference ArbesfeldArb21, § 6.2] and [Reference Liu, Yan and ZhouLYZ02, § 3.2].
Suppose we are interested in the integral

for a
$\mathsf{T}$
-equivariant bundle
$V$
and some multiplicative genus
$f$
. On
$S$
where
$V$
has Chern roots
$w_1,\ldots , w_r$
and
$E$
has Chern roots
$m_1,\ldots ,m_N$
, this can be computed as

We write
$T^{\prime}=({\mathbb{C}}^*)^2$
with cohomological variables
$\lambda _1^{\prime},\lambda _2^{\prime}$
so that
$H_{{\mathsf{T}}^{\prime}}^*({\rm pt})={\mathbb{C}}[\lambda _1^{\prime},\lambda _2^{\prime}]$
. Let
$S^{\prime}$
be a toric projective surface with a natural action by
${\mathsf{T}}^{\prime}=({\mathbb{C}}^*)^2$
. Say the fixed points are
$p_1,\ldots ,p_M$
and the Chern roots of the tangent space of
$S^{\prime}$
at
$p_i$
are
$\lambda _1^{(i)},\lambda _2^{(i)}$
, which live in
$H_{{\mathsf{T}}^{\prime}}^*({\rm pt})={\mathbb{C}}[\lambda _1^{\prime},\lambda _2^{\prime}]$
. Suppose that
$E^{\prime}$
is a
${\mathsf{T}}^{\prime}$
-equivariant bundle with Chern roots
$m_1^{(i)},\ldots ,m_N^{(i)}\in H_{{\mathsf{T}}^{\prime}}^*({\rm pt})$
at each
$p_i$
such that the fixed locus of
$\textrm {Quot}_{S^{\prime}}(E^{\prime},n)$
is a finite collection of reduced points for all
$n$
. Let
$V^{\prime}$
be a
${\mathsf{T}}^{\prime}$
-equivariant bundle on
$S^{\prime}$
with Chern roots
$w_1^{(i)},\ldots ,w_r^{(i)}\in H_{{\mathsf{T}}^{\prime}}^*({\rm pt})$
. By virtual equivariant localization, we have

Since
$S^{\prime}$
is compact, the sum on the right-hand side lives in
$H^*_{\mathsf{T}}({\rm pt})={\mathbb{C}}[\lambda _1^{\prime},\lambda _2^{\prime}]$
. Therefore, it is indeed valid to set
$\lambda _1^{\prime}=\lambda _2^{\prime}=0$
, and the equality follows from the virtual Bott residue formula.
If we define
$I_S(E,V;q):=\sum _{n=0}^\infty q^n\int _{[\textrm {Quot}_S(E,n)]} f(V)$
, then from the above expansions, we obtain

Suppose furthermore that the series in the above expression have the following universal structure

where
$\int _S$
is given by (4). The exponents
$c_\mu (V^{\prime})c_\nu (S^{\prime})c_\xi (E^{\prime})$
are regarded as numbers in
$H^2(S^{\prime})$
. Exponentiating a series to a power of an element of
$H^*_{\mathsf{T}}({\rm pt})_{{\rm loc}}$
follows the recipe in Remark 1.6. Then (12) implies

The final sum is only over
$|\mu |+|\nu |+|\xi |=2$
because if
$|\mu |+|\nu |+|\xi |\neq 2$
, then the term

vanishes for degree reasons. By universality, we conclude that if the series structure (13) exists, then the series
$U^{{\rm vir}}_{\mu ,\nu ,\xi }(q)$
of
$I_{S^{\prime}}(E^{\prime},V^{\prime};q)$
for the projective case coincide with the series
$U_{\mu ,\nu ,\xi }(q)$
of
$I_S(E,V;q)$
whose exponents lie in
$H^0_{\mathsf{T}}({\rm pt})$
.
3.3 Universal series expansion
A general tactic for studying Segre and Verlinde series is by using a more general genus. In the non-virtual setting for Hilbert schemes of surfaces, this could be [Reference Göttsche and MellitGM22, (1.1)]. In the virtual case, we use the invariant (14) defined below. For Calabi–Yau 4-folds, we consider the Nekrasov genus (35) introduced by [Reference Nekrasov and PiazzalungaNP19].
For projective surfaces, we define an auxiliary virtual invariant as follows

The variable
$z$
is considered as the weight of an extra
${\mathbb{C}}^*$
-action that is trivial on
$S$
and
$\textrm {Quot}_S(E,n)$
. We shall refer to this as the Nekrasov genus for Quot schemes of surfaces (cf. (35)).
We generalize this to the equivariant setting using virtual equivariant localization. On
$S={\mathbb{C}}^2$
for vector bundles, this is given by:

The following Chern and Verlinde limits are satisfied, analogous to [Reference Göttsche and MellitGM22, Proposition 3.5].
Lemma 3.3.
For
$S={\mathbb{C}}^2$
, the Chern series and the Verlinde series can be retrieved from
${\mathcal{N}}_S$
by taking limits. We have

Here we use the identification of Remark 2.4 with
$c_1(t_i)=\lambda _i,c_1(y_i)=m_i,c_1(v_i)=w_i$
.
Proof.
For the Chern limit, first consider the substitutions
$\lambda _i\leadsto -{\varepsilon }\lambda _i,w_i\leadsto -{\varepsilon } w_i,m_i\leadsto -{\varepsilon } m_i$
. This turns the term
$\prod _{j=1}^N\prod _{\square \in \mu ^{(j)}}\prod _{i=1}^r(1-t_1^{-c(\square )}t_2^{-r(\square )}v_iy_j(1+{\varepsilon })^{-1})$
into

For the denominator in the sum (15), we note that for a Chern root
$x$
, substituting it by
$-{\varepsilon } x$
turns
$1-e^{-x}=x-\frac {x^2}2+\ldots$
into
$1-e^{{\varepsilon } x}=-{\varepsilon }(x+O({\varepsilon }))$
. Therefore, after the substitution, the denominator
$\textrm {ch}(\Lambda _{-1}(T^{{\rm vir}}|_{Z_\mu })^\vee )$
becomes

Substituting back into (15), the Chern limit becomes the limit of

which converges to
${\mathcal{C}}_S(E,V;q)$
by (11).
For the Verlinde series, we have

Before starting the proof for universal series expressions, let us discuss what the expansion of
${\mathcal{N}}_S(E,V;q,z)$
as a formal Laurent series in the variables
$\vec {\lambda },\vec {m},\vec {w},q,z$
would look like. In [Reference ArbesfeldArb21, Proposition 3.2], Arbesfeld shows that invariants such as
$[q^n]{\mathcal{N}}_S(E,V;q,z)$
can be written as a quotient whose numerator is a Laurent polynomial in
$\vec {t},\vec {y},\vec {v},z$
and whose denominator is of the form
$\prod _{\mathsf{w}}(1-\mathsf{w})$
for some non-compact weights
$\mathsf{w}$
in the sense of the following definition.
Definition 3.4 [Reference ArbesfeldArb21, Definition 3.1]. Let
$M$
be a quasi-projective scheme with an action by some torus
$\mathsf{T}$
. For a weight
$\mathsf{w}\in {\mathsf{T}}^\vee$
, denote by
${\mathsf{T}}_{\mathsf{w}}$
the maximal subtorus of
$\mathsf{T}$
containing
$\ker \mathsf{w}$
. If the fixed locus
$M^{{\mathsf{T}}_{\mathsf{w}}}$
is proper, then
$\mathsf{w}$
is a compact weight; otherwise, it is a non-compact weight.
Fasola, Monavari and Ricolfi used this to prove that the K-theoretic Donaldson–Thomas partition functions on
${\mathbb{C}}^3$
are Laurent polynomials with respect to the variables
$y_1,\ldots ,y_N$
[Reference Fasola, Monavari and RicolfiFMR21, Theorem 6.5]. We give an outline of their argument, applied to the invariant
${\mathcal{N}}_S$
for
$S={\mathbb{C}}^2$
. First note that, by (9), for any
$N$
-colored partition
$\mu$
, we have

for some series
$A(\vec {t}\,)\in {\mathbb{Q}}[\![t_1,t_2]\!]_{{\rm loc}}$
and some sets of weights
$A_{ij},B_{ij}$
. We shall show that the denominator of
${\mathcal{N}}_S$
does not have factors of the form
$(1-y_i^{-1}y_jt^b)$
for any
$i\neq j$
and
$b\in {\mathbb{Z}}^2$
. By [Reference ArbesfeldArb21, Proposition 3.2], we need to prove that
$\mathsf{w}=y_i^{-1}y_jt^b$
is a compact weight. Since

is itself a torus, we have
${\mathsf{T}}_{\mathsf{w}}=\ker \mathsf{w}$
. By definition, it suffices to show that
$\textrm {Quot}_S(E,n)^{{\mathsf{T}}_{\mathsf{w}}}$
is proper. With the automorphism
${\mathsf{T}}\rightarrow {\mathsf{T}}$
defined by

we identify the subgroup
${\mathsf{T}}_{\mathsf{w}}$
as
${\mathsf{T}}_0\times \{(w_1,\ldots ,w_N):w_i=w_j\}\times {\mathsf{T}}_2$
, which contains the subgroup
${\mathsf{T}}_0={\mathsf{T}}_0\times \{(1,\ldots ,1)\}$
. This gives us an inclusion

The quotients in the fixed locus
$\textrm {Quot}_S(E,n)^{{\mathsf{T}}_0}$
are all supported at the origin
$0\in {\mathbb{C}}^2$
, so the fixed locus lies inside the punctual Quot scheme
$\textrm {Quot}_S(E,n)_0$
. The punctual Quot scheme is proper since it is a fiber of the Quot-to-Chow map
$\textrm {Quot}_S(E,n)\rightarrow \textrm {Sym}^nS$
, which is a proper morphism [Reference Fasola, Monavari and RicolfiFMR21, Remark 3.4]. In conclusion,
$[q^n]{\mathcal{N}}_S(E,V;q,z)$
is a Laurent polynomial with respect to the variables
$y_1,\ldots ,y_N$
, so it can be expanded into a power series with respect to the cohomological parameters
$m_1,\ldots ,m_N$
.
Furthermore, if
$\mathsf{w}$
is a weight that contains both
$t_1$
and
$t_2$
, then we have
${\mathsf{T}}_{\mathsf{w}}\cong \{(t_1,t_2):t_1t_2=1\}\times {\mathsf{T}}_1\times {\mathsf{T}}_2$
. The fixed locus of this subgroup remains the same as that of
$\mathsf{T}$
, as explained in the next section for reduced invariants. Therefore,
$\mathsf{w}$
is a compact weight, and the denominator of
${\mathcal{N}}_S$
will not contain factors of the form
$(1-t_1^at_2^b)$
for any
$a\neq 0$
,
$b\neq 0$
. This means that in cohomology
$[q^n]{\mathcal{N}}_S(E,V;q,z)$
can be expanded into a Laurent series in
$\lambda _1,\lambda _2$
whose coefficients are power series in
$\vec {m},\vec {w},z$
, where the degrees on
$\lambda _1,\lambda _2$
are bounded below individually. We shall see the importance of this lower bound in the proof of the following theorem.
Theorem 3.5.
Let
$S={\mathbb{C}}^2$
. For any
$r\in {\mathbb{Z}}$
,
$N\gt 0$
, there exist universal power series
$A_{\mu ,\nu ,\xi }(q),B_{\mu ,\nu ,\xi }(q),C_{\mu ,\nu ,\xi }(q)$
, dependent on
$N$
and
$r$
, such that for
$E=\oplus _{i=1}^{N}{\mathcal{O}}_S\langle y_i\rangle$
and
${\alpha }\in K_{\mathsf{T}}(S)$
of rank
$r$
, the equivariant virtual Segre, Verlinde and Chern series on
$\textrm {Quot}_S(E,n)$
can be written as the following infinite products

Proof.
We begin with the case where
$\alpha$
is a vector bundle
$V$
. Assume
$V=\oplus _{i=1}^r{\mathcal{O}}_S\langle v_i\rangle$
. At the end of the proof, we can generalize this to arbitrary
$\mathsf{T}$
-equivariant bundles by substituting
$\mathsf{T}$
-weights into the variables
$v_1,\ldots ,v_r$
.
Begin by expanding
$\log {\mathcal{N}}_S(E,V;q,z)$
as a Laurent series in
$\lambda _1,\lambda _2$
as follows:

for some series
$H_{j,k}\in {\mathbb{Q}}[\![q,z;m_1,\ldots ,m_N;w_1,\ldots ,w_r]\!]$
. By the symmetry in
$w_1,\ldots ,w_r$
and the symmetry in
$m_1,\ldots , m_N$
, this expands to

for some series
$G_{\mu ,\xi ,j,k}\in {\mathbb{Q}}[\![q,z]\!]$
.
Our goal is to get a universal series expression by exponentiating the above equality. To do so, we show that the terms in the second summation vanish, using the relations from § 3.2. This proves that
${\mathcal{N}}_S(E,V;q,z)$
is admissible, from which we deduce the desired expressions by taking the limits of Lemma 3.3.
Let
$S^{\prime}$
be a toric projective surface with
${\mathsf{T}}^{\prime}=({\mathbb{C}}^*)^2$
-action, fixed points
$p_1,\ldots , p_M$
and vector bundles
$E^{\prime},V^{\prime}$
as in § 3.2. Then (12) applied to
$\mathcal{N}$
becomes

Instead of taking
$\vec m^{(i)},\vec w^{(i)}$
as elements in
$H_{{\mathsf{T}}^{\prime}}^*({\rm pt})={\mathbb{C}}[\vec \lambda ^{\prime}]$
, we could regard them as formal variables
$\vec m,\vec w$
, independent of
$i$
. This can be done by adding trivial actions by
${\mathsf{T}}_1,{\mathsf{T}}_2$
to
$S^{\prime}$
, and replacing
$E^{\prime},V^{\prime}$
by

which are trivial bundles whose equivariant Chern roots at each fixed point
$p_i$
are, respectively,
$\vec m$
and
$\vec w$
, independent of
$i$
. In this setting, the substitutions
$\vec m=\vec m^{(i)},\vec w=\vec w^{(i)}$
are no longer necessary in the above equation for
$\mathcal{N}_{S^{\prime}}$
, which gives

Substituting the previous expansion of
$\log {\mathcal{N}}_S(E,V;q,z)$
, we see that

Again, the reason that we can evaluate
$\vec {\lambda }^{\prime}=\vec {w}=\vec {m}=0$
on the right-hand side is that, by compactness of
$S^{\prime}$
, the terms in the brackets are power series in
$\vec \lambda ^{\prime},\vec w,\vec m$
. Since the terms
$c_\mu (V),c_\xi (E)$
form a basis for symmetric polynomials in the variables
$\vec w,\vec m$
, we conclude that each summand

is a power series in
$\lambda _1^{\prime},\lambda _2^{\prime}$
.
Let
$S^{\prime}={\mathbb{P}}^1\times {\mathbb{P}}^1$
, with
${\mathsf{T}}_0$
-action

We refer to [Reference Liu, Yan and ZhouLYZ02, § 3.7] for the following computations of equivariant weights. The fixed points are


The corresponding weights are
$\vec {\lambda }^{(1)}=\vec {\lambda }^{(00)},\vec {\lambda }^{(2)}=\vec {\lambda }^{(01)},\vec {\lambda }^{(3)}=\vec {\lambda }^{(10)},\vec {\lambda }^{(4)}=\vec {\lambda }^{(11)}$
where

for
$i,j\in \{0,1\}$
for each
$\mu ,\xi$
. Substituting into (16), we see that the summands with odd
$j$
or
$k$
would cancel each other out, leaving us with

By our previous observation, this expression is a power series in
$\lambda _1^{\prime},\lambda _2^{\prime}$
for each
$\mu ,\xi$
and does not contain negative-degree terms. Therefore, if
$\min \{j,k\}\leqslant -2$
, we have
$G_{\mu ,\xi ,j,k}(q)=0$
.
Having dealt with the case where
$j,k$
are both even, we would like to apply the same argument to the other cases. To do so, we need to solve the problem that the summands vanish whenever one of
$j,k$
is odd. We claim that if each
$w_i$
is replaced by
$w_i(l+\lambda _1+\lambda _2)^2$
in
$\mathcal{N}_{S}$
for
$i=1,\ldots , r$
, where
$l$
is a fixed number, then the resulting coefficients in front of
$q,z$
are of the form

for some series
$\tilde f$
. The number
$l$
here will be substituted by integers later to get more specific conditions. Then we may repeat the above argument and get that
$\sum _{i=1}^M\tilde F(\vec \lambda ^{(i)},\vec m^{(i)},\vec w)\in H^*_{\mathsf{T}}({\rm pt})$
is a power series in the variables
$\vec \lambda ^{\prime}, \vec m, \vec w$
. Separating out the coefficients of
$c_\mu (V),c_\xi (E)$
as we did in (16), this gives that the coefficients of
$q,z$
in

are power series in
$\lambda _1^{\prime},\lambda _2^{\prime}$
.
Let us prove the claim above. Note that the coefficients of
$q,z$
in
$\mathcal{N}_S(E,V;q,z)$
are of the form

for some series
$f$
. Set
$\vec Y=(m_j-c(\square )\lambda _1-r(\square )\lambda _2)_{j=1,\ldots ,N}^{\square \in \mu ^{(j)}}$
the collection of Chern roots of
${\mathcal{O}}_S^{[n]}$
and
$\vec X=(\lambda _1+\lambda _2+m_j-c(\square )\lambda _1-r(\square )\lambda _2)_{j=1,\ldots ,N}^{\square \in \mu ^{(j)}}$
the collection of Chern roots of
$K_S^{[n]}$
. Let us also denote
$x=\lambda _1+\lambda _2$
. To prove the claim, it suffices to show that, for any series
$f$
, there exists some series
$\tilde f$
such that

We may rewrite the left-hand side and see that this is equivalent to showing

for any series
$g$
. This follows from Lemma 3.6 by taking
$\vec x=x,\vec y=\vec Y$
.
As a result of the previous paragraph, the coefficients of

are power series in
$\lambda _1^{\prime},\lambda _2^{\prime}$
for any integer
$l\geqslant 0$
. When
$\mu \neq (0)$
, the matrix formed by the vectors

for
$l=1,2,3,\ldots$
has maximal rank. We may take a linear combination of the above expression evaluated at different integer values of
$l$
, and get that

is a power series in
$\lambda _1^{\prime},\lambda _2^{\prime}$
for each
$s=0,1,\ldots ,2|\mu |$
.
Taking
$s=2$
, we get that the following series is a power series in
$\lambda _1^{\prime},\lambda _2^{\prime}$

This means the coefficients
$G_{\mu ,\xi ,j,k}(q,z)$
vanish for terms
$(\lambda _1^{\prime})^{j+1}(\lambda _2^{\prime})^{k+1}$
whenever the degree
$j+1$
or
$k+1$
is negative, which happens exactly when
$\min \{j,k\}\leqslant -2$
. Hence, we conclude
$G_{\mu ,\xi ,j,k}=0$
whenever
$j,k$
are odd,
$\min \{j,k\}\leqslant -2$
, and
$\mu \neq (0)$
.
If one of
$j,k$
is odd and the other is even, continuing to assume
$\mu \neq (0)$
, we take
$s=1$
and get

Although these are not polynomials when
$\min \{j,k\}\leqslant -2$
, we see there might be some linear dependence because we may only conclude that

for
$j$
odd and
$k$
even. To solve this issue, we further apply the argument to the
$s=3$
case and obtain the following dependencies

for all
$j$
odd,
$k$
even and
$\min \{j,k\}\leqslant -4$
. Consequently,
$G_{\mu ,\xi ,j,k}=G_{\mu ,\xi ,j+2,k-2}$
for any
$j$
even and
$k$
odd with
$\min \{j,k\}\leqslant -3$
. Combining these relations we see that, for
$\min \{j,k\}\leqslant -2$
, there exist some constants
$C_{\mu ,\xi ,a,b,l}^{\pm }$
, labeled by the partitions
$\mu ,\xi$
, integers
$a,b,l$
and a sign
$\pm$
, such that

The reason for the superscript
$\pm$
is due to cases such as
$j=-1,k=0$
, where we would have
$\min \{j,k\}\gt -2$
, so the dependence does not necessarily hold. Because of this gap, we cannot always relate the coefficient when
$j\geqslant 0$
to
$j\lt 0$
, resulting in separated cases. By the paragraphs preceding this theorem, for a fixed
$a$
, the degrees
$j,k$
on
$\lambda _1^{\prime},\lambda _2^{\prime}$
of the
$[q^{a}]$
coefficient are bounded below. However the above indicates that the constants
$C_{\mu ,\xi ,a,b,l}^\pm$
only depend on the value
$l=j+k$
, and we can make
$j$
or
$k$
arbitrarily small. Hence
$C_{\mu ,\xi ,a,b,l}^\pm=0$
whenever
$\mu \neq (0)$
.
With all the vanishings of
$G_{\mu ,\xi ,j,k}$
, we write

To deal with the terms
$G_{(0),\xi ,j,k}(q,z)$
for
$\min \{j,k\}\leqslant -2$
, we apply Lemma 4.1 and find

so
$G_{(0),j,k}$
is constant with respect to the variable
$z$
. Let us attempt to extract the
$[q^{n}\lambda _1^j\lambda _2^kc_{(0)}(V)c_\xi (E)]$
coefficient of the Chern series from
$G_{(0),\xi ,j,k}$
using the Chern limit of Lemma 3.3. This results in a limit
${\varepsilon }\rightarrow 0$
of the term
${\varepsilon }^{n(N-r)}{\varepsilon }^{|\xi |}{\varepsilon }^{j+k}$
, which does not make sense when the rank
$r$
is sufficiently large, so we must have
$G_{(0),\xi ,j,k}=0$
for such
$r$
. To generalize this to arbitrary ranks, we apply [Reference Göttsche and MellitGM22, Lemma 3.3] to
${\mathcal{N}}_S$
, which says that the coefficients of
${\mathcal{N}}_S$
are polynomials in
$r$
when
$r\geqslant 0$
. Now we can write

As noted in [Reference Oprea and PandharipandeOP22, (31)], the obstruction on
$\textrm {Hilb}^n(S)$
at a fixed point
$[Z_\mu ]$
is
$(K_S^{[n]})^\vee |_{Z_\mu }$
. From (10), we see a copy of
$K_S^\vee =t_1t_2$
is in
$K_S^{[n]}|_{Z_\mu }$
. By (9), the obstruction bundle on
$\textrm {Quot}_S(E,n)$
at any fixed point has at least one copy of
$K_S^\vee$
as a direct summand. For a line bundle
$L$
, we have

where
$\ldots$
are some omitted terms in
$H_{\mathsf{T}}^{\gt 0}({\rm pt})$
. Therefore,
$1/\textrm {ch}(\Lambda _{-1}(T^{{\rm vir}}_Z)^\vee )$
has a factor of
$e(K_S^\vee )=c_1(S)=\lambda _1+\lambda _2$
in its numerator. We also note that this factor does not appear in the denominator, because if we pass to the subtorus
$\{(t_1,t_2):t_1t_2=1\}$
, the Zariski tangent space has no
$\mathsf{T}$
-fixed parts: by (7), the fixed part can only come from the direct summands with
$i=j$
, which correspond to the Hilbert scheme case, but, by [Reference ArbesfeldArb21, (1.6)], these summands have no fixed parts because
$a(\square ),l(\square )\geqslant 0$
for any box
$\square$
. Therefore, we may extract this factor of
$c_1(S)$
and obtain

for some series
$H_{\mu ,\xi ,j,k}\in {\mathbb{Q}}[\![q,z]\!]$
. Furthermore, since
$j,k$
are now bounded below by
$-1$
, multiplying by
$\lambda _1\lambda _2$
would give us a power series expansion in
$\lambda _1,\lambda _2$
, allowing us to use the symmetry in
$\lambda _1,\lambda _2$
and write

for some series
$H_{\mu ,\nu ,\xi }\in {\mathbb{Q}}[\![q,z]\!]$
.
Finally, taking Chern and Verlinde limits of
$H_{\mu ,\nu ,\xi }$
as in Lemma 3.3 then exponentiating gives us series
$C_{\mu ,\nu ,\xi },B_{\mu ,\nu ,\xi }$
such that

and the fact that
${\mathcal{S}}_S(E,V;q)={\mathcal{C}}_S(E,-V;q)$
implies that there exists series
$A_{\mu ,\nu ,\xi }$
such that

To generalize this to arbitrary K-theory classes
${\alpha }=[V^{\prime}]-[V^{\prime\prime}]\in K_{\mathsf{T}}(S)$
for equivariant bundles
$V^{\prime},V^{\prime\prime}$
of rank
$m,l$
, respectively, we apply [Reference Göttsche and MellitGM22, Lemma 3.3] once more; it states that the invariants for
$\alpha$
are obtained by substituting

where the
$p_n$
are the power-sum symmetric polynomials. Hence, the above universal series expressions hold for all
${\alpha }\in K_{\mathsf{T}}(S)$
.
Lemma 3.6.
Let
$F(\vec x,\vec y)$
be a polynomial symmetric in
$\vec x=(x_1,\ldots , x_n)$
and symmetric in
$\vec y=(y_1,\ldots , y_m)$
. Then
$F$
can be written as a polynomial expression of symmetric functions in
$\{y_1,\ldots ,y_m\}$
and symmetric functions in
$\{x_i+y_j\}_{i=1,\ldots ,n}^{j=1,\ldots ,m}$
.
Proof.
Note that this lemma is additive and multiplicative on
$F$
, meaning that, if the result holds for
$F_1$
and
$F_2$
, then it holds for
$F_1+F_2$
and
$F_1\cdot F_2$
. By the additive property, we assume that
$F$
is homogeneous of degree
$d$
. We proceed by induction on
$d$
, where the base case of
$d=0$
is trivial.
We expand
$F$
in the elementary symmetric polynomial basis with respect to the variables
$y_1,\ldots , y_m$
:

Note that each
$f_\mu (\vec x)$
is symmetric in
$\vec x$
and trivially symmetric in
$\vec y$
, and has degree strictly less than
$F$
if
$\mu \neq (0)$
. By the induction hypothesis, we can assume that
$f_\mu (\vec x)$
satisfies the lemma when
$\mu \neq (0)$
. Then, by the additive and multiplicative properties, it remains to prove
$f_{(0)}(\vec x)$
satisfies the lemma.
Since
$f_{(0)}(\vec x)$
is a symmetric polynomial, which can be written as a sum of products of the elementary symmetric polynomials
$e_k(\vec x)$
, we may apply the additive and multiplicative properties again and conclude that the result follows once we prove it for
$e_k(\vec x)$
. Since we are performing an induction on the degree
$d$
of
$F$
, we may assume that the result holds for
$k\lt d$
and proceed to prove it for
$k=d$
.
We have

for some constant
$K\in {\mathbb{Z}}$
, and we have that
$G$
is a polynomial symmetric in
$\vec x$
and in
$\vec y$
and that every monomial term in
$G$
contains some
$y_j$
. Moreover,
$G$
is homogeneous of degree
$d$
, and we can repeat the previous argument and write

Now each
$g_\mu (\vec x)$
has degree strictly less than
$d$
, and induction shows the result holds for each
$g_\mu (\vec x)$
, and consequently for
$G(\vec x,\vec y)$
and
$e_d(\vec x)$
.
By the non-equivariant Segre–Verlinde correspondence [Reference BojkoBoj21a, Theorem 1.7] and the relations between the non-equivariant series and equivariant series illustrated in § 3.2, we have a weak Segre–Verlinde correspondence as the following corollary. The same argument as in the following proof also gives us a weak symmetry in the form of Corollary 1.14.
Corollary 3.7. In the setting of Theorem 3.5 , we have the following correspondence

whenever one of
$\mu ,\nu ,\xi$
is
$(1)$
and the other two are
$(0)$
. In particular, the degree 0 part satisfies

for some terms
$f_n\in H_{\mathsf{T}}^{2n-2}({\rm pt})$
dependent on
$\alpha$
through its rank and Chern classes.
Proof.
By the last paragraph of § 3.2, the universal series in Theorem 3.5, when passed to a toric projective surface, must give the Segre–Verlinde correspondence of [Reference BojkoBoj21a, Theorem 1.7] in degree 0. Since the degree 0 terms occur only when one of
$\mu ,\nu ,\xi$
is
$(1)$
and the other two are
$(0)$
, we have

in those cases.
Note that when we take
$\exp$
of (17), the total degree 0 part might come from the product of a negative-degree term and a positive-degree term, but since each term in the integrand is accompanied by a copy
$c_1(S)$
, we know this difference must be a multiple of
$c_1(S)^2$
. We also see that the
$[q^n]$
coefficients are sums of products of at most
$n$
such integrals, giving a denominator of
$\lambda _1^n\lambda _2^n$
, so we are done.
For illustration, we shall extract this difference, and express it explicitly. This is just a standard computation. For a partition
$\mu =(\mu _1,\mu _2,\ldots ,\mu _L)$
of size
$n$
with length
$L$
, and a sequence of positive integers
$\kappa =(k_1,k_2,\ldots ,k_L)$
, write
$\kappa |\mu$
if each
$k_i|\mu _i$
. We also associate a set of tuples of partitions to
$\kappa$
by

For each
$n\gt 0$
, we would like to find the degree 0 part of the
$[q^n]$
coefficient of
$\exp$
of (17). By expanding the exponential using definition, we observe that these terms come from products of integrals labeled by
$\mu ^{(i)},\nu ^{(i)},\xi ^{(i)}$
in
$M_\kappa$
for some tuples
$\kappa |\pi$
for some partition
$\pi$
of size
$n$
. A more precise description is given by the following equation. Suppose
$\log (A_{\mu ,\nu ,\xi }(q))=\sum _{i=1}^\infty a_{\mu ,\nu ,\xi ,i}q^i$
and
$\log (B_{\mu ,\nu ,\xi }(q))=\sum _{i=1}^\infty b_{\mu ,\nu ,\xi ,i}q^i$
. Then we have

In the summation we have
$\ell (\pi )\gt 1$
because
$M_\kappa$
is empty for any
$\kappa |\pi$
if
$\pi =(n)$
, by definition. Therefore
$2\leqslant \sum _{i=1}^{\ell (\pi )}k_i\leqslant n$
. By multiplying the denominator and numerator of the right-hand side by some appropriate power of
$\lambda _1\lambda _2$
, we can express it as a rational function in
$\lambda _1,\lambda _2$
, with denominator
$\lambda _1^n\lambda _2^n$
and numerator a multiple of
$c_1(S)^2$
. Setting this multiple as
$f_n\in {\mathbb{C}}[\lambda _1,\lambda _2]$
gives the desired expression.
Example 3.8. The universal series for
${\mathcal{N}}_S$
are known explicitly in the compact case [Reference BojkoBoj21a, Theorem 1.2]. For a smooth projective surface
$S$
and
$\alpha$
of rank
$r$
, we apply [Reference Arbesfeld, Johnson, Lim, Oprea and PandharipandeAJL+21, Theorem 17, (16), (17)] for
$f(x)=1-ze^x,g(x)=\frac {x}{1-e^{-x}}$
and get

via the substitution

As mentioned in the introduction, when
$N=1$
, we shall set
$y_1=1$
, and omit the subscript
$\xi$
for
$H_{\mu ,\nu ,\xi }$
. The formula above allows us to compute
$H_{(1),(0)}(q,z)$
and
$H_{(0),(1)}(q,z)$
. For a smooth projective surface
$S$
and
$\alpha$
of rank
$0$
, we have

The exponent is interpreted as an intersection product, which in the toric case corresponds to the equivariant push-forward

Therefore, for rank 0, we have the following series from expansion (17)

We take the Chern limit of Lemma 3.3 by substituting
$q\leadsto -q{\varepsilon }$
,
$z\leadsto (1+{\varepsilon })^{-1}$
and get

Replacing
$\alpha$
by
$-{\alpha }$
in the Chern series to get the Segre series for
$\alpha$
, we see

On the other hand, the Verlinde limit yields

The Segre–Verlinde correspondence of Corollary 3.7 is indeed satisfied.
Example 3.9. Let
$S={\mathbb{C}}^2$
,
$n=N=2$
,
$E={\mathcal{O}}_S\langle y_1\rangle \oplus {\mathcal{O}}_S\langle y_2\rangle$
and
$L={\mathcal{O}}_S\langle v\rangle$
. The
${\mathsf{T}}_1$
-fixed locus of
$\textrm {Quot}_S(E,n)$
is the disjoint union of

Denote by
$Z_\mu$
the point in
$\textrm {Hilb}^i(S)^{{\mathsf{T}}_0}$
corresponding to a partition
$\mu$
, then the
$\mathsf{T}$
-fixed points of
$\textrm {Quot}_S({\mathbb{C}}^2,2)$
are

Therefore, by (7), the virtual tangent bundles at these five points are, respectively,

The equivariant Chern roots of
$\alpha ^{[n]}$
at these points are, respectively,

The contribution to the Segre numbers at each of these fixed points are





Summing them up, we have

A similar computation yields another complicated expression for the Verlinde number. We are interested in the total degree 0 part of their difference in the variables
$\vec \lambda ,\vec m,\vec w$
. This computes to

Note that even though the expressions for the Segre and Verlinde numbers are complicated, their difference in degree 0 simplifies tremendously and satisfies Corollary 3.7.
3.4 Reduced virtual classes and invariants
As mentioned previously, the obstruction for
$\textrm {Quot}_S(E,n)$
at
$Z$
contains at least one copy of
$K_S^\vee$
. For
$K$
-trivial surfaces, this causes the Euler class of
$T^{{\rm vir}}$
to vanish. Therefore, the virtual Verlinde and Segre numbers both vanish. One can instead study the ‘reduced’ versions of these invariants. By [Reference LimLim20, Proposition 9], when
$S$
is a
$K$
-trivial surface,
$n\gt 0$
, and
$E$
a torsion-free sheaf, there is a reduced obstruction theory that is perfect in the sense of Definition 3.1. The reduced (virtual) tangent bundle in this case is given by adding a trivial summand to the usual virtual tangent bundle:

In this section, we study the equivariant analogue where
$S={\mathbb{C}}^2$
with the natural action of the 1-dimensional torus

We write


Using the argument of [Reference Cao and KoolCK17, Lemma 3.1], we see that the
${\mathsf{T}}=({\mathsf{T}}_0\times {\mathsf{T}}_1\times {\mathsf{T}}_2)$
-fixed locus of
$\textrm {Quot}_S(E,n)$
stays unchanged, and the Zariski tangent space at each of the fixed points has no fixed parts, by (9). The equivariant reduced Segre and Verlinde series
$\mathcal{S}^{{\rm red}}_S$
and
$\mathcal{V}^{{\rm red}}_S$
are defined in the same way as the virtual ones by replacing
$T^{{\rm vir}}$
with
$T^{{\rm red}}$
. Here we omit the subscript since we are only interested in the
$S={\mathbb{C}}^2$
case.

Note that we do not include the
$n=0$
term because condition 2 of [Reference LimLim20, Proposition 9] is only satisfied when
$n\gt 0$
.
The same strategy as used in the previous section can be applied to study these invariants. For
$E=\oplus _{i=1}^N{\mathcal{O}}_S\langle y_i\rangle$
and
$V=\oplus _{i=1}^r{\mathcal{O}}_S\langle v_i\rangle$
, we define

Again note that the
$[q^0]$
coefficient is 0. We can think of the reduced obstruction as removing a copy of
$K_S^\vee$
from the usual obstruction in
${\mathbb{Z}}[t_1^{\pm 1},t_2^{\pm 1}]$
, then passing to the quotient ring
${\mathbb{Z}}[t_1^{\pm 1},t_2^{\pm 1}]/(t_1t_2-1)$
. This gives us the following corollary.
Corollary 3.10.
For
$n\gt 0$
, the
$[q^n]$
coefficient of
$\mathcal{N}^{{\rm red}}$
can be obtained from the non-reduced version by taking the following limit:

Using Remark 1.6, we may expand the universal series expression from Theorem 3.5 and obtain

Using the above corollary to extract the reduced coefficients, we see the terms with
$i\gt 1$
all vanish and

The Chern and Verlinde cases are similar, and thus we have the following result.
Theorem 3.11.
When
$S={\mathbb{C}}^2$
, the equivariant reduced Segre, Verlinde and Chern series for
$E=\oplus _{i=1}^N{\mathcal{O}}_S\langle y_i\rangle$
and
${\alpha }\in K_{\mathsf{T}}(S)$
are

where
$A_{\mu ,\nu ,\xi }, B_{\mu ,\nu ,\xi }$
and
$C_{\mu ,\nu ,\xi }$
are the same series as in Theorem 3.5
.
The integrals in the above theorem labeled by
$\mu ,\nu ,\xi$
have degree
$|\mu |+|\nu |+|\xi |-2$
, which is one degree lower than the integrals in the non-reduced expressions. Therefore, we have a Segre–Verlinde correspondence in degree
$-1$
for the reduced setting. However, the results for degree
$-1$
have no compact analogues since they automatically vanish in the compact setting. In § 4.2, we compute some of the universal series explicitly, giving us some Segre–Verlinde relations in non-negative degrees for the reduced case.
4. Explicit computations of universal series
4.1 Virtual Segre number in rank
$r=-1$
In this section, we first consider the non-virtual equivariant Chern series
$I^{\mathcal{C}}({\alpha };q)$
at rank
$r={\rm rank}({\alpha }) = 2$
for Hilbert schemes of
$S={\mathbb{C}}^2$
, given by (1). We will assume that
${\alpha }=[V]$
is the class of a rank 2 vector bundle
$V$
. After the proof of Lemma 4.1, a closed expression for
$I^{\mathcal{C}}({\alpha };q)$
will be used to compute the equivariant virtual Segre series
${\mathcal{S}}_S({\mathcal{O}}_S,{\beta };q)$
at rank
$r={\rm rank} ({\beta })=-1$
. This latter series is equivalent to the virtual Chern series
${\mathcal{C}}_S({\mathcal{O}}_S,\gamma ;q)$
at rank
$r={\rm rank}(\gamma )=1$
, where we can take
$\gamma =[L]$
to be the class of some line bundle
$L$
.
Consider the Chern series
$I^{\mathcal{C}}(V;q)$
for
$r={\rm rank}(V)=2$
, which has universal series structure

for some series
$G_{\mu ,\nu }$
, as a result of [Reference Göttsche and MellitGM22, (2.6)] and Proposition 2.7. On the other hand, expanding (1) using (4) gives us

If
$Z$
corresponds to a partition
$\mu$
, the numerator
$c(V^{[n]}|_Z)$
in
$H^*_{\mathsf{T}}({\rm pt})$
lies in degrees
$0$
to
$2|\mu |=2n$
, while the denominator
$e(T_Z)$
has degree exactly
$2|\mu |=2n$
. Hence, the total expression
$I^{\mathcal{C}}(V;q)$
ranges over the degrees
$-2n$
to
$0$
. From this observation, we conclude that the series
$G_{\mu ,\nu }$
must vanish for
$|\mu |+|\nu |-2\gt 0$
. Otherwise, expanding (18) using Remark 1.6 would give us a positive-degree term in
$I^{\mathcal{C}}(V;q)$
.
For a smooth projective surface
$S^{\prime}$
, let
$I_{S^{\prime}}^{\mathcal{C}}(V;q)=\sum _{n=0}^\infty q^n\int _{\textrm {Hilb}^n(S^{\prime})}s(V^{[n]})$
denote the ordinary (non-equivariant, non-virtual) Chern series for Hilbert schemes. By [Reference Marian, Oprea and PandharipandeMOP21, Remark 6], if
${\rm rank}(V)=2$
, then

By the argument of § 3.2, the universal series for
$I^{\mathcal{C}}_{S^{\prime}}$
coincides with the degree 0 universal series for
$I^{\mathcal{C}}$
, which are
$G_{\mu ,\nu }$
for
$|\mu |+|\nu |-2=0$
. Therefore
$G_{(2),(0)}=(1+q)$
and
$G_{(0),(2)}=G_{(1),(1)}=G_{(1,1),(0)}=G_{(0),(1,1)}=0$
. We have seen before that
$G_{\mu ,\nu }$
vanishes for degrees
$|\mu |+|\nu |-2\gt 0$
. Thus in order to compute
$I^{\mathcal{C}}$
for
$S={\mathbb{C}}^2$
, it remains to find the universal series
$G_{\mu ,\nu }$
with
$|\mu |+|\nu |-2\lt 0$
, i.e.
$G_{(1),(0)},G_{(0),(1)},G_{(0),(0)}$
.
As previously mentioned, Göttsche and Mellit used an invariant similar to the Nekrasov genus
$\mathcal{N}_S({\mathcal{O}}_S,V;q,z)$
, denoted by
$I_{S^{\prime},V}(q,z)$
in [Reference Göttsche and MellitGM22, (1.1)], for
$S^{\prime}$
a smooth projective toric surface. To match the notation we use, we shall write this as

The equivariant analogue on
$S={\mathbb{C}}^2$
is given by
$\Omega (q;ze^{w_1},\ldots ,ze^{w_r};e^{\lambda _1},e^{\lambda _2})$
in [Reference Göttsche and MellitGM22, § 3.2], which we denote by
$I^{\mathcal{N}}(V;q,z)=I^{\mathcal{N}}(v_1,\ldots ,v_r;q,z)$
where
$v_i=e^{w_i}$
. As in Lemma 3.3, the Chern series
$I^{\mathcal{C}}$
is recovered from
$I^{\mathcal{N}}$
by taking the Chern limit as in the first equation of [Reference Göttsche and MellitGM22, Proposition 3.5]; that is,

Since we are interested in the universal series for
$I^{\mathcal{C}}$
in negative degrees, let us compute
$I^{\mathcal{N}}$
in negative degrees.
The invariant
${{I}}^{\mathcal{N}}$
admits a universal series structure by [Reference Göttsche and MellitGM22, § 3.4] given as follows:

Here,
$H_{j,k}(q;z_1,\ldots ,z_r)$
are series labeled by integers
$j,k$
, symmetric in
$z_1,\ldots ,z_r$
. The expansion [Reference Göttsche and MellitGM22, (3.12)–(3.13)] states that

where the omitted terms have positive degrees in
$H_{\mathsf{T}}^*({\rm pt})$
. The series
$C_{2},C_{1,1},D_1$
can be computed explicitly using [Reference Göttsche and MellitGM22, (3.16)], which expresses them in terms of another collection of series
$G_0,G_1,G_3$
given by [Reference Göttsche and MellitGM22, Theorem 1.1]. We have

We are now in a situation to apply Lemma 4.1. We take the series
$H(z_1,\ldots ,z_r)$
in the lemma to be
$H_{-1,-1}(q;z_1,\ldots ,z_r)$
and the variables
$x_1,\ldots ,x_r$
to be
$w_1,\ldots ,w_r$
. The above formula for
$H_{-1,-1}$
shows that
$H(q;ze^{w_1},\ldots ,ze^{w_r})$
satisfies the expansion
$\sum _{\mu\, \mathrm{partition}}H_{\mu }(z)\prod _{i=1}^{\ell (\mu )}e_{\mu _i}(w_1,\ldots ,w_r)$
for
$H_{(0)}=C_0, H_{(1)}=C_1, H_{(2)}=C_2$
,
$H_{(1,1)}=C_{1,1}$
and
$H_\mu=0$
for any other index
$\mu$
. Lemma 4.1 states that

Similarly, applying this to
$H_{-1,0}$
yields
$D_zD_0=2D_1$
. Therefore

Here
${\rm Li}_s(z)={\rm Li}_s(z)=\sum _{k=1}^\infty z^k/k^s$
denote the polylogarithm functions. The terms
$C_0(q,0)$
(respectively
$D_0(q,0)$
) are obtained by extracting the coefficients of
$\lambda _1^{-1}\lambda _2^{-1}$
(respectively
$\lambda _1^{-1}$
or
$\lambda _2^{-1}$
) of
$\log I^{\mathcal{N}}$
via the expression [Reference Göttsche and MellitGM22, (2.5)]. We get

Finally, by the Chern limit (19) for rank
$r=2$
, we see

Similar limits for
$G_{(1),(0)}$
(respectively
$G_{(0),(1)}$
) using
$C_1$
(respectively
$D_0$
) yield
$\log G_{(1),(0)}=\log (1+q)$
and
$\log G_{(0),(1)}=0$
. Exponentiating, we get
$G_{(0),(1)}=1$
and
$G_{(0),(0)}(q)=G_{(1),(0)}=$
$1+q$
. Hence

for any
$\mathsf{T}$
-equivariant bundle
$V$
of rank 2.
Lemma 4.1.
Suppose
$r\geqslant 0$
and let
$H(z_1,\ldots ,z_r)$
be a power series in
$z_1,\ldots z_r$
whose coefficients are series in some other variables
$q_1,q_2,\ldots$
. If
$H$
is symmetric in
$z_1,\ldots ,z_r$
with an expansion

then for any
$k\geqslant 0$
, we have

where
$\binom {r}{\mu }$
denotes
$\prod _{i=1}^{\ell (\mu )}\binom {r}{\mu _i}$
, and
$D_z=z\frac {\partial }{\partial z}$
.
Proof.
We begin by claiming that the statement is closed under polynomial expressions; that is, if the equality holds for both
$F(z_1,\ldots ,z_r)$
and
$G(z_1,\ldots ,z_r)$
, then it holds for
$F\cdot G$
and
$aF+G$
for any
$a\in {\mathbb{Q}}[\![q_1,q_2,\ldots ]\!]$
. The additive part is straightforward, and we shall prove the multiplicative part of this claim. Expand

then
$H=F \cdot G$
can be expanded as

where by
$\nu +\xi$
we mean combining them as sequences to get a partition of size
$|\nu |+|\xi |$
with length
$\ell (\nu )+\ell (\xi )$
. Suppose the statement holds for both
$F$
and
$G$
, then we have

Since
$H(z_1,\ldots ,z_r)$
is symmetric, by the above observation, it suffices to prove the statement when
$H$
is the power-sum symmetric polynomial
$p_n(z_1,\ldots ,z_r)=z_1^n+\ldots +z_r^n$
. For each
$n\geqslant 0$
, we expand

This means
$H_{(0)}(z)=rz^n$
and

Fixing
$r\geqslant 1$
, we write

for some constant terms
$C_\mu$
. Evaluating at
$x_1=\ldots =x_r=1$
, we get

Hence

A quick calculation for the
$k=0$
and
$r=0$
cases finishes the proof.
Now we proceed with the virtual case. Recall that, on Hilbert schemes, the obstruction theory at a fixed point
$[Z_\mu ]$
is given by
$(K^{[n]}_S)^\vee |_{Z_\mu }$
, so

Let
$L={\mathcal{O}}_S\langle v_1\rangle$
be an equivariant line bundle, and
$V=L\oplus {\mathcal{O}}_S\langle v_2\rangle$
. Apply (19) to
$V$
and we have

Set
$w_2=c_1(K_S)-1$
and replace
$q$
by
$-q$
. Then this becomes

On the other hand, we have by definition

Therefore we conclude

In particular, restricting to the lowest degree part in the variables
$\lambda _1,\lambda _2,w_1$
, we obtain the following Corollary.
Corollary 4.2.
For
$S={\mathbb{C}}^2$
, the following equality holds:

4.2 Segre–Verlinde correspondence in non-zero degrees
Recall that we use the notation
$H_{\mu ,\nu ,\xi }$
for the series from (17) describing the virtual Nekrasov genus for Quot schemes on
$S={\mathbb{C}}^2$
. Note that
$\mathcal{N}_S$
satisfies

Applying Lemma 4.1 to
$H_{\mu ,\nu ,\xi }$
in the variables
$w_1,\ldots , w_r$
gives us, for
$r\gt 0$
,

and applying in the variables
$m_1,\ldots ,m_N$
yields

When the rank
$r$
is negative, we consider
${\alpha }=-[V]$
where
$V=\oplus _{i=1}^{-r}{\mathcal{O}}_S\langle v_i\rangle$
. Write

Then the same argument as used in Lemma 4.1 applies, and, for all
$r\neq 0$
,

Observe that when
$r\gt 0$
, the Chern limit of
$H_{\mu ,\nu ,\xi }$
returns the universal series for the Chern invariants
$\log C_{\mu ,\nu ,\xi }$
, and the Verlinde limit returns
$\log B_{\mu ,\nu ,\xi }$
. On the other hand, when
$r\lt 0$
, the Chern series retrieves the rank
$-r$
Segre invariants
$\log A_{\mu ,\nu ,\xi }$
, but the Verlinde limit does not give the Verlinde series. This is because
$H_{\mu ,\nu ,\xi }$
is associated to
$c_\mu (V)$
, whereas
$\log B_{\mu ,\nu ,\xi }$
is associated to
$c_\mu ({\alpha })=c_\mu (-[V])$
. This results in a change of basis for the symmetric series in the Chern roots of
$V$
, and Lemma 4.1 no longer applies for these negative ranks.
Our goal in this section is to apply Chern and Verlinde limits to (20) and (21) for
$k\gt 0$
, then obtain relations for the Chern and Verlinde series for various
$\mu$
,
$\nu$
, and
$\xi$
. We may obtain explicit expressions for when
$|\mu |+|\nu |+|\xi |=1$
from the compact setting. For a smooth projective surface
$S^{\prime}$
, a torsion-free sheaf
$E^{\prime}$
of rank
$N$
, and a K-theory class
$\alpha$
of rank
$r$
, the universal series structure for
$\mathcal{N}_S$
can be obtained from [Reference BojkoBoj21a, (5.1), Theorem 5.1] by setting

This gives us

where

The series
$H_i(q)$
are Newton–Puiseux solutions to

and
$G(R)$
is given explicitly by [Reference BojkoBoj21a, (4.24)]. Therefore

The sign
$\frac {r}{|r|}$
in the first line appears as a result of
$c_1(-V)=-c_1(V)$
.
4.2.1 The Chern limit.
We begin with the case
$\nu =\xi =(0)$
. Apply [Reference BojkoBoj21a, (4.17)], which gives a term-by-term expression for (22), by setting
$f(t)=1-ze^t$
and
$g(t)=\frac {t}{1-e^{-t}}$
. We have

By (20) and (21), we have, for
$k_1,k_2\geqslant 0$
and
$k:=k_1+k_2\gt 0$
,


where
$p_{n,k}$
is a polynomial of degree
$k$
. We may show inductively that
$p_{n,k}(1)=(nr-1)_{(k-1)}$
. With this expansion, we would like to apply the Chern limit of Lemma 3.3. In the expansion of universal series (17), the series
$H_{\mu ,(0),(0)}$
are multiplied by a term in
$\vec {\lambda },\vec {w},\vec {m}$
of homogeneous degree
$|\mu |-1=k-1$
. When taking the Chern limit, we need to make substitutions

Therefore, we need to multiply by a factor of
$(-{\varepsilon })^{k-1}$
when taking the limit of
$H_{\mu ,(0),(0)}$
. Furthermore, we substitute
$q\leadsto (-1)^Nq{\varepsilon }^{N-r}(1+{\varepsilon })^r$
and
$z\leadsto (1+{\varepsilon })^{-1}$
, and the right-hand side of the above expansion becomes

Since we are extracting the
$[t^{nN-1}]$
coefficient of the function inside the curly bracket, we may replace
$t\leadsto -{\varepsilon } t$
and divide the function by
$(-{\varepsilon })^{nN-1}$
, which gives us

Taking
${\varepsilon }\rightarrow 0$
, the term
$p_{n,k}\left (\frac {\exp (-{\varepsilon } t)}{1+{\varepsilon }}\right )$
converges to
$p_{n,k}(1)=(nr-1)_{(k-1)}$
. Thus the limit gives us

We further apply the following identity:

The Chern limit of the left-hand side of (24) will be the Chern series of rank
$r$
for
$r\gt 0$
and the Segre series of rank
$-r$
for
$r\lt 0$
. Therefore, for all
$r\gt 0$
,


4.2.2 The Verlinde limit.
We apply a similar argument for the Verlinde limit using (22). To simplify the computation, we consider a different change of variable. Let

Then the
$\tilde {H}_i$
are the Newton–Puiseux solutions to

Also,

so

Apply the Lagrange inversion theorem of [Reference Behrend and FantechiBoj23, Corollary 5] with
${\varphi }(t)=\log ((1-t-z)/$
$(1-t)(1-z))$
and
$R=(1-t-z)^r/(1-t)^{r}$
, we have

Here
$q_{n,k}(z)$
is a polynomial of degree
$k$
in
$z$
whose coefficients involve the variable
$t$
, and its leading coefficient is
$(nr)^{k-1}$
. To take the Verlinde limit for
$r\gt 0$
, we substitute

then take
${\varepsilon }\rightarrow 0$
and get

Observe that the Verlinde series for
${\alpha }\in K_{\mathsf{T}}(S)$
only depends on
$c_1({\alpha })$
by definition, so the universal series are non-trivial only when
$\mu =(1)_k:=(1,\ldots , 1)$
has
$k$
copies of 1. We can therefore simplify the left-hand side of the above equation and get

By [Reference Göttsche and MellitGM22, Lemma 3.3], the universal series are polynomials in
$r$
for
$r\geqslant 0$
. Therefore, the above results hold for the rank
$r=0$
case as well.
4.2.3 The degree
$-1$
case.
When
$k=0$
, the argument from above still applies, where
$D_z^{-1}(-)$
denotes the anti-derivative of
$\frac 1z(-)$
with respect to
$z$
. However, this would result in an undetermined constant term from the integration, so we deal with this case separately. We compute
$A_{(0),(0),(0)},B_{(0),(0),(0)},C_{(0),(0),(0)}$
using expressions for
$H_{(0),(0),(0)}$
, which we shall denote as
$A,B,C$
and
$H$
, respectively. Since these series are associated to the part of their respective invariants independent of the weights of
$\alpha$
, we have

and by the second part of [Reference Göttsche and MellitGM22, Lemma 3.3],
$A,B,C,H$
are polynomials with respect to
$r$
for all
$r\in {\mathbb{Z}}$
.
By Lemma 4.1,
$D_zH=|r|H_{(1),(0),(0)}$
. Taking
$D_z^{-1}$
of (23) with respect to
$z$
, we get

for some
$H_0(q)$
independent of the variable
$z$
.
We apply the result of § 4.2.1 with
$k=0$
. We see that
$H_1(q)$
admits the following Chern limit:

Write
$H_0(q)=\sum _{n=1}^\infty h_nq^n$
. Then its Chern limit is

When
$N-r\leqslant 0$
, we must have
$h_n=0$
for all
$n$
, since otherwise we would have negative powers on
$\epsilon$
and the limit does not make sense. As each
$h_n$
is polynomial in
$r$
, we conclude that
$H_0(q)=0$
. Hence, for all
$r\in {\mathbb{Z}}$
,


When
$0\leqslant r\leqslant N-1$
, the formula for
$C(q)$
is consistent with Conjectures 1.20 and 1.21.
Similarly by § 4.2.2, the Verlinde limit of
$H_{(0),(0),(0)}$
is

Combining the results of the above sections, we have the following theorem.
Theorem 4.3.
For rank
$r\geqslant 0$
and integers
$k_1,k_2\geqslant 0$
with
$k:=k_1+k_2$
, the universal series of Theorem 3.5 satisfy

Furthermore, we have

which can be compared to the identities above by replacing
$r$
with
$-r$
.
When
$k=2$
, we have the following Segre–Verlinde correspondences in degree
$1$
.
Corollary 4.4.
For rank
$r\geqslant 0$
, the universal series of Theorem 3.5 satisfy the following correspondences

Remark 4.5. As mentioned in the introduction, combining Theorem 4.3 with Theorem 3.11 yields the corresponding relations for reduced invariants. In particular, Corollary 4.4 implies a correspondence in degree 0 for reduced invariants, which could provide insight into the reduced invariants for K3-surfaces in the compact setting.
4.3 Universal series via Lagrange inversion
Let
$r\gt 0$
. Consider (25) and (26) from the previous section. The right-hand sides of these identities are linear combinations of series of the form

for
${\varphi }(t)=\log (1+t)$
and
$k\gt 0$
. The goal of this section is to express these series without the process of extracting coefficients.
By Lagrange inversion theorem [Reference Behrend and FantechiBoj23, Corollary 2], we have

where
$D_q=q\frac {d}{d q}$
and
$H_i$
are the Newton–Puiseux solutions to

Note that

Here
$H_i(q^{\frac 1N})$
is a Puiseux series, and by
$dH_i/dq$
we mean
$(dq/dH_i)^{-1}$
. Differentiating both sides of (27) with respect to
$H_i$
yields

Let
$\psi (t)=Nt^{-1}+r(1+t)^{-1}$
, then
$D_qH_i=\frac 1{\psi (H_i)}$
. Define
$D_\psi =\frac 1\psi \cdot \frac {d}{dt}$
. We conclude that

for arbitrary power series
${\varphi }(t)$
. Therefore

Theorem 4.6 follows directly by applying this to (25) and (26).
Theorem 4.6.
Let
${\varphi }(t)=\log (1+t)$
and
$\psi (t)=Nt^{-1}+r(1+t)^{-1}$
. Define the differential operator

Furthermore, use the notation

for
$k\geqslant 0$
, where
$D_\psi ^{-1}(-)$
denotes integrating
$\psi \cdot (-)$
assuming a constant term 0. In the setting of Theorem 4.3
, we have the following relations

where the
$H_i$
are the Newton–Puiseux solutions to
$H_i^N=q(1+H_i)^{-r}$
.
Example 4.7. We shall compute
$\log B_{(1,1,1),(0),(0)}$
using the above theorem. Set
$k_1=3,k_2=0$
. We have

Hence

Example 4.8. Similarly, we may compute
$\sum _{|\mu |=3}\binom {r}{\mu }\log A_{\mu ,(0),(0)}$
:

Again, the
$H_i$
are the Newton–Puiseux solutions to
$H_i^N=q(1+H_i)^{-r}$
.
4.4 Strong Segre symmetry and weak Verlinde symmetry
Let
$r\gt 0$
. According to Theorem 4.3, the following identity holds for the Segre series:

The right-hand side does not change if we swap
$r$
and
$N$
; that is

This is consistent with the strong Segre symmetry of Conjecture 1.15. We have checked that this conjecture holds when

As for the Verlinde series, the weak Segre symmetry of Corollary 1.14 together with the weak Segre–Verlinde correspondence of Corollary 1.9 would imply a weak Verlinde symmetry. However, calculations using a computer program show the ‘strong’ Verlinde symmetry does not hold for
$S={\mathbb{C}}^2$
. This can also be observed from the fact that the equivariant Verlinde series only depends on
$V$
through
$c_1(V)$
but depends on
$E$
through
$c_1(E),c_2(E),\ldots ,c_N(E)$
, thus breaking the symmetry.
5. Segre and Verlinde invariants on
${\mathbb{C}}^4$
Consider
$X={\mathbb{C}}^4$
with a
$({\mathbb{C}}^4)^*$
-action by scaling coordinates

Let
${\mathsf{T}}_0=\{(t_1,t_2,t_3,t_4):t_1t_2t_3t_4=1\}\subseteq ({\mathbb{C}}^4)^*$
be the subtorus which preserves the usual volume form on
$X$
, making
$X$
a smooth quasi-projective toric Calabi–Yau 4-fold. As in the surface case, we include two additional tori

where
${\mathsf{T}}_1$
acts naturally on
${\mathbb{C}}^N$
, and
${\mathsf{T}}_2$
acts naturally on
${\mathbb{C}}^{r}\times {\mathbb{C}}^s$
. Set
${\mathsf{T}}={\mathsf{T}}_0\times {\mathsf{T}}_1\times {\mathsf{T}}_2$
, and consider
$E=\oplus _{i=1}^N{\mathcal{O}}_X\langle y_i\rangle , {\alpha }=[\oplus _{i=1}^r{\mathcal{O}}_X\langle v_i\rangle ]-[\oplus _{i=r+1}^{r+s}{\mathcal{O}}_X\langle v_i\rangle ]$
. Write

By [Reference Huybrechts and ThomasHT08, Theorem 4.1] the truncated Atiyah class of the universal subsheaf
$\mathcal{I}$
defines an obstruction theory

where
${\bf R}\mathscr{H}om_{q}={\bf R}q_*\circ {\bf R}\mathscr{H}om$
, and
$(\cdot )_0$
denotes the trace-free part. Note that the obstruction theory is
$\mathsf{T}$
-equivariant by [Reference RicolfiRic21, Theorem B]. The virtual tangent bundle is then

5.1 Cohomological virtual invariants
When
$X$
is a projective Calabi–Yau 4-fold, the virtual fundamental class involves a choice of orientation on
$\textrm {Quot}_X(E,n)$
. Let
$\mathcal{L}={\rm det}{\bf R}\mathscr{H}om_{q}(\mathcal{I},\mathcal{I})$
be the determinant line bundle. An orientation
$o(\mathcal{L})$
is a choice of square root of the isomorphism

induced by Serre duality. A virtual class
$[\textrm {Quot}_X(E,n)]^{{\rm vir}}_{o(\mathcal{L})}\in H_{2nN}(\textrm {Quot}_X(E,n),{\mathbb{Z}})$
was constructed in [Reference BojkoBoj21a, § 2.1] for
$X$
a strict Calabi–Yau 4-fold and
$E$
a simple rigid locally free sheaf. For
$\gamma \in H^{2nN}(\textrm {Quot}_X(E,n))$
, the holomorphic Donaldson invariants are defined to be

For the non-compact
$X={\mathbb{C}}^4$
, similar to the surface case, we have, by [Reference Cao and KoolCK17, Lemma 3.6], that
$T^{{\rm vir}}$
has no fixed parts and the
$\mathsf{T}$
-fixed locus of
$\textrm {Hilb}^n(X)$
consists of only finitely many reduced points. Thus one can define these invariants equivariantly using Oh and Thomas’ localization formula [Reference Oh and ThomasOT20, Theorem 7.1]. For a fixed orientation and any
$SO(2k)$
-bundle
$B$
, one can define its Edidin–Graham square-root Euler class
$\sqrt {e}(B)$
[Reference Oh and ThomasOT20, § 3]. Consider the self-dual resolution (33) of
$T^{{\rm vir}}$

where
$B^\bullet = (T\rightarrow B\rightarrow T^*)$
and
$B$
is a
$SO(6n^2)$
-bundle. Set [Reference Oh and ThomasOT20, (115)]

For computational purposes, we consider a square root
$\sqrt {T^{{\rm vir}}|_Z}\in K_{\mathsf{T}}({\rm pt})$
for each fixed point
$Z$
such that

where
$\overline {(\cdot )}$
denotes the involution
$t_i\mapsto t_i^{-1}$
. The existence of this square root follows from
$T^{{\rm vir}}|_Z$
being self-dual and containing no fixed part (see discussion before [Reference Cao and KoolCK17, Definition 3.15]). This allows us to compute the square-root Euler class at the cost of a sign dependent on the choice of orientation
$o(\mathcal{L})$

As each fixed point is reduced, Kool and Rennemo [Reference Kool and RennemoKR] show that their virtual fundamental classes constructed with respect to the fixed parts of the obstruction theory are given by further signs determined by an orientation on
$B^\bullet$
. We will denote the product of the two signs at each
$Z\in \textrm {Quot}_X(E,n)^{\mathsf{T}}$
by
$(-1)^{o(\mathcal{L})|_Z}$
.
Definition 5.1. For
$n\gt 0$
,
$\gamma \in H^{*}_{\mathsf{T}}(\textrm {Quot}_X(E,n))$
, the holomorphic Donaldson invariants are

The authors of [Reference Kool and RennemoKR] also gave an explicit description of the moduli space
$\textrm {Quot}_X(E,n)$
when
$X={\mathbb{C}}^4$
as a vanishing locus of an isotropic section. This allowed them to derive the signs
$(-1)^{o(\mathcal{L})|_Z}$
by knowing
$\mathcal{L}$
and
$T^{{\rm vir}}$
. For the purpose of the Verlinde invariants from the introduction, we recall their approach here.
Consider the quiver with four loops and
$N$
framings as in Figure 3. After imposing the relations

its representations with dimension vector
$(1,n)$
consist of a one-dimensional complex vector space
$\mathbb{C}$
and an
$n$
-dimensional vector space
$V$
together with
$N$
morphisms
$f_i: {\mathbb{C}}\to V$
for
$i=1,\ldots ,N$
and 4 morphisms
$x_i: V\to V$
satisfying (30).

Figure 3. Framed quiver with four loops at one node.
The space of all representations without requiring the relations is

where we use the suggestive notation
$E={\mathbb{C}}^N$
. This carries an
$SO(6n^2)$
vector bundle
$B$
that is trivial with fiber

and pairing

The existence of an isotropic section

connects it to the local toy model of [Reference Oh and ThomasOT20, (1)].
To construct the Quot scheme, we need to take a quotient by the
$GL(V)$
-action on
$R$
defined for each
$g\in GL(V)$
by
$\big (\vec {x},\vec {f}\big )\mapsto \big (g\circ \vec {x}\circ g^{-1}, g\circ \vec {f}\big )$
. One can also extend it in a natural way to an action on
$B$
. In particular, after defining
$R^0\subset R$
as the open subscheme of representations satisfying

the vector bundle
$B$
and its section
$s$
restrict and then descend to the non-commutative Quot scheme

The usual Quot scheme is identified with the zero locus

We will always use the letters
$B$
and
$s$
to denote the vector bundle and section on
$R$
,
$R^0$
, their descent to
${\rm NC}\textrm {Quot}_{X}(E,n)$
and their restriction to
$\textrm {Quot}_{X}(E,n)$
without distinguishing the four cases.
To make it all equivariant, one uses the action of
${\mathsf{T}}_0\times {\mathsf{T}}_1$
on
$R$
by

which commutes with the action of
$\textrm {GL}(V)$
so that it descends to one on
$\textrm {Quot}_{X}(E,n)$
. If we use
$T$
to denote the tangent bundle of
$A$
, which at each point is the cokernel of some injective map

obtained by differentiating the action of
$\textrm {GL}(V)$
on
$R$
, then there is a
${\mathsf{T}}_0\times {\mathsf{T}}_1$
-equivariant resolution of
${\bf R}\mathscr{H}om_{p}(\mathcal{I},\mathcal{I})^\vee _0[-1]$
given by

which gives the natural
${\mathsf{T}}_0\times {\mathsf{T}}_1$
-equivariant obstruction theory

Note that the first term in (32) has trivial weights.
The choice of orientations
$o({\mathcal{L}})$
was shown in [Reference Oh and ThomasOT20, Proposition 4.2] to be equivalent in this setting to choosing a positive isotropic subbundle
$\mathsf{I}$
of
$B$
. This is done explicitly in [Reference Kool and RennemoKR] by constructing
$ {\mathsf{I}}$
as a trivial bundle with the fiber

for some non-zero vector
$v\in {\mathbb{C}}^4$
.Footnote
1
We will not go further in recalling the explicit derivation of
$o({\mathcal{L}})|_Z$
in [Reference Kool and RennemoKR], as we only do computations up to a fixed order
$n$
to formulate conjectures, the proof of which we leave for the future.
Definition 5.2. Let
$X={\mathbb{C}}^4$
,
$\alpha =[\oplus _{i=1}^r{\mathcal{O}}_X\langle v_i\rangle ]-[\oplus _{i=r+1}^{r+s}{\mathcal{O}}_X\langle v_i\rangle ]\in K_{\mathsf{T}}(X)$
, and
$E=\oplus _{i=1}^N$
${\mathcal{O}}_X\langle y_i\rangle$
. The equivariant Segre and Chern series for a choice of signs
$o(\mathcal{L})$
are, respectively,

5.2 K-theoretic virtual invariants
In the setting where the moduli space
$M$
is a zero locus of an isotropic section
$s$
of an
$SO(2m)$
bundle
$B$
on some ambient space
$A$
, just as in (31) above, [Reference Oh and ThomasOT20] give a simpler construction of
$\hat {{\mathcal{O}}}^{{\rm vir}}$
relying on their equivariant localized K-theoretic square-root Euler class

defined after choosing orientations on
$M$
.
When
$B$
admits an isotropic subbundle
$\mathsf{I}$
compatible with the choice of orientation, then the push-forward under the inclusion

becomes just tensoring with the equivariant K-theoretic square-root Euler class

where
$\mathfrak{e}_{{\mathsf{T}}}({\mathsf{I}}^*) = \Lambda _{-1}{\mathsf{I}}^*$
. In fact,
$\sqrt {\mathfrak{e}_{{\mathsf{T}}}}(E,s)$
can also be written as the product

where
$\mathfrak{e}_{{\mathsf{T}}}({\mathsf{I}}^*,s)$
is some localization of
$\mathfrak{e}_{{\mathsf{T}}}({{\mathsf{I}}^*})$
to
$M=s^{-1}(0)$
constructed using Kiem and Li’s cosection localization in K-theory [Reference Kiem and LiKL20]. Their equivariant twisted virtual structure sheaf
$\hat {{\mathcal{O}}}^{{\rm vir}}$
is then constructed as

For our case of
$M=\textrm {Quot}_{{\mathbb{C}}^4}(E,n)$
, we can use (32) and (33) to show that

and
$\det ({\mathsf{I}}) = t_4^{2n^2}$
. This implies that the only term in the construction of
$\hat {{\mathcal{O}}}^{{\rm vir}}$
that does not admit a square root is

This gives further motivation for the definition of the untwisted virtual structure sheaf

that appeared in [Reference Behrend and FantechiBoj21b, Definition 5.10], [Reference BojkoBoj21a, § 1.4]. Here, the superscript
$(-)^{\frac {1}{2}}$
refers to taking the uniquely defined square root of a line bundle (see [Reference Oh and ThomasOT20, Remark 5.2]). From the above discussion it is clear that the following integrality statement holds.
Proposition 5.3. The untwisted virtual structure sheaf is an integral class:

Using
$\hat {{\mathcal{O}}}^{{\rm vir}}$
and
${\mathcal{O}}^{{\rm vir}}$
, we define the following twisted and untwisted Euler characteristics:


For compact
$X$
and
$\alpha \in K^0(X)$
, the Verlinde series [Reference BojkoBoj21a, § 1.3] is then defined by

Using the virtual Riemann–Roch formula and the equivariant localization of Oh and Thomas [Reference Oh and ThomasOT20, Theorem 6.1, Theorem 7.3], we have the following twisted equivariant virtual Euler characteristic for
$X={\mathbb{C}}^4$
:

where
$\sqrt {{\textrm {td}}}$
is the the square-root Todd class satisfying

Following the notation used in [Reference Cao, Kool and MonavariCKM22, § 0.1], we write

Substituting into the above equation, we have

Now including the twist of (34), we may define the equivariant Verlinde series as follows.
Definition 5.4. The equivariant Verlinde series for a choice of signs
$o(\mathcal{L})$
is

The relation between Segre and Verlinde numbers in the compact case is studied in [Reference BojkoBoj21a] using the Nekrasov genus for Hilbert schemes, introduced for the 3-fold case by [Reference Nekrasov and OkounkovNO14]. We consider the following Quot scheme version, originally defined as an instanton partition function in [Reference Nekrasov and PiazzalungaNP19, (2.7)], then rephrased as follows in [Reference Cao, Kool and MonavariCKM22, Remark 0.4]:

Remark 5.5. Recall that in [Reference Cao, Kool and MonavariCKM22], the Nekrasov genus for Hilbert schemes involves a variable
$y$
coming from a trivial
${\mathbb{C}}^*$
-action on
$X$
. This is exactly the
$N=1$
case for the above definition, where we have the parameter
$y_1$
from the
${\mathsf{T}}_1$
-action on
$E$
.
5.3 Vertex formalism
The invariants in the previous section can be calculated for Hilbert schemes using a vertex formalism developed by [Reference Cao and KoolCK20], based on the method introduced in [Reference Maulik, Nekrasov, Okounkov and PandharipandeMNOP06a] for Calabi–Yau 3-folds. We generalize this to Quot schemes using the computations from [Reference BojkoBoj21a, § 2.1]. First, when
$N=1$
, the
$\mathsf{T}$
-fixed points of
$\textrm {Hilb}^n(X)$
correspond to monomial ideals of
${\mathbb{C}}[x_1,x_2,x_3,x_4]$
[Reference Cao and KoolCK17, Lemma 3.1], which are labeled by solid partitions
$\pi$
of size
$n$
where

We denote by
$Q_\pi$
the character of
${\mathcal{O}}_{Z_\pi }$
:

As in the surface case, for
$E=\oplus _{i=1}^N{\mathcal{O}}_X\langle y_i\rangle$
the
$\mathsf{T}$
-fixed points for
$\textrm {Quot}_X(E,n)$
are labeled by
$N$
-colored solid partitions
$\pi =(\pi ^{(1)},\ldots ,\pi ^{(n)})$
of size
$n$
, i.e. sequences of the form

such that each
$Z_i$
corresponds to the solid partition
$\pi ^{(i)}$
.
Let
$Q_i$
be the character of
${\mathcal{O}}_{Z_i}$
. The virtual tangent bundle at
$Z_\pi$
is

where
$P(I)$
is the Poincaré polynomial of
$I$
defined analogously to (8), and
$P_{I}:=\prod _{i\in I}(1-t_i^{-1})$
for the set of indices
$I$
. Specializing to
$t_1t_2t_3t_4=1$
, we get the following (non-unique) square root

in the sense that it satisfies (29). The reason for this choice of square root is that

matches our twist in (34), and this simplifies our computation as we now have

In this case, the signs
$(-1)^{o(\mathcal{L})}$
are described in [Reference MonavariMon22, Reference Kool and RennemoKR] as follows: for any solid partition
$\pi$
,

and for any
$N$
-colored solid partition
$\pi$
,

The fiber of
$V^{[n]}=\oplus _{i=1}^r {\mathcal{O}}_X^{[n]}\langle v_i\rangle$
over
${Z_\pi }=(Z_1,\ldots Z_N)$
is the
$rn$
-dimensional representation

Therefore, for any point
$Z$
corresponding to an
$N$
-colored solid partition
$\pi$
, we have

Using these expressions, we see that the Chern and Verlinde series can be extracted by taking limits of the Nekrasov genus, as in the surface case. Also, it follows that

The argument of [Reference Cao, Kool and MonavariCKM22, Propositions 1.13 and 1.15] can be applied to show that
${\mathcal{N}}_X(E,V;q)$
in fact lives in
$\frac {{\mathbb{Q}}(t_1,t_2,t_3,t_4)}{(t_1t_2t_3t_4-1)}[\![q,y_1^{\pm \frac 12},\ldots ,y_N^{\pm \frac 12},v_1^{\pm \frac 12},\ldots ,v_r^{\pm \frac 12}]\!]$
. This enables us to talk about admissibility (up to specializing to
$t_1t_2t_3t_4=1$
) in the sense of Definition 2.6.
5.4 Factor of
$c_3(X)$
In the surface case, we saw that the powers in the universal series of virtual invariants are multiples of
$c_1(S)$
. In the
$X={\mathbb{C}}^4$
case, we shall show that if the universal expressions exist, then they are multiples of
$c_3(X)$
, by showing that

has
$c_3(X)=-(\lambda _1+\lambda _2)(\lambda _1+\lambda _3)(\lambda _2+\lambda _3)$
in its numerator. This factor of
$c_3(X)$
and the weak Segre–Verlinde correspondence and Segre symmetry of Corollaries 1.7 and 1.14 in the surface case motivate Conjecture 1.19. This is because the only degree zero contribution linear in
$c_3(X)$
is expected to come from
$\int _X c_3(X)c_1(\alpha )$
, which was already studied in the compact case in [Reference BojkoBoj21a, § 5.3]. We do not expect any additional terms with exponential
$\int _X c_3(X)c_1(\alpha )$
coming from the equivariant setting, just as we did not have any in the case of a surface.
It suffices to show that this term vanishes when we set
$\lambda _i=-\lambda _j$
for
$i\neq j$
in
$\{1,2,3\}$
. By symmetry, we may assume
$i=1,j=2$
. Recall that
$e$
is the top equivariant Chern class, which vanishes if its input has a trivial summand. The process of setting
$\lambda _1=-\lambda _2$
in cohomology is the same as setting
$t_1=t_2^{-1}$
in K-theory. Therefore, we would like to show that
$-\sqrt {T^{{\rm vir}}|_{Z_\pi }}$
has a trivial summand when we set
$t_1=t_2^{-1}$
, i.e. the character of
$\sqrt {T^{{\rm vir}}|_{Z_\pi }}$
in
$K_{\mathsf{T}}({\rm pt})$
having a strictly negative constant term. This occurs if and only if the image of
$T^{{\rm vir}}_{Z_\pi }$
in

has a strictly negative constant term (which is necessarily a negative even integer). From (36), we see that it suffices to show this for the term

whenever
$\pi$
is a non-trivial solid partition.
Lemma 5.6.
For any non-trivial solid partition
$\pi$
, the expression

has a strictly negative constant term when viewed in the quotient ring

Proof.
Let
$x=t_1=\frac 1{t_2}$
,
$y=t_3=\frac 1{t_4}$
, so that

Let
$P_\pi$
be the image of
$Q_\pi$
in
${\mathbb{Z}}[x^{\pm 1},y^{\pm 1}]$
. Then

Write

The image of
$\overline {Q_\pi }$
is then

We see that the constant terms of
$P_\pi$
and
$\overline {P_\pi }$
are both
$p_{0,0}$
. By definition, all monomial terms in
$Q_\pi$
have positive coefficients, and
$Q_\pi$
has constant term 1, so
$p_{0,0}\gt 0$
. We need to find the constant term of
$P_\pi \overline {P_\pi } (1-x)(1-\frac 1x)(1-y)(1-\frac 1y)$
.
Observe that

We write
$F=\sum f_{i,j}x^iy^j$
. The constant term of
$F\cdot (1-x)(1-\frac 1x)(1-y)(1-\frac 1y)$
is equal to

If we set
$F=P_\pi \overline {P_\pi }$
, then

In particular,

We write


Then (38) becomes

For the remainder of this proof, we shall show that
$S^{++}\geqslant p_{0,0}$
. The same will hold for the summands
$S^{+-},S^{-+},S^{{-}{-}}$
by symmetry. We conclude that the value of (38) is at least
$4p_{0,0}$
. Hence, by (37), the constant term of
$T^{{\rm vir}}|_{Z_\pi }$
is at most
$-2p_{0,0}\lt 0$
, and we are done.
Recall that

so

and

Fix
$k$
and
$l$
, and then the set
$\{(i,j): (i,j,k,l)\in \pi \}$
is a plane partition. By property (5) of solid partitions, for fixed
$b=k-l$
, we have
$p_{a,b}\geqslant p_{{a+1},b}$
when
$a\geqslant 0$
. For the same reason, we have
$p_{a,b}\geqslant p_{{a},b+1}$
when
$b\geqslant 0$
. Therefore, the numbers
$(p_{a,b})_{a,b\geqslant 0}$
are non-increasing as the pair
$(a,b)$
move away from the origin.
We apply induction on
$\max \{a:p_{a,0}\neq 0\}$
. Suppose for all sequences
$(q_{a,b})$
with
$\max \{a:q_{a,0}\neq 0\}\lt \max \{a:p_{a,0}\neq 0\}$
, we have

whenever the sequence
$(q_{a,b})_{a,b\geqslant 0}$
satisfies the property that
$q_{a,b}$
is non-increasing in
$a,b$
. The base case is simply when
$q_{a,b}=0$
for all
$a,b$
, which sums to 0. Let
$q_{a,b}=p_{{a+1},b}$
, then

where the first equality follows from the definition and the inequality is by the induction hypothesis.
Now apply another induction on the value of
$\max \{b:p_{0,b}\neq 0\}$
. The induction hypothesis is that for any sequences
$(q_{a,b})$
with
$q_{a,b}$
non-increasing in
$a,b$
and
$\max \{b:q_{0,b}\neq 0\}\lt$
$\max \{b:p_{0,b}\neq 0\}$
, we have

Again, the base case is trivial, and we can apply the hypothesis to
$q_{a,b}=p_{a,b+1}$
, giving us

So we have the following inequalities

where the last inequality is due to
$p_{0,0}-p_{1,0}-p_{0,1}+p_{1,1}$
being an integer. Therefore,

which finishes the second induction. By (39),

which finishes the first induction and the proof.
We checked the statements in Conjecture 1.19 using a computer program. The source code is available at [Reference HuangHua23]. The correspondence part was checked with a computer program for

The symmetry part was checked for

5.5 Cohomological limits
Recall that the proof of Theorem 3.5 mainly involved showing that the genus
$\mathcal{N}_S$
on the surface
$S={\mathbb{C}}^2$
is admissible in the sense of Definition 2.6. Also, by Proposition 2.7, universal series expressions for the Nekrasov genus, and therefore the Segre and Verlinde series, can be obtained if and only if the Nekrasov genus is admissible. Thus one might ask when the Nekrasov genus
$\mathcal{N}_X$
for the Calabi–Yau 4-fold
$X={\mathbb{C}}^4$
is admissible. For the rank
$r=N$
case, we shall show that admissibility is a consequence of the following explicit formula, as conjectured by Nekrasov and Piazzalunga [Reference Nekrasov and PiazzalungaNP19, § 2.5]. We write

The conjecture is in regards to any toric Calabi–Yau 4-fold
$X$
(see [Reference Cao, Kool and MonavariCKM22, p. 3, footnote 1]). Say the maximal dense open torus
$({\mathbb{C}}^4)^*\subseteq X$
contains the torus
${\mathsf{T}}_0\cong ({\mathbb{C}}^*)^3$
that preserves the volume form. This gives a
$\mathsf{T}$
-action on
$X$
similar to the
${\mathbb{C}}^4$
case with weights labeled by
$t_1,\ldots ,t_4,y_1,\ldots ,y_N,v_1,\ldots v_r$
as before.
Conjecture 5.7 (Nekrasov–Piazzalunga). Let
$X$
be a toric Calabi–Yau 4-fold. There exists some choice of signs
$o(\mathcal{L})$
such that for
$E=\oplus _{i=1}^N{\mathcal{O}}_X\langle y_i\rangle$
,
$V=\oplus _{i=1}^N{\mathcal{O}}_X\langle v_i\rangle$
,

with a change of variables
$s=\prod _{i=1}^Ny_iv_i$
.
Proposition 5.8.
Nekrasov and Piazzalunga’s Conjecture 5.7 implies that the Nekrasov genus
$\mathcal{N}_X$
of rank
$r=N$
is admissible with respect to the variables
$t_1,t_2,t_3,t_4$
.
Proof.
Expanding the term inside the plethystic exponential, and specializing with the relation
$t_1t_2t_3t_4=1$
, we have

Recalling Definition 2.6, we have that

is a series in
$q,y_1^{\pm \frac 12},\ldots ,y_N^{\pm \frac 12},v^{\pm \frac 12}_1,\ldots ,v^{\pm \frac 12}_r$
whose coefficients are polynomials in
$t_1,t_2,t_3,t_4$
, as required.
Lastly, we prove the claim made in the introduction that Conjecture 1.20 is a consequence of Conjecture 5.7 in the
$X={\mathbb{C}}^4$
case.
Proposition 5.9.
Let
$X={\mathbb{C}}^4$
. If Conjecture 5.7 holds for some choice of signs, then Conjecture 1.20 holds for
$Y=X$
.
In particular, we may retrieve the following well-known identity from Conjecture 5.7 :

Remark 5.10. One can compare this to the 3-fold case, where [Reference Fasola, Monavari and RicolfiFMR21, Theorem 7.2] states

Here
$M$
denotes the MacMahon function.
Proof. We shall compute the following limit using both the definition and the expression from Conjecture 5.7, and then compare the two sides:

Let
$V=\oplus _{i=1}^N{\mathcal{O}}_X\langle v_i\rangle$
be a rank
$N$
bundle, then for any
$Z_\pi \in \textrm {Quot}_X(E,n)^{\mathsf{T}}$
, we have

Take the limit
${\varepsilon } \rightarrow 0$
and let
$Q=m_Nq$
, then

Now take
$w_N\rightarrow \infty$
and substitute into (35). Let
$V^{\prime}=\oplus _{i=1}^{N-1}{\mathcal{O}}_X\langle v_i\rangle$
, then

On the other hand, we apply the same procedure to

For
$n\geqslant 1$
, we have

Note that the right-hand side is independent of the weights on
$V$
. Together with (42), we have

This is exactly Conjecture 1.20.
With the same method, we can take limits

for
$1\lt i\leqslant N$
and
$V$
of rank
$N-i$
, and get

In particular, when
$i=N$
and
$N\gt 1$
, we have

Acknowledgements
We would like to thank Rahul Pandharipande for asking about equivariant Segre–Verlinde correspondences during the first author’s seminar talk, and Henry Liu whose box-counting code sped up the progress of this project. We further wish to express gratitude towards Martijn Kool and Sergej Monavari for helpful discussions on the topic. We also thank the anonymous referee for their helpful comments to improve the paper.
Financial Support
A.B. was supported by ERC-2017-AdG-786580-MACI. This project has received funding from the European Research Council (ERC) under the European Union Horizon 2020 research and innovation program (grant agreement no. 786580).
Conflicts of Interest
None.
Journal Information
Moduli is published as a joint venture of the Foundation Compositio Mathematica and the London Mathematical Society. As not-for-profit organisations, the Foundation and Society reinvest
$100\%$
of any surplus generated from their publications back into mathematics through their charitable activities.