Supersymmetric Shimura operators and interpolation polynomials
Abstract.
The Shimura operators are a certain distinguished basis for invariant differential operators on a Hermitian symmetric space. Answering a question of Shimura, Sahi–Zhang showed that the Harish-Chandra images of these operators are specializations of certain -symmetric interpolation polynomials that were defined by Okounkov.
We consider the analogs of Shimura operators for the Hermitian symmetric superpair where and and we prove their Harish-Chandra images are specializations of certain -supersymmetric interpolation polynomials introduced by Sergeev–Veselov.
1. Introduction
The Harish-Chandra homomorphism for a symmetric space gives an explicit isomorphism of the algebra of invariant differential operators with a certain polynomial algebra, and knowing the Harish-Chandra image of an operator allows one to determine its spectrum. In [Shi90] Shimura introduced a basis for the algebra of invariant differential operators on a Hermitian symmetric space, and formulated the problem of determining their eigenvalues. Shimura’s problem was solved in [SZ19] where it was shown that the Harish-Chandra images of Shimura operators are specializations of certain interpolation polynomials of Type BC introduced by Okounkov [Oko98].
Our main result, Theorem A below, solves the analogous problem for the Hermitian symmetric superpair with and .
Let be the universal enveloping algebra of , then the algebra of invariant differential operators is the quotient of -invariants . In Section 3.3 we describe a basis of indexed by the set of -hook partitions, which are partitions that satisfy . These are the super analogs of Shimura operators, and their definition involves the Cartan decomposition , and the multiplicity-free -decompositions of and , which have -summands and , naturally indexed by [CW01].
The Harish-Chandra homomorphism (Section 2) is an algebra map , where is an even Cartan subspace of and is the algebra of polynomials on . For our pair the image of can be identified with the ring of even supersymmetric polynomials in variables. In Section 6 we introduce a basis of . These are suitable specializations of the supersymmetric interpolation polynomials of Type BC introduced by Sergeev–Veselov [SV09], which are characterized up to multiple by certain vanishing properties.
Theorem A.
The Harish-Chandra image of is where is explicitly given in Eq. (55).
We will deduce this from two other results. First, let be the center of , then we have a natural map , and we prove the following result.
Theorem B.
The map is surjective. In particular, there exist such that .
For let be the generalized Verma module for obtained by parabolic induction from of the -summand of (Section 5).
Theorem C.
The central element acts on by unless for all .
The special case of Theorem A was proved previously in [Zhu22], where it was also shown that the general case would follow if one knew that certain irreducible finite dimensional -modules are -spherical [Zhu22, Conjecture 1, Theorem A]. For ordinary Lie algebras, the Cartan–Helgason Theorem [Hel00] provides necessary and sufficient conditions for -sphericity. However its precise analog for Lie superalgebras remains open and seems quite hard. Therefore in this paper we give a different argument for Theorem A, which does not rely on the -sphericity conjecture.
The Shimura problem discussed in this paper is closely related to the Capelli problem studied in [KS93, Sah94] for usual Lie algebras, in [SS16, ASS18], and [SSS20] for Lie superalgebra, and more recently [LSS23, LSS22b, LSS22a] for quantum groups. Thus it is possible that the results of this paper can be generalized to many of the Hermitian symmetric superpairs constructed using Jordan superalgebras, e.g. in [SSS20, Theorem 1.4], and also to the quantum setting, and we hope to address this in future work.
We now briefly discuss the proofs of these results.
The proof of Theorem C is a generalization of an analogous argument in [SZ19] and relies on two key properties of , namely that it is -spherical and has a natural grading by the center of .
For Theorem B let be a Cartan subalgebra of containing , let be the usual Harish-Chandra homomorphism [Mus12], and consider the diagram
(1) |
where is the restriction map induced by the decomposition . We first show that the diagram commutes and then we prove that the image of surjects onto the image of using an explicit description of generators of from [Ste85]. This implies Theorem B.
For Theorem A the main point is to show that satisfies the vanishing properties which characterize . The action of on can be expressed in terms of its Harish-Chandra image and thus Theorem C, implies certain vanishing properties for . Now we use Diagram 1 to deduce the necessary vanishing properties for . This last step requires an additional result that we deduce from [AS15]: consider the set of for which the irreducible quotient of is -spherical, and let be the set of highest -weights of these , then is a Zariski dense subset of .
The structure of the paper is as follows. In Section 2, we review some known results related to the pair . In particular, we discuss the Iwasawa decomposition, the two Harish-Chandra homomorphisms, the Cheng–Wang decomposition of , and Alldridge’s necessary conditions for -sphericity of finite dimensional irreducible -modules. We also present the Sergeev–Veselov supersymmetric interpolation polynomials of Type BC. In Section 3, we define the supersymmetric Shimura operators. In Section 4, we describe the image of as the ring of even supersymmetric polynomials. We prove the surjectivity of in Subsection 4.1 (Proposition 4.1) and Theorem B in Subsection 4.2. In Section 5 we discuss the generalized Verma modules and prove Theorem C, which provides the necessary representation theoretic machinery for the vanishing properties. Finally in Section 6, we reformulate the Type BC interpolation polynomials in our context. Then we prove Theorem A using the results obtained in previous sections.
2. Preliminaries
Throughout the paper, we assume that , the set of natural numbers, includes 0. For a finite dimensional vector space , we identify the polynomial algebra on , , with the symmetric algebra on the dual of . We denote the -vector superspace of superdimension by where the even subspace is and the odd subspace is . A -graded algebra is called a superalgebra. For any superalgebra , we denote its even subspace as and the odd subspace as . For , we write for its parity . The space of linear endomorphisms is naturally -graded and identified with the space of matrices. We denote the Lie superalgebra structure defined on as . In this subsection, we denote the standard diagonal Cartan subalgebra of as , that is, where denotes the matrix with 1 in the -th entry and 0 elsewhere. We also let and be the coordinate functions of for and of for respectively.
2.1. Symmetric superpairs
In this subsection, we let be a symmetric superpair. Specifically, the Cartan decomposition is given by an involution where is the fixed point subalgebra and is the -eigenspace of . We assume the chosen nondegenerate invariant form on is -invariant. Let be a maximal toral subalgebra. Let be the restricted root system of with respect to , and be a positive system. We denote the form on by induced from . As in Subsection 2.2, we say a restricted root is anisotropic if , and isotropic otherwise. We write for the root space of , and let
be the sets of even and odd restricted roots respectively. Set . If , but , we say is indivisible. We define the multiplicity of as . For a positive system , we define the nilpotent subalgebra for as . We denote the restricted Weyl vector for as .
We assume
(2) |
which is called an Iwasawa decomposition for the pair . For an in-depth discussion and recent developments on Iwasawa decomposition, see [She22].
The Poincaré–Birkhoff–Witt theorem applied to Eq. (2) yields the following identity:
Let be the respective projection onto , and define on . The map is called the Harish-Chandra homomorphism associated with (with respect to the choice ). The quotient isomorphism
is called the Harish-Chandra isomorphism associated with . Note .
Next, we introduce Alldridge’s result on spherical representations. Suppose is a -invariant Cartan subalgebra extended from . If is a Borel subalgebra containing , then we denote the irreducible representation of highest weight with respect to as . A -module is said to be -spherical if is non-zero. A non-zero vector in is called a -spherical vector. When the context is clear, we simply say spherical instead of -spherical. Define for anisotropic . We say is high enough if
-
(1)
for any isotropic root ,
-
(2)
and for any odd anisotropic indivisible root .
Then by [AS15, Theorem 2.3, Corollary 2.7], we have the following assertion.
Theorem 2.1.
If where , , and is high enough, then is spherical.
Let be the nilpotent subalgebra for with the corresponding Weyl vector . We may then write down the opposite Iwasawa decomposition
(3) |
From now on, we set
(4) |
The following Lemma is [Zhu22, Lemma 5.1].
Lemma 2.1.
Let be a -module. If , then acts on by a scalar.
A result of the second author [Zhu22, Proposition 4.1] says that any finite dimensional irreducible -spherical -module has a spherical vector unique up to constant. Another result in the same work [Zhu22, Theorem 5.2] relates this scalar with the highest weight restricted on .
Theorem 2.2.
If is spherical then , and acts on by the scalar given by .
2.2. The Harish-Chandra homomorphism for
Let , and be the standard Cartan subalgebra. We denote the root system of with respect to as , and fix a choice of positive system in . Let and be the sums of positive root spaces and negative root spaces respectively. Then is the triangular decomposition of . We denote the Weyl vector for , as , where is the multiplicity of .
We denote the universal enveloping algebra as , and the center of as . Following [Hum78, Mus12], by the Poincaré–Birkhoff–Witt theorem, we have
(5) |
Let be the respective projection onto . On , we define an automorphism so that for all . This automorphism can be equivalently extended from for . The Harish-Chandra isomorphism is then defined as
Denote by the central character afforded by a -module of highest weight . Then
(6) |
The last equality is due to identification between and .
To describe the image of , we first introduce the supersymmetric polynomials. Let and be two sets of independent variables, and set . We write for the polynomial obtained by the substitutions and . A polynomial in is said to be supersymmetric if
-
(1)
is invariant under permutations of and of separately.
-
(2)
is independent of .
Let denote the -algebra of supersymmetric polynomials in . By [Ste85, Theorem 1] (c.f. [Mus12, Theorem 12.4.1]), is generated by the power sums:
(7) |
We record the following standard result about where we identify as standard basis for . Thus for , and for . We refer to [Mus12, Theorem 13.1.1, Theorem 13.4.1], which are based on the original works by Kac, Serganova, and Gorelik [Kac84, Ser99, Gor04].
Theorem 2.3.
The homomorphism is an isomorphism and .
We also use the notation for the algebra of supersymmetric polynomials on when we suppress the choice of coordinates. Then we have:
(8) |
2.3. Cheng–Wang decomposition
Let . We recall a multiplicity-free -module decomposition, known as Howe duality in [CW01], which generalizes Schmid’s decomposition [Sch70, FK90]. It will allow us to define the supersymmetric Shimura operators in Section 3.
We first introduce some notation. A partition is a sequence of non-negative integers with only finitely many non-zero terms and (c.f. [Mac95]). Let denote the size of , the length of , and for which the transpose of . When viewed as the corresponding Young diagram, is the collection of “boxes”
A -hook partition is a partition such that . We define
For , we define a -tuple
(9) |
where for . The last coordinates can be viewed as the lengths of the remaining columns after discarding the first rows of .
Let be any Borel subalgebra of containing a Cartan subalgebra . Then can be described by an -chain, , a sequence consisting of characters where exhaust all the simple roots defining . Therefore,
gives the standard Borel subalgebra , while gives the opposite one, denoted as . For an -tuple , we associate an irreducible -module of highest weight with respect to , denoted as .
Let be a vector superspace and be the supersymmetric algebra on , so has a natural -grading , and explicitly, as vector spaces. The natural action of on gives an action of on , which extends to an action on . We record the following result regarding the -module structure on . See [CW01, Theorem 3.2].
Theorem 2.4.
The supersymmetric algebra as a -module is completely reducible and multiplicity-free. In particular,
2.4. Supersymmetric Polynomials
We devote this subsection to an overview of known results of supersymmetric polynomials, including the Type BC interpolation polynomials introduced and studied by Sergeev and Veselov in [SV09]. These are super analog for the now-classic Okounkov polynomials [Oko98, OO06]. We specialize them to (as in Theorem A) in Section 6.
Let be the standard basis for , and and be the coordinate functions of and , for . Let and be two parameters. Following [SV09], we assume that (generic), and set by . We also set
As in [SV09, Section 6], we let where
(10) |
This is in fact the deformed Weyl vector calculated with the deformed root multiplicities in [SV09]. We identify with , and define as the subalgebra of polynomials which: (1) are symmetric separately in shifted variables and , and invariant under their sign changes; (2) satisfy the condition
on the hyperplane . If we equip with an inner product defined by
Then Condition (2) is equivalent to: (2’) for any on the hyperplane
For , we set and . Equivalently, (see Eq. (9)). Then [SV09, Proposition 6.3] says the following.
Theorem 2.5.
For each , there exists a unique polynomial of degree such that
and satisfies the normalization condition . Moreover, is a basis for .
In the same work [SV09], a closely related polynomial is introduced
(11) |
for
(12) |
A tableau formula is provided in [SV09, Proposition 6.4] as follows.
Theorem 2.6.
For , we have
(13) |
Here is any reverse bitableau of type and shape , with a filling by symbols (see [SV09, Section 6]). The weight is defined as in [SV05, Eq. (41)], and has leading term if for and leading term if for . In fact, [SV09] uses Eq. (13) as the definition for and Eq. (11) as a proposition. Here we present the tableau formula as a theorem in parallel with the following result regarding super Jack polynomials.
We recall the theory of super Jack polynomials from [SV05]. The (monic) Jack symmetric functions are a linear basis for the ring of symmetric functions, and the power sums are free generators , see e.g. [Mac95]. Let be the homomorphism from to the polynomial ring which is defined on the generators as follows
Then the super Jack polynomial is defined by
Theorem 2.7.
3. Supersymmetric Shimura Operators
From now on, we fix and . This section is devoted to the description of the pair and important subspaces therein. We then define the supersymmetric Shimura operators and specialize related results introduced above. Throughout the section we set indices , and is either or depending on the context.
3.1. Realization
We fix the following embedding of into , and identify with its image
(17) |
We let , and . Then has fixed point subalgebra . We also have the Harish-Chandra decomposition
where (respectively ) consists of matrices with non-zero entries only in the upper right (respectively bottom left) sub-blocks in each of the four blocks. Set . Then .
In our theory, we need to work with a -stable Cartan subalgebra of extended from the toral subalgebra . Note in [AHZ10, Section 4], it is showed that is of even type if and only if , satisfied by our pair with . We present a construction of and using a certain Cayley transform as follows.
We let , , , and be the characters on . Let
These are the Harish-Chandra strongly orthogonal roots, and we denote the set of 111Here B indicates the Boson–Boson block (top left) and F the Fermion–Fermion block (bottom right), c.f. [All12]. as . We set for . Associated with each is an -triple spanned by and (similarly for with ). It is not hard to see that all -triples commute. We write for the imaginary unit to avoid confusion. Define
The product
(18) |
is thus a well-defined automorphism on as all terms commute. We set
Then by a direct (rank 1) computation, we see that under :
(19) |
We now define
(20) |
In , the space is the orthogonal complement of with respect to the Killing form on . Also, on we let
be dual to respectively. Then and vanish on and respectively. We identify with its restriction to . We also have , and we identify them with their restrictions on . On , we set
(21) |
which is induced from , the one-half of the supertrace form. For future purposes, we also consider the following basis for :
(22) |
3.2. Root Data
Recall the superdimension of a super vector space is denoted as . Recall that for a (restricted) root space of , the root multiplicity is defined as and the deformed root multiplicity is defined as
We first give an explicit description of using the coordinates defined in Eq. (22). The order on the indices is interpreted as the natural order on , where is identified with . Then we have
(23) | ||||||
(24) | ||||||
This can be seen by (I) viewing as a Weyl group conjugate of the standard Cartan subalgebra and identifying the roots expressed using and , or (II) considering the Cayley transform which in turn gives a correspondence between positive roots on and those on . Alternatively, the chain in Eq. (35) gives the above choice of positive roots.
We now record restricted root data taken from [Zhu22]. The restricted root system is of Type C. In the following table, we give the standard choice of positivity, together with the superdimensions of the restricted root spaces. The parameters are , and these are the 5 parameters first introduced by Sergeev and Veselov in [SV09] to study interpolation polynomials. Note only the last column gives the odd restricted roots.
/ | / |
3.3. Supersymmetric Shimura operators
Throughout the subsection, we set and . Recall is the irreducible -module of highest weight
with respect to , and it is of Type M ([CW12]). In this case, Schur’s Lemma indicates that , and is actually irreducible as a -module as . If we let act on the second component contragrediently (via a negative supertranspose), then we define the irreducible -module as . Note both and are -modules by the short grading. We identify as via the form also used for Eq. (21).
Proposition 3.1.
The symmetric superalgebras and are completely reducible and multiplicity free as -modules. Specifically,
(27) |
Proof.
By duality, it suffices to show the first equation. First we have , by identifying and as spaces of column and row vectors respectively. The contragredient -module structure on is obtained by applying the negative supertranspose on . Then Theorem 2.4 implies
proving the claim, c.f. [SSS20, Theorem 1.4] and notations therein. ∎
We write down the highest weight of with respect to the Borel subalgebra of :
(28) | ||||
(29) |
The part with negative terms is indeed dominant since we take the opposite Borel subalgebra for the second copy of in .
As are supercommutative, the respective universal enveloping algebras are just . The direct summand embedded in is then multiplied into . We write for the element in corresponding to . Let .
Definition 3.2.
For each , we let be the image corresponding to under the composition of the multiplication and the embedding
(30) |
The element is called the supersymmetric Shimura operator associated with the partition .
Remark 2.
For an irreducible Hermitian symmetric space , such can be similarly defined (c.f. [SZ19]) in , the algebra of differential operators on , identified with the universal enveloping algebra of the Lie algebra of . As commutes with , its right action descends to , the algebra of differential operators on . These are the original Shimura operators. As we will study the action of on -modules, we call as (supersymmetric) Shimura operators as well, by a slight abuse of name. Working with the lift of gives tremendous flexibility. By the definitions of , , and , we see
(31) |
3.4. Specialized Results
We now specialize some of the results we introduced in Section 2. First, we set the choice of positive roots according to the chain (c.f. Subsection 2.3)
(32) |
which equivalently gives a Borel subalgebra of , denoted as . In fact, we have
(33) |
Note the Cayley transform as in Eq. (18) allows us to send the positivity on one Cartan subalgebra to that of the other Cartan subalgebra. Specifically, we have , and the dual map . It is then a direct computation to check that
(34) |
Thus, the induced choice of positivity of is given by
(35) |
We denote the corresponding Borel subalgebra as . Specifically, when restricted to , , , , and are positive restricted roots in .
For the Harish-Chandra isomorphism on the non-diagonal , we obtain it via the Cayley transform . By Eq. (19), sends the standard basis for the basis for . Thus by a change of variable using , Theorem 2.3 and Eq. (8) imply
(36) |
The central character for is now given by
(37) |
For the Harish-Chandra isomorphism , we simply apply Eqs. (3, 4) to the choice of and in this context. See Eqs. (3.2, 26) and , the Weyl vectors of and respectively.
We also need the specialization of the spherical highest weights. Consider the weight (Eq. (28)) on and (Eq. (33)). For , we define
This is the in the introduction. To detect the -sphericity (used in Section 4), we specialize Theorem 2.1, which turns out to be quite simple (see [Zhu22, Proposition 4.4]).
Proposition 3.3.
Let and be as above. If , then .
Proof.
It is showed in [Zhu22, Theorem 4.3] that is always finite dimensional via some combinatorial machinery. This allows us to apply the Cayley transform to rewrite -weights as -weights by [Zhu22, Proposition 3.3]. The resulting weight is
(38) |
which vanishes on . The high enough condition degenerates to for . Then a computation shows the claim. ∎
Note that by definition of our notation, we have
(39) |
We write for the restriction of to . Then by Theorem 2.2, for spherical , acts on by the scalar .
Remark 3.
For , it is proved that every corresponds to a spherical [Zhu22, Theorem 1] utilizing Kac induction which appears to be difficult for higher ranks.
4. Harish-Chandra Isomorphisms and a Commutative Diagram
Let , , and be as in Section 3. We have , the ring of supersymmetric polynomials on . The image of (equivalently of ), however, is less understood in such terms. Alldridge’s work [All12] formulates as the intersection of some subalgebras of , while the second author proved a Weyl groupoid invariance formulation in Type A in [Zhu22]. We will study the for and in this section. Specifically, we will give an independent proof that and .
4.1. Image of
We introduce even supersymmetric polynomials on par with what we did in Subsection 2.2. Let and be two sets of independent variables, and set . We write for the polynomial obtained by the substitutions and . A polynomial in is said to be even supersymmetric if (1) is invariant under permutations of and of separately, and invariant under sign changes of and (2) is independent of . Denote the algebra of even supersymmetric polynomials in as . It is not hard to see that an even supersymmetric polynomial is supersymmetric in , so is generated by in Eq. (7) for .
Let be the standard basis for , equipped with an inner product such that and for some . Denote the coordinate functions for this basis as and . Then Condition (2) is equivalent to (see [Mus12]) (2’) if . As such, is identified as a subalgebra of , and denoted as (c.f. ).
Let be the Weyl group associated with the restricted root system (see Subsection 3.2). Then is in fact of Type BC and equals . Originally in [All12], is given by the intersection of certain subspaces using explicitly chosen elements. In [Zhu22], it is proved that this description is in fact independent of any choice, and we have
(40) |
We will offer a different proof that and below in Subsection 4.2.
Remark 4.
See [SV11, Zhu22] for an alternative way of writing down (even) supersymmetric polynomials. Namely, we can consider a so-called Weyl groupoid that captures both the usual double symmetry on two sets of coordinates, and the translational invariance on hyperplanes for all isotropic (restricted) roots (as in Condition (2’)). If (respectively ) denotes the Weyl groupoid associated with (respectively ), then while .
In this section, we prove Theorem B. The strategy is to use a commutative diagram (1) and the surjectivity of a certain restriction map we introduce below. This map allows us to show that surjects onto . Loosely speaking, the non-restricted supersymmetry should cover the restricted supersymmetry entirely.
Recall . Let be the respective projection from to . Then we extend to a projection homomorphism from to , again denoted as . Under the identification between and , we let be the “restriction” homomorphism corresponding to . Specifically, the adjoint map gives a pullback of a polynomial defined on to a polynomial defined on , and . We denote its restriction on as too.
Proposition 4.1.
The restriction map on surjects onto .
Proof.
We show that maps generators of to all the generators of . For this, we choose coordinate systems for both and as in Section 3. Throughout the proof we set indices . Recall , , , , , and from Subsection 3.1. Also, and span while and span by Eq. (20). Thus by definition, the homomorphism on makes the substitution
(41) |
The generators in Eq. (7) of become . By Eq. (41), we have
The preimage of under is , meaning the generators of are in the image of . Therefore, . ∎
The substitution in Eq. (41) also tells us that and are related by
(42) |
4.2. A commutative diagram
As pointed out in the Introduction, the center acts on the unique up to constant spherical vector by the character via . Considered as a subalgebra of , there is another way to compute such a character using the Harish-Chandra map . We aim to relate the two Harish-Chandra isomorphisms and .
We denote the quotient map as . Note by definition, is contained in the kernel of the Harish-Chandra projection, hence in . So the homomorphism is well-defined. Then . Of course, the bijectivity of would imply . We form the following diagram first.
(43) |
Then Theorem B follows from the two assertions below.
Proposition 4.2.
Diagram (43) is commutative.
Proposition 4.3.
The homomorphism is the isomorphism and .
Proof of Theorem B.
Proof of Proposition 4.2.
The set of all the hook partitions satisfying the condition in Proposition 3.3 give us a family of finite dimensional, irreducible, and spherical -module . In notation of Eq. (22) (c.f. proof to Proposition 4.1), the corresponding highest weight (see Eq. (38)) is:
(44) |
whose restriction to is
(45) |
Let us denote the set of weights in Eq. (45) as . Now, considered as a subset of the integral lattice in , is fully determined by many inequalities
Then is Zariski dense in as it is the intersection of an open cone with .
For , it acts by a scalar by Lemma 2.1, and by Theorem 2.2, this scalar is
On the other hand, by Eq. (37), acts on the entirety of by . Hence
(46) |
But vanishes on . Therefore, the left side of Eq. (46) is just . This can be verified directly using Eq. (42) with inputs Eqs. (3.2, 26, 44). Thus Eq. (46) implies
on . As the two sides agree on a Zariski dense subset of , they must be equal everywhere. This proves the commutativity of (43). ∎
Proof of Proposition 4.3.
Now we show that , and is injective. For the sake of simplicity, write for , and let . We also let , and .
Step 1. Let . By the Poincaré–Birkhoff–Witt Theorem, we have . This is in fact a -module decomposition, and its -invariant counterpart is
By Definition 3.2, we see that constitute a basis for
(47) |
indexed by . Let be the equivalent class of . Identifying , we see that gives a basis for .
Step 2. Furthermore, we see that is graded by non-negative integers according to in Eq. (47) (c.f. Eq. (27)) as a vector superspace. The basis is homogeneous. We further define
This gives a vector superspace filtration. The reason why we consider filtration instead of grading is that only preserves filtration due to the shift in its definition. For each filtered degree, let
Step 3. By Proposition 4.1, is surjective onto , from which we have
We also showed that the diagram commutes. Thus the right side is , which gives . By Step 2., the filtered version of the assertion is
(48) |
5. Generalized Verma Modules
In this section, we study certain generalized Verma modules, denoted as . These are the -modules induced on () from a parabolic subalgebra containing to . If we choose a Borel subalgebra of as the parabolic subalgebra, then we obtain the usual Verma module, which explains the name. We show that is spherical. We investigate a natural grading on , and use it to show that certain central elements act trivially on trivially whenever . Further strengthening is also obtained, asserting the same vanishing action of for . This is Theorem C.
5.1. Basics
Consider the parabolic subalgebra in . The associated set of roots is given by . Let be as in the Cheng–Wang decomposition. We extend the -action trivially to to obtain a -module structure on . We define the generalized Verma module as
By the Poincaré–Birkhoff–Witt theorem, we see that , and we have
both as super vector spaces and as -modules. Clearly, is a weight module and the highest weight is Eq. (28), given with respect to (see Eq. (33)). Then is the irreducible quotient of .
Next, we introduce a grading on . Recall that in the one-dimensional center of , there is a special element (see Section 3) such that the Harish-Chandra decomposition corresponds to -eigenspaces of , and gives rise to the short grading. We may extend the action of to . Let and be homogeneous bases for , , and respectively. Then the Poincaré–Birkhoff–Witt basis is:
where a strict inequality of subindices occurs when the corresponding basis vectors are odd. We let
Lemma 5.1.
The universal enveloping algebra is graded by , and .
Proof.
The action of on the basis is given by the scalar in . The bases for for all give a partition of the Poincaré–Birkhoff–Witt basis, and the sum is hence direct. Furthermore, the multiplication in respects the Lie superbracket, which in turn respects the short grading on . Therefore the grading is well-defined. ∎
Let . Then is also endowed with the grading. Explicitly,
has degree for . In particular, its top homogeneous component is of degree , isomorphic to . By Lemma 5.1, since the acts by left multiplication on , we have . This leads to the following lemma.
Lemma 5.2.
Let . Then .
Next, we show that is in fact spherical.
Proposition 5.3.
The module is spherical. Furthermore, and .
Proof.
As a -module, we have and each component can be identified as . Only the degree 0 component has a one-dimensional -invariant subspace by Schur’s Lemma for Type M modules. Hence is spherical with a unique up to constant spherical vector. ∎
By the above Proposition, in is unique up to constant. Let us fix such in the following discussion.
5.2. Vanishing actions
Recall that the supersymmetric Shimura operator in is defined as the image of under the following composition of maps (Definition 3.2)
In particular . If we take a homogeneous basis for and its dual basis for , then can be written as
(49) |
Proposition 5.4.
If , then for any .
Proof.
Viewing , we have the representation given by for and .
Lemma 5.5.
The image of restricted to is homomorphic to .
Proof.
Let . For each , if we apply a homogeneous , we get
(50) |
as . Thus, the map intertwines with the actions of on and .
Let be the subspace of spanned by . Then is a homomorphic image of in by Eq. (50). ∎
Proposition 5.6.
If but , then .
Proof.
Corollary 5.7.
For any such that , , we have in .
Recall . By Theorem B, is surjective. Then it is possible to find such that . We show that acts by 0 on when satisfies certain condition.
Proposition 5.8.
The central element acts on by when , and .
Proof.
By Lemma 2.1, acts on by a scalar for any . On the other hand, the action of on a spherical vector in a -module descends to an action of by setting . Hence, in , there is no ambiguity of writing
(51) |
Therefore, by Corollary 5.7,
(52) |
Since acts on the entirety of a highest weight module by a scalar and is a highest weight module, such scalar has to be which vanishes according to Eq. (52). ∎
5.3. A stronger result
What is proved in the previous subsection turns out to be enough for the purposes of this paper (Theorem A). Nonetheless, we give a better description of such vanishing actions, proving Theorem C.
First, we recall the following “weight decomposition” proposition from [Zhu22], parallel to Proposition 3.2 in [Kum10], and is first due to Kostant [Kos59]. We let be a Lie (super)algebra, be a Borel subalgebra.
Proposition 5.9.
Let and be two finite dimensional irreducible -modules with highest weight and , and be a weight module of . If , then for some weight of .
We would like to show the following branching statement, i.e., occurs in as a homomorphic image if . This result is similar to the one regarding the irreducible representation in [Zhu22].
Proposition 5.10.
If , then .
Proof.
The -module structure on is given by . Let us denote the representation map on (respectively ) as (respectively ). Then the braiding map gives a canonical module isomorphism, meaning that if occurs in , then it also occurs in .
By Proposition 5.9, is a weight of . Here, and are weights in . As both and are integral combinations of , so is . Since , such integral coefficients in must be non-positive. Equivalently, . ∎
Proof of Theorem C.
As in the proof of Proposition 5.8, is the scalar by which acts on . Again by Eq. (51) and the fact that has to act uniformly by a scalar on , is the scalar by which acts on the entirety of . By Lemma 5.5, the representation of on always gives a homomorphic image of itself in . If , then this image is zero by Proposition 5.10. Spelling out as in Eq. (49), we see that . Thus,
(53) |
and this proves that acts on by 0 unless for all . ∎
Eq. (53) gives more vanishing points of than we actually need for proving Theorem A in Section 6. It is stronger than Proposition 5.8 which concerns only zeros at lower degrees. In fact, this is much closer to the original formulation of the vanishing properties, introduced in Theorem 2.5, c.f. Proposition 6.1 below.
6. Type BC Interpolation Polynomials
In this section, we specify Sergeev and Veselov’s results (Subsection 2.4) with all the parameters in Table 1 (Subsection 3.2), and then an explicit change of variables. For the sake of simplicity, write for , for . Throughout the section, we set , .
We now specify using our restricted root system . In particular, we set , , , , . By Table 1, , while and do not exist. By comparing Eq. (10) with as in Eq. (26), we have . Thus we have the two parameters
for . Note our form Eq. (21) is the negative of what is defined in Subsection 2.4. But this does not matter when . We set (c.f. Eq. (9))
We write for , and define . We define .
Proposition 6.1.
For each , there is a unique polynomial of degree such that
and that . Moreover, is a basis for .
Proof.
Proposition 6.2.
Every is determined by its values on .
Proof.
Let be the space of functions on . Then . In particular, has a Kronecker-delta basis . Next, the evaluation of on gives a restriction map
To prove the statement, we show that is an isomorphism.
Fix a total order on such that implies . Consider the matrix for with respect to the bases for and for arranged by . Since , and for any such that , we see that is upper triangular with non-zero diagonal entries. Therefore is invertible, proving the statement. ∎
To prove Theorem A, we first show that is proportional to defined above. Then we pin down the scalar
(55) |
by comparing the top homogeneous degrees of both sides. Here
(c.f. Subsection 2.4). The factor of is due to the definition of , which is a result of the definition of in [SV05, Eq. (41)].
Proof of Theorem A.
We show that
Step 1. From the definition of and , we see that . By (1), we have . The left side is just . The right side, when applied to is just by Eq. (42) (c.f. Eq. (46)). Hence
(56) |
By Theorem C, acts by 0 on for all , , which exactly gives . Thus
for all , , proving the vanishing properties.
Alternatively, by Proposition 5.8, we have the “relaxed” vanishing properties when, . By Proposition 6.2, this also implies the full “extra” vanishing properties.
Therefore, is proportional to by its vanishing properties.
Step 2. We first pick a basis for extended from
and a dual basis for (identified via the bilinear form ) extended from
Then the identity map corresponds to
(57) |
in . On the other hand, we may extend the basis for (see Subsection 3.1) to a basis for according to . This is used to define and . Then it is a direct computation to see that the top homogeneous degree of
is precisely
Note the shift does not change this top degree.
Now we let . The identity map on corresponds to an element in . Thus the sum corresponds to an element in by the decomposition in Proposition 3.1. On , the bilinear form is induced from on by normalizing by . Then similarly, we see that the top homogeneous degree of is given by
(58) |
Note (Eq. (12)) specializes to . From Eq. (11), Theorems 2.6 and 2.7, we see that
and the top homogeneous degree of is . For , this becomes
Thus by setting in Eq. (16),
(59) |
Comparing the coefficients, we have
(60) |
Therefore the coefficient is precisely given by Eq. (55). ∎
Additionally, we can answer the interesting question of by which scalar acts on the spherical vector , without any direct consideration on the module itself!
Corollary 6.3.
The operator acts on by .
Proof.
This scalar is the value of at which equals the value of at :
∎
References
- [AHZ10] Alexander Alldridge, Joachim Hilgert, and Martin R. Zirnbauer, Chevalley’s restriction theorem for reductive symmetric superpairs, J. Algebra 323 (2010), no. 4, 1159–1185. MR 2578599
- [All12] A. Alldridge, The Harish-Chandra isomorphism for reductive symmetric superpairs, Transform. Groups 17 (2012), no. 4, 889–919. MR 3000475
- [AS15] Alexander Alldridge and Sebastian Schmittner, Spherical representations of Lie supergroups, J. Funct. Anal. 268 (2015), no. 6, 1403–1453. MR 3306354
- [ASS18] Alexander Alldridge, Siddhartha Sahi, and Hadi Salmasian, Schur -functions and the Capelli eigenvalue problem for the Lie superalgebra , Representation theory and harmonic analysis on symmetric spaces, Contemp. Math., vol. 714, Amer. Math. Soc., [Providence], RI, [2018] ©2018, pp. 1–21. MR 3847241
- [CW01] Shun-Jen Cheng and Weiqiang Wang, Howe duality for Lie superalgebras, Compositio Math. 128 (2001), no. 1, 55–94. MR 1847665
- [CW12] by same author, Dualities and representations of Lie superalgebras, Graduate Studies in Mathematics, vol. 144, American Mathematical Society, Providence, RI, 2012. MR 3012224
- [FK90] J. Faraut and A. Korányi, Function spaces and reproducing kernels on bounded symmetric domains, J. Funct. Anal. 88 (1990), no. 1, 64–89. MR 1033914
- [Gor04] Maria Gorelik, The Kac construction of the centre of for Lie superalgebras, J. Nonlinear Math. Phys. 11 (2004), no. 3, 325–349. MR 2084313
- [Hel00] Sigurdur Helgason, Groups and geometric analysis, Mathematical Surveys and Monographs, vol. 83, American Mathematical Society, Providence, RI, 2000, Integral geometry, invariant differential operators, and spherical functions, Corrected reprint of the 1984 original. MR 1790156
- [Hum78] James E. Humphreys, Introduction to Lie algebras and representation theory, Graduate Texts in Mathematics, vol. 9, Springer-Verlag, New York-Berlin, 1978, Second printing, revised. MR 499562
- [Kac84] Victor G. Kac, Laplace operators of infinite-dimensional Lie algebras and theta functions, Proc. Nat. Acad. Sci. U.S.A. 81 (1984), no. 2, 645–647. MR 735060
- [Kos59] Bertram Kostant, A formula for the multiplicity of a weight, Trans. Amer. Math. Soc. 93 (1959), 53–73. MR 109192
- [KS93] Bertram Kostant and Siddhartha Sahi, Jordan algebras and Capelli identities, Invent. Math. 112 (1993), no. 3, 657–664. MR 1218328
- [Kum10] Shrawan Kumar, Tensor product decomposition, Proceedings of the International Congress of Mathematicians. Volume III, Hindustan Book Agency, New Delhi, 2010, pp. 1226–1261. MR 2827839
- [LSS22a] Gail Letzter, Siddhartha Sahi, and Hadi Salmasian, The Capelli eigenvalue problem for quantum groups, 2022, arXiv.
- [LSS22b] by same author, Weyl algebras for quantum homogeneous spaces, 2022, arXiv.
- [LSS23] by same author, Quantized Weyl algebras, the double centralizer property, and a new First Fundamental Theorem for , 2023, arXiv.
- [Mac95] I. G. Macdonald, Symmetric functions and Hall polynomials, second ed., Oxford Mathematical Monographs, The Clarendon Press, Oxford University Press, New York, 1995, With contributions by A. Zelevinsky, Oxford Science Publications. MR 1354144
- [Mus12] Ian M. Musson, Lie superalgebras and enveloping algebras, Graduate Studies in Mathematics, vol. 131, American Mathematical Society, Providence, RI, 2012. MR 2906817
- [Oko98] A. Okounkov, -type interpolation Macdonald polynomials and binomial formula for Koornwinder polynomials, Transform. Groups 3 (1998), no. 2, 181–207. MR 1628453
- [OO06] Andrei Okounkov and Grigori Olshanski, Limits of -type orthogonal polynomials as the number of variables goes to infinity, Jack, Hall-Littlewood and Macdonald polynomials, Contemp. Math., vol. 417, Amer. Math. Soc., Providence, RI, 2006, pp. 281–318. MR 2284134
- [Sah94] Siddhartha Sahi, The spectrum of certain invariant differential operators associated to a Hermitian symmetric space, Lie theory and geometry, Progr. Math., vol. 123, Birkhäuser Boston, Boston, MA, 1994, pp. 569–576. MR 1327549
- [Sch70] Wilfried Schmid, Die Randwerte holomorpher Funktionen auf hermitesch symmetrischen Räumen, Invent. Math. 9 (1969/70), 61–80. MR 259164
- [Ser99] Alexander Sergeev, The invariant polynomials on simple Lie superalgebras, Represent. Theory 3 (1999), 250–280. MR 1714627
- [She22] Alexander Sherman, Iwasawa decomposition for Lie superalgebras, J. Lie Theory 32 (2022), no. 4, 973–996. MR 4512155
- [Shi90] Goro Shimura, Invariant differential operators on Hermitian symmetric spaces, Ann. of Math. (2) 132 (1990), no. 2, 237–272. MR 1070598
- [SS16] Siddhartha Sahi and Hadi Salmasian, The Capelli problem for and the spectrum of invariant differential operators, Adv. Math. 303 (2016), 1–38. MR 3552519
- [SSS20] Siddhartha Sahi, Hadi Salmasian, and Vera Serganova, The Capelli eigenvalue problem for Lie superalgebras, Math. Z. 294 (2020), no. 1-2, 359–395. MR 4054814
- [Sta89] Richard P. Stanley, Some combinatorial properties of Jack symmetric functions, Adv. Math. 77 (1989), no. 1, 76–115. MR 1014073
- [Ste85] John R. Stembridge, A characterization of supersymmetric polynomials, J. Algebra 95 (1985), no. 2, 439–444. MR 801279
- [SV05] A. N. Sergeev and A. P. Veselov, Generalised discriminants, deformed Calogero-Moser-Sutherland operators and super-Jack polynomials, Adv. Math. 192 (2005), no. 2, 341–375. MR 2128703
- [SV09] Alexander N. Sergeev and Alexander P. Veselov, Calogero-Moser operator and super Jacobi polynomials, Adv. Math. 222 (2009), no. 5, 1687–1726. MR 2555909
- [SV11] by same author, Grothendieck rings of basic classical Lie superalgebras, Ann. of Math. (2) 173 (2011), no. 2, 663–703. MR 2776360
- [SZ19] Siddhartha Sahi and Genkai Zhang, Positivity of Shimura operators, Math. Res. Lett. 26 (2019), no. 2, 587–626. MR 3999556
- [Yan92] Zhi Min Yan, A class of generalized hypergeometric functions in several variables, Canad. J. Math. 44 (1992), no. 6, 1317–1338. MR 1192421
- [Zhu22] Songhao Zhu, Shimura operators for certain Hermitian symmetric superpairs, arXiv, submitted, 2022.