1 Introduction
Equivariant cohomology of an affine Grassmannian has been a topic of intensive investigations for decades. For the small torus action, it can be identified with a certain commutative subalgebra of the associated nil–Hecke algebra of a Kac–Moody root system called the Peterson subalgebra [Reference Peterson16]. One of its remarkable properties says that after taking localization it becomes isomorphic to the (small) quantum cohomology of the respective finite part (flag variety) [Reference Lam and Shimozono12, Reference Peterson16]. A parallel isomorphism for the K-theory was conjectured and discussed in [Reference Lam, Li, Mihalcea and Shimozono10, Reference Lam, Schilling and Shimozono11] and is known as the Peterson Conjecture. This conjecture was recently proven by Kato in [Reference Kato7] using a language of semi-infinite flag varieties.
In the present notes, we study a generalization of the Peterson subalgebra to an oriented (generalized) cohomology theory
$h(\text {-})$
, e.g., algebraic cobordism
$\Omega (\text {-})$
of Levine–Morel. Such a cohomology theory was first introduced and studied in [Reference Levine and Morel15], and extended to the torus-equivariant setup in [Reference Heller and Malagón-López6, Reference Krishna8] for arbitrary smooth varieties. As for flag varieties associated with root systems, it can be described using the Kostant–Kumar localization approach (for finite root systems see [Reference Calmès, Zainoulline and Zhong3–Reference Calmès, Zainoulline and Zhong5], and for Kac–Moody see [Reference Calmès, Zainoulline and Zhong2]). The respective generalization of the nil Hecke algebra is called the formal affine Demazure algebra (FADA). The generalization of the Peterson algebra introduced recently in [Reference Zhong18] which we call a formal Peterson subalgebra is then the centralizer of the equivariant coefficient ring in the small torus FADA.
To state our first result, let
$R=h(pt)$
denote the coefficient ring of the oriented theory
$h(\text {-})$
, let
$S=h_T(pt)$
denote the respective small torus T equivariant coefficient ring, let
$\mathbf {D}_{W_{\mathrm {a}}}$
denote the small torus FADA, and let
$\mathbf {D}_{Q^{\vee }}$
denote the formal Peterson subalgebra as constructed in [Reference Zhong18]. We then obtain the following important property of the centre of FADA.
Theorem 1.1 (cf. Theorem 4.4)
If
$\mathbb {Q} \subseteq R$
, then the centre
$Z(\mathbf {D}_{W_{\mathrm {a}}})$
of the small torus FADA generates the formal Peterson subalgebra
$\mathbf {D}_{Q^{\vee }}$
as an S-module. Moreover, the centre
$Z(\mathbf {D}_{W_{\mathrm {a}}})$
generates
$\mathbf {D}_{W_{\mathrm {a}}}$
as a
$\mathbf {D}_W$
-module, where
$\mathbf {D}_W$
stands for the FADA associated with the finite part of the Kac–Moody root system.
Our next result can be viewed as an extension of the Peterson conjecture.
Theorem 1.2 (cf. Theorem 5.6)
The localization
$\mathbf {D}_{Q^{\vee }\!,\mathrm {loc}}$
of the formal Peterson subalgebra
$\mathbf {D}_{Q^{\vee }}$
with respect to an affine root system of type
$\hat A_1$
has the following presentation:

where
$x_{-1}$
is a certain characteristic class in h and
$\mu $
is an element depending on
$x_{-1}$
.
For cohomology and K-theory, this presentation gives quantum cohomology and quantum K-theory of
$\mathbb {P}^1,$
respectively. Hence,
$\mathbf {D}_{Q^{\vee }\!,\mathrm {loc}}$
can be viewed as the “quantum” oriented cohomology of the projective line
$\mathbb {P}^1$
.
As for our last result, observe that the S-linear dual
$\mathbf {D}_{Q^{\vee }}^{*}$
of the formal Peterson subalgebra is a natural model for the (small torus) equivariant oriented cohomology of the affine Grassmannian [Reference Zhong18]. We obtain the following “Kac–Moody” analog of results of [Reference Lanini, Xiong and Zainoulline13].
Theorem 1.3 (cf. Theorem 6.2)
The S-linear dual
$\mathbf {D}_{Q^{\vee }}^{*}$
of the formal Peterson subalgebra is isomorphic to the
$0$
-th Hochschild homology of the dual
$\mathbf {D}_{W_{\mathrm {a}}}^{*}$
of the small torus FADA.
Here, the dual
$\mathbf {D}_{W_{\mathrm {a}}}^{*}$
can be interpreted as a model for the T-equivariant oriented cohomology of the respective affine flag variety. Therefore, it has two commuting actions by the equivariant coefficient ring S. Following the ideas of [Reference Lanini, Xiong and Zainoulline13] one defines its zeroth Hochschild homology as the quotient obtained by merging these two S-module structures. To prove this result, we introduce a special filtration on the dual
$\mathbf {D}_{Q^{\vee }}^{*}$
(to reduce it to finite cases). This approach seems to be new even for cohomology and K-theory.
The article is organized as follows: Section 2 revisits the definition of the formal Peterson subalgebra
$\mathbf {D}_{Q^{\vee }}$
from [Reference Zhong18]. In Section 3, we establish some basic properties of
$\mathbf {D}_{Q^{\vee }}$
and study the action of
$\mathbf {D}_{W_{\mathrm {a}}}$
on it. In Section 4, we study Borel isomorphisms involving the FADA and the Peterson subalgebra, and prove our first main result Theorem 4.4. In Section 5, we focus on the example of type
$\hat A^1$
and prove our second result, Theorem 5.6. In the last section, we investigate the dual of the formal Peterson subalgebra, and prove our third main result Theorem 6.2. In the appendix, we prove several combinatorial properties of the affine Weyl group that are used in the proof of Theorem 6.2.
2 The formal Peterson subalgebra
In this section, we recall the definition of a small torus FADA and the formal Peterson algebra following [Reference Calmès, Zainoulline and Zhong2, Reference Zhong18].
Given an oriented algebraic cohomology theory
$h(\text {-}),$
in the sense of Levine–Morel (see [Reference Levine and Morel15]) there is an associated formal group law F over a commutative ring R with characteristic 0. Here,
$R=h(pt)$
is the coefficient ring, and F is defined from the Quillen formula for the characteristic class of a tensor product of line bundles. For example, for connective K-theory (see, e.g., [Reference Richmond and Zainoulline17]) we have
$F_\beta (x,y)=x+y-\beta xy$
over the polynomial ring
$R=\mathbb {Z}[\beta ]$
. Specializing to
$\beta =1$
(resp.
$\beta =0$
), one obtains the usual K-theory (resp. cohomology). In these notes, by usual cohomology, we always mean its algebraic part: the Chow ring (modulo rational equivalence) with rational coefficients.
Given a lattice
$\Lambda $
(free abelian group of finite rank) and a formal group law F, consider the associated formal group algebra S of [Reference Calmés, Petrov and Zainoulline1] that is the quotient of the power series ring

where
$\mathcal {J}_F$
is the closure of the ideal of relations

In the case
$F=F_\beta ,$
we set S to be the quotient of the polynomial ring

Let
$\Phi $
be a finite irreducible root system with a fixed subset
$I=\{\alpha _1,\ldots ,\alpha _n\}$
of simple roots. Let Q and
$Q^{\vee }$
denote the root and the coroot lattice, respectively. Let W denote the Weyl group, generated by simple reflections
$s_{\alpha _i}$
,
$\alpha _i\in I$
. Consider an affine root system corresponding to the extended Dynkin diagram for
$\Phi $
with the extra simple root

where
$\theta \in \Phi $
is the highest root and
$\delta $
is the so called null root so that
$s_{\alpha _0}= t_{\theta ^{\vee }} s_\theta $
is an extra generator of the respective affine Weyl group
$W_{\mathrm {a}}= Q^{\vee } \rtimes W$
. Recall that the latter is generated by reflections
$s_{\alpha +k\delta }= t_{-k\alpha ^{\vee }} s_\alpha $
, where
$s_\alpha \in W$
is a reflection and
$t_\lambda $
,
$\lambda \in Q^{\vee }$
is a translation. The affine Weyl group
$W_{\mathrm {a}}$
acts on the lattice Q via W that is

Therefore, it also acts on the formal group algebra
$S=R[\![Q]\!]_F$
.
Suppose
$x_\alpha $
is a regular (not a zero-divisor) element in S for each
$\alpha \in \Phi $
. In particular, this holds if
$2$
is not a zero-divisor in R (see [Reference Calmès, Zainoulline and Zhong4, Lemma 2.2]). Let
$\mathfrak {Q}=S[\tfrac {1}{x_\alpha }\colon \alpha \in \Phi ]$
be the localization of S at
$x_\alpha $
s. Consider the twisted group algebra
$\mathfrak {Q}_{W_{\mathrm {a}}}$
associated with the affine Weyl group
$W_{\mathrm {a}}$
. By definition, it is a free left
$\mathfrak {Q}$
-module
$\mathfrak {Q}_{W_{\mathrm {a}}}=\mathfrak {Q}\otimes _RR[W_{\mathrm {a}}]$
with basis
$\{\eta _u\}_{u\in W_{\mathrm {a}}}$
and the product given by

For each
$\alpha \in \Phi $
, define elements

All these elements satisfy the quadratic relations (e.g.,
$X_\alpha ^2=\kappa _\alpha X_\alpha $
) and the twisted braid relations (see, e.g., [Reference Calmès, Zainoulline and Zhong2]). For simplicity of notation, we will omit
$\alpha $
or s in the indices, i.e., we will write
$x_i=x_{\alpha _i}$
,
$s_i=s_{\alpha _i}$
,
$\eta _i=\eta _{s_{i}}$
,
$X_i=X_{\alpha _i}$
, and
$Y_i=Y_{\alpha _i}$
.
Similarly, consider the twisted group algebra
$\mathfrak {Q}_{Q^{\vee }}=\mathfrak {Q}\otimes _RR[Q^{\vee }]$
. It is a free
$\mathfrak {Q}$
-module with basis
$\{\eta _{t_\lambda }\}_{\lambda \in Q^{\vee }}$
. Observe that
$\mathfrak {Q}_{Q^{\vee }}$
is commutative since
$t_\lambda (c)=c$
,
$c\in \mathfrak {Q}$
. Consider two homomorphisms of left
$\mathfrak {Q}$
-modules

By definition,
$\imath $
is a section of
$\mathrm {pr}$
, and it is a ring homomorphism. Set
$\psi =\imath \circ \mathrm {pr}$
, so
$\psi |_{\imath (\mathfrak {Q}_{Q^{\vee }})}=\mathrm {id}$
. Define elements

Lemma 2.1 We have
$\psi (zX_i)=0$
for any
$z\in \mathfrak {Q}_{W_{\mathrm {a}}}$
,
$\alpha _i\in I$
, and
$\psi (X_0)=Z_\theta $
.
Proof For
$z=c\eta _u$
,
$c\in \mathfrak {Q}$
,
$u\in W_{\mathrm {a}}$
, and
$\alpha _i\in I,$
we have

As for the
$X_0$
, we obtain

and the proof is finished.
Denote
$I_{\mathrm {a}}=\{\alpha _0,\ldots ,\alpha _n\}$
. Following [Reference Calmès, Zainoulline and Zhong2] define the small torus FADA
$\mathbf {D}_{W_{\mathrm {a}}}$
to be the subring of
$\mathfrak {Q}_{W_{\mathrm {a}}}$
generated by S and the elements
$X_i$
,
$\alpha _i\in I_{\mathrm {a}}$
. Set
$\mathbf {D}_{W_{\mathrm {a}}/W}=\mathrm {pr}(\mathbf {D}_{W_{\mathrm {a}}})$
to be its image in
$\mathfrak {Q}_{Q^{\vee }}$
. We called it the relative FADA. Assuming
$\mathbb {Q} \subset R$
and using the small torus GKM descripition it is proven in [Reference Zhong18, Lemma 5.1] that the map
$\imath $
induces a map
$\mathbf {D}_{W_{\mathrm {a}}/W}\to \mathbf {D}_{W_{\mathrm {a}}}$
. We then define the formal Peterson subalgebra to be the image of the relative FADA

According to [Reference Zhong18, Theorem 5.7] the formal Peterson subalgebra
$\mathbf {D}_{Q^{\vee }}$
is a Hopf subalgebra in
$\mathfrak {Q}_{Q^{\vee }}$
. Moreover,
$\mathbf {D}_{Q^{\vee }}$
coincides with the centralizer
$C_{\mathbf {D}_{W_{\mathrm {a}}}}(S)$
of the formal group algebra S in the FADA
$\mathbf {D}_{W_{\mathrm {a}}}$
.
3 Properties of the FADA and the formal Peterson subalgebra
In the present section, we establish several properties of the FADA and the formal Peterson subalgebra. For the K-theory, some of these properties were proven in [Reference Kato7] using different arguments. We start with the following version of the projection formula.
Lemma 3.1 For any z,
$z'\in \mathfrak {Q}_{W_{\mathrm {a}}}$
and
$\xi \in \mathfrak {Q}_{Q^{\vee }}$
we have in
$\mathfrak {Q}_{Q^{\vee }}$
-
(i)
$\mathrm {pr}(\imath (\xi )z)=\xi \mathrm {pr}(z)$ , and
-
(ii)
$\mathrm {pr}(z\sigma z')=\mathrm {pr}(z)\mathrm {pr}(\sigma z')$ , where
$\sigma =\sum _{w\in W}\eta _w$ .
Observe that for the K-theory the property (ii) played a key role in [Reference Kato7, Theorem 1.7].
Proof (i) Let
$\xi =c_1\eta _{t_{\lambda _1}}$
and
$z=c_2\eta _{t_{\lambda _2}w}$
, where
$w\in W$
,
$c_i\in \mathfrak {Q}$
,
$\lambda _i\in Q^{\vee }$
. Then, we obtain

(ii) Let
$z=c\eta _{t_{\lambda }v}$
and
$z'=c'\eta _{t_{\lambda '}v'}$
,
$v,v'\in W$
. Then, we get

Since
$t_\lambda vw t_{\lambda '} v'=t_\lambda (vwt_{\lambda '}(vw)^{-1}) vw v'=t_\lambda t_{vw(\lambda ')}vwv'$
, reindexing the sum by
$w'=vw$
we obtain

On the other side,
$\mathrm {pr}(z)=c\eta _{t_\lambda }$
and

The result then follows.
We now extend the Hecke action on the Peterson algebra for the K-theory introduced in [Reference Kato7, Section 2] to the action on the formal Peterson algebra.
We define an action of
$\mathfrak {Q}_{W_{\mathrm {a}}}$
on
$\mathfrak {Q}_{Q^{\vee }}$
by

More explicitly, we have

Direct computation shows that
$W_{\mathrm {a}}$
is an action.
For
$w\in W$
,
$\xi \in \mathfrak {Q}_{Q^{\vee },}$
and
$\alpha \in \Phi $
define

We then have the following.
Lemma 3.2 For any
$z,z'\in \mathfrak {Q}_{W_{\mathrm {a}}}$
and
$\xi \in \mathfrak {Q}_{Q^{\vee }}$
we have
-
(i)
$\mathrm {pr}(z z')=z \diamond \mathrm {pr}(z')$ , and, in particular,
$$\begin{align*}\mathrm{pr}(X_iz)=X_i\diamond \mathrm{pr}(z)=\Delta_i( \mathrm{pr}(z)), ~\alpha_i\in I,\end{align*}$$
-
(ii)
$X_0\diamond \xi =\Delta _{-\theta }(\xi )+Z_\theta s_{\theta }(\xi )$ .
Proof (i) Let
$z=c\eta _{t_{\lambda }w}$
and
$z'=c'\eta _{t_{\lambda '}w'}$
. Then, we obtain

(ii) For
$\xi =c\eta _{t_\lambda ,}$
we get

and the result follows.
Lemma 3.3 The
$\diamond $
-action of
$\mathfrak {Q}_{W_{\mathrm {a}}}$
on
$\mathfrak {Q}_{Q^{\vee }}$
induces an action of
$\mathbf {D}_{W_{\mathrm {a}}}$
on
$\mathbf {D}_{W_{\mathrm {a}}/W}$
.
Proof Let
$z,z'\in \mathbf {D}_{W_{\mathrm {a}}}$
, and let
$\xi =\mathrm {pr}(z')$
. Then, we have

and the lemma follows.
Identifying the formal Peterson algebra
$\mathbf {D}_{Q^{\vee }}$
(resp.
$\imath (\mathfrak {Q}_{Q^{\vee }})$
) with
$\mathbf {D}_{W_{\mathrm {a}}/W}$
(resp.
$\mathfrak {Q}_{Q^{\vee }}$
) via the ring homomorphism
$\imath $
we obtain an action of
$\mathbf {D}_{W_{\mathrm {a}}}$
on
$\mathbf {D}_{Q^{\vee }}$
and an action of
$\mathfrak {Q}_{W_{\mathrm {a}}}$
on
$\imath (\mathfrak {Q}_{Q^{\vee }})$
. From this point on, we write
$\xi $
as both an element in
$\mathbf {D}_{W_{\mathrm {a}}/W}$
(resp.
$\mathfrak {Q}_{Q^{\vee }}$
) and in
$\mathbf {D}_{Q^{\vee }}=\imath (\mathbf {D}_{W_{\mathrm {a}}/W})$
(resp.
$\imath (\mathfrak {Q}_{Q^{\vee }})$
). If we consider a product
$\xi _1\xi _2$
with
$\xi _i\in \mathfrak {Q}_{Q^{\vee }}$
, we may assume it is in
$\mathfrak {Q}_{W_{\mathrm {a}}}$
. However, for the product
$z\xi $
with
$z\in \mathfrak {Q}_{W_{\mathrm {a}}}$
and
$\xi \in \mathfrak {Q}_{Q^{\vee }}$
, we view
$\xi $
as an element in
$\mathfrak {Q}_{W_{\mathrm {a}}}$
via the map
$\imath $
. Following these identifications we obtain

and Lemma 3.1 gives

Example 3.4 Consider the affine root system of extended Dynkin type
$\hat A_2$
. It has three simple roots
$\alpha _0,\alpha _1,\alpha _2$
and the highest root
$\theta =\alpha _1+\alpha _2$
. Denote
$X_{ij}=X_{i}X_j$
for simplicity. Direct computations then give:

Finally, we describe the centre of FADA.
Lemma 3.5 (i) For any
$\xi \in \mathfrak {Q}_{Q^{\vee }}$
and
$\alpha _i\in I$
, we have

Moreover, if this condition is satisfied, we have

(ii) The centres of
$\mathfrak {Q}_{W_{\mathrm {a}}}$
and
$\mathbf {D}_{W_{\mathrm {a}}}$
can be described as follows:

(iii) There are ring homomorphisms

Proof (i) For a given
$\xi =\sum _\lambda c_{\lambda } \eta _{t_\lambda }$
,
$c_\lambda \in \mathfrak {Q,}$
we get

On the other side, we have

Therefore,
$\eta _i\diamond \xi =\xi $
if and only if
$s_i(c_{s_i(\lambda )})=c_\lambda $
for any
$\lambda \in Q^{\vee }$
, which is equivalent to say that
$\eta _i\xi =\xi \eta _i$
.
Now if this condition is satisfied, then

(ii) Since
$\mathfrak {Q}_{Q^{\vee }}=\imath (\mathfrak {Q}_{Q^{\vee }})=C_{\mathfrak {Q}_{W_{\mathrm {a}}}}(\mathfrak {Q})$
, we have
$Z(\mathfrak {Q}_{W_{\mathrm {a}}})\subset \mathfrak {Q}_{Q^{\vee }}$
, and the first identity then follows. By part (i), we know that
$\eta _i\diamond \xi =\xi $
,
$\forall \alpha _i\in I$
is equivalent to
$\eta _w \xi =\xi \eta _w$
,
$\forall w\in W$
that is equivalent to
$z\xi =\xi z$
,
$\forall z\in \mathfrak {Q}_{W}$
. Since
$\xi $
already commutes with
$\eta _{t_\lambda }$
,
$\lambda \in Q^{\vee }$
,
$\xi $
belongs to the centre
$Z(\mathfrak {Q}_{W_{\mathrm {a}}})$
. Conversely, if
$\xi \in Z(\mathfrak {Q}_{W_{\mathrm {a}}})\cap \mathfrak {Q}_{Q^{\vee }}$
, then it is invariant under all
$\eta _i$
,
$\alpha _i\in I$
. The description of the centre
$Z(\mathbf {D}_{W_{\mathrm {a}}})$
follows similarly.
(iii) Follows from parts (i) and (ii).
4 Borel isomorphisms
In this section, we study Borel isomorphisms involving the FADA and the formal Peterson subalgebra. We assume
$\mathbb {Q} \subset R$
throughout this section.
Consider the left S-linear dual
$\mathbf {D}_W^{*}$
embedded into
$\mathfrak {Q}_W^{*}$
. The latter has a
$\mathfrak {Q}$
-basis
$\{f_w\}_{w\in W}$
. Following [Reference Calmès, Zainoulline and Zhong4, Section 11] there is an (equivariant) characteristic map

which induces the Borel isomorphism (see [Reference Calmès, Zainoulline and Zhong4, Theorem 11.4])

Recall that
$\sigma = \sum _{w\in W}\eta _{w}\in \mathbf {D}_{W_{\mathrm {a}}}$
. Denote
$\mathbf {x}=\prod _{\alpha \in \Phi ^+}x_{-\alpha }$
and
$Y = \sigma \tfrac {1}{\mathbf {x}}$
.
By [Reference Calmès, Zainoulline and Zhong3, Lemma 10.12] (
$Y=Y_\Pi $
) we have
$Y\in \mathbf {D}_W$
. We denote by
$X_{I_u}$
,
$Y_{I_u}$
products corresponding to a reduced sequence
$I_u$
of
$u\in W_{\mathrm {a}}$
.
Lemma 4.1 We have
$\sigma \mathbf {D}_W=Y\mathbf {D}_W=Y S$
.
Proof Observe that
$Y=\tfrac {1}{|W|} \sigma Y$
, so
$Y S\subset Y\mathbf {D}_{W}\subset \sigma Y \mathbf {D}_{W}\subset \sigma \mathbf {D}_W$
. Conversely,

Note that
$\mathbf {D}_W$
is also a right S module with basis
$X_{I_v}, v\in W$
, and
$Y X_{I_v}=\delta _{v,e} Y$
. So given
$X_{I_v}b\in \mathbf {D}_{W}$
with
$v\in W$
,
$b\in S$
, we have

So
$\sigma \mathbf {D}_W\subset Y S$
, and the result follows.
Lemma 4.2 We have
$SYS=\mathbf {D}_W$
. So
$\mathbf {D}_W$
is a cyclic S-S-bimodule.
Proof According to [Reference Calmès, Zainoulline and Zhong3, Lemma 10.3]
$\mathbf {x} f_e\in \mathbf {D}_W^{*}$
. Let
$\sum _i a_i\otimes b_i\in S\otimes _{S^W}S$
so that
$\rho (\sum _i a_i\otimes b_i)=\mathbf {x} f_e$
. Then,

Therefore,

Finally, by Lemma 4.1, for any
$z\in \mathbf {D}_W,$
we can write
$z=\sum _i a_iY b_iz=\sum _i a_iY b^{\prime }_i$
for some
$b^{\prime }_i\in S$
.
Lemma 4.3 We have
$\psi (\sigma \mathbf {D}_{W_{\mathrm {a}}}) = Z(\mathbf {D}_{W_{\mathrm {a}}})$
.
Proof Let
$z=c\eta _{t_{\lambda }u}$
, where
$c\in \mathfrak {Q}$
,
$u\in W$
. We have

so it obviously belongs to
$Z(\mathbf {D}_{W_{\mathrm {a}}})$
.
As for the opposite inclusion, take
$z\in Z(\mathbf {D}_{W_a})$
. Since
$Z(\mathbf {D}_{W_{\mathrm {a}}})\subset C_{\mathbf {D}_{W_{\mathrm {a}}}}(S)=\mathbf {D}_{Q^{\vee }}$
, we get
$\psi (z)=z$
. Observe that
$\mathrm {pr}(z'\sigma )=|W|\mathrm {pr}(z')$
for any
$z'\in \mathbf {D}_{W_{\mathrm {a}}}$
. So we obtain

Thus,
$z\in \psi (\sigma \mathbf {D}_{W_{\mathrm {a}}})$
, and the proof is finished.
Consider two ring homomorphisms induced by the usual mutiplication:


Note that in the definition of
$\Theta $
and
$\Xi ,$
one can switch the tensor factors. Moreover, both homomorphisms are left S-linear. The following is our first main result.
Theorem 4.4 Assume
$\mathbb {Q}\subset R$
. The maps
$\Theta $
and
$\Xi $
are ring isomorphisms.
Remark 4.5 The geometric interpretation of this theorem is well-known for equivariant homology and equivariant K-homology. As explained in [Reference Lam, Lee and Shimozono9, Reference Lam and Shimozono12], we have the following isomorphisms of algebras

where
$G/B$
is the flag variety,
$Fl_G$
is the affine flag variety, and
$Gr_G$
is the affine Grassmannian. By (ii) of Lemma 3.5, we could identify

As a result, the morphisms
$\Theta $
and
$\Xi $
can be viewed as the isomorphisms of algebras

The isomorphism
$\Xi $
can be also rewritten in a more familiar form

which is what Corollary 4.8 implies.
For a general equivariant oriented cohomology theory h, it follows from [Reference Calmès, Zainoulline and Zhong5] that
$\mathbf {D}_W^{*}$
is isomorphic to the equivariant oriented cohomology
$h_T(G/B)$
. However, the respective results for
$Fl_G$
and
$Gr_G$
are not known since they are not varieties of finite type. Therefore, our isomorphism in this case serves as the algebraic analogs of the potentially-correct geometric result. Observe also that in general, the isomorphisms in Corollary 4.8 are only isomorphisms of modules.
The proof of the theorem will occupy the rest of this section. We start proving the surjectivity first.
Lemma 4.6 The map
$\Theta \colon S\otimes _{S^W} Z(\mathbf {D}_{W_{\mathrm {a}}}) \to \mathbf {D}_{Q^{\vee }}$
given in (4.3) is surjective.
Proof Consider the following diagram

Since
$\psi $
is an S-module homomorphism, this diagram commutes.
By Lemma 4.2, we can write
$1=\sum _i a_iY b_i$
for some
$a_i$
,
$b_i\in S$
. For any
$z\in \mathbf {D}_{W_{\mathrm {a}}}$
, we then have
$z=\sum _i a_i Y b_i z$
. This shows that elements of
$Y \mathbf {D}_{W_{\mathrm {a}}}$
generate
$\mathbf {D}_{W_{\mathrm {a}}}$
as a left S-module. Similarly to the proof of Lemma 4.1, we obtain that
$Y \mathbf {D}_{W_{\mathrm {a}}}=\sigma \mathbf {D}_{W_{\mathrm {a}}}$
. So the top horizontal map is surjective, and the result follows.
Lemma 4.7 The map
$\;\Xi \colon \mathbf {D}_W\otimes _{S^W} Z(\mathbf {D}_{W_{\mathrm {a}}}) \to \mathbf {D}_{W_{\mathrm {a}}}$
given in (4.4) is surjective.
Proof Since the elements of
$\mathbf {D}_W$
and of
$Z(\mathbf {D}_{W_{\mathrm {a}}})$
commute with each others, the image of
$\Xi $
is the subalgebra generated by
$\mathbf {D}_W$
and
$Z(\mathbf {D}_{W_{\mathrm {a}}})$
. It contains S and
$X_i$
for
$\alpha _i\in I$
by definition, so it suffices to show that it contains
$X_0$
as well. Observe that

Since
$Z_{\theta }\in \mathbf {D}_{Q^{\vee }}$
by [Reference Zhong18, Lemma 4.1], we have
$X_\theta \in \mathbf {D}_W$
and
$\eta _{s_\theta }\in \mathbf {D}_W$
. So
$X_0$
belongs to the subalgebra generated by
$\mathbf {D}_W$
and
$Z(\mathbf {D}_{W_{\mathrm {a}}})$
.
Corollary 4.8 The maps
$\mathbf {D}_W\otimes _{S}\mathbf {D}_{Q^{\vee }}\to \mathbf {D}_{W_{\mathrm {a}}}$
and
$\mathbf {D}_{Q^{\vee }}\otimes _{S}\mathbf {D}_W\to \mathbf {D}_{W_{\mathrm {a}}}$
induced by the usual multiplication are isomorphisms of left S-modules (In the first map,
$\mathbf {D}_W$
is viewed as an S-S-bimodule, and in the second map,
$\mathbf {D}_W$
is viewed as a left S-module.)
Proof Since
$\mathbf {D}_{Q^{\vee }}\supset Z(\mathbf {D}_{W_{\mathrm {a}}})$
, by Lemma 4.7, both maps are surjective. To prove the injectivity, we change the base to
$\mathfrak {Q}$
-modules by applying the exact functors
$\text {--}\otimes _S\mathfrak {Q}$
and
$\mathfrak {Q}\otimes \text {--}$
. It then suffices to show that the induced maps


are injective. But these are even isomorphisms. So, the conclusion follows.
We now discuss injectivity of the maps in the theorem.
For any parabolic subgroup
$W_P$
of
$W,$
we denote by
$W^P$
the subset of minimal length left coset representatives. Consider the
$\mathfrak {Q}$
-linear dual
$\mathfrak {Q}^{*}_{W^P}=\mathrm {Hom}(W^P,\mathfrak {Q})$
with a basis
$\{f_{w}\}_{w\in W^P}$
. One can also identify it with the invariants
$(\mathfrak {Q}_W^{*})^{W_P}$
by identifying each
$f_w$
,
$w\in W^P$
with
$\sum _{v\in W_P}f_{wv}\in (\mathfrak {Q}_W^{*})^{W_P}$
(see [Reference Calmès, Zainoulline and Zhong3, Section 11] for more details).
Lemma 4.9 The map
$\rho _{P,\mathfrak {Q}}\colon S\otimes _{S^W}\mathfrak {Q}^{W_P}\to \mathfrak {Q}_{W^P}^{*}$
defined by

is an isomorphism.
Proof Assume first that
$P=B$
(the Borel case). Then, the map
$\rho _{P,\mathfrak {Q}}$
is obtained from the isomorphism
$\rho $
by the base change with the functor
$\text {--}\otimes _S\mathfrak {Q}$
. So
$\rho _{B,\mathfrak {Q}}$
is an isomorphism.
For a general parabolic
$W_P,$
there is a commutative diagram

Both vertical maps identify the top with the
$W_P$
-invariant subsets of the bottom, so the top horizontal map is an isomorphism.
Lemma 4.10 The map
$\Theta \colon S\otimes _{S^W} Z(\mathbf {D}_{W_{\mathrm {a}}}) \to \mathbf {D}_{Q^{\vee }}$
defined in (4.3) is injective.
Proof Let
$z=\sum _{\lambda \in Q^{\vee }} c_{\lambda }\eta _{t_{\lambda }}\in Z(\mathbf {D}_{W_{\mathrm {a}}})\subset \mathbf {D}_{Q^{\vee }}$
with
$c_\lambda \in \mathfrak {Q}$
. Since
$\eta _u z=z\eta _u$
for any
$u\in W$
, we have

These properties give us an injective map:

where
$Q^{\vee }_{\ge 0}$
is the set of dominant coroots and
$W_{\lambda }$
is the stabilizer of
$\lambda $
, which is a parabolic subgroup of W.
Let
$W^\lambda $
denote the set of minimal length representatives of the cosets
$W/W_\lambda $
. Consider the following diagram

where
$\Theta '$
is the direct sum of maps

for all
$\lambda \in Q_{\geq 0}^{\vee }$
. Since by Lemma 4.9, each such component map is injective, so is
$\Theta '$
.
By direct computations and by the property (*), the diagram is commutative. Since both maps
$\operatorname {id}\otimes \phi $
and
$\Theta '$
are injective, so is
$\Theta $
.
Lemma 4.11 The map
$\Xi \colon \mathbf {D}_W\otimes _{S^W} Z(\mathbf {D}_{W_{\mathrm {a}}}) \to \mathbf {D}_{W_{\mathrm {a}}}$
defined in (4.4) is injective.
Proof It follows from the combination of previous results:

5 The
$\hat A_1$
-case
In this section, we discuss an example of the formal Peterson subalgebra for the affine root system of type
$\hat A_1$
. We show that it provides a natural model for “quantum” oriented cohomology of
$\mathbb {P}^1$
.
Recall that a root system of type
$\hat A_1$
has two simple roots,
$\alpha _1=\theta =\alpha $
and
$\alpha _0=-\alpha +\delta $
, and each
$w\in W_{\mathrm {a}}$
has a unique reduced decomposition. We follow the notation of [Reference Lam, Schilling and Shimozono11, Section 4.3] and define for
$i\ge 1$
:

The set of minimal length coset representatives of
$W_{\mathrm {a}}/W$
is then
$W_{\mathrm {a}}^{-}=\{\sigma _i\colon i\ge 0\}$
.
The root lattice is
$Q=\mathbb {Z} \alpha $
, and in the formal group algebra
$S=R[\![Q]\!]_F,$
we have
$x_n=x_{n\alpha }=n\cdot _F x_{\alpha }$
, where
$n\cdot _F x$
,
$n\in \mathbb {Z}$
is the n-fold formal sum (inverse) of x.
As for the Demazure elements, we have

Set
$\mu =-\frac {x_{-1}}{x_1}$
. Observe that if F is of the form
$F(x,y)=\frac {x+y-\beta xy}{g(x,y)}$
for some power series
$g(x,y)$
, we have
$x_{-1}=\tfrac {x_1}{\beta x_1-1}$
, hence,
$\mu =\tfrac {1}{1-\beta x_1}$
. Given a reduced expression
$w=s_is_j\ldots $
, we will use the notation
$Y_{ij\cdots }$
for
$Y_w=Y_{I_w}$
. Denote
$\mathfrak {X}_w=\mathrm {pr}(X_w)$
and
$\mathfrak {Y}_w=\mathrm {pr}(Y_w)$
.
Example 5.1 Direct computations give:

and

These computations also show that
$\mathfrak {X}_{\sigma _{i}}$
,
$i=1,2,3$
satisfy identities similar to those of [Reference Lam, Li, Mihalcea and Shimozono10, Lemma 3].
We now look at various products of elements
$\mathfrak {Y}_w\in \mathbf {D}_{Q^{\vee }}$
.
Lemma 5.2 For each
$i\ge 1$
and
$w\in W_{\mathrm {a}}^{-}$
we have
$\mathfrak {Y}_{w\sigma _{2i}}=\mathfrak {Y}_{w}\mathfrak {Y}_{\sigma _{2i}}$
.
In particular,
$\mathfrak {Y}_{\sigma _{2i}}=\mathfrak {Y}_{\sigma _2}^i=\mathfrak {Y}_{10}^i$
and, therefore,
$\{\mathfrak {Y}_{\sigma _{2i}}\colon i\ge 1\}$
is a multiplicative set.
Proof Observe that
$\sigma _{2i}=s_1s_0s_1\cdots s_0$
, so
$Y_{\sigma _{2i}}=Y_1Y_{\sigma _{2i-1}}=(1+\eta _1)\frac {1}{x_{-\alpha }}Y_{\sigma _{2i-1}}$
. By Lemma 3.1, we obtain

and the result follows.
Lemma 5.3 For each
$i\ge 1$
we have
$\mathfrak {Y}_{\sigma _{2i}}\in (\mathbf {D}_{Q^{\vee }})^W$
, and for
$j=0,1$

In particular, we have

Proof Since
$\sigma _{2i}=s_1\sigma _{2i-1}$
, we get
$Y_{\sigma _{2i}}=(1+\eta _1)\tfrac {1}{x_{-\alpha }}Y_{\sigma _{2i-1}}$
, which implies that
$\eta _1Y_{\sigma _{2i}}=Y_{\sigma _{2i}}$
. By Lemma 3.1, we then obtain

therefore,
$\mathfrak {Y}_{\sigma _{2i}}\in (\mathbf {D}_{Q^{\vee }})^W$
. Similarly, we obtain
$Y_j\diamond \mathfrak {Y}_{w}=\mathrm {pr}(Y_jY_{w})$
, and the formula for the action follows.
Corollary 5.4 The set
$\mathbf {D}_{Q^{\vee }}$
is a cyclic
$\mathbf {D}_{W_{\mathrm {a}}}$
-module, generated by
$\mathfrak {Y}_{0}=\mathfrak {Y}_{\sigma _1}$
.
Moreover, the kernel of the map
$\pi \colon \mathbf {D}_{W_{\mathrm {a}}}\to \mathbf {D}_{Q^{\vee }}$
defined by
$z\mapsto z\diamond \mathfrak {Y}_{\sigma _1}$
is
$\mathbf {D}_{W_{\mathrm {a}}} X_0$
.
Proof The first part follows from (5.1). For the second part, we have

so
$X_0\in \ker \pi $
.
Conversely, let
$z=\sum _{i\ge 1}a_iY_{\sigma _i}+\sum _{j\ge 0}b_jY_{\sigma _{-j}}\in \ker \pi $
. We then obtain

Therefore, if
$z\diamond \mathfrak {Y}_{\sigma _1}=0$
, then
$b_{i-1}=-a_i\kappa _\alpha $
for all
$i\ge 1$
, and we obtain

Remark 5.5 Observe that the map
$\mathrm {pr}:\mathbf {D}_{W_{\mathrm {a}}}\to \mathbf {D}_{W_{\mathrm {a}}/W}\simeq \mathbf {D}_{Q^{\vee }}$
has the kernel
$\mathbf {D}_{W_{\mathrm {a}}} X_1=\oplus _{i<0}SX_{\sigma _i}$
, while the map
$\pi $
has the kernel
$\mathbf {D}_{W_{\mathrm {a}}} X_0=\oplus _{i> 0}SX_{\sigma _i}$
.
For a general affine root system (for an extended Dynkin diagram), one can show that
$\mathbf {D}_{Q^{\vee }}$
is a cyclic left
$\mathbf {D}_{W_{\mathrm {a}}}$
-module, generated by
$\mathrm {pr}(Y_0)$
.
From the identities of Example 5.1 it follows that

and we obtain the following presentation of the formal Peterson algebra in terms of generators and relations:

According to Lemma 5.2, we may define the localization

From (5.2), we then obtain our second main result.
Theorem 5.6 We have the following presentation

In particular, the action of
$\mathbf {D}_{W_{\mathrm {a}}}$
on
$\mathbf {D}_{Q^{\vee }}$
extends to an action on the localization
$\mathbf {D}_{Q^{\vee }\!,\mathrm {loc}}$
by

Observe that it is well-defined since for any
$i,j\ge 1,$
we have

It then follows from Corollary 5.4 that
Corollary 5.7 The localized algebra
$\mathbf {D}_{Q^{\vee }\!,\mathrm {loc}}$
is a cyclic
$\mathbf {D}_{W_{\mathrm {a}}}$
-module generated by
$\mathfrak {Y}_0$
.
Remark 5.8 Let
$F(x,y)=x+y-\beta x y$
. Observe that for cohomology (
$\beta =0$
) and K-theory (
$\beta =1$
) the localization
$\mathbf {D}_{Q^{\vee }\!,\mathrm {loc}}$
computes quantum cohomology and quantum K-theory of
$\mathbb {P}^1,$
respectively. For instance, for K-theory the presentation (5.2) recovers that of [Reference Lam, Li, Mihalcea and Shimozono10, Equation 17]. Therefore, it makes sense to think of
$\mathbf {D}_{Q^{\vee }\!,\mathrm {loc}}$
as an algebraic model for “quantum” oriented cohomology of the projective line
$\mathbb {P}^1$
.
Finally, by the result of [Reference Zhong18]
$\mathbf {D}_{Q^{\vee }}$
is a Hopf algebra with coproduct defined by

In our case, we obtain

Example 5.9 In particular, for the cohomology we get

and for the K-theory (identifying
$x_\alpha =1-e^{-\alpha }$
) we get

6 The dual of the formal Peterson subalgebra
In this section, we study the dual of the formal Peterson subalgebra.
Consider the
$\mathfrak {Q}$
-linear dual of the twisted group algebra
$\mathfrak {Q}_{W_{\mathrm {a}}}^{*}=\mathrm {Hom}_{\mathfrak {Q}}(\mathfrak {Q}_{W_{\mathrm {a}}}, \mathfrak {Q})$
. It is generated by
$f_w$
,
$w\in W_{\mathrm {a}}$
. Following [Reference Lenart, Zainoulline and Zhong14] there are two actions of
$\mathfrak {Q}_{W_{\mathrm {a}}}$
on the dual
$\mathfrak {Q}_{W_{\mathrm {a}}}^{*}$
defined as follows:

Indeed, the
$\odot $
-action comes from left multiplication in
$\mathfrak {Q}_{W_{\mathrm {a}}}$
, and the
$\bullet $
-action comes from right multiplication. Observe that these two actions commute, which makes
$\mathfrak {Q}_{W_{\mathrm {a}}}^{*}$
into a
$\mathfrak {Q}$
-
$\mathfrak {Q}$
-bimodule. Moreover,
$(z\bullet f)(z')=f(zz')$
,
$z,z'\in \mathfrak {Q}_{W_{\mathrm {a}}}$
,
$f\in \mathfrak {Q}_{W_{\mathrm {a}}}^{*}$
.
We now define two tensor products.
The first one is the tensor product
$\mathfrak {Q}_{W_{\mathrm {a}}}\otimes \mathfrak {Q}_{W_{\mathrm {a}}}$
of left
$\mathfrak {Q}$
-modules that is

There is a canonical map
$\Delta \colon \mathfrak {Q}_{W_{\mathrm {a}}}\to \mathfrak {Q}_{W_{\mathrm {a}}}\otimes \mathfrak {Q}_{W_{\mathrm {a}}}$
given by
$a\eta _w\mapsto a\eta _w\otimes \eta _w$
. This map defines a co-commutative coproduct structure on
$\mathfrak {Q}_{W_{\mathrm {a}}}$
with the co-unit
$\mathfrak {Q}\to \mathfrak {Q}_{W_{\mathrm {a}}}$
,
$a\mapsto a\eta _e$
.
The second tensor product
$\hat \otimes $
was introduced in [Reference Lanini, Xiong and Zainoulline13]. Here, we provide a different but equivalent definition:

Observe that
$\mathfrak {Q}_{W_{\mathrm {a}}}\hat \otimes \mathfrak {Q}_{W_{\mathrm {a}}}$
is also a left
$\mathfrak {Q}$
-module.
Similarly, we define:

By definition, there is an isomorphism of
$\mathfrak {Q}$
-modules:

There is a left
$\mathfrak {Q}$
-module homomorphism defined by the product structure of
$\mathfrak {Q}_{W_{\mathrm {a}}}$
:

whose dual is given by

Indeed, given any element
$a\eta _u\hat \otimes b\eta _v=au(b)\eta _u\hat \otimes \eta _v$
, we have

Recall (see also [Reference Zhong18, Section 1.7]) that there is the Borel map defined via the characteristic map

Similar to [Reference Lanini, Xiong and Zainoulline13] one obtains the following commutative diagram:

Definition 6.1 We define the
$0$
-th Hochschild homology of the bimodule
$\mathfrak {Q}_{W_{\mathrm {a}}}^{*}$
to be the quotient

Consider the dual of the map
$\imath \colon \mathfrak {Q}_{Q^{\vee }}\to \mathfrak {Q}_{W_{\mathrm {a}}}$
,
$\eta _{t_\lambda }\mapsto \eta _{t_\lambda } $
. It gives a surjection

We have

So it induces a surjection

On the other hand,
$\ker \imath ^{*}=\prod _{w\neq e, \lambda \in Q^{\vee }}\mathfrak {Q} f_{t_\lambda w}$
. Now for any w, let
$x_\mu \in S$
so that
$w(\mu )\neq \mu $
, then we obtain

Therefore, we have proven the following lemma.
Lemma 6.1 There is an isomorphism
$\mathop {\textrm {HH}_0}(\mathfrak {Q}_{W_{\mathrm {a}}}^{*})\simeq \mathfrak {Q}_{Q^{\vee }}^{*}$
which fits into the following commutative diagram

By definition, we have the commutative diagram of left S-modules:

It is also easy to see that the surjection
$\imath ^{*}\colon \mathfrak {Q}_{W_{\mathrm {a}}}^{*}\twoheadrightarrow \mathfrak {Q}_{Q^{\vee }}^{*}$
induces a surjective map

Our goal is to show that the isomorphism and the diagram of Lemma 6.1 can be restricted to the formal Peterson subalgebra
$\mathbf {D}_{Q^{\vee }}$
. Namely, we want to prove the following.
Theorem 6.2 The map
$\imath ^{*\prime }$
gives an isomorphism of Hopf algebras
$ \mathop {\textrm {HH}_0}(\mathbf {D}_{W_{\mathrm {a}}}^{*})\simeq \mathbf {D}_{Q^{\vee }}^{*}$
.
Since the product structure on both the domain and the codomain is induced by the coproduct structure

where the codomain is the tensor product of left
$\mathfrak {Q}$
-modules, the map
$\imath ^{*\prime }$
is a ring homomorphism.
Moreover, since the coproduct structure on both the domain and the codomain is induced by the product structure in
$\mathfrak {Q}_{W_{\mathrm {a}}}$
and
$\mathfrak {Q}_{Q^{\vee }}$
, the map
$\imath ^{*\prime }$
is a coalgebra homomorphism. Therefore, it only suffices to prove the injectivity of
$\imath ^{*\prime }$
.
To prove the latter, we introduce the following filtration on the dual
$\mathbf {D}_{W_{\mathrm {a}}}^{*}$
of the FADA.
Definition 6.2 Let
$w_\lambda $
be as defined in the appendix. Set
$F_i=\bigcup _{\ell (w_{\lambda })\geq i} t_\lambda W$
. For any
$f\in \mathfrak {Q}_{W_{\mathrm {a}}}^{*}=\mathrm {Hom}(W_{\mathrm {a}}, \mathfrak {Q})$
set
$\mathrm {supp} f=\{w\in W_{\mathrm {a}}\colon f(w)\neq 0\}$
. Define the i-th stratum of the filtration to be

We have
$\mathcal {Z}_i=\prod _{w\in F_i} S\cdot Y_{I_{w}}^{*}$
. Each
$\mathcal {Z}_i$
is a S-bimodule, since
$x\bullet af_u=u(x)af_u$
and
$x\odot af_u=xa f_u$
,
$x,a\in S$
,
$u\in W_{\mathrm {a}}$
.
Consider the following two conditions:


Proof Let
$f\in \mathcal {Z}_i$
and
$\ell (w_\lambda )=i$
. Assume
$\left <\lambda ,\alpha ^{\vee }\right>\leq 0$
. For
$k\in [1, \ell _{\alpha }(w_{\lambda })]$
by Lemma 7.2 of the appendix, we conclude that
$w_{\lambda +k\alpha ^{\vee }}<w_{\lambda }$
and, in particular,
$\ell (w_{\lambda +k\alpha ^{\vee }})<\ell (w_{\lambda })$
. So
$f(\eta _{\lambda +k\alpha ^{\vee }})=0$
, and we have for any
$u\in W$
,

Note that
$\frac {x_\alpha }{x_{-\alpha }}$
is invertible in S, so we can replace
$x_\alpha $
by
$x_{-\alpha }$
whenever needed.
Similarly, if
$\left <\lambda ,\alpha ^{\vee }\right>
<0$
, then
$\ell (w_{\lambda -k\delta })<\ell (w_{\lambda })$
for
$k\in [1,\ell _{\alpha }(w_{\lambda })]$
. Thus,

The result then follows.
Define

Here,
$f(\eta _{v_1}-\eta _{v_2}):=f(v_1)-f(v_2)$
. Note that
$\mathcal {Y}_{(i)}$
is a S-bimodule in the usual sense, that is
$(a\odot f)(v)=af(v)$
and
$(a\bullet f)(v)=v(a)f(v)$
. By Lemma 6.3, we have a natural projection
$\mathcal {Z}_i\to \mathcal {Y}_{(i)}$
, which induces an injective S-bimodule map

Lemma 6.4 The map
$\mathrm {res}$
is an isomorphism of S-bimodules. In particular,
$\mathcal {Z}_i/\mathcal {Z}_{i+1}$
is free of rank
$|F_i\backslash F_{i+1}|$
.
Proof We only need to prove that
$\mathrm {res}$
is surjective. Let
$f\in \mathcal {Y}_{(i)}$
. We pick a minimal element
$w\in \mathrm {supp}(f)$
. We first show that

As
$x_{\alpha }$
’s are relatively prime, it reduces to show

for each root
$\alpha $
.
Let
$w=t_{\lambda }u$
for
$\lambda \in Q^{\vee }$
,
$u\in W$
and
$\ell _\alpha (w_\lambda )=i$
. By Lemma 7.1 of the appendix,
$\ell _{\alpha }(w)\in \{\ell _{\alpha }(w_{\lambda }),\ell _{\alpha }(w_{\lambda })+1\}$
.
If
$\ell _{\alpha }(w)=\ell _{\alpha }(w_{\lambda })$
, then (6.3) follows from (6.1) directly. If
$\ell _{\alpha }(w)=\ell _{\alpha }(w_{\lambda })+1$
, by (7.2) and (7.3), we have

It implies
$t_{\lambda }s_{\alpha }u<w$
(note that
$t_{\lambda }s_{\alpha }u$
and w are always comparable under the Bruhat order). By (6.2), we have

Observe that the images of
$Y_{I_w}^{*}$
for
$w\in Z_i\setminus Z_{i+1}$
form a basis of
$\mathcal {Z}_i/\mathcal {Z}_{i+1}$
, and we have
$Y_{I_{w}}^{*}(\eta _{w})=\prod _{\alpha>0} x_{\alpha }^{\ell _{\alpha }(w)}$
.
The conclusion then follows after replacing f by
$f-\frac {f(\eta _w)}{\prod _{\alpha \in \Phi ^+}x_\alpha ^{\ell _\alpha (w_\lambda )}}Y_{I_w}^{*}$
.
Denote for each
$\lambda \in Q^{\vee }$
,
$\Delta _{\lambda }= \prod _{\alpha>0} x_{\alpha }^{\ell _{\alpha }(w_{\lambda })}\in S$
. It is clear that we have an S-bimodule isomorphism

To finish the proof of Theorem 6.2, we define a filtration on
$ \mathbf {D}_{Q^{\vee }}^{*}$
by

Then,
$Y_{I_w}^{*}$
with
$\ell (w_\lambda )=i$
is a S-basis of
$\mathcal {X}_i/\mathcal {X}_{i+1}$
. So the rank of
$\mathcal {X}_i/\mathcal {X}_{i+1}$
is
$|F_i\backslash F_{i+1}|$
. Moreover, by definition, we know that
$\imath ^{*\prime }$
induces a map on each associated graded piece:

From Lemma 6.4, the rank of
$\mathcal {Z}_i/\mathcal {Z}_{i+1}$
is
$|F_i\backslash F_{i+1}|$
, therefore,
$\imath ^{*\prime }$
is an isomorphism.
7 Appendix
Here, we prove several combinatorial properties of the affine Weyl group that are used in the proof of Theorem 6.2.
For
$w\in W_{\mathrm {a}}$
, denote

It is clear that
$\ell (w)=\sum _{\alpha>0} \ell _\alpha (w)$
. Also denote by
$w_{\lambda }\in W_a^{-}$
the minimal representative of
$t_{\lambda }W$
. Then,
$w_{\lambda }\leq w_{\mu }$
if and only if there exists
$w\in t_{\lambda }W$
and
$y\in t_{\mu }W$
such that
$w\leq y$
. Note that
$w\leq y$
implies
$\ell _{\alpha }(w)\leq \ell _{\alpha }(y)$
for all
$\alpha \in \Phi _+$
, so after fixing
$\alpha $
$\ell _\alpha (w_\lambda )$
becomes minimal for elements w from
$t_\lambda W$
.
Lemma 7.1 We have the following property:

Proof For
$w=t_{\lambda }u\in t_{\lambda }W$
,
$\beta =\pm \alpha +k\delta>0$
(so
$k\ge 0$
), we have

If
$\langle \lambda ,\alpha \rangle \le 0$
, then
$w^{-1}(\beta )<0$
implies
$\beta =\alpha +k\delta $
, and moreover,
$k\in [0,\ell _\alpha (w)-1]$
(since
$\ell _\alpha (w)=|Inv_\alpha (w)|$
). So we have

and the minimal value is
$-\langle \lambda ,\alpha \rangle $
.
If
$\left <\lambda ,\alpha \right>> 0$
, then
$w^{-1}(\beta )<0$
if and only if
$\beta =-\alpha +k\delta $
and
$k\in [1,\ell _{\alpha }(w)]$
, in which case we have

The minimal value is
$\langle \lambda ,\alpha \rangle -1$
.
Lemma 7.2 Let
$\alpha \in \Phi ^+$
,
$\lambda \in Q^{\vee }$
, and
$k\in [0,\ell _\alpha (w_\lambda )]$
.
If
$\langle \lambda , \alpha \rangle \le 0$
, then
$w_\lambda>w_{\lambda +k\alpha ^{\vee }}$
. If
$\langle \lambda ,\alpha \rangle>0$
, then
$w_\lambda>w_{\lambda -k\alpha ^{\vee }}$
.
Proof If
$\left <\lambda ,\alpha \right>\leq 0$
, consider
$w\in t_\lambda W$
such that
$w=t_{\lambda }u$
with
$u^{-1}(\alpha )<0$
. We have

so
$w>s_\alpha w$
. From (7.2), we get
$\ell _{\alpha }(w)=\ell _{\alpha }(w_{\lambda })+1=-\left <\lambda ,\alpha \right>+1$
.
Since
$1\leq k\leq \ell _{\alpha }(w_{\lambda })=-\left <\lambda ,\alpha \right>$
, we get

which implies

Therefore,
$w>t_{k\alpha ^{\vee } }w$
.
Since
$w\in t_{\lambda }W$
and
$t_{k\alpha ^{\vee }}w\in t_{\lambda +k\alpha ^{\vee }}W$
, we get
$w_{\lambda }>w_{\lambda +k\alpha ^{\vee }}$
.
If
$\left <\lambda ,\alpha \right>> 0$
, consider
$w=t_{\lambda }u$
with
$u^{-1}(\alpha )>0$
. We have

so
$w>s_{-\alpha +\delta }w$
. From (7.3), we have
$\ell _{\alpha }(w)=\langle \lambda ,\alpha \rangle =\ell _{\alpha }(w_{\lambda })+1. $
Since
$1\leq k\leq \ell _{\alpha }(w)-1=\ell _{\alpha }(w_{\lambda })=\left <\lambda ,\alpha \right>-1$
, we get

So
$w>t_{-k\alpha ^{\vee }}w$
. Since
$w\in t_{\lambda }W$
and
$t_{k\alpha ^{\vee }}w\in t_{\lambda +k\alpha ^{\vee }}W$
, we have
$w_{\lambda }>w_{\lambda +k\alpha ^{\vee }}$
.
Lemma 7.3 If
$\langle \lambda ,\alpha \rangle =0$
, then we have the following sequence:

If
$\langle \lambda ,\alpha \rangle =1$
, then we have the following sequence:

The lengths
$\ell _\alpha $
are given by
$(0,1,2,3,4,\ldots )$
.
Proof We only prove the case when
$\langle \lambda ,\alpha \rangle =0$
. Consider
$\lambda +k\alpha ^{\vee }$
and
$\lambda -k\alpha ^{\vee }$
with
$k> 0$
, then
$\langle \lambda -k\alpha ^{\vee },\alpha \rangle =-2k<0$
, and
$2k=\ell _\alpha (w_{\lambda -k\alpha ^{\vee }})$
, so by Lemma 7.2,

Finally, consider
${\lambda -k\alpha ^{\vee }}$
and
$\lambda +(k+1)\alpha ^{\vee }$
,
$k\ge 0$
, then
$\langle \lambda +(k+1)\alpha ^{\vee },\alpha \rangle =2(k+1)\ge 2$
, and
$\ell _\alpha (w_{\lambda +(k+1)\alpha ^{\vee }})=2k+1$
, so by Lemma 7.2,
