Symmetric functions and Springer representations111MSC2010: 14N15,20G44
Abstract
The characters of the (total) Springer representations are identified with the Green functions by Kazhdan [Israel J. Math. 28 (1977)], and the latter are identified with Hall-Littlewood’s -functions by Green [Trans. Amer. Math. Soc. (1955)]. In this paper, we present a purely algebraic proof that the (total) Springer representations of are -orthogonal to each other, and show that it is compatible with the natural categorification of the ring of symmetric functions.
Dedicated to the memory of Tonny Albert Springer
Introduction
Let be a connected reductive algebraic group over an algebraically closed field with a Borel subgroup . Let be the Weyl groups of , and let denote the variety of nilpotent elements. The cohomology of a fiber of the Springer resolution
affords a representation of . This is widely recognized as the Springer representation [24], and it is proved to be an essential tool in representation theory of finite and -adic Chevalley groups [16, 13, 17, 18, 12]. Here and below, we understand that the Springer representation refers to the total cohomology of a Springer fiber instead of the top cohomology, commonly seen in the literature.
In [10], we found a module-theoretic realization of Springer representations that is axiomatized as Kostka systems. For , it takes the following form: Let
be a graded ring obtained by the smash product of the symmetric group and a polynomial algebra such that and (). Let be the category of finitely generated graded -modules. Let , , and denote the graded versions of , , and , respectively. The set of simple graded -modules is parametrized by (up to grading shift), and is denoted as . We have a projective cover as graded -modules.
Theorem A.
For each , we have two modules and in with the following properties:
-
1.
We have a sequence of -module surjections , where the first map is obtained by annihilating all graded Jordan-Hölder components such that with respect to the dominance order on ;
-
2.
The graded ring is a polynomial ring. The unique graded quotient yields ;
-
3.
We have the following -orthogonality:
Remark B.
Theorem A follows from works of many people ([9, 8, 27, 15, 14, 2, 5]) in several different ways as well as an exact account ([10, 11]) that works for an arbitrary . All of these proofs utilize some structures (geometry, cells, or affine Lie algebras) that is hard to see in the category of graded -modules.
The main goal of this paper is to give a new proof of Theorem A based on a detailed analysis of due to Garsia-Procesi [6] and some algebraic results from [14, 10]. This completes author’s attempt [10, Appendix A] to give a proof of Theorem A inside the category of graded -modules.
As a byproduct, we obtain an interesting consequence: We call (resp. ) to be -filtered (resp. -filtered) if admits a decreasing separable filtration (resp. finite filtration) whose associated graded is isomorphic to the direct sum of (resp. direct sum of ) up to grading shifts.
Theorem C ( Theorem 2.37).
The induction of graded -modules sends the external tensor product of and a -filtered module to a -filtered module. Dually, the restriction of graded -modules sends a -filtered module of to a -filtered module of .
Recall that the graded modules
are Hopf algebras by Zelevinsky [28], that is identified with the ring of symmetric functions up to scalar extensions (Theorem 1.1). In particular, this ring is equipped with four bases , and , usually referred to as the Schur functions, the Hall-Littlewood -functions, the Hall-Littlewood -functions, and the big Schur functions, respectively ([19]). We exhibit a natural character identification (that we call the twisted Frobenius characteristic)
(0.1) |
that intertwines the products with inductions, and the coproducts with restrictions. (The complete symmetric functions and the elementary symmetric functions are expanded positively by the Schur functions, and hence corresponds to a direct sum of projective modules in this table).
Under this identification, Theorem C implies that the multiplication of a Schur function in exhibits positivity with respect to the Hall-Littlewood functions (Corollary 2.39). In addition, we deduce a homological interpretation of skew Hall-Littlewood functions (Corollary 2.40).
In a sense, our exposition here can be seen as a direct approach to an algebraic avatar of the Springer correspondence. We note that interpreting sheaves appearing in the Springer correspondence as constructible functions produces totally different algebraic avatar of the Springer correspondence via Hall algebras (as pursued in Shimoji-Yanagida [22]). Although our Hopf algebra structure is closely related to the Heisenberg categorification (cf. [1]), the author was not able to find a result of this kind in the literature. Nevertheless, he plans to write a follow-up paper that covers the relation with the Heisenberg categorification in an occasion.
Finally, the author was very grateful to find related [25] during the preparation of this paper.
1 Preliminaries
A vector space is always a -vector space, and a graded vector space refers to a -graded vector space whose graded pieces are finite-dimensional and its grading is bounded from the below. Tensor products are taken over unless stated otherwise. We define the graded dimension of a graded vector space as
In case , we set , where for each . We set for each .
For a -algebra , let denote the category of finitely generated left -modules. If is a graded algebra in the sense that and (), then we denote by the category of finitely generated graded -modules. We also have a full subcategory of consisting of finite-dimensional modules.
For a graded algebra , the category admits an autoequivalence for each such that is sent to , where . For , we set
In particular, and are graded vector spaces if for each . Moreover, consists of graded -module homomorphisms that raise the degree by .
For , the head of (that we denote by ) is the maximal semisimple graded quotient of , and the socle of (that we denote by ) is the maximal semisimple graded submodule of .
For a decreasing filtration
of graded vector spaces, we define its -th associated graded piece as (). We call such a filtration separable if .
For an exact category , let denote its Grothendieck group. For , we have its class . In case admits the grading shift functor (), an element () defines the direct sum
We may represent a number that is not important by .
1.1 Partitions and the ring of symmetric functions
We employ [19] as the general reference about partitions and symmetric functions. We briefly recall some key notions there. The set of partitions is denoted by , and the set of partitions of is denoted by . Each of is equipped with a partial order such that is the largest element. We extend the order to the whole by declaring that elements of and are comparable only if . Let be the multiplicity of , let be the partition length, and let be the partition size of . The conjugate partition of is denoted by . We set
For and , let be the partition of obtained by rearranging , and for , we set be the partition of obtained by rearranging . We set
Let be the ring of symmetric functions with their coefficients in . Let be its scalar extension to . We have direct sum decompositions and into the graded components. The ring is equipped with four distinguished bases
called (the sets of) complete symmetric functions, Schur functions, elementary symmetric functions, and monomial symmetric functions, respectively. We have equalities
We have a symmetric inner product on such that
The ring has a structure of a Hopf algebra with the coproduct satisfying
and the antipode satisfying
The antipode preserves the inner product .
1.2 Zelevinsky’s picture for symmetric groups
For a (not necessarily non-increasing) sequence such that , we define the subgroup
We usually omit in . Each defines an irreducible representation of of . We normalize such that
For , we have induction/restriction functors
where the latter is the natural restriction. They induce corresponding maps between the Grothendieck groups that we denote by the same letter.
Theorem 1.1 (Zelevinsky [28]).
We have a -module isomorphism
with the following properties: For and , we have
In particular, we have
1.3 The algebra and its basic properties
We follow [10, §2] here. We set
where acts on the ring by
We usually denote in place of , and in place of . The ring acquires the structure of a graded ring by
The grading of the ring is non-negative, and the positive degree part defines a graded ideal such that . In particular, each can be understood to be a graded -module concentrated in degree .
The assignments () and () define an isomorphism . Therefore, if , then acquires the structure of a graded -module. We have for each as is a real reflection group.
For each , we have an idempotent such that . We set .
Proposition 1.2 (see [10] §2).
The modules is the complete collection of simple objects in . In addition, is the projective cover of in for each .
We define
For each , we set
In case the specialization of makes sense, we denote it by .
Lemma 1.3 (see [10] §2).
For each , we have
Proof.
Immediate from the definition. ∎
For , we consider the subalgebra
We have induction/restriction functors
Since is free of rank over , we find that the both functors are exact, and preserves finite-dimensionality of the modules. We sometimes omit the functor from notation in case there are no possible confusion.
We consider the category . We define
Lemma 1.4.
We embed into by regarding as a semisimple graded -module concentrated in degree for each . Then, we have
on . In particular, can be understood as a Hopf subalgebra of by extending the scalar in Theorem 1.1.
The following three theorems are quite well-known to experts.
Theorem 1.5 (Frobenius-Nakayama reciprocity).
For and , it holds
Proof.
This follows from the fact that is a free -module by the classical Frobenius reciprocity as sends a projective resolution of to a projective resolution of . ∎
Theorem 1.6.
For , it holds
Proof.
We borrow terminology from [7, §2.2]. We have natural isomorphism
Since the derived functors of the both sides (defined in an appropriate ambient categories) are -functors in each variables, it suffices to see that they are universal -functors. By approximating by its injective envelope (and hence by its projective cover), we find that the both sides are effacable on the second variables. Thus, they must coincide by [7, 2.2.1 Proposition]. ∎
Theorem 1.7.
The global dimension of is finite. In particular, every admits a graded projective resolution of finite length.
Proof.
See McConnell-Robson-Small [21, 7.5.6]. ∎
We have a -bilinear inner product on prolonging
Lemma 1.8.
The pairing is a well-defined symmetric form on .
Proof.
Since the Euler-Poincaré form respects the short exact sequences, the form must be additive with respect to the both variables.
By the arrangement of duals in the definition of , we find that replacing with and replacing with both result in multiplying (). As the category has finite direct sums, we conclude that must be -bilinear.
2 Main results
Keep the setting of the previous section.
Definition 2.1.
Fix . A -filtration (resp. -filtration) of is a decreasing separable filtration
of graded -modules (resp. graded -modules) such that
for each . In case admits a -filtration, then we set
where takes value if the proposition is true, and otherwise.
The following theorem is not new (see Remark 2.4). Nevertheless, the author feels it might worth to report a yet another proof based on Garsia-Procesi [6], that differs significantly from other proofs and is carried out within the category of -modules:
Theorem 2.3.
Let . We have the following:
-
1.
For each , the graded ring is a polynomial ring generated by homogeneous polynomials of positive degrees;
-
2.
The module is free over , and we have
where is the unique graded one-dimensional quotient of ;
-
3.
We have the -orthogonality:
-
4.
Each admits a -filtration, and we have .
Proof.
Postponed to §2.4. ∎
Remark 2.4.
Theorem 2.3 is originally proved in [10, 11] essentially in this form by using the geometry of Springer correspondence (that works for arbitrary Weyl groups with arbitrary cuspidal data). Theorem 2.3 also follows from results of Haiman [9, 8] that employ the geometry of Hilbert schemes of points on . We also have two algebraic proofs of Theorem 2.3, one is to use a detailed study of two-sided cells of affine Hecke algebras by Xi [27] together with König-Xi [15] and Kleshchev [14], and another is an analogous result for affine Lie algebras (Chari-Ion [2]) together with Feigin-Khoroshkin-Makedonskyi [5].
2.1 Garsia-Procesi’s theorem
For each and , let be the -th elementary symmetric function with respect to the variables . For , we set
We set
Let be the ideal generated by (originally introduced in [26]).
Definition 2.5.
We set , and call it the Garsia-Procesi module.
Lemma 2.6 ([6] §3).
The algebra admits a structure of graded -module generated by . In addition, .
Proof.
Since is the quotient of , it suffices to see that the ideal is graded and -stable. Since consists of homogeneous polynomials and it is stable under the -action, we conclude the first assertion. For the second assertion, it suffices to notice that contains all the elementary symmetric polynomials in , and hence contains all the positive degree part of . ∎
Theorem 2.7 (Garsia-Procesi [6] §1).
Let . The -module admits a decreasing filtration
(2.1) |
such that for . In addition, this filtration respects the -action, and hence can be regarded as an -module filtration.
Theorem 2.8 ([6] Theorem 3.1 and Theorem 3.2).
Let . It holds:
-
1.
We have ;
-
2.
We have a -module isomorphism .
In particular, we have only if .
In view of [19, III (2.1)], we have the Hall-Littlewood - and - functions in indexed by , that we denote by and , respectively (we changed notation of -functions to in order to avoid confusion with projective modules). They satisfy the following relation:
(2.2) |
We also have the big Schur function ([19, III (4.6)])
where are the raising operators.
Theorem 2.9 ([6] §5, particularly (5.24)).
For each , the polynomial
is the Hall-Littlewood’s -function.
Theorem 2.10 ([19] III (4.9)).
There exists a -linear bilinear form on referred to as the Hall inner product characterized as
(2.3) |
for each .
Lemma 2.11.
For each , we have .
Proof.
Proposition 2.12 ([10] Theorem A.4 and Corollary A.3).
We have
For each , the head of is , and the socle of is .
Proof.
Proposition 2.13 (De Concini-Procesi [4], Tanisaki [26]).
We have an isomorphism as graded -modules.
Proof.
By Lemma 2.11, is a graded -module such that and if and . Thus, we obtain a map of graded -modules. This map is injective as they share as their socles.
We prove that is an equality for every by induction on . The case is clear as the both are . Thanks to Theorem 2.7 and the induction hypothesis, we deduce that a (graded) direct summand of the head of as -module must be of the shape for and . The module arises as the restriction of a (graded) -module () such that for . In case , then forces .
From this, it is enough to assume to conclude that does not yield a non-zero module of . By Theorem 2.8 2), we can assume . Hence, is obtained from by moving one box in the Young diagram to some strictly larger entries.
In case is not the shape , there exists such that for every . It follows that contains a -module that is not in the head of as -modules. Thus, this case does not occur.
In case is of the shape , then we have and . In this case, we have . In particular, grading shifts of appears in the filtration of afforded by Theorem 2.7 only once, and its head is a part of by counting the degree. Therefore, contributes zero in .
From these, we conclude that by induction hypothesis. This forces , and the induction proceeds. ∎
2.2 Identification of the forms
Consider the twisted (graded) Frobenius characteristic map
(2.4) |
By Theorem 2.9, we have
(2.5) |
Lemma 2.14.
For , we have
Proof.
This is a straight-forward consequence of Lemma 1.4. The detail is left to the reader. ∎
Proposition 2.15.
We have
In particular, we have
(2.6) |
Remark 2.16.
Proof of Proposition 2.15.
The equations in Theorem 2.10, that are equivalent to the Cauchy identity [19, (4.4)], are special cases of [23, Corollary 4.6]. It is further transformed into the main matrix equality of the so-called Lusztig-Shoji algorithm in [23, Theorem 5.4]. The latter is interpreted as the orthogonality relation with respect to in [10, Theorem 2.10]. In particular, Kostka polynomials defined in [19] and [23] are the same (for symmetric groups and the order on ). This implies the first equality in view of (2.5). The second equality is read-off from the relation between and . The last assertion follows as forms a -basis of , and the Hall inner product is non-degenerate. ∎
Proposition 2.17.
For each , we have .
Proof.
Corollary 2.18.
For each , we have
Proof.
Corollary 2.19.
For each , we have
2.3 An -estimate
Lemma 2.20.
For each , the -module contains a unique non-zero -fixed vector up to scalar.
Proof.
This follows from Theorem 2.8 2) and the Frobenius reciprocity. ∎
For each , we set
(2.7) |
Lemma 2.21.
We have , where is the quotient of the polynomial ring by the submodule generated by degree one part that is complementary to as -modules.
Proof.
We have . Its degree one part is as -modules, and quotient out by yields a polynomial ring generated by the image of . ∎
Lemma 2.22.
Let . We have a unique graded -module map of degree up to scalar.
Proof.
We have , in which appears without multiplicity. All the -modules appearing in are trivial. It follows that if and only if . The latter implies (Theorem 2.8). Therefore, a -module map extends uniquely to a graded -module map by the definition of . ∎
In the setting of Lemma 2.22, we set
For each , we have an endomorphism on extending
obtained by the multiplication of . Consider the group
Lemma 2.23.
The group yields automorphisms of as -modules.
Proof.
The group permutes s in (2.7) in such a way the size of the factors (i.e. the values of ) are invariant. These are -module endomorphisms, and hence inherits these endomorphisms as required. ∎
Let denote the subring of generated by . The action of permutes and such that . Thus, acts on as automorphisms. The invariant part is a polynomial ring.
Lemma 2.24.
For each , we have
where is the multiplicity one copy as -modules.
Proof.
By construction, is a direct sum of (grading shifts of) copies of as a -module. We have . The action of preserves the -isotypic part. As the action of sends to all the contributions of in , we conclude the assertion. ∎
Proposition 2.25.
For each , we have
Proof.
Since has as its unique simple graded quotient, is determined by the image of . In addition, is stable under the action of as is. Therefore, Lemma 2.24 implies . Thus, we have the inequality in the assertion by
(see Corollary 2.18 for the second equality). We have an identification
(2.8) |
where . The actions of on the first term of (2.8) are induced by the multiplication of . Hence, the action of on the first two terms of (2.8) are realized by the multiplication of . Thus, the inequality must be in fact an equality and . ∎
Let us consider the image of the center in and by and , respectively.
Lemma 2.26.
For each , we have a quotient map
as an algebra that induces a surjection . In addition, is precisely the image of in .
Proof.
Lemma 2.27.
For each , the algebra is a finitely generated module over .
Proof.
Since we have a surjection , it suffices to see that is a finitely generated module over . We have
The RHS is a finitely generated module over as required. ∎
For two power series with integer coefficients
we say if we have for every . We say if
(2.9) |
Theorem 2.28 ([20]).
Let be a finitely generated graded integral algebra with , and let be its proper graded quotient algebra. For a finitely generated graded -module , we have
Proof.
This follows form [20, Theorem 13.4] if we take into account the Krull dimension inequality , and the completion with respect to the grading makes and into local rings. ∎
Lemma 2.29.
For each and an algebra quotient , the actions of on and have joint eigenvalues of shape
(2.10) |
up to -permutation.
Proof.
Theorem 2.30.
For each , we have
Proof.
We set and during this proof. The specialization with respect to a maximal ideal decomposes into the generalized eigenspaces with respect to , whose set of joint eigenvalues in have multiplicities that constitute a partition of . We have
(2.11) |
by the definition of and the fact that has semi-simple representation theory. Being the cyclic -module generator, we have .
We can choose a non-zero generalized eigenspace
of that can be regarded as an (ungraded) -module. We choose
(2.12) |
as -modules with partitions of , respectively. Since each piece of the external tensor products of (2.12) have distinct -eigenvalues, we deduce
(2.13) |
By the Littlewood-Richardson rule, the smallest label (with respect to ) of -module that appears in the LHS of (2.13) is attained by such that
For an appropriate choice in (2.12), we attain by Lemma 1.3. It follows that defines a division of entries of into small groups. In view of (2.10), the maximal ideal is the pullback of a maximal ideal of . In other words, we find that shares with the same support as in .
We define graded -modules () as:
We show that each is supported in a proper subset of . Equivalently, we show that general specializations of and with respect to (2.10) are the same. In view of the above construction of the partitions and , we have necessarily and for each as otherwise smaller partitions arise. By Lemma 2.21, a thickening of (2.13) as (ungraded) -modules must be achieved by the actions of
(2.14) |
As these are contained in the action of , we conclude that general specializations of and are the same.
Proposition 2.31.
For each , the module admits a decreasing separable filtration whose associated graded is the direct sum of grading shifts of .
Proof.
Consider the submodule generated by the unique copy . In view of Lemma 2.24, we find
(2.15) |
Consequently, we have .
Let and be the specializations of and with respect to a maximal ideal of such that the joint eigenvalues in Lemma 2.29 are distinct. Let be a joint -eigenspace of or , that is a -module. The distinct eigenvalue condition implies
(2.16) |
The -module appears in or only if . Applying the Littlewood-Richardson rule to the middle term of (2.16), we deduce . In particular, we have . By the semi-continuity of the specializations, we deduce
(2.17) |
From this, we conclude . Thus, the torsion free -action on yields the assertion. ∎
Corollary 2.32.
Keep the setting of Proposition 2.31. We have .
Corollary 2.33.
For each and an -module proper quotient of , we have .
Proof.
We borrow the setting of the proof of Proposition 2.31. Since , we find for some . As all the copies of and in are obtained by the -action from unique copies at degree zero, we find such that
This forces to be zero in . Therefore, we have
This implies the assertion. ∎
Corollary 2.34.
Let . Assume that is a graded -module generated by the subspace
Then, we have .
Proof.
We have a surjection
Consider the quotient of by . Let be the map induced from . Let us choose a maximal subset such that injects into by . We take the quotient of by this image. Then, the image of () in under the induced map must be a proper quotient of .
2.4 Proof of Theorem 2.3
We assume the assertion for all and prove the assertion for . We fix and set
For each and , Theorem 1.5 implies
(2.18) |
Since is projective as -modules, Theorem 2.7 implies that
(2.19) |
by the short exact sequences associated to (2.1). In other word, we have
and it is nonzero if and only if for some . This is equivalent to for some . We set .
Note that , and hence every satisfies . In view of Lemma 1.3, we further deduce . Therefore, the image of the map
obtained by taking the sum of all the maps of satisfies
-
•
is the direct sum of ();
-
•
.
We consider an -submodule generated by the preimage of (considered as the direct sum of grading shifts of ), that we denote by . Although the module might depend on the choice of a lift, the number of its -module generators is unambiguously determined.
We have for by inspection. In particular, is a totally ordered set with respect to . Moreover, is generated by as an -module, and an irreducible constituent of is of the form for by the Littlewood-Richardson rule. As a consequence, we find that . For each , we set . We have for and .
By the Littlewood-Richardson rule and Lemma 1.3, we find that
(2.20) |
Claim A.
We have for .
Proof.
Assume to the contrary to deduce contradiction. We have some and such that . Here we have by (2.20). By rearranging , we assume that is the minimal number with this property. In particular, we have
(2.21) |
This in turn implies that for for every . By rearranging if necessary, we can assume that the -submodule generated by -isotypic components such that satisfies and the value is minimum among all . Then, the lift of to is uniquely determined as graded -module. It follows that the maximal quotient of (and hence also a quotient of ) such that is finite-dimensional (as the grading must be bounded) and if . By Proposition 2.12 and Theorem 1.6, we find
by a repeated applications of the short exact sequences. In particular, the non-zero map prolongs to , and hence it gives rise to a map . By (2.21), we additionally have
Therefore, we deduce a non-zero map from our assumption that does not come from the generator set of for every . This is a contradiction, and hence we conclude the result. ∎
We return to the proof of Theorem 2.3. Note that Claim A guarantees that () is defined unambiguously as all the possible generating -isotypical components of (i.e. for ) must belong to . In view of the above and Corollary 2.19, we deduce
(2.22) |
This expansion exhibits positivity (as a formal power series in ).
Claim B.
For each , the module is the direct sum of grading shifts of .
Proof.
We assume that the assertion holds for all the larger (or ), and (and hence ). We apply Claim A, and compare Lemma 1.3 and Theorem 2.9 with (2.22) to find
Since must be the sum of for () by the induction hypothesis and the above formulae, Theorem 2.9 implies
It follows that admits a surjection from direct sum of with its multiplicity (as this latter number counts the number of generators of ). Applying Corollary 2.34, we conclude that is the direct sum of grading shifts of . These proceed the induction, and we conclude the result. ∎
Claim C.
Let us enumerate as . We have a finite increasing filtration
as -modules such that each is isomorphic to the direct sum of grading shifts of . In addition, each contains a copy of as its -module direct summand.
Proof.
The first part is a rearrangement of Claim B.
Claim D.
For each , we have
(2.23) |
Proof.
We return to the proof of Theorem 2.3. All elements of appear as for suitable and . By rearranging if necessary, we conclude (2.23) for every . A repeated use of short exact sequences decomposes into by starting from (see Lemma 1.3). Substituting these to the second factor of (2.23), we deduce
This implies for all . Therefore, Proposition 2.25 and Proposition 2.31 imply Theorem 2.3 1) and 2) for , and (2.23) is Theorem 2.3 3) for .
In view of the above arguments, we find that each admits a -filtration. Since preserves projectivity, we deduce that admits a filtration by by the induction hypothesis. Therefore, admits a -filtration. Since each is generated by its simple head, applying an idempotent does not separate them out non-trivially. Therefore, we conclude that each projective module of also admits a -filtration. Given this and Theorem 2.3 2) and 3), the latter assertion of Theorem 2.3 4) is standard (see e.g. [11, Corollary 3.12]). This is Theorem 2.3 4) for .
This completes the proof of Theorem 2.3.
2.5 Applications of Theorem 2.3
Note that is a Noetherian ring as a finitely generated -module is also finitely generated by . The global dimension of is finite (Theorem 1.7). We have by inspection.
We introduce a total order on that refines and set for each . The two sided ideals satisfies if . By Lemma 1.3, we deduce that
is a surjection. By Proposition 2.12 and Theorem 2.3 2), we further deduce
Theorem 2.3 1) implies that is a graded polynomial ring for each . In conjunction with Theorem 2.3 2), we find that
is a free module over for each . In particular, that the graded algebra is an affine quasi-hereditary in the sense of [14, Introduction] with and .
Theorem 2.35 ([14] Theorem 7.21 and Lemma 7.22).
A module admits a -filtration if and only if
A module admits a -filtration if and only if
Corollary 2.36 ([14] §7, particularly Lemma 7.5).
Let . If admits a -filtration, then the multiplicity space of in is given by
If the module admits a -filtration, then the multiplicity space of of in is given by
Theorem 2.37.
Fix , and . Let . We have the following:
-
1.
(Garsia-Procesi [6]) The module admits a -filtration;
-
2.
The module admits a -filtration.
Remark 2.38.
Proof of Theorem 2.37.
We prove the first assertion. By the second part of Theorem 2.35, it suffices to check the -vanishing with respect to () as a module over (equivalently, we can check the -vanishing with respect to as a module of ; see below). In particular, we do not need to mind the first factor as the -action is granted by construction. Therefore, the first assertion is just a -times repeated application of Theorem 2.7.
We prove the second assertion. For each and , we have
(2.27) |
by Theorem 1.5. Applying Theorem 2.7 to as many as -times, we find that the restriction of to admits a filtration whose associated graded is the direct sum of grading shifts of . Since is free over a polynomial ring of -variables, we have
Thus, we derive a natural isomorphism
(2.28) |
By Theorem 2.3 3) and Theorem 2.7, the RHS of (2.28) is zero. By the first part of Theorem 2.35, we conclude the second assertion. ∎
Corollary 2.39.
Let . We have the following:
-
1.
We have ;
-
2.
We have . In case , it is the Pieri rule.
Proof.
Corollary 2.40.
The skew Hall-Littlewood -function expands positively with respect to the big Schur function. In addition, we have a graded -module defined as
such that its image under is .
Proof.
Let . The Hall-Littlewood -polynomial corresponds to the module by Theorem 2.9. Therefore, its restriction admits a -filtration. In particular, we have
Here corresponds to in by Lemma 2.14. Taking [Mac98, III (5.2)], (2.3), and (2.5) into account, we conclude that
This is the first assertion by , read off from (2.4). In view of Theorem 2.37 1) and Corollary 2.36, we conclude the second assertion. ∎
Acknowledgement: The author presented an intensive lecture course based on the contents of this paper at Nagoya University on Fall 2020. The author thanks the hospitality of Shintarou Yanagida and Nagoya University. He also thanks Kota Murakami for pointing out some typos. This research was supported in part by JSPS KAKENHI Grant Number JP19H01782.
References
- [1] Sabin Cautis and Anthony M. Licata “Heisenberg categorification and Hilbert schemes” In Duke Mathematical Journal 161.13, 2012, pp. 2469–2547 DOI: 10.1215/00127094-1812726
- [2] Vyjayanthi Chari and Bogdan Ion “BGG reciprocity for current algebras” In Compos. Math. 151.7, 2015, pp. 1265–1287 DOI: 10.1112/S0010437X14007908
- [3] Claude Chevalley “Invariants of finite groups generated by reflections” In Amer. J. Math. 77, 1955, pp. 778–782
- [4] Corrado De Concini and Claudio Procesi “Symmetric functions, conjugacy classes and the flag variety” In Invent. Math. 64, 1981, pp. 203–219
- [5] Evgeny Feigin, Anton Khoroshkin and Ievgen Makedonskyi “Peter-Weyl, Howe and Schur-Weyl theorems for current groups”, arXiv:1906.03290
- [6] A. M. Garsia and C. Procesi “On certain graded -modules and the -Kostka polynomials” In Adv. Math. 94.1, 1992, pp. 82–138
- [7] Alexander Grothendieck “Sur quelques points d’algèbre homologique” In Tohoku Math. J. (2) 9, 1957, pp. 119–221
- [8] Mark Haiman “Combinatorics, symmetric functions, and Hilbert schemes” In Current developments in mathematics, 2002 Int. Press, Somerville, MA, 2003, pp. 39–111
- [9] Mark Haiman “Hilbert schemes, polygraphs and the Macdonald positivity conjecture” In J. Amer. Math. Soc. 14.941–1006, 2001
- [10] Syu Kato “A homological study of Green polynomials” In Ann. Sci. Éc. Norm. Supér. (4) 48.5, 2015, pp. 1035–1074
- [11] Syu Kato “An algebraic study of extension algebras” arXiv:1207.4640 In Amer. J. Math. 139.3, 2017, pp. 567–615
- [12] Syu Kato “An exotic Deligne-Langlands correspondence for symplectic groups” In Duke Math. J. 148.2, 2009, pp. 305–371
- [13] David Kazhdan and George Lusztig “Proof of the Deligne-Langlands conjecture for Hecke algebras” In Invent. Math. 87.1, 1987, pp. 153–215
- [14] Alexander S. Kleshchev “Affine highest weight categories and affine quasihereditary algebras” In Proc. Lond. Math. Soc. (3) 110.4, 2015, pp. 841–882 DOI: 10.1112/plms/pdv004
- [15] Steffen König and Changchang Xi “Affine cellular algebras” In Adv. Math. 229.1, 2012, pp. 139–182
- [16] G. Lusztig “Intersection cohomology complexes on a reductive group” In Invent. Math. 75.2, 1984, pp. 205–272
- [17] George Lusztig “Cuspidal local systems and graded Hecke algebras. I” In Inst. Hautes Études Sci. Publ. Math. 67, 1988, pp. 145–202
- [18] George Lusztig “Green functions and character sheaves” In Ann. of Math. (2) 131, 1990, pp. 355–408
- [19] I. G. Macdonald “Symmetric functions and Hall polynomials” With contributions by A. Zelevinsky, Oxford Science Publications, Oxford Mathematical Monographs The Clarendon Press, Oxford University Press, New York, 1995, pp. x+475
- [20] Hideyuki Matsumura “Commutative ring theory” Translated from the Japanese by M. Reid 8, Cambridge Studies in Advanced Mathematics Cambridge University Press, Cambridge, 1986, pp. xiv+320
- [21] John C. McConnell and James C. Robson “Noncommutative Noetherian rings.” 30, Graduate Studies in Math. AMS, 2001
- [22] Ryosuke Shimoji and Shintarou Yanagida “A study of symmetric functions via derived Hall algebra” In Communications in Algebras to appear, 2020
- [23] Toshiaki Shoji “Green functions associated to complex reflection groups” In J. Algebra 245.2, 2001, pp. 650–694
- [24] T. A. Springer “Trigonometric sums, Green functions of finite groups and representations of Weyl groups” In Invent. Math. 36, 1976, pp. 173–207
- [25] T. A. Springer and A. V. Zelevinsky “Characters of and Hopf algebras” In J. London Math. Soc. (2) 30.1, 1984, pp. 27–43
- [26] Toshiyuki Tanisaki “Defining ideals of the closures of the conjugaty classes and representations of the Weyl groups” In Tohoku Mathematical Journal 34, 1982, pp. 575–585
- [27] Nanhua Xi “The based ring of two-sided cells of affine Weyl groups of type ” In Mem. Amer. Math. Soc. 157.749, 2002, pp. xiv+95
- [28] Andrey V. Zelevinsky “Representations of finite classical groups” 869, Lecture Notes in Mathematics Springer-Verlag, Berlin-New York, 1981