1. Introduction
There is a well-studied connection between the combinatorics of the nabla operator of [Reference Bergeron and GarsiaBG99, Reference Bergeron, Garsia, Haiman and TeslerBGHT99], and the homology or cohomology of the affine Springer fibers
$\mathcal{B}_{\gamma}$
of the sort studied in [Reference Goresky, Kottwitz and MacphersonGKM03], see for instance [Reference Lusztig and SmeltLS91, Reference Goresky, Kottwitz and MacPhersonGKM04, Reference Gorsky and MazinGM13, Reference Gorsky, Oblomkov, Rasmussen and ShendeGORS14, Reference Gorsky, Mazin and VaziraniGMV14, Reference HikitaHik14, Reference Oblomkov and YunOY14, Reference Carlsson and OblomkovCO18, Reference KivinenKiv20]. In this picture, objects such as parking functions
$(\pi,w)$
are seen to be in bijection with cells in an affine paving of
$\mathcal{B}_\gamma$
, and the well-studied statistics such as
$\mathrm{dinv}(\pi,w)$
that appear in the shuffle theorem [Reference HaglundHag08, HHL
$^{+}$
05a, Reference Carlsson and MellitCM18] and other nabla-type formulas are essentially the dimensions of the corresponding cells. In this way, combinatorial formulas may be interpreted as graded characters of the homology of some
$\mathcal{B}_\gamma$
, with the q-degree representing twice the homological degree, the t-degree being more subtle.
For example, consider the following power series:

Here,
$W^+_n$
is a set of extended affine permutations in which
$w(i)\geq 1$
for
$1\leq i \leq n$
. In other words, the positivity condition says that in ‘window notation,’ the entries of
$w=(w_1,\ldots ,w_n)$
are all positive, where
$w_i=w(i)$
. In the second condition, we say that w said to be m-stable if
$w_{i+m}> w_i$
for all i. The statistics area and
$\mathrm{dinv}_m$
are also defined in § 5.1. A slightly different version of this series was presented in [Reference Gorsky, Mazin and VaziraniGMV14] in the case when n,m are relatively prime, which the authors showed describes the combinatorics of the rational version of the shuffle theorem [Reference Bergeron, Garsia, Sergel Leven and XinBGSLX16, Reference MellitMel21]. The corresponding Springer fiber in this case is
$\mathcal{B}_\gamma=\mathcal{B}_{n,m}\subset \mathcal{Y}_n$
, which is the one associated to the nil-elliptic operator
$\gamma=N^m$
, where
$N(e_i)=e_{i+1}$
for
$i\leq n$
,
$N(e_n)=te_1$
, contained in the affine flag variety
$\mathcal{Y}_n=G(\mathbb{C}((t)))/G(\mathbb{C}[[t]])$
for
$G=GL_n(\mathbb{C})$
.
There is an extension of
$\mathcal{B}_{n,m}$
for n,m not relatively prime, which, in the case of
$m=kn$
, becomes the unramified affine Springer fiber studied in [Reference Goresky, Kottwitz and MacPhersonGKM04], and is defined in § 5.3. In this case, the equivariant homology
$H_*^T(\mathcal{B}_{n,kn})$
for a standard torus action
is equipped with two commuting actions of the symmetric group, generally known as the ‘dot’ and ‘star’ actions which act on the left and right, respectively, due to Knutson and Tymoczsko [Reference KnutsonKnu03, Reference TymoczkoTym08]. The dot action comes from a space-level action on the affine flag variety, which permutes different fibers
$\mathcal{B}_{n,kn}$
, whereas the right action comes from the Springer action. In this paper, we present an LLT-type expansion

for
$k\geq 1$
, where the quantities in the summand are defined in § 2.3. We predict that
$\Omega_k[X,Y]$
corresponds to a Frobenius character extension of (1), namely
$\Omega_k[X,Y]=\mathcal{F}_{Y,X}(H_*^T(\mathcal{B}_{n,kn}))$
, where the Y-variables represent the dot action, and the X-variables correspond to the star action. In particular, the coefficient of the monomial with all exponents equal to one in
$\Omega_k[X,Y]$
is shown (see Corollary 5.7) to agree with
$\mathcal{H}_{n,kn}(q,t)$
.
Our main theorem is that
$\Omega_k[X,Y]$
is computed by powers of the
$\nabla$
-operator applied to the Cauchy product for the modified Macdonald polynomials, shown in plethystic notation.
Theorem A For
$k\geq 1$
, we have

Notice that, unlike most combinatorial formulas involving the nabla operator, the one in Theorem A completely determines
$\nabla^k$
, and could therefore be taken as a definition. This is the key point to our first proof, which is done by verifying that
$\Omega_k[X,Y]$
satisfies the defining properties of
$\nabla^k$
, similar to the approach taken in [Reference Haglund, Haiman and LoehrHHL05b]. The most difficult part turns out to be showing that the candidate operator
$\nabla'_k$
, which is defined in terms of
$\Omega_k[X,Y]$
, is triangular in the modified Schur basis
$s_{\lambda}[X/(1-q)]$
, as
$\nabla^k$
is. This is equivalent to showing that the coefficients of the monomials
$X^\lambda Y^\mu$
in the plethystically transformed function
$\Omega[X(t-1),Y(q-1)]$
are zero unless
$\lambda\trianglelefteq \mu'$
in the dominance order, where
$\mu'$
is the transposed partition. This is done in Proposition 3.7 below, by defining a combinatorial sign-reversing involution which cancels out all other terms.
We then give a second proof based on a method developed by the second author for counting bundles on
$\mathbb{P}^1$
over a finite field, which we hope will lead to further connections with geometry and number theory.
We also deduce as corollaries some well-known formulas involving the
$\nabla$
-operator, namely the generalization of the shuffle theorem [HHL
$^{+}$
05a, Reference Carlsson and OblomkovCO18] for arbitrary powers
$\nabla^k e_n$
, and the Elias–Hogancamp expression for
$\nabla^k p_1^n$
[Reference Elias and HogancampEH16, Reference Gorsky and HogancampGH22]. Our formulas also motivate further conjectures corresponding to the Frobenius character of more general modules over the smash product algebra
$\mathbb{C}[{\mathbf{x}},{\mathbf{y}}] \rtimes S_n$
, of the sort which appear in Haiman’s polygraph theory. See for instance Conjecture 5.10 below.
One source of motivation for the axiomatic approach was to develop methods that would extend to other Macdonald-type operators, such as the delta operator. Another has to do with an ongoing study of Tor groups of certainly polygraph-type modules, in connection with the nabla positivity conjecture of Bergeron, Garsia, Haiman, and Tesler [Reference Bergeron, Garsia, Haiman and TeslerBGHT99], which predicts that the coefficients of the Schur expansion of
$\nabla^k s_\lambda$
are polynomials in
$c_{\lambda,\mu}(q,t)$
whose coefficients are entirely positive or entirely negative, which has been followed up on in [Reference Carlsson and MellitCM21].
2. Premilinary definitions and notation
In this section we give general background on plethysm, affine permutations, and the combinatorial constructions that appear in our main theorem.
2.1 Macdonald polynomials
Given a symmetric function f, we will adopt the usual plethysm notation of f[X] when X is an element of some
$\lambda$
-ring, so that
$f[x_1+\cdots+x_N]$
is the substitution
$f(x_1,\ldots ,x_N)$
. If
$X=(x_1,x_2,\ldots )$
is some alphabet, we will use the same letter X to denote the sum in plethystic formulas. For details, we refer the reader to [Reference HaimanHai01b].
Let
$\tilde{H}_\lambda=\tilde{H}_\lambda(X;q,t)$
denote the modified Macdonald polynomial [Reference Bergeron, Garsia, Haiman and TeslerBGHT99], defined by

Alternatively, they are uniquely characterized by the following axioms [Reference Haglund, Haiman and LoehrHHL05b]:
-
(1) they are orthogonal with respect to the modified Macdonald inner product
(2)where l is the length of\begin{equation}(p_\lambda,p_\mu)_*=(p_\lambda[X], \omega p_\mu[(1-q)(1-t)X]) = \delta_{\lambda,\mu} \mathfrak{z}_\lambda(-1)^{|\lambda|-l} \prod_{i=1}^{l} (1-q^{\lambda_i})(1-t^{\lambda_i}),\end{equation}
$\lambda$ and
$\mathfrak{z}_\lambda =(p_\lambda, p_\lambda)$ ,
$\omega$ is the involution
$\omega p_\lambda = (-1)^{|\lambda|-l} p_\lambda$ ;
-
(2) they are triangular in the modified Schur basis
(3)and similarly with t in place of q;\begin{equation}\tilde{H}_\lambda[X;q,t]= \sum_{\mu \trianglelefteq \lambda'} a_{\lambda',\mu}(q,t) s_{\mu}[X(1-q)^{-1}]\end{equation}
-
(3)
$(\tilde{H}_\lambda,h_n)=1$ .
Now let
$\nabla$
be the Garsia–Haiman–Bergeron–Tesler operator

where

is the usual statistic from Macdonald’s book [Reference MacdonaldMac98]. In this paper,
$\nabla$
will always denote an operator applied to the X variables.
2.2 Cauchy identities
The standard identities

imply

Using
$(p_\lambda, p_\lambda) = \mathfrak{z}_\lambda$
and
$p_k[XY]=p_k[X] p_k[Y]$
this implies

By a standard argument, for any orthogonal basis
$(f_\lambda)_{\lambda\vdash n}$
of symmetric functions of degree n, for instance for Schur functions, we obtain what we call the Cauchy identity:

Similarly, for the modified scalar product we have

which implies that the same identity holds for any basis of symmetric functions of degree n which is orthogonal with respect to the modified scalar product. In particular, for the Macdonald basis we have

Suppose L is an interesting operator on the space of symmetric functions of degree n, for instance
$L=\nabla^k$
. Then we have, for any symmetric function g of degree n,

where the subscripts/superscripts X, respectively Y, indicate that we apply the operator/scalar product with respect to the alphabet X, respectively Y. So computing the expression
$L_Xe_n\left[{XY}/({(1-q)(1-t)})\right]$
allows one to compute the operator L applied to any function simply by taking the scalar product. Equivalently, we may write using the standard scalar product

2.3 Combinatorial definitions
Fix n and define a label to be an n-tuple of positive integers
${\mathbf{a}} =(a_1,\ldots ,a_n)$
with
$a_i\geq 1$
. We will write
$\mathrm{labs}(n)$
for the set of all labels of length n, and will also call the individual
$a_i$
labels. For any label
${\mathbf{a}}$
, we have a multiset
$A=A({\mathbf{a}})=(|A|,m_A)$
where
$|A|=\{a_1,\ldots ,a_n\}$
is the total set, and
$m_A:|A|\rightarrow \mathbb{Z}_{\geq 1}$
is the multiplicity. We define a (strict) composition of n

In other words,
$\alpha({\mathbf{a}})$
is the result of sorting
${\mathbf{a}}$
in increasing order, and reading off the sizes of the groups, for instance

The multiset in this example would be
$A=\{1^4,2,4^3\}$
. We may also define the corresponding partition
$\mu({\mathbf{a}})=\mu(\alpha({\mathbf{a}}))$
which is the result of sorting
$\alpha({\mathbf{a}})$
in decreasing order, so
$\mu({\mathbf{a}})=(4,3,1)$
in the above example. Given a multiset A, let
$\mathrm{labs}(A)$
denote the set of labels
${\mathbf{a}}$
with
$A({\mathbf{a}})=A$
, with similar definitions for
$\mathrm{labs}(\alpha)$
and
$\mathrm{labs}(\mu)$
.
If
$\mathcal{A},\mathcal{B},\ldots $
are totally ordered sets, we define the ordering on
$\mathcal{A}\times \mathcal{B}\times \cdots $
as the corresponding lexicographic order, breaking ties from left to right. If
${\mathbf{a}} \in \mathcal{A}^n$
,
${\mathbf{b}} \in \mathcal{B}^n,\ldots$
are some elements, we define
$[{\mathbf{a}},{\mathbf{b}},\ldots ]$
to be the sorted representative of the simultaneous action of
$S_n$
on all components. In other words, view
$({\mathbf{a}},{\mathbf{b}},\ldots ,)$
as a matrix, transpose the matrix, sort according to the order on
$\mathcal{A}\times \mathcal{B}\times \cdots $
, and transpose back. For instance, in the case
${\mathbf{a}}\in \mathcal{A}^n$
,
${\mathbf{b}}\in \mathcal{B}^n$
for
$\mathcal{A}=\mathcal{B}=\mathbb{Z}_{\geq 1}$
, we have

We can then define
$\alpha({\mathbf{a}},{\mathbf{b}},\ldots )$
using the same rules as above, so in the above example
$\alpha({\mathbf{a}},{\mathbf{b}})=(1,3,1,1)$
. We make a similar definition for
$\mu$
, which also applies when the sets are unordered.
2.4 The dinv statistic
Let
${\mathbf{a}},{\mathbf{b}}$
be labels, let
${\mathbf{m}} \in \mathbb{Z}_{\geq 0}^n$
, with the decreasing order on the
$m_i$
, so that a triple
$[{\mathbf{m}},{\mathbf{a}},{\mathbf{b}}]$
means one sorted as in the following way.
Definition 2.1. Let
${\mathbf{m}} \in \mathbb{Z}_{\geq 0}^n$
and let
${\mathbf{a}},{\mathbf{b}}$
be labels. We will say that
$({\mathbf{m}},{\mathbf{a}},{\mathbf{b}})$
is sorted if for every
$i<j$
we have:
-
(1)
$m_i \geq m_j$ ;
-
(2) if
$m_i=m_j$ then
$a_i\leq a_j$ ; and
-
(3) if
$m_i=m_j$ and
$a_i=a_j$ then
$b_i\leq b_j$ .
For instance,

We will often write such lists as arrays, as in Example 2.5 below.
We now define a statistic
$\mathrm{dinv}_k({\mathbf{m}},{\mathbf{a}},{\mathbf{b}})$
on triples which are sorted according to Definition 2.1.
Definition 2.2. Let
${\mathbf{a}},{\mathbf{b}}\in \mathrm{labs}(n)$
, and suppose that
$({\mathbf{m}},{\mathbf{a}})$
are sorted, although
$({\mathbf{m}},{\mathbf{a}},{\mathbf{b}})$
may not be. We define

where

and
$\delta(a_1>a_2)$
is one if
$a_1>a_2$
, zero otherwise.
We similarly define
$\mathrm{dinv}_k({\mathbf{m}},{\mathbf{a}})$
as the result of removing
$\delta(b_i>b_j)$
, which is the same as setting
${\mathbf{b}}=(1^n)$
by default.
Recall that a Dyck path is a path of North and East steps in the
$n\times n$
grid beginning at the origin (0,0), placed in the South-West, or lower left corner, and ending at (n,n), which never goes below the diagonal. It is determined uniquely by the set

Definition 2.3. Fix
$k\geq 0$
, suppose
$({\mathbf{m}},{\mathbf{a}})$
is sorted, and let
$i<j$
. We will say that i k-attacks j (or just attacks) if

In other words, i k-attacks j if switching the order of
$b_i,b_j$
has an effect on
$\mathrm{dinv}_k$
. For instance, for
$k=1$
we have that i attacks j if:
-
(1)
$m_i=m_j+1$ and
$a_i>a_j$ ; or
-
(2)
$m_i=m_j$ and
$a_i\leq a_j$ .
Definition 2.4. Let
$\pi=\pi_k({\mathbf{m}},{\mathbf{a}})$
denote the Dyck path such that the elements of
$D(\pi)$
, are the pairs
$i<j$
for which i k-attacks j.
We now have that

where

To see this, we just check that, for any
$i<j$
, we have

Example 2.5. Let
${\mathbf{m}},{\mathbf{a}}$
be given in array notation by

which is a sorted term for
$n=6$
. Then we find that
$\pi_2({\mathbf{m}},{\mathbf{a}})$
is the Dyck path given in Figure 1, as the attacking pairs are the elements of
$D(\pi)$
.

Figure 1: A Dyck path of size
$(6,\,6)$
with area sequence
$a(\pi)=(0,\, 1,\, 2,\, 3,\, 1,\, 1)$
, and
$D(\pi)\,=\{(1,2), (1,3),(1,4),(2,3),(2,4),(3,4),(4,5),(5,6)\}$
.
2.5 Examples
A sum over all
${\mathbf{a}}$
will mean the infinite sum over all labels, unless some upper bound is specified,
$a_i\leq N$
. We will adopt a convenient convention that a sum over
$[{\mathbf{a}},{\mathbf{b}},\ldots ]$
means a sum over orbits with respect to the diagonal
$S_n$
action, with the assumption that
$({\mathbf{a}},{\mathbf{b}},\ldots )$
is the sorted representative in the summand. We will also allow for some summands in which only some of the summands are grouped, which means that just those terms are sorted. For instance, the symbol

indicates the sum over quadruples
$({\mathbf{a}},{\mathbf{b}},{\mathbf{c}},{\mathbf{d}})$
so that for every
$i<j$
we have
$a_i\leq a_j$
,
$b_i\leq b_j$
if
$a_i=a_j$
,
$d_i\leq d_j$
, and there are no constraints on
${\mathbf{c}}$
. We also define automorphism factors for the orbits

where
$\mu=\mu({\mathbf{a}},{\mathbf{b}},\ldots )$
, and

are the q-number and q-factorial.
We give some examples in symmetric functions. Let

be the associated monomial to
${\mathbf{a}}$
, where (A,m) is the associated multiset.
Example 2.6. The complete and monomial symmetric functions are given by

We also have the quasi-symmetric monomials defined by

Example 2.7. We have

which can be deduced from the Cauchy product,

at
$Y=(1-q)^{-1}$
and the well-known specializations for
$e_k(1,q,\ldots )$
. Replacing
$e_n$
with
$h_n$
simply removes the
$q^{n(\mu({\mathbf{a}})')}$
factor.
Example 2.8. A more involved example has the same form as our main theorem, but without the nabla operator:

This can be deduced by successively applying the methods of the previous example. For instance, we would have

We then apply the same expansion to each of the
$e_{\mu_i}[\ldots]$
factors, extracting the expressions of X, followed by
$(1-t)^{-1}$
. Applying

and the above expression for
$e_n[(1-q)^{-1}]$
, we arrive at (10).
3. Main results
We can now state and prove our main theorem, and some consequences.
3.1 Main theorem
Recall the conventions for summations of sorted representatives described in § 2.5. We have our main theorem.
Theorem 3.1. For any
$k\geq 1$
, we have

Notice that we may recover
$\nabla^k$
as an operator from (11) using

where
$\Omega_k[X,Y]$
denotes the expression on the right-hand side of (11), and n is the homogeneous degree of f.
Before proving Theorem 3.1, we state a few consequences. Let

where
$\pi$
is a Dyck path, and
$\mathrm{inv}_{\pi}({\mathbf{b}})$
is defined in § 2.4.
Proposition 3.2. The right-hand side of (11) is given by

Proof. Notice that whenever
$m_i=m_j$
,
$a_i=a_j$
, and
$b_i<b_j$
, switching the order of
$b_i$
and
$b_j$
always increases
$\mathrm{dinv}_k({\mathbf{m}},{\mathbf{a}},{\mathbf{b}})$
by one. Therefore we may remove the sorting condition in
${\mathbf{b}}$
, i.e. replace
$[{\mathbf{m}},{\mathbf{a}},{\mathbf{b}}]$
with
$[{\mathbf{m}},{\mathbf{a}}],{\mathbf{b}}$
in the sum, and remove it from the automorphism factor without changing the answer. Then we apply equation (8).
We have the following interpretation of
$\xi_{\pi}[Y;q]$
. Let

be Stanley’s chromatic symmetric function.
Proposition 3.3. We have that

In particular, it is a symmetric function.
Proof. We have that
$\xi_{\pi}[Y;q]$
is the same as the LLT polynomial
$\chi_{\pi}[Y;q]$
in [Reference Carlsson and MellitCM18], and the statement follows from Proposition 3.5 of that paper.
Proposition 3.4. The expression
$\Omega_k[X,Y]$
is symmetric under exchanging the X and Y variables,
$X\leftrightarrow Y$
.
In particular, since we have already shown that
$\Omega_k[X,Y]$
is a symmetric function in the Y-variables in Proposition 3.3, it is symmetric in the X variables as well.
Proof. It suffices to check that the bijection
$[{\mathbf{m}},{\mathbf{a}},{\mathbf{b}}]\leftrightarrow[{\mathbf{m}},{\mathbf{b}},{\mathbf{a}}]$
, which is well defined on the collection of diagonal orbits, preserves the
$\mathrm{dinv}_k$
statistic. For this, we may assume that
${\mathbf{m}}=(0,\ldots ,0)$
, since the
$\mathrm{dinv}_k$
pairs between different blocks of
${\mathbf{m}}$
is not changed by this operation. The symmetry follows from

for sorted pairs
$[{\mathbf{a}},{\mathbf{b}}]$
, which is unchanged by simultaneous reordering. We may also see the symmetry using Proposition 5.4 below.
Remark 3.5. This can also be seen using a conjecture of Shareshian and Wachs [Reference Shareshian, Wachs, Bjorner, Cohen, De Concini, Procesi, Salvetti and SpacesSW12], later proved in two different ways in [Reference Brosnan and ChowBC15, Reference Guay-PaquetGP16], which would show that both sides of (15) are equal to the Frobenius character of the equivariant cohomology of the regular semisimple Hessenberg variety. This should have a geometric interpretation, and we expect that it corresponds to the paving of the affine Springer fiber by Hessenberg varieties from [Reference Goresky, Kottwitz and MacphersonGKM03].
As a corollary, we have the expression for
$(\nabla^k e_1^n,e_n)$
from [Reference Elias and HogancampEH16]. It was later proved in [Reference Gorsky and HogancampGH22], where it was shown that both sides equal the Poincaré polynomial for the Khovanov–Rozansky knot homology of kth power of the full twist, and also the Hilbert series for the kth power of Haiman’s alternant ideal
$J_n^k$
. Here,
$J_n \subset\mathbb{C}[{\mathbf{x}},{\mathbf{y}}]$
is defined as the ideal generated by the alternating elements under the diagonal action.
Corollary 3.6. We have that

where
$M=(1-q)(1-t)$
,
${\mathbf{m}} \in \mathbb{Z}_{\geq 0}$
ranges over all compositions (not just the sorted ones), and

Proof. Taking
$f=e_1^n$
and applying (12), we obtain that

where
$[X_{\mathbf{a}}]$
is the coefficient for
${\mathbf{a}}=(1,\ldots ,n)$
. We then obtain that

Notice now that the compositions
${\mathbf{m}}\in \mathbb{Z}^n$
are in bijection with sorted pairs
$[{\mathbf{m}}',{\mathbf{a}}]$
when
${\mathbf{a}}$
has distinct elements. The second equality of (16) then follows from checking that
$d_k({\mathbf{m}})=\mathrm{dinv}_{k+1}({\mathbf{m}}',{\mathbf{a}})$
when
$({\mathbf{m}}',{\mathbf{a}})$
is sorted, and
${\mathbf{m}}={\mathbf{m}}'|_{\sigma^{-1}}\in \mathbb{Z}_{\geq 0}^n$
for
$\sigma=\mathrm{Std}({\mathbf{a}})$
.
The first equality is simple identity in
$\nabla$
which holds for all symmetric functions in place of
$e_1^n$
. It follows from the formulas

Then it is also the same as taking the coefficient of
$X_{{\mathbf{a}}}$
and
$Y_{{\mathbf{b}}}$
for
${\mathbf{a}}=(1,\ldots ,n)$
and
${\mathbf{b}}=(1,\ldots ,1)$
on the right-hand side. Since the entries of
${\mathbf{a}}$
are distinct, there is no automorphism factor. Now notice that the compositions
${\mathbf{m}}$
are in bijection with sorted pairs
$[{\mathbf{m}}',{\mathbf{a}}]$
where
${\mathbf{a}}$
has distinct elements, meaning it is a permutation, and that
$d_k({\mathbf{m}})=\mathrm{dinv}_k({\mathbf{m}}',{\mathbf{a}})$
.
3.2 Proof of Theorem A
As above, let
$\Omega_k[X,Y]$
denote the expression on the right-hand side of equation (11). Since we know that
$\Omega_k[X,Y]$
is a symmetric function in each set of variables, we may define an operator
$\nabla'_{k}$
on symmetric functions by

Then, in light of the expression (12), Theorem 3.1 is equivalent to the statement that
$\nabla'_k=\nabla^k$
. We will give our first proof of Theorem 3.1 by verifying that
$\nabla'_k$
satisfies the defining properties of
$\nabla^k$
, similar to the approach in [Reference Haglund, Haiman and LoehrHHL05b].
We will prove the following proposition.
Proposition 3.7. We have the following properties of
$\Omega_k[X,Y]$
.
-
(1) It is symmetric in the two sets of variables,
$\Omega_k[X,Y]=\Omega_k[Y,X]$ .
-
(2) If
$\lambda,\mu$ are partitions, then the coefficient of
$X^{\lambda} Y^{\mu}$ in
$\Omega_k[X(t-1),Y(q-1)]$ is zero unless
$\lambda \trianglelefteq \mu'$ in the dominance order.
-
(3) The leading coefficient in front of
$X^{\lambda}Y^{\lambda'}$ is
$q^{kn(\lambda')}t^{kn(\lambda)}$ .
Then we have the following.
Corollary 3.8. We have that
$\nabla'_{k}=\nabla^k$
, proving Theorem 3.1.
Proof. First, we claim that
$\nabla^k$
is the unique operator satisfying the following.
-
(1) It is self-adjoint with respect to the Macdonald inner product,
\[(\nabla^k f,g)_*=(f,\nabla^k g)_*.\]
-
(2) It is triangular in the modified Schur basis,
\[\nabla s_\lambda[X(q-1)^{-1}]=\sum_{\mu \trianglelefteq \lambda} a_{\lambda,\mu}(q,t) s_{\mu}[X(q-1)^{-1}], \]
-
(3) The leading coefficient is given by
$a_{\lambda,\lambda}(q,t)=q^{kn(\lambda')}t^{kn(\lambda)}$ .
That
$\nabla^k$
satisfies these properties is a consequence of the defining properties of the modified Macdonald polynomials from § 2.1. To see the uniqueness, the triangularity statement implies the triangularity of the matrix of
$\nabla^k$
in the modified Macdonald basis. On the other hand, the self-adjointness implies that the matrix is symmetric, so it is also diagonal.
We therefore need to check that
$\nabla_k'$
satisfies these properties as well, using the corresponding parts from Proposition 3.7. First, item (1) implies that the operator given by
$T(f)=(f,\Omega_k[X,Y])_Y$
is self-adjoint in the usual Hall inner product. It follows that the operator
$f\mapsto T(f[-X(1-q)$
$(1-t)])$
is self-adjoint with respect to the modified Macdonald inner product (2), and therefore so is
$\nabla_k'$
.
To see the triangularity statement, we check that

Then part (2) of the proposition says that the expansion of the resulting symmetric function in the modified monomial basis
$m_{\mu}[X(t-1)^{-1}]$
contains only nonzero terms for
$\mu\trianglelefteq \lambda'$
. Since
$e_{\lambda'},s_\lambda$
, and
$m_\lambda$
are all triangular with respect to each other, their modified versions are as well, and the triangularity statement follows. The leading term statement is immediate from item (3) of the proposition.
To prove Proposition 3.7, we first note that the symmetry statement is just Proposition 3.4. We now turn to the hard part which is showing triangularity. We first evaluate the plethystic substitution
$Y\mapsto Y(q-1)$
. Recall the definition of k-attacks from Definition 2.4, which depends only on the sorted pair
$[{\mathbf{m}},{\mathbf{a}}]$
.
Lemma 3.9. We have

Proof. Apply equation (14) to equation (15), noting that the automorphism factors all disappear because we never have a nonzero term with
$(m_i,a_i,b_i)=(m_j,a_j,b_j)$
for
$i\neq j$
.
We next give a combinatorial formula for the additional substitution
$X\mapsto X(t-1)$
, so as to arrive at
$\Omega_k[X(t-1),Y(q-1)]$
. We first define the combinatorial objects and their corresponding statistics.
Definition 3.10. For any pair
$({\mathbf{m}},{\mathbf{b}})$
, we define

and
$\delta$
is the delta function, 1 for true, 0 for false.
Definition 3.11. Let
$\mathcal{A}(n,k)$
be the set of all quadruples
$(l,{\mathbf{a}},{\mathbf{m}},{\mathbf{b}})$
satisfying:
-
1. the terms
$(a_i,m_i,b_i)$ are sorted for
$l+1 \leq i \leq n$ , and in reverse order for
$1\leq i \leq l$ ;
-
2.
$m_i>0$ for
$1 \leq i\leq l$ ;
-
3. if
$m_i \in \{m_j-k+1,\ldots ,m_j+k\}$ for
$i<j$ , then
$b_i\neq b_j$ .
Example 3.12. For instance, we would have

where we are drawing a dividing line to indicate that
$l=2$
. Below is the table of the contributions (before taking the max with zero) to
$d_2({\mathbf{m}},{\mathbf{b}})$
:

We see that
$d_2({\mathbf{m}},{\mathbf{b}})=16$
, by adding up the positive entries above the diagonal.
Lemma 3.13. We have

Proof. We must write equation (18) as a quasi-symmetric function in the X-variables. We first sort the triples in a different order, so that
$[{\mathbf{a}},{\mathbf{m}},{\mathbf{b}}]'$
is a triple in which the
$a_i$
are in descending order, the
$m_i$
are in increasing order to break ties, and the
$b_i$
are in descending order to breaking ties. This is the reverse of the usual order, modified so that
${\mathbf{a}}$
has priority over
${\mathbf{m}}$
. Define two conditions W (wrong) and NW (not wrong) on pairs
${\mathbf{m}},{\mathbf{b}}$
:

We can reconstruct the condition of when a nonzero term in (18) must have the inequality
$a_i>a_{i+1}$
based on the ordering
${\mathbf{m}},{\mathbf{b}}$
, to produce a quasi-symmetric expansion.
For
$({\mathbf{a}},{\mathbf{m}},{\mathbf{b}})=[{\mathbf{a}},{\mathbf{m}},{\mathbf{b}}]'$
reverse sorted, we have that

for every nonzero term in (18).
Now let

for all pairs
$({\mathbf{m}},{\mathbf{b}})$
, not necessarily sorted. We can now write

which is a quasi-symmetric expansion. We may apply the operator of the substitution
$F[X]\mapsto F[(t-1)X]$
using the standardization approach from [Reference Haglund, Haiman and LoehrHHL05b] to obtain

Before proving the vanishing, it will be helpful to write equation (23) in a more convenient form by collecting powers of t. Define the rotation operator
$\rho$
on pairs
$({\mathbf{m}},{\mathbf{b}})$
by
$\rho({\mathbf{m}}, {\mathbf{b}}) = ({\mathbf{m}}', {\mathbf{b}}')$
, where

which satisfies

where
$\mathrm{area}({\mathbf{m}},{\mathbf{b}})=|{\mathbf{m}}|$
. Moreover, for
$1\leq i<n-1$
,
$\mathrm{W}_i$
for
$({\mathbf{m}}, {\mathbf{b}})$
is equivalent to
$\mathrm{W}_{i+1}$
for
$\rho({\mathbf{m}}, {\mathbf{b}})$
. The triples in (23) are then bijectively mapped via
$\rho^l$
to triples satisfying
$m_1,\ldots,m_l\geq 1$
, so the right-hand side of (23) becomes equation (20).
We now demonstrate the triangularity of equation (20) by finding an involution
$\iota_k:\mathcal{A}(n,k)\rightarrow \mathcal{A}(n,k)$
which sends a quadruple
$(l,{\mathbf{a}},{\mathbf{m}},{\mathbf{b}})$
to itself, or sends it to one which cancels it in (20). We then show that the set of fixed points are empty unless the dominance order property is satisfied.
Definition 3.14. For any i and any quadruple
$(l,{\mathbf{a}},{\mathbf{m}},{\mathbf{b}})$
, we define

where
$l'=l-1$
if
$i\leq l$
or
$l+1$
if
$i>l$
, and
$({\mathbf{a}}',{\mathbf{m}}',{\mathbf{b}}')$
is the result of inserting
$(a_i,m_i,b_i)$
in the unique position on the opposite side of the dividing line l so that
$(l',{\mathbf{a}}',{\mathbf{m}}',{\mathbf{b}}')$
is sorted as in condition (1) of Definition 3.11.
Notice that for any element of
$\mathcal{A}(n,k)$
, we never have
$(a_i,m_i,b_i)=(a_j,m_j,b_j)$
unless
$i=j$
, because of condition (3). We therefore have a unique permutation
$\sigma$
so that
$({\mathbf{a}}_{\sigma},{\mathbf{m}}_{\sigma},{\mathbf{b}}_{\sigma})$
is overall sorted, not in reverse order for
$i\leq l$
.
Definition 3.15. Given
$A=(l,{\mathbf{a}},{\mathbf{m}},{\mathbf{b}})\in \mathcal{A}(n,k)$
, we will say that i is k-movable if
$\mathrm{move}_i(A)\in \mathcal{A}(n,k)$
, and for any j with
$\sigma_j<\sigma_i$
, we have
$d_k^{i,j}({\mathbf{m}},{\mathbf{b}}),d_k^{j,i}({\mathbf{m}},{\mathbf{b}})\leq 0$
. Let
$\iota_k : \mathcal{A}(n,k)\rightarrow \mathcal{A}(n,k)$
be the involution defined by setting

Lemma 3.16. The map
$\iota_k$
is an involution. For any non-fixed element given by
$(l',{\mathbf{a}}',{\mathbf{m}}',{\mathbf{b}}')=\iota_k(l,{\mathbf{a}},{\mathbf{m}},{\mathbf{b}})$
, we have

Proof. Denote the elements by A,A’ so that
$A'=\iota_k(A)$
, and let i be the index for which
$A'=\mathrm{move}_i(A)$
. Let
$\sigma,\sigma'$
be the corresponding permutations as above, and let i’ be the index with
$\sigma_i=\sigma'_{i'}$
, so that
$A=\mathrm{move}_{i'}(A')$
. Then it is clear that i’ is k-movable in A’, so the only thing that needs to be checked is that it is the element with lowest value of
$\sigma'_{i'}$
.
To see this, suppose that j’ is movable in A’ and that
$\sigma_{j}<\sigma_{i'}$
. Let j be the index with
$\sigma_j=\sigma'_{j'}$
, which by assumption is not movable in A. Then we must have that
$\mathrm{move}_j(A) \notin \mathcal{A}(n,k)$
, which can only happen by condition (3) of Definition 3.11 being satisfied. The second index for which the condition is violated must be i, otherwise j’ would not have been movable in A’. But since j,j’ are both on the same side of i, this cannot happen either.
Equation (25) is clear.
Example 3.17. Let us compute the involution on the term
$A\in \mathcal{A}(6,2)$
from Example 3.12. We have that
$\sigma=(5,3,1,2,4,6)$
. The smallest element is therefore
$i=3$
, which is not moveable because
$m_3=0$
, so we cannot move it to the left of the dividing line
$l=2$
without violating condition (2) of the definition of
$\mathcal{A}(n,k)$
. The next smallest values of
$i=4,2,5$
cannot be moved because we have
$d_2^{i,j}({\mathbf{m}},{\mathbf{b}})>0$
or
$d_2^{j,i}({\mathbf{m}},{\mathbf{b}})>0$
for some j earlier in the list. However,
$i=1$
is moveable, and we end up with

Let
$\mathcal{A}'(n,k)$
denote the fixed points of
$\iota_k$
.
Proposition 3.18. If
$\lambda,\mu$
are partitions, then the set

is empty unless
$\lambda\trianglelefteq \mu'$
in the dominance order. If
$\lambda=\mu'$
, then it contains just one element, namely

Example 3.19. If
$\lambda=\mu'=(2,1,1)$
, then
$\mathcal{A}'_{\lambda,\mu}(4,2)$
contains only the element

We will prove Proposition 3.18 through some lemmas. Let
$T(A)=T({\mathbf{a}},{\mathbf{m}},{\mathbf{b}})$
be the result of filling the rth row of the composition
$\alpha=\alpha({\mathbf{a}})$
with the pairs
$(m_i,b_i)$
for which
$a_i=\alpha_r$
, in sorted order left to right. For instance, we have the following.

Notice that the diagram does not depend on the ordering or on l.
Let us refer to the numbers on the left and right in each box of the diagram as the ‘m’ and ‘b’-numbers, respectively.
Lemma 3.20. If
$A\in \mathcal{A}'(n,k)$
, then every m-number in row r of T(A) is at most
$(r-1)k$
.
Proof. Since the rows of the diagram are decreasing, it suffices to check the inequality for the first element of each row. The first element of the first row must be zero, otherwise the lowest element would be movable by simply switching the position of the dividing line, which can only violate condition (2) of Definition 3.11. If some element is greater than
$(r-1)k$
, then there must be a rows whose first entry exceeds all previous entries by more than k, in which case it is movable according to (19).
Lemma 3.21. If
$A\in \mathcal{A}'(n,k)$
, then the same b-number can appear in the first r rows of T(A) at most r times. If it appears the maximum r times, then they all occur in different rows, and all occurrences
$b_i$
are to the right of the dividing line,
$i\geq l+1$
.
Proof. Suppose the number b appears
$r+1$
times in rows 1 through r. Let
$x_1,\ldots ,x_r$
denote the set of the corresponding values of
$m_i$
in the order they appear in
${\mathbf{m}}$
, for instance (1,0,1,0) for the b-value of 1 in (27). Then we must have that

for
$1\leq s<t \leq r+1$
by condition (3) of Definition 3.11. Now let
$0\leq y_1\leq \cdots \leq y_{r+1}$
denote the same set of numbers as the
$x_s$
but in sorted order. By (28) we have that
$y_{s+1}\geq y_s+k$
, so that
$y_{r+1}\geq rk$
, which contradicts Lemma 3.20.
To prove the second statement, define
$0\leq y_1\leq \cdots \leq y_r$
as above. Then, by the same reasoning, we have
$y_r\geq (r-1)k$
and also
$y_r\leq (r-1)k$
by the same lemma, so we must have
$y_s=sk$
. Then only the first case is possible in (28), and so all the
$x_s$
are same order
$x_s=y_s$
. Since
$x_1=0$
, it must be to the right of the dividing line because of condition (2), and so the rest are as well. Then if two b-values appear in the same row, there will be increasing m-values for the same a-value, so in the wrong order for the entries to the right of the dividing line.
We now prove Proposition 3.18.
Proof. The shape of
$T(l,{\mathbf{a}},{\mathbf{m}},{\mathbf{b}})$
is just
$\lambda$
, and the first statement of Lemma 3.21 easily shows that
$\lambda\trianglelefteq \mu'$
.
For the second statement, if
$\lambda=\mu'$
, then the ith lowest b-number appears
$\lambda'_i$
times. By the second statement of Lemma 3.21, it appears once in every row up to
$\lambda'_i$
, to the right of the dividing line, and (by the proof) with corresponding m-values
$0,k,2k,\ldots$
. The unique term with these properties is precisely the one from (26).
Finally, we can prove Proposition 3.7, and therefore Theorem 3.1.
Proof. The first statement is clear from the symmetry of
$\mathrm{dinv}_k$
in
${\mathbf{a}}$
and
${\mathbf{b}}$
. The second statement follows from Proposition 3.18, since all terms in (20) corresponding to
$A\in \mathcal{A}(n,k)-\mathcal{A}'(n,k)$
cancel with
$\iota_k(A)$
by (25). Finally, the leading term from the proof is the contribution from (26), which is easily seen to be
$q^{kn(\lambda)}t^{kn(\lambda')}$
.
3.3 A new proof of the shuffle conjecture
We now show how to recover the shuffle theorem from Theorem 3.1.
Notice that if we have
$b_i=b_j$
for any
$i\neq j$
in (18), then we cannot have
$m_i=m_j$
, or
$m_i=m_j+1$
and
$a_i>a_j$
. By the first condition, we can uniquely sort the orbits so that the
$b_i$
are sorted in reverse order,
$b_1\geq \cdots \geq b_n$
, and if
$b_i=b_{i+1}$
then
$m_{i}<m_{i+1}$
. Then the second condition says

For a pair of sequences
$({\mathbf{m}},{\mathbf{a}})$
and a position i we will define two conditions, ‘parking function at i,’ and ‘not parking function at i’, noting that one is the negation of the other:

We can reformulate dinv as we did in § 3.2:

noting that
$d({\mathbf{m}},{\mathbf{a}})=\mathrm{dinv}([{\mathbf{m}},{\mathbf{a}},{\mathbf{b}}])$
.
We can now write

We would like to evaluate the substitution
$Y\rightarrow (1-t)Y$
. To do this, notice that (29) is a sum of quasi-symmetric functions in Y. We can therefore compute the substitution using the standardization approach from [Reference Haglund, Haiman and LoehrHHL05b]. The result is

Finally, let us make the evaluation
$Y=-1$
in (30) adding an extra sign
$(-1)^n$
, which amounts to counting only the terms in which the quasi-symmetric functions have strict inequalities. We obtain

We would like to cancel certain terms in the right-hand side of (31), this time using the rotation operator
$\rho$
defined in (24). For
$1\leq i<n-1$
, notice that
$\mathrm{PF}_i$
for
$({\mathbf{m}}, {\mathbf{a}})$
is equivalent to
$\mathrm{PF}_{i+1}$
for
$\rho({\mathbf{m}}, {\mathbf{a}})$
. Consider those triples
$(l,{\mathbf{m}},{\mathbf{a}})$
satisfying:
-
(1A)
$l>0$ ;
-
(2A)
$PF_1$ for
$\rho({\mathbf{m}}, {\mathbf{a}})$ if
$l<n$ .
The image of
$\rho$
on these triples is the set of triples
$(l,{\mathbf{m}},{\mathbf{a}})$
satisfying:
-
(1B)
$l<n$ ;
-
(2B)
$NPF_{n-1}$ for
$\rho^{-1}({\mathbf{m}},{\mathbf{a}})$ if
$l>0$ ;
-
(3B)
$m_1>0$ .
We can now check the following proposition, which implies that the two sets have no elements in common, and so the terms coming from the two sets cancel each other out in (31).
Proposition 3.22. The set of triples satisfying (1A), (2A), (1B), (2B), and (3B) is empty.
We can now give a new proof of the shuffle theorem [Reference Haglund, Haiman and LoehrHHL05b, Reference Carlsson and MellitCM18], noting that the conditions of the summation in (32) mean that
${\mathbf{m}}$
is the area sequence of a Dyck path, and
${\mathbf{a}}$
is a word parking function, see [Reference HaglundHag08].
Theorem 3.23. We have

Proof. Using Proposition 3.22 to cancel terms in (31), the terms that remain are the ones that fail to satisfy at least one out of (1A) and (2A), and also fail at least one of (1B), (2B), and (3B). If a term does not satisfy (2A), it means
$l<n$
and
$\mathrm{NPF}_1$
holds for
$\rho({\mathbf{m}},{\mathbf{a}})$
. In particular, we have
$0<m_n+1<m_1$
, so the only property that can fail among (1B), (2B), and (3B) is (2B). Therefore,
$l>0$
and
$\mathrm{PF}_{n-1}$
holds for
$\rho^{-1}({\mathbf{m}},{\mathbf{a}})$
, which is equivalent to
$\mathrm{PF}_1$
for
$\rho({\mathbf{m}},{\mathbf{a}})$
, a contradiction. Then among (1A) and (2A), the property (1A) is the one that fails, so we have
$l=0$
. The only property among (1B), (2B), and (3B) that can fail in the case
$l=0$
is (3B), so we have
$m_1=0$
.
4. Parabolic bundles
In this section we give second proof of Theorem 3.1 by counting parabolic bundles in two different ways.
4.1 Counting formula
On the first side, we will need a result from [Reference MellitMel20, §5] for counting bundles on
$\mathbb{P}^1$
over a finite field. Let q be a prime power, and let
$\mathbf{k}$
be the finite field with
$|\mathbf{k}|=q$
elements. Let
$S=\{s_1,\ldots,s_k\}\subset \mathbb{P}^1(\mathbf{k})$
be a collection of distinct rational points. Let N be a big integer (this will correspond to the number of variables in each alphabet). We need k alphabets
$X_1,\ldots,X_k$
. The variables in alphabet
$X_i$
are denoted
$x_{i,j}$
(
$1\leq i\leq k$
,
$1\leq j\leq N$
).
Definition 4.1. A parabolic bundle is a pair
$(\mathcal{E}, \mathbf{F})$
, where
$\mathcal{E}$
is a vector bundle on
$\mathbb{P}^1$
over
$\mathbf{k}$
, and
$\mathbf{F}=(F_{i,j})_{1\leq i\leq k,0\leq j\leq N}$
is a collection of vector spaces so that for each i we have

An endomorphism of
$(\mathcal{E}, \mathbf{F})$
is an endomorphism of
$\mathcal{E}$
preserving each
$F_{i,j}$
. An endomorphism
$\theta$
is nilpotent if
$\theta^n=0$
for some n.
Here,
$\mathcal{E}(s_i)$
is the fiber of
$\mathcal{E}$
over
$s_i$
. If
$\mathcal{E}$
had rank n, then
$\mathcal{E}(s_i)$
is an n-dimensional vector space.
Parabolic bundles have the following discrete invariants:
-
–
$\mathrm{rank}(\mathcal{E})$ = rank of
$\mathcal{E}$ ;
-
–
$\deg(\mathcal{E})$ = degree of
$\mathcal{E}$ ;
-
–
$r_{i,j}=\dim(F_{i,j}/F_{i,j-1})$ .
Note that
$r_{i,\bullet}$
is a composition of n for each
$i=1,\ldots,k$
(of length N with zeros allowed). These invariants are packaged in the following weight:

It is well known that over
$\mathbb{P}^1$
every vector bundle is a sum of line bundles, so we can write
$\mathcal{E}=O(m_1)\oplus\cdots\oplus O(m_n)$
. We write
$\mathcal{E}\geq 0$
if all
$m_i\geq 0$
and
$\mathcal{E}\leq 0$
if all
$m_i\leq 0$
.
Fix
$n\geq 0$
and
$\lambda\vdash n$
. In [Reference MellitMel20, §5.2] the second author introduced the counting functionFootnote
1

The summation runs over the isomorphism classes of pairs
$(\mathcal{E},\theta)$
of a vector bundle over
$\mathbb{P}^1$
of rank n and an endomorphism,
$\mathrm{Nilp}(\mathcal{E})$
denotes the set of all nilpotent endomorphisms of
$\mathcal{E}$
, and
$\mathrm{Aut}(\mathcal{E},\theta)$
is the set of automorphisms of
$\mathcal{E}$
which commute with
$\theta$
. The notation
$\mathrm{type}\ \theta(s_i)$
respectively
$\mathrm{type}\ \theta$
stands for the partition whose conjugate specifies the sizes of the Jordan blocks of
$\theta$
restricted to the fiber
$\mathcal{E}(s_i)$
respectively over the generic point.
The specialized Macdonald polynomials
$\tilde H_{\mu}[X;q,0]$
, also called the Hall–Littlewood polynomials have an interpretation as counting partial flags preserved by a nilpotent matrix [Reference MellitMel20, Corollary 2.13]. Let M be a nilpotent matrix of type
$\mu\vdash n$
over
$\mathbb{F}_q$
. Then

Thus we can rewrite the counting function as followsFootnote 2 :

This is essentially [Reference MellitMel20, (5.2)]. Now the summation runs over the isomorphism classes of parabolic bundles with an endomorphism,
$\mathrm{Nilp}(\mathcal{E},\mathbf{F})$
denotes the set of all nilpotent endomorphisms of
$\mathcal{E}$
which preserve
$\mathbf{F}$
, and
$\mathrm{Aut}(\mathcal{E},\mathbf{F},\theta)$
is the set of automorphisms of
$(\mathcal{E},\mathbf{F})$
commuting with
$\theta$
.
Between (33) and (34) one can also stop midway. Assume
$s_1=0$
and
$s_2=\infty$
and expand the Hall–Littlewood polynomials
$\tilde H_{\mathrm{type}\,\theta(s_i)}[X_i;q,0]$
only for
$i=1,2$
. We obtain

Here, we are summing over parabolic bundles with flags at 0 and
$\infty$
, denoted
$\mathbf{F}^0$
and
$\mathbf{F}^\infty$
.
The following explicit formula has been proved in [Reference MellitMel20, Corollary 5.9]Footnote 3 :

Corollary 4.2. We have

where the first summation on the right-hand side runs over the isomorphism classes of parabolic bundles with marked points
$0,\infty$
.
Proof. Combine (35) with (36), sum over all n and
$\lambda$
, and replace
$\mathcal{E}$
by the dual bundle
$\mathcal{E}^*$
noting that
$\mathcal{E}\leq 0\Leftrightarrow \mathcal{E}^*\geq 0$
,
$\deg\mathcal{E}^*=-\deg \mathcal{E}$
, and the endomorphisms and flags in fibers of
$\mathcal{E}$
are in a natural bijection with those of
$\mathcal{E}^*$
.
Below we are interested in expressions of the form

Recall that

and by setting
$t=0$

Applying
$(-,s_{1^n})$
in the alphabets
$X_3,\ldots,X_k$
to both sides of Corollary 4.2, and then replacing k by
$k+2$
and relabeling
$s_i$
we obtain the following.
Corollary 4.3. Let
$k\geq 0$
, and let
$\{s_1,\ldots,s_k\}$
be an arbitrary collection of distinct points on
$\mathbb{P}(\mathbf{k})\setminus\{0,\infty\}$
. We have

where
$\mathrm{Nilp}_k(\mathcal{E},\mathbf{F}^0,\mathbf{F}^\infty)$
denotes the set of nilpotent endomorphisms
$\theta$
satisfying
$\theta(s_i)=0$
for
$i=1,\ldots,k$
.
4.2 Parabolic bundles with two marked points
Next we will use an explicit classification of triples
$(\mathcal{E},\mathbf{F}^0,\mathbf{F}^\infty)$
to give an alternative formula for the generating function in Corollary 4.3. The building blocks of the classification will be parabolic bundles of rank 1, i.e. parabolic line bundles.
Example 4.4. Consider Definition 4.1 in the case
$\mathrm{rank}(\mathcal{E})=1$
. Then
$\mathcal{E}(s_i)$
is a vector space of dimension 1, so the sequence of vector spaces
$0=F_{i,0}\subseteq\cdots\subseteq F_{i,N}=\mathcal{E}(s_i)$
is determined by an integer
$j_i$
such that
$F_{i,j_i-1}=0$
,
$F_{i,j_i}\neq 0$
. Since we are on
$\mathbb{P}^1$
, the line bundle
$\mathcal{E}$
is uniquely determined by its degree m. So a parabolic line bundle is uniquely determined by an integer m and a tuple
$(j_1, j_2, \ldots, j_k)$
,
$1\leq j_i\leq N$
. In the case
$k=2$
, we will denote
$a=j_1$
,
$b=j_2$
. The corresponding parabolic line bundle is denoted by
$O(m;a,b)$
.
Proposition 4.5. Let
$(\mathcal{E},\mathbf{F}^0,\mathbf{F}^\infty)$
be a parabolic vector bundle of rank n on
$\mathbb{P}^1$
with two marked points. There exists a unique multiset of triples
$(m_1,a_1,b_1)$
, …,
$(m_n,a_n,b_n)$
such that

This can be thought of as a generalization of the classical Bruhat decomposition for
$GL_n$
. There is a tedious direct proof based on several applications of the standard Bruhat decomposition, but we will use homological algebra instead. The proof will occupy the rest of the section.
Of course, the category of parabolic bundles is not an abelian category, but it can be embedded as a full subcategory into the abelian category of parabolic coherent sheaves (see [Reference HeinlothHei04] or [Reference MellitMel20, §6.1]), which has global dimension 1, so all
$\mathrm{Ext}^i$
vanish for
$i>1$
. The Euler form is given by

We denote by
$\overline{\mathcal{E}}$
the pair
$(\mathcal{E},\mathbf{F})$
. In the case
$k=2$
we write
$(\mathcal{E}, \mathbf{F}^0, \mathbf{F}^\infty)$
.
Lemma 4.6. The dimension of Hom between two parabolic line bundles is given by

Proof. This is clear from the direct description: a homomorphism from O(m) to O(m’) is a polynomial of degree
$m'-m$
, and in order to respect the parabolic structures, for each i such that
$j_i<j_i'$
the polynomial must vanish at
$s_i$
(the leading term has to vanish if
$s_i=\infty$
). Thus we have the dimension of the space of polynomials of degree
$m'-m$
, which have to vanish in
$\#\{i:j_i<j_i'\}$
distinct points.
Combining with the formula for the Euler form we obtain

Introduce a total order on parabolic line bundles in such a way that
$O(m;j_1,\ldots,j_k)<O(m';j_1',\ldots,j_k')$
precisely when

This order clearly satisfies the following proposition.
Proposition 4.7. For two parabolic line bundles L, L’ if
$\mathrm{Hom}(L, L')\neq 0$
, then
$L\leq L'$
.
Proof of Proposition 4.5
. Let
$k=2$
. We prove the existence first. The proof goes by induction on the rank n. The case
$n=1$
is clear. Assume
$n>1$
and suppose
$\overline{\mathcal{E}}$
is a parabolic bundle of rank n. Consider the set of all parabolic line bundles L such that
$\mathrm{Hom}(L, \overline{\mathcal{E}})\neq 0$
. This set is non-empty because any vector bundle has a line subbundle, and we can simply induce the parabolic structure from
$\overline{\mathcal{E}}$
to make it into a parabolic line subbundle. Among these choose one that is maximal in our order. The maximal one exists because the degrees of line subbundles in a given vector bundle are bounded from above.
Now consider the short exact sequence

Let us show that
$\overline{\mathcal{E}}'$
is a parabolic bundle. If it is not, it has some nonzero torsion part
$\overline{\mathcal{E}}'_\mathrm{tor} \subset \overline{\mathcal{E}}'$
. Let K be the kernel of the map
$\overline{\mathcal{E}} \to \overline{\mathcal{E}}'/\overline{\mathcal{E}}'_\mathrm{tor}$
. It is a parabolic line bundle, there is a nonzero map
$L\to K$
, so by Proposition 4.7 we have
$L\leq K$
. Thus
$K=L$
by the maximality of L and therefore
$\overline{\mathcal{E}}'_\mathrm{tor}$
has to be zero.
By the induction assumption,
$\overline{\mathcal{E}}'\cong \bigoplus_{l=1}^{n-1} L_l$
. If the short exact sequence above splits, we are done. Suppose it does not split. This implies that
$\mathrm{Ext}(\overline{\mathcal{E}}', L)\neq 0$
, so there exists l such that for
$L'=L_l$
we have

where we let
$L=O(m;j_1,\ldots,j_k)$
and
$L'=O(m';j_1',\ldots,j_k')$
. Note that, since
$k=2$
, this implies
$m'\geq m$
. Our plan is to construct a parabolic line bundle L” such that:
-
(1)
$\mathrm{Hom}(L'', L')\neq 0$ ;
-
(2)
$\mathrm{Ext}(L'', L)=0$ ;
-
(3)
$L'' > L$ .
By the exact sequence

these conditions would guarantee that any nonzero homomorphism
$h\in\mathrm{Hom}(L'', L')\subset \mathrm{Hom}(L'', \mathcal{E}')$
can be lifted to a nonzero homomorphism
$L''\to \overline{\mathcal{E}}$
, and we would obtain a contradiction with the maximality of L.
If
$m'\geq m+1$
, we pick
$L''=O(m+1;N,N)$
. This guarantees that
$\dim \mathrm{Hom}(L'', L') = m'-m>0$
,
$\dim \mathrm{Ext}(L'', L) = 0$
and
$L''>L$
, so the required conditions are satisfied.
Otherwise, we must have
$m'=m$
,
$j_1'<j_1$
and
$j_2'<j_2$
. Picking
$L'' = O(m;j_1',j_2)$
(or
$O(m;j_1,j_2')$
) satisfies
$\dim \mathrm{Hom}(L'', L') = 1$
,
$\dim \mathrm{Ext}(L'', L) = 0$
and
$L''> L$
. So the existence have been proven.
Note that we have in particular demonstrated that the maximal line subbundle is a direct summand. By Proposition 4.7 it must be present in any direct sum decomposition, and by successively splitting away the maximal subbundle we deduce the uniqueness.
Remark 4.8. For
$k>2$
the statement does not hold. For a counter-example for
$k=3$
, pick trivial bundle of rank 2 and three lines in general position over the marked points.
4.3 Computations
We are ready to identify all the ingredients on the left-hand side of Corollary 4.3. By Proposition 4.5, the summation runs over the set of sorted triples
$[{\mathbf{m}}, {\mathbf{a}}, {\mathbf{b}}]$
(see § 2.3 for combinatorial notation). Each sorted triple corresponds to a direct sum of line bundles
$L_i=O(m_i;a_i,b_i)$
, which satisfy
$L_1\geq \cdots \geq L_n$
. Denote

Proposition 4.9. Suppose
$({\mathbf{m}},{\mathbf{a}},{\mathbf{b}})$
is sorted. The number of automorphisms of
$O({\mathbf{m}};{\mathbf{a}},{\mathbf{b}})$
is given by

Proof. By Proposition 4.7, automorphisms are given by block-upper-triangular matrices with block sizes equal to the multiplicities of triples
$(m_i,a_i,b_i)$
. A block-upper-triangular matrix is invertible precisely if the blocks are. So we obtain that the number of automorphisms is given by

where
$\mu_1, \mu_2, \ldots$
denote the multiplicities,
$\sum_i \mu_i = n$
. The number of elements of
$GL_r(\mathbf{k})$
is given by

Since
$\dim \mathrm{Hom}(L_i, L_i)=1$
, we have

and applying Lemma 4.6 we obtain the formula.
Below we include the case
$k=0$
for completeness. We have the following.
Proposition 4.10. Suppose
$({\mathbf{m}},{\mathbf{a}},{\mathbf{b}})$
is sorted and
$k\geq 0$
. We have

where
$\mu=(\mu_1,\ldots,\mu_l)$
are the multiplicities of the triples
$(m_i,a_i,b_i)$
.
Proof. As in the proof of Proposition 4.9, the endomorphisms are given by block-upper-triangular matrices, where the block structure is governed by repetitions among the triples
$(m_i, a_i, b_i)$
. A block-upper-triangular matrix is nilpotent if each diagonal block is nilpotent. The diagonal blocks consist of constant functions. So in the case
$k>0$
the diagonal blocks are automatically zero. The space of off-diagonal entries in position (i,j) is given by
$\mathrm{Hom}(L_j, L_i)$
(
$i<j$
). This is the space of polynomials of bounded degree. When polynomials are forced to have zeroes at k further points, the dimension drops down by k, similarly to the proof of Lemma 4.6. This completely describes the case
$k>0$
. For the case
$k=0$
we need to count the number of nilpotent matrices in each diagonal block. This is given by
$q^{r^2-r}$
for a block of size
$r\times r$
(see e.g. [Reference Rodríguez-VillegasRV07]). For each block, the factor
$q^{\sum_{i<j} \cdots}$
already contains
$q^{\binom{r}2}$
, so extra factor
$q^{\binom{r}2}$
has to be added.
The remaining pieces of the left-hand side of Corollary 4.3 are identified as follows:

Example 4.11. Let
$k=0$
. The left-hand side of Corollary 4.3 becomes

So in each summand each triple (m,a,b) with multiplicity
$\mu$
contributes a factor of

Summing over all n we obtain every possible triple with every multiplicity, so the result can be written as an infinite product

which matches the right-hand side of Corollary 4.3.
Our main conclusion is as follows.
Theorem 4.12. For
$k\geq 1$
we have

Proof. Applying Corollary 4.3, we write the left-hand side as a summation over isomorphism classes of parabolic bundles with two marked points. These are identified with multisets of triples
$({\mathbf{m}}, {\mathbf{a}}, {\mathbf{b}})$
by Proposition 4.5. We claim that for each such multiset the corresponding terms match, i.e. we have

The monomials
$X_{\mathbf{a}}$
and
$Y_{{\mathbf{b}}}$
and the t-degree match naturally. Using Propositions 4.9 and 4.10 to express the right-hand side and throwing away the resulting common factor
${1}/({(q-1)^n \mathrm{aut}_q({\mathbf{m}},{\mathbf{a}},{\mathbf{b}})})$
, we are reduced to show that the q-degree
$\mathrm{dinv}_k({\mathbf{m}}, {\mathbf{a}}, {\mathbf{b}})$
equals

For each pair
$i<j$
let
$c_{i,j} = m_i-m_j - \delta_{a_j<a_i} - \delta_{b_j<b_i}$
and note that
$c_{i,j}\geq -1$
(case-by-case analysis using the assumption that the sequence of triples is ordered). Then the above sum can be written as

Each summand matches the corresponding summand in Definition 2.2,

and therefore the q-degrees also match.
5. Geometric interpretations
We explain the underlying motivation behind Theorem A, which was discovered experimentally using conjectural relations between Haiman’s polygraph rings and the equivariant Borel–Moore homology of the unramified affine Springer fiber
$H^T_*(\mathcal{B}_{n,kn})$
. We connect the combinatorics of § 2.3 to cells in
$\mathcal{B}_{n,kn}$
.
5.1 Affine permutations
We describe the connection between affine permutations and the combinatorics of the dinv statistic and rational slope parking functions, following [Reference Gorsky, Mazin and VaziraniGMV14].
The set of affine permutations is defined as

Each one is determined by its values in window notation,
$w=(w_1,\ldots ,w_n)$
, where as usual we denote
$w_i=w(i)$
. Note that we are using the unconstrained ‘
$GL_n$
’ version of affine permutations as opposed to the ‘
$SL_n$
’ types, which requires that
$w_1+\cdots +w_n=n(n+1)/2$
. We define the set of positive affine permutations as

Let

We may multiply
$w\in W_{n,d}$
and
$w'\in W_{n,d'}$
to obtain a permutation in
$W_{n,d+d'}$
. For each d we have the Bruhat order
$\leq_{bru}$
on
$W^+_{n,d}$
, as defined in [Reference Bjorner and BrendiBB05, Reference Lam, Lapointe, Morse, Schilling, Shimozono and ZabrockiLLM+14].
Following [Reference Gorsky, Mazin and VaziraniGMV14], we have the following.
Definition 5.1. An affine permutation is said to be m-stable if
$w_{i+m}>w_i$
for all i, and is said to be m-restricted if
$w^{-1}$
is m-stable.
For integers
$a,b \in \mathbb{Z}$
which are not congruent modulo n, there is a unique affine transposition
$t_{a,b}$
which switches the two. Note that
$t_{a,b}=t_{b,a}$
, and
$t_{a,b}=t_{a+kn,b+kn}$
, so the map taking pairs of incongruent integers to W is many-to-one. Given an m-restricted permutation, let

The statistic
$|a-b|$
does not depend on the representatives a,b or their order, and is called the height of the transposition. The set
$E_m(w)$
represent directed edges
$w\rightarrow v$
with
$v=t_{a,b}w$
in the Goresky–Kottwitz–MacPherson (GKM) graph of
$\mathcal{B}_{n,m}$
, corresponding to the one-dimensional orbits under
$\widetilde{T}$
.
Recall that an (n,m)-rational slope Dyck path is one that begins at (0,0) and ends at (m,n), never crossing the line of slope
$n/m$
. Again, we have the area and coarea sequences
$\mathrm{area}(\pi),\mathrm{coarea}(\pi)$
, and also
$D(\pi)$
. For any m-restricted permutation w, there is a rational (n,m)-Dyck path with coarea sequence

where

This is the underlying Dyck path of sequence
${\mathbf w}_m(w)=\mathcal{PS}_{w^{-1}}$
of [Reference Gorsky, Mazin and VaziraniGMV14], which was shown to define a bijection from the set of m-stable affine permutations in
$W_n$
to rational parking functions for (n,m) coprime in [Reference Thomas and WilliamsTW15].
We define the following.
Definition 5.2. Let

where
$\pi_{n,m}=(1^n0^m)$
is the (n,m)-Dyck path of maximal area.
We now explain the connection with the statistic
$\mathrm{dinv}_k({\mathbf{m}},{\mathbf{a}},{\mathbf{b}})$
of § 2.3. Recall the definition of Standardization from [Reference Haglund, Haiman and LoehrHHL05b].
Definition 5.3 The standardization of a label is the unique permutation
$\sigma=\mathrm{Std}({\mathbf{a}})$
such that
${\mathbf{a}}_{\sigma^{-1}}$
is weakly increasing, and the sub-list
$(\sigma_{i_1},\ldots ,\sigma_{i_k})$
is increasing on those indices
$\{i_1,\ldots ,i_k\}$
for which
$a_{i_j}=x$
.
We will also define
$\mathrm{Std}_<({\mathbf{a}})$
and
$\mathrm{Std}_>({\mathbf{a}})$
with respect to the usual, and reverse order on
$\mathbb{Z}_{\geq 1}$
, so that
$\mathrm{Std}=\mathrm{Std}_<$
. For instance, if
${\mathbf{a}}=(3,3,3,1,2,3,1)$
, then

In particular,
$\mathrm{dinv}_k$
respects standardization, i.e.
$\mathrm{dinv}_k({\mathbf{m}},\mathrm{Std}({\mathbf{a}}),\mathrm{Std}({\mathbf{b}}))=\mathrm{dinv}_k({\mathbf{m}},{\mathbf{a}},{\mathbf{b}})$
.
Now given a tuple
$({\mathbf{m}},{\mathbf{a}},{\mathbf{b}})$
which is sorted, we define an affine permutation

where
$t({\mathbf{m}})=(n+m_1 n,\ldots ,1+m_nn)$
is the maximal representative of its coset in
$S_n\backslash W^+_n /S_n$
, and
$\mathrm{rev}({\mathbf{b}})$
is the result of writing
${\mathbf{b}}$
in the reverse order. We similarly define
$\mathrm{aff}({\mathbf{m}},{\mathbf{a}})$
as the left coset
$S_n \mathrm{aff}({\mathbf{m}},{\mathbf{a}},{\mathbf{b}})$
, which is independent of
${\mathbf{b}}$
. The proof of the following proposition is tedious, and will be omitted.
Proposition 5.4. Fix multisets A,B of size n with
$|A|,|B|\subset \mathbb{Z}_{\geq 1}$
. Then
$\mathrm{aff}({\mathbf{m}},{\mathbf{a}},{\mathbf{b}})$
defines a bijection from the set of sorted triples

to the double coset
$S_{\mathrm{rev}(\alpha(B))}W^+_n S_{\alpha(A)}$
, where
$S_\alpha$
is the Young subgroup, and
$\mathrm{aff}({\mathbf{m}},{\mathbf{a}},{\mathbf{b}})$
is the unique representative of its double coset of maximal length. Moreover, we have

and the Dyck path is determined by

where
$w_{\min},w_{\max}$
are the unique representatives of the coset
$\mathrm{aff}({\mathbf{m}},{\mathbf{a}})$
which are minimal and maximal in the Bruhat order.
Example 5.5. Take
$({\mathbf{m}},{\mathbf{a}},{\mathbf{b}})=((2,1,0,0),(2,3,1,1),(1,2,1,1))$
, which is sorted. Then we have

which gives
$w=\mathrm{aff}({\mathbf{m}},{\mathbf{a}},{\mathbf{b}})=(3,2,12,5)$
. This is the maximal length element in the double coset


Now, for
$k=1$
, we have

so that
$\mathrm{dinv}_{4}(w)=6-4=2$
. On the other hand,
$({\mathbf{m}},{\mathbf{a}})$
has three attacking pairs,
$\left\{(2,3),(2,4),(3,4)\right\}$
. Since (2,3) and (2,4) are the pairs for which
$b_i>b_j$
, we see that
$\mathrm{dinv}_1({\mathbf{m}},{\mathbf{a}},{\mathbf{b}})=2$
, in agreement with Proposition 5.4.
Example 5.6 Let us compute the Dyck path for the terms
${\mathbf{m}},{\mathbf{a}}$
from Example 2.5, using Proposition 5.4. Then we have that

are the minimal and maximal representatives of the left coset of
$\mathrm{aff}({\mathbf{m}},{\mathbf{a}})\in S_n\backslash W^+_n$
. Then

and the unique (n,n)-Dyck path whose area sequence is the difference (0, 1, 2, 3, 1, 1) is the expected one from Figure 1.
In this language, we have a corollary of Theorem A.
Corollary 5.7. The coefficient of the monomial symmetric functions associated to
$\lambda=\mu=(1^n)$
is given by

5.2 Polygraphs and the Hilbert scheme
If M is a representation of
$S_n\times \cdots \times S_n$
with k factors, we will denote the Frobenius character by

which is a function in k sets of variables,
$X_i=(x_{i,1},x_{i,2},\ldots )$
, individually symmetric in each one. For doubly graded modules, the Frobenius character encodes the degrees with the q,t variables, namely

where
$M^{(i,j)}$
is the homogeneous component of the bigrading.
In Haiman’s theory [Reference HaimanHai01b], the expression in Theorem A is the equivariant index of a sheaf on the Hilbert scheme of points in the complex plane
$\mathrm{Hilb}_n \mathbb{C}^2$
, with respect to the usual torus action
. Let P be the Procesi bundle of rank
$n!$
whose fibers carry an action of
$S_n$
isomorphic to the regular representation. The modified Macdonald polynomial is the Frobenius character
$\tilde{H}_\lambda=\mathcal{F} P\big|_{\lambda}$
of the fiber of P at a monomial ideal, which are the torus-fixed points of
$\mathrm{Hilb}_n \mathbb{C}^2$
. Explicitly, they are given by the Garsia–Haiman module, as in Haiman’s proof of the
$n!$
conjecture [Reference HaimanHai01a]. Then by a noncompact version of the localization theorem due to Nakajima of [Reference Nakajima and YoshiokaNY03, Proposition 4.1], we have

We next define a variant of Haiman’s polygraph modules [Reference HaimanHai01a]. Fix n and let
${\mathbf{x}}$
denote the set of variables
$(x_1,\ldots ,x_n)$
, for some variable x. Let
$\mathbb{C}[{\mathbf{x}},{\mathbf{y}}]\cdot S_n$
denote the free left
$\mathbb{C}[{\mathbf{x}},{\mathbf{y}}]$
-module with one free generator for each permutation
$\tau\in S_n$
. Consider the following variant of Haiman’s map from [Reference HaimanHai01a] equation (152):

We define a module M as the image of
$\phi$
, as a
$\mathbb{C}[{\mathbf{x}},{\mathbf{y}}]$
-module. We have the usual biograding on M compatible with the grading on the ring
$\mathbb{C}[{\mathbf{x}},{\mathbf{y}},{\mathbf{z}},{\mathbf{w}}]$
, in which the degree of the
${\mathbf{x}},{\mathbf{z}}$
variables are (1,0), and the
${\mathbf{y}},{\mathbf{w}}$
variables have degree (0,1). Note that
${\mathbf{x}},{\mathbf{y}}$
have nothing to do with the symmetric function variables X,Y.
There is an action of
$S_n\times S_n$
on M, which may also be interpreted as a commuting left and right action by

Then
$\phi$
intertwines this action with the one where the first factor simultaneously permutes
${\mathbf{x}},{\mathbf{y}}$
, and the second factor permutes
${\mathbf{z}},{\mathbf{w}}$
. Notice that the left
$S_n$
-action on M is compatible with the action on the ground ring
$\mathbb{C}[{\mathbf{x}},{\mathbf{y}}]$
by simultaneously permuting the indices of the variables, whereas the right
$S_n$
-action does not act on the variables. Another way to say this is that M is a bigraded module over the smash product
$\mathbb{C}[{\mathbf{x}},{\mathbf{y}}]\rtimes S_n$
, which is the noncommutative ring by adjoining a generator for each
$\sigma \in S_n$
with the relation

and that the right action of
$S_n$
acts by automorphisms of M.
The following conjecture was proved in [Reference Alvarez and LosevAL24].
Conjecture 5.8. As a module over
$\mathbb{C}[{\mathbf{x}},{\mathbf{y}}] \rtimes S_n$
, M is the image of the Procesi bundle P under the Haiman–Bridgeland–King–Reid isomorphism

The higher derived functors
$R^i\Gamma(P\otimes P)$
vanish, and so
$M\cong \Gamma_{\mathrm{Hilb}_n}(P\otimes P)$
. Moreover, we have that M is free when regarded as a module over
$\mathbb{C}[{\mathbf{x}}]$
, in other words forgetting the
$\mathbb{C}[{\mathbf{y}}]$
-action.
Remark 5.9. Haiman identified the module in the case
$F= B^{\otimes l}$
as the polygraph module R(n,l) defined in [Reference HaimanHai01b], where B is the tautological bundle, and he proved the second two statements for R(n,l). To the best of our knowledge Conjecture 5.8 is not known. We point out that the vanishing statement is definitely false for three powers of the Procesi bundle
$P^{\otimes 3}$
, which may be seen by observing the Atiyah–Bott localization actually has negative terms.
To connect this with Theorem A, observe that combining (42) at
$k=1$
with Conjecture 5.8 gives the following.
Conjecture 5.10. We have

Remark 5.11. The conjecture is motivated by the following geometric picture. Recall the following commutative diagram [Reference HaimanHai01a].

The diagram is a reduced cartesian product and the space
$X_n$
is the isospectral Hilbert scheme. The map
$\pi$
is finite and
$P=\pi_* \mathcal{O}_{X_n}$
. Thus the ring
$\Gamma_{\mathrm{Hilb}_n}(P\otimes P)$
is the ring of functions on
$X_n\times_{\mathrm{Hilb}_n\mathbb{C}^2} X_n$
, which is a closed subscheme of
$X_n\times X_n$
. On the other hand,
$X_n\times_{\mathrm{Hilb}_n\mathbb{C}^2} X_n$
is reduced, so it coincides with the reduced fiber product

The space
$\mathbb{C}^{2n}\times_{\mathbb{C}^{2n}/S_n}\mathbb{C}^{2n}$
is the union of graphs of permutations viewed as maps
$\mathbb{C}^{2n}\to \mathbb{C}^{2n}$
. This induces a covering of
$X_n\times_{\mathrm{Hilb}_n\mathbb{C}^2} X_n$
by
$n!$
copies of
$X_n$
. Passing to the rings of functions we obtain ring homomorphisms,

whose composition becomes the map
$\phi$
of (43) under the identification
$\Gamma(\mathcal{O}_{X_n})=\Gamma_{\mathrm{Hilb}_n \mathbb{C}^{2}} (P)=\mathbb{C}[{\mathbf{x}},{\mathbf{y}}]$
, see [Reference HaimanHai01b]. The second map above is injective because the functor
$\Gamma$
is left exact. If we knew that the first map is surjective, we would have

The conjecture is then reduced to the vanishing of the higher cohomologies of the ideal sheaf of
$X_n\times_{\mathrm{Hilb}_n\mathbb{C}^2} X_n$
in
$X_n \times X_n$
.
5.3 Connection with affine Springer fibers
The m-restricted permutations can be interpreted as the torus-fixed points of a certain affine Springer fiber
$\mathcal{B}_{n,m}=\mathcal{B}_{\gamma}$
of the type studied in [Reference Goresky, Kottwitz and MacphersonGKM03] as follows. Define
$v_i \in \mathbb{C}^n((t))$
by

where
$e_i$
is the standard basis vector in
$\mathbb{C}^n$
and
$i_1$
is the unique element in
$\{1,\ldots ,n\}$
congruent to i modulo n. We can then describe affine permutations as invertible matrices with values in
$\mathbb{C}[t,t^{-1}]$
by the action
$w\cdot v_i=v_{w_i}$
. For instance, we would have

Now let
$\gamma=\gamma_{n,m}$
be the topologically nilpotent operator

The
$a_i$
are distinct nonzero complex numbers for
$1\leq i \leq d$
, and
$a_{i+d}=a_i$
for
$d=\gcd(n,m)$
. For
$m=kn$
, we have
$\gamma=\mathrm{diag}(a_1t^k,\ldots ,a_nt^k)$
, corresponding to the unramified case studied in [Reference Goresky, Kottwitz and MacPhersonGKM04]. The affine Springer fiber associated to
$\gamma$
is defined by

where
$I_n$
is the Iwahori subgroup. Then an affine permutation w is m-restricted precisely when
$wI_n \in \mathcal{B}_{n,m}=\mathcal{B}_{\gamma}$
, realizing w as a matrix as above. They are in fact the fixed points of
$\mathcal{B}_{n,m}$
for the action of a certain restricted torus. In this case the n-dimensional torus
$T \subset GL_n(C)$
acts by multiplication on the left via, as well as the extended
$(n+1)$
-dimensional torus T, which includes loop rotation, both having discrete fixed points described by
$W^+_{n,d}$
.
The space
$\mathcal{B}_{n,m}$
has an affine paving by the results of [Reference Lusztig and SmeltLS91] in the coprime case, and [Reference Goresky, Kottwitz and MacphersonGKM03] for the general case, including
$\mathcal{B}_{n,m}$
for general (n,m) in type A. In the unramified case of
$m=kn$
studied in [Reference Goresky, Kottwitz and MacPhersonGKM04], every component in the paving is GKM with respect to the torus
$\tilde{T}\cong (\mathbb{C}^*)^{n+1}$
consisting of the maximal torus
$T\subset GL_n(\mathbb{C})\subset GL_n(\mathbb{C}((t)))$
together with the ‘loop rotation.’ The torus fixed points are the entire set
$W_n$
, and the outgoing edges in the GKM graph correspond precisely to the elements of
$E_{nk}(w)$
defined above. The equivariant Borel–Moore homology
$H_{\tilde{T}}^*(\mathcal{B}_{n,kn})$
is a submodule of the equivariant homology of the fixed point set, which is a free
$\mathbb{C}[{\mathbf{x}},\epsilon]$
-module with basis
$W_n$
, where
$\mathbb{C}[{\mathbf{x}},\epsilon]$
is identified with the equivariant cohomology of a point. Moreover, the fundamental classes of the closures of the elements of the affine paving are a free basis of
$H_{\tilde{T}}^*(\mathcal{B}_{n,nk})$
[Reference GrahamGra01, Reference Edidin and GrahamEG96, Reference BrionBri98].
By [Reference GrahamGra01, Proposition 2.1], we have a dual basis of equivariant cohomology, which must be triangular in the Bruhat order in the opposite direction. By the GKM property, the leading terms must be

where
$E^{i,j}_{m}(w)$
is the set of transpositions defined above. By restricting the torus and setting
$\epsilon=0$
, we lose the GKM property but have localization, as well as the given basis. The standard description of the corresponding homology is then given as a subspace of the
$\mathbb{C}({\mathbf{x}})$
-vector space with the same fixed point basis, which is different from the description of [Reference Goresky, Kottwitz and MacPhersonGKM04]. For instance, in the case of
$k=\infty$
, we can compare the coefficient
$a_{w,\infty}({\mathbf{x}})$
with the leading terms in Kostant and Kumar’s nil Hecke ring [Reference Kostant and KumarKK86, Reference Lam, Lapointe, Morse, Schilling, Shimozono and ZabrockiLLM+14], which encodes the equivariant homology of the affine flag variety

Theorem A was discovered by attempting to embed the above module M as a submodule of the
$GL_n$
version of
$H_*^T(\mathcal{B}_{n,kn})$
, in which the fixed points only consist of positive permutations

A construction of this type was used in [Reference Carlsson and OblomkovCO18] for instance, in which the authors exhibited an isomorphism
$DR_n\cong H_*(\mathcal{B}_{n,n+1})$
related to the ones studied in [Reference Oblomkov and YunOY14], and used it to study the diagonal coinvariant algebra
$DR_n$
as a module over
$\mathbb{C}[{\mathbf{x}}]$
. In another example, Kivinen showed that Haiman’s alternant ideal
$J_n\subset \mathbb{C}[{\mathbf{x}},{\mathbf{y}}]$
in general Lie type satisfies a suitable version of the GKM relations, and therefore injects into the equivariant Borel–Moore homology of the Grassmannian version of
$\mathcal{B}_{n,kn}$
. In type A, when combined with Haiman’s results, it follows that the map
$J_n^k \rightarrow H^T_*(\mathcal{B}_{n,kn}^{Grass})$
is an isomorphism when the y-variables are inverted, see [Reference KivinenKiv20, Theorem 1.1].
Now let

whose degree is
$\mathrm{dinv}_{kn}(w)$
. Here
$E^{i,j}_{kn}(w)$
is the set of transpositions
$t_{a,b}\in E_{kn}(w)$
for which
$\{\bar{a},\bar{b}\}=\{\bar{i},\bar{j}\}$
, where the bar is the congruence class modulo n. The following conjecture illustrates the connection with Theorem A in the case
$k=1$
.
Conjecture 5.12. There exist free generators
$A_w=\sum_{v} c_{v,w}({\mathbf{x}}) e_v \in M$
as
$v,w\in W^+_n$
satisfying the following properties.
-
(1) The
$A_w$ freely generate M as a
$\mathbb{C}[{\mathbf{x}}]$ -module.
-
(2) The coefficients satisfy
$c_{v,w}({\mathbf{x}})=0$ unless
$v\leq_{bru} w$ , and the leading term is given by
$c_{w,w}({\mathbf{x}})=b_{w,1}({\mathbf{x}})$ .
-
(3) For any compositions
$\alpha,\beta$ , if w is the element of maximal length in
$S_\alpha \backslash W^+_n / S_\beta$ , then
$A_w \in M^{\alpha,\beta}$ , the invariant subspace with respect to the product of the corresponding Young subgroups.
In particular, there is the expected freeness of M over
$\mathbb{C}[{\mathbf{x}}]$
. On the other hand, in light of Conjecture 5.12, we expect that
$\mathcal{F}_{Y,X} M=\Omega_1[X,Y]$
with respect to the dot and star actions mentioned in the introduction. To see this in the case of the Hilbert series, we take the contribution to
$\Omega_1[X,Y]$
for which
${\mathbf{a}},{\mathbf{b}}$
have all distinct entries. Then the automorphism factor is trivial, and we obtain the sum in (1), using Proposition 5.4 to relate the corresponding dinv statistics.
Financial support
E. Carlsson was supported by NSF DMS-1802371 during part of this project. A. Mellit was supported by the projects Y963-N35 and P31705 of the Austrian Science Fund, as well as the ERC grant ‘Refined invariants in combinatorics, low-dimensional topology and geometry of moduli spaces’ No. 101001159.
Conflicts of interest
None.
Journal information
Compositio Mathematica is owned by the Foundation Compositio Mathematica and published by the London Mathematical Society in partnership with Cambridge University Press. All surplus income from the publication of Compositio Mathematica is returned to mathematics and higher education through the charitable activities of the Foundation, the London Mathematical Society and Cambridge University Press.