On the intertwining differential operators from a line bundle to a vector bundle over the real projective space
Abstract.
We classify and construct -intertwining differential operators from a line bundle to a vector bundle over the real projective space by the F-method. This generalizes a classical result of Bol for . Further, we classify the -type formulas for the kernel and image of . The standardness of the homomorphisms corresponding to the differential operators between generalized Verma modules are also discussed.
Key words and phrases:
Bol operator, intertwining differential operator, generalized Verma module, F-method, -type formula, standard map2020 Mathematics Subject Classification:
22E46, 17B101. Introduction
Gerrit van Dijk worked on harmonic analysis on -adic groups and real Lie groups, inspired by Harish-Chandra as a post-doc at IAS Princeton. He studied both abstract questions such as the nature of convolution algebras of functions on symmetric spaces, but also concrete special functions and distributions on such spaces. He also considered induced representations from a parabolic subgroup to a reductive Lie group and the explicit structure of such in concrete cases. In particular, he jointly with Molchanov studied the case of being the real projective space ([71]). In this paper we aim to study intertwining differential operators between such induced representations.
1.1. Main problems
To state the main problems of this paper, we first introduce some notation. Let be a real reductive Lie group with complexified Lie algebra . Let be a parabolic subgroup with Langlands decomposition . We write for the set of equivalence classes of finite-dimensional irreducible representations of . Likewise, let denote the set of characters of . Then, for the outer tensor product of , , and the trivial representation of , we put
for an (unnormalized) parabolically induced representation of . Let denote the space of -intertwining differential operators .
There are two main problems in this paper. The first problem concerns the classification and construction of intertwining differential operators as follows.
Problem A.
Do the following.
-
(A1)
Classify such that
-
(A2)
For classified in (A1), determine the dimension
-
(A3)
For classified in (A1), construct generators
Let be a maximal compact subgroup of . For , the kernel and image are -invariant subspaces of and , respectively. Let and denote the space of -finite vectors of and , respectively. The following is the other main problem of this paper.
Problem B.
Classify the -type formulas of and .
For instance, if , then the maximal compact subgroup is . Problem B then asks how and decompose as -modules. We shall answer this question in Corollary 6.7 for with .
Problem A for the first order intertwining differential operators for general was done around 2000 independently by Ørsted [63], Slovák–Souček [66], and Johnson–Korányi–Reimann [29], which generalize the work of Fegan [20] for . See also the works of Korányi–Reimann [49] and Xiao [72] as relevant works for this case. The higher order case is still a work in progress. For some recent studies, see, for instance, Barchini–Kable–Zierau [2, 3], Kable [32, 33, 34, 35, 36, 37], Kobayashi–Ørsted–Somberg–Souček [47], and the first author [54, 55].
Problem A has been paid attention especially in conformal geometry. The Yamabe operator, also known as the conformal Laplacian, is a classical example of a conformally covariant differential operator (cf. [46, 59]). In this paper we consider the projective structure in parabolic geometry. In other words, our aim is to classify and construct “projectively covariant” differential operators.
On the study of intertwining differential operators, the BGG sequence is an important background (cf. [8, 11]). The kernel of the first BGG operators has also been studied carefully from a geometric point of view (cf. [9, 10, 22]). Intertwining differential operators are also studied intensively by Dobrev in quantum physics (cf. [14, 15, 16, 17, 18, 19]).
We next describe the concrete setup of this paper.
1.2. Specialization to
In this paper we consider Problems A and B for , where is the maximal parabolic subgroup of corresponding to the partition so that is diffeomorphic to the real projective space . The representation denotes a one-dimensional representation of . The details will be discussed in Section 3.1. We shall solve Problem A in Theorems 4.2 and 4.5, and Problem B in Corollary 6.7.
In 1949, Bol from projective differential geometry showed that -intertwining differential operators for an equivariant line bundle over are only the powers of the standard derivative (cf. [5] and [64, Thm. 2.1.2]). The operator is sometimes called the Bol operator. For recent results on analogues of the Bol operator for Lie superalgebras, see, for instance, [6, 7]. We successfully generalize the classical result of Bol on , which is a double cover of , to on .
It is classically known that the Bol operators are the residue operators of the Knapp–Stein operator. In contrast to the case of , the Knapp–Stein operator does not exist for for (cf. [62]). Thus, the differential operators obtained in Theorem 4.5 are not the residue operators of such. In the sense of Kobayashi–Speh [48], our differential operators are all sporadic operators for .
1.3. The F-method
Our main tool to work on Problem A is the so-called F-method. This is a fascinating method invented by Toshiyuki Kobayashi around 2010 in the course of the study of his branching program (see, for instance, [40], [47, Introduction], and [69, Sect. B]). Since then, Problem A has been intensively studied by the F-method especially in symmetry breaking setting (cf. [21, 38, 39, 42, 44, 45, 47, 51, 52, 53, 61, 65, 67]).
The F-method makes it possible to classify and construct intertwining differential operators (or more generally speaking, differential symmetry breaking operators) by solving a certain system of partial differential equations. To describe the main idea more precisely, recall that it follows from a fundamental work of Kostant [50] that the space of intertwining differential operators is isomorphic to the space of -homomorphisms between generalized Verma modules (see also [12, 14, 23, 27, 49]). Schematically, we have
(1.1) |
The precise statement will be given in Theorem 2.5. In general, it is easier to work with the Verma module side “” than the differential operator side . Thus, the standard strategy to tackle Problem A is to convert the problem into the one for generalized Verma modules. Nonetheless, even in a case that the unipotent radical is abelian, it requires involved combinatorial computations (see, for instance, [30, Chap. 5]).
The novel idea of the F-method is to further identify with the space of polynomial solutions to a system of PDEs by applying a Fourier transform to a generalized Verma module. In other words, the F-method puts another picture “” to (1.1) as follows.
(1.2) |
Then, in the F-method, one achieves the classification and construction of intertwining differential operators simultaneously by solving the system of PDEs. We shall explain more details in Section 2.
1.4. The counterpart of generalized Verma modules
As observed above, one can obtain -homomorphisms in “” from the solutions in “”. We thus study the algebraic counterpart of Problem A for . It is achieved in Theorems 4.6 and 4.8. We further consider a variant of Problem A for -homomorphisms between generalized Verma modules in Theorem 5.23. We also classify in Corollary 5.24 the reducibility of the generalized Verma module in consideration; this recovers a result of [1, 24, 25] for the pair .
A -homomorphism between generalized Verma modules is called a standard map if it is induced from a -homomorphism between the corresponding (full) Verma modules; otherwise, it is called a non-standard map ([4, 60]). In this paper we also show that the resulting -homomorphisms in Theorem 5.23 are all standard.
1.5. The -type formulas for and
Kable [31] and the authors [58] recently showed a Peter–Weyl type theorem for the kernel of an intertwining differential operator . The theorem allows us to compute the -type formula of explicitly by solving the hypergeometric/Heun differential equation ([57, 58]). Tamori [68] independently used a similar idea to determine the -type formula of for his study of minimal representations.
The Peter–Weyl type theorem works nicely for first and second order differential operators; nonetheless, it requires a certain amount of computations for higher order cases. Since the differential operators that we obtained in Theorem 4.5 have arbitrary order, we take another approach in this paper.
In 1990, van Dijk–Molchanov [71] and Howe–Lee [26] independently showed among other things that the degenerate principal series representations in consideration have length two and have a unique finite-dimensional irreducible subrepresentation . (See Möllers–Schwarz [62] for recent development of this matter.) Since the -type structures of the induced representations are also known, the determination of the -type formula of is equivalent to showing . As the length is two, it simultaneously determines the -type formula of .
1.6. Organization of the paper
Now we outline the rest of this paper. There are seven sections including the introduction. In Section 2, we review a general idea of the F-method. In particular, a recipe of the F-method will be given in Section 2.9. Since we do not find a thorough exposition of the F-method in the English literature (except the original papers of Kobayashi with his collaborators [44, 45, 47]), we decided to give some detailed account. We hope that it will be helpful for a wide range of readers. It is remarked that part of the section is an English translation of the Japanese article [56] of the first author.
Section 3 is for the specialization of the framework discussed in Section 2 to the case . In this section we fix some notation and normalizations for the rest of the sections. Then, in Section 4, we give our main results for Problem A for the triple . These are accomplished in Theorems 4.2 and 4.5. Further, a variant of Problem A for -homomorphisms between generalized Verma modules is also discussed in this section. These are stated in Theorems 4.6 and 4.8. We give proofs of these theorems in Section 5 by following the recipe of the F-method.
We consider Problem B in Section 6. In this section we classify the -type formulas of the kernel and the image of the intertwining differential operators classified in Section 4. These are obtained in Corollary 6.7.
Section 7 is an appendix discussing the standardness of the -homomorphisms obtained in Section 4 between generalized Verma modules. We first quickly review the definition of the standard map. Then, by applying a version of Boe’s criterion, we show that the homomorphisms are all standard maps. This is achieved in Theorem 7.4.
2. Quick review on the F-method
The aim of this section is to review the F-method. In particular, a recipe of the F-method is given in Section 2.9. In Section 5, we shall follow the recipe to classify and construct intertwining differential operators . We mainly follow the arguments in [41] and [44] in this section.
2.1. General framework
Let be a real reductive Lie group and a Langlands decomposition of a parabolic subgroup of . We denote by and the Lie algebras of and , respectively.
For a real Lie algebra , we write and for its complexification and the universal enveloping algebra of , respectively. For instance, , and are the complexifications of , and , respectively.
For , we denote by the one-dimensional representation of defined by . For a finite-dimensional irreducible representation of and , we denote by the outer tensor representation . As a representation on , we define . By letting act trivially, we regard as a representation of . Let be the -equivariant vector bundle over the real flag variety associated with the representation of . We identify the Fréchet space of smooth sections with
the space of -invariant smooth functions on . Then, via the left regular representation of on , we realize the parabolically induced representation on . We denote by the right regular representation of on .
Similarly, for a finite-dimensional irreducible representation of , we define an induced representation on the space of smooth sections for a -equivariant vector bundle . We write for the space of intertwining differential operators .
Let be the Gelfand–Naimark decomposition of , and write . We identify with the open Bruhat cell of via the embedding , . Via the restriction of the vector bundle to the open Bruhat cell , we regard as a subspace of .
We view intertwining differential operators as differential operators such that the restriction to is a map (see (2.1) below).
(2.1) |
In particular, we regard as
(2.2) |
To describe the F-method, one needs to introduce the following:
After reviewing these objects, we shall discuss the F-method in Section 2.7.
2.2. Infinitesimal representation
We start with the representation of on derived from the induced representation of on .
For a representation of , we denote by the infinitesimal representation of . For instance, and denote the infinitesimal representations of the left and right regular representations of on . As usual, we naturally extend representations of to ones for its universal enveloping algebra of its complexification . The same convention is applied for closed subgroups of .
For , we write
where and . Similarly, for with , we write
where and .
Let . Since is an open dense subset of , we have for sufficiently small . Thus, we have
which implies
Further, for and sufficiently small , we have
Observe that
Then, for , we have
We remark that the -invariance property that for is applied in the first line. Since
the representation on the image of the inclusion is given by
(2.3) |
Equation (2.2) shows that, for , the formula on can be extended to the whole space . By extending the formula complex linearly to , we have
(2.4) |
for and .
2.3. Duality theorem
For a finite-dimensional irreducible representation of , we write and for the contragredient representation of . By letting act on trivially, we regard the infinitesimal representation of as a -module. The induced module
is called a generalized Verma module. Via the diagonal action of on , we regard as a -module.
The following theorem is often called the duality theorem. For the proof, see, for instance, [12, 44, 49].
Theorem 2.5 (duality theorem).
There is a natural linear isomorphism
(2.6) |
where, for and , the element is given by
(2.7) |
where .
Remark 2.8.
By the Frobenius reciprocity, (2.6) is equivalent to
(2.9) |
Remark 2.10.
In general is not connected. If it is connected, then
Thus, in the case, the isomorphism (2.9) is equivalent to
(2.11) |
2.4. Algebraic Fourier transform of Weyl algebras
Let be a complex finite-dimensional vector space with . Fix a basis of and let denote the coordinates of with respect to the basis. Then the algebra
with relations , , and is called the Weyl algebra of , where is the Kronecker delta. Similarly, let denote the coordinates of the dual space of with respect to the dual basis of . We write for the Weyl algebra of . Then the map determined by
(2.12) |
gives a Weyl algebra isomorphism
(2.13) |
The map (2.13) is called the algebraic Fourier transform of Weyl algebras ([44, Def. 3.1]). We remark that the minus sign for in (2.12) is put in such a way that the resulting map in (2.13) is indeed a Weyl algebra isomorphism. We remark that, for the Euler homogeneity operators for and for , we have .
The Weyl algebra naturally acts on the space of distributions on . In particular, the action of preserves the subspace of distributions supported at . Let be the annihilator of the Dirac delta function , that is, the kernel of the homomorphism
Then is the left ideal of generated by the coordinate functions . Since the map is surjective (see, for instance, [70, Thm. 5.5]), it induces the isomorphism
(2.14) |
On the other hand, by applying the algebraic Fourier transform in (2.13) to , the space is isomorphic to the space of polynomials on . Thus, we have
(2.15) |
It is remarked that the composition maps .
2.5. Fourier transformed representation
For , we denote by the one-dimensional representation of defined by . For the contragredient representation of , we put . As for , we regard as a representation of . Define the induced representation on the space of smooth sections for the vector bundle associated with , which is isomorphic to the tensor bundle of the dual vector bundle and the bundle of densities over . Then the integration on gives a -invariant non-degenerate bilinear form for and .
As for , the space can be regarded as a subspace of . Then, by replacing with in (2.4), we have
(2.16) |
for and . Via the exponential map , one can regard as a representation on . It then follows from (2.16) that gives a Lie algebra homomorphism
where are coordinates of with . We extend the coordinate functions for holomorphically to the ones for . Thus we have
Now we fix a non-degenerate symmetric bilinear form on and . Via , we identify with the dual space of . Then the algebraic Fourier transform of Weyl algebras (2.13) gives a Weyl algebra isomorphism
In particular, it gives a Lie algebra homomorphism
(2.17) |
2.6. Algebraic Fourier transform of generalized Verma modules
Our aim is to construct a -module isomorpshism
where denotes the space of polynomial functions on . It is classically known that the map gives a -module isomorphism
(2.18) |
where denotes the identity coset of and denotes the Dirac delta function supported at (see, for instance, [44, 50]). It thus suffices to construct a -module isomorphism
First, observe that the restriction of the dual vector bundle to the open Bruhat cell gives an isomorphism
Via the isomorphism , one can induce a -module structure on from . In particular, the Lie algebra acts on via the infinitesimal representation . Consequently, by the isomorphism (2.15), one obtains the following chain of -isomorphisms:
(2.19) | ||||||
We denote by the composition of the three -module isomorphisms in (2.19).
Definition 2.20 ([44, Sect. 3.4]).
We call the -module isomorphism
(2.21) |
the algebraic Fourier transform of the generalized Verma module .
2.7. The F-method
Observe that the algebraic Fourier transform in (2.21) gives an -representation isomorphism
(2.22) |
which induces a linear isomorphism
(2.23) |
Here acts on via the action
(2.24) |
Recall from (2.17) that we have
Since the nilpotent radical acts trivially on , and since the algebraic Fourier transform is a -module isomorphism, the Fourier image must act on trivially. Therefore, the operator for is a differential operator of order at least 1; in particular, the space is the space of polynomial solutions to the system of partial differential equations defined by for . It is remarked that the operators are not vector fields; for instance, if is abelian, then are second order differential operators (cf. Proposition 5.4 and [44, Prop. 3.10]).
To emphasize that is the space of polynomial solutions to a system of differential equations, we let
(2.25) |
Via the identification , we have
(2.26) |
(2.27) |
We call the system of PDEs (2.27) the F-system ([42, Fact 3.3 (3)]). Since
the isomorphism (2.23) together with (2.25) shows the following.
Theorem 2.28 (F-method, [44, Thm. 4.1]).
There exists a linear isomorphism
Equivalently, we have
2.8. The case of abelian nilradical
Now suppose that the nilpotent radical is abelian. In this case intertwining differential operators have constant coefficients. Indeed, observe that, as is abelian, so is . Thus, we have
For a given ordered basis with , one may identify with via the exponential map
(2.30) |
Then, for , we have
Since intertwining differential operators are given by linear combinations of the differential operators over by Theorem 2.5, this shows that the coefficients of must be constant.
2.9. A recipe of the F-method for abelian nilradical
-
Step 1
Compute and for .
-
Step 2
Classify and construct .
-
Step 3
Solve the F-system (2.27) for .
-
Step 4
For obtained in Step 3, do the following.
-
Step 4a
Apply to to obtain .
-
Step 4b
Apply to to obtain .
-
Step 4a
In Section 5, we shall carry out the recipe for the case .
3. Specialization to
The aim of this short section is to specialize the general framework in Section 2 to the case , where denotes the maximal parabolic subgroup of corresponding to the partition . Throughout this section, we assume .
3.1. Notation
Let with Lie algebra for . We put
and
where denote the matrix units. We normalize as , namely,
Let
(3.1) | ||||
Then we have
For , let denote the trace form of . Then and satisfy . In what follows, we identify the dual of with via the trace form .
Let and
(3.2) |
Here is regarded as . We have and the decomposition is a Gelfand–Naimark decomposition of . The subalgebra is a maximal parabolic subalgebra of . It is remarked that are abelian.
Let be the normalizer of in . We write for the Langlands decomposition of corresponding to . Then and . The group is given by
Here is regarded as . As is not connected, let denote the identity component of . We write
Then and
where is the identity matrix and .
For a closed subgroup of , we denote by and the sets of equivalence classes of irreducible representations of and finite-dimensional irreducible representations of , respectively.
For , we define a one-dimensional representation of by
(3.3) |
Then is given by
For , a one-dimensional representation of is defined by
where
Then is given by
Since , the set can be parametrized by
For , we write
(3.4) |
for unnormalized parabolically induced representation of . For instance, the unitary axis of is , where denotes the trivial representation of . Similarly, for the representation space of , we write
(3.5) |
In the next section we classify and construct intertwining differential operators
and -homomorphisms
4. Classification and construction of and
The aim of this section is to state the classification and construction results for intertwining differential operators as well as -homomorphisms . These are given in Theorems 4.2 and 4.5 for and Theorems 4.6 and 4.8 for . The proofs of the theorems will be discussed in Section 5. We remark that the case of is studied in [56].
4.1. Classification and construction of intertwining differential operators
We start with the classification of the parameters such that .
For and , we mean by
Then we define as
(4.1) |
where denotes the irreducible representation on of induced by the standard action on . We regard as for all .
Theorem 4.2.
The following three conditions on are equivalent.
-
(i)
.
-
(ii)
.
-
(iii)
One of the following two conditions holds:
-
(iii-a)
.
-
(iii-b)
.
-
(iii-a)
We next consider the explicit formula of for in (iii) of Theorem 4.2. We write
Then, as in (2.2) and (2.30), we understand for as a map
via the diffeomorphism
(4.3) |
For , we put
For , we write
We define for by
(4.4) |
For , we understand as the identity operator .
Theorem 4.5.
We have
4.2. Classification and construction of -homomorphisms
Let . We regard as .
Define as
where denotes the irreducible representation on of . As for , we regard as for all .
The classification of the parameters such that is given as follows.
Theorem 4.6.
The following three conditions on are equivalent.
-
(i)
.
-
(ii)
.
-
(iii)
One of the following two conditions holds:
-
(iii-a)
.
-
(iii-b)
.
-
(iii-a)
To give the explicit formula of , we write
where are the standard basis elements of .
For , we write
Observe that we have
We then define in such a way that , which gives for .
We define by means of
(4.7) |
Since as linear spaces, we have
Further, the following holds.
Theorem 4.8.
We have
Here, by abuse of notation, we regard as a map
defined by
5. Proofs of the classification and construction
The aim of this section is to give proofs of Theorems 4.2 and 4.5 and Theorems 4.6 and 4.8. As the nilpotent radical is abelian, we achieve them simultaneously by proceeding with Steps 1–4 of the recipe of the F-method in Section 2.9.
5.1. Step 1: Compute and for
For and , we simply write
with . We put for the Euler homogeneity operator for .
Proposition 5.1.
For , we have
(5.2) |
Proof.
Since and via the diffeomorphism (4.3), this shows the proposition. ∎
For later convenience, we next give the formula for instead of . In the following, we silently extend the coordinate functions on in (5.2) holomorphically to the ones on as in Section 2.5.
For , we write for the Euler operator for . We also write for the Euler homogeneity operator for . Observe that we have .
Proposition 5.4.
For , we have
(5.5) |
Proof.
5.2. Step 2: Classify and construct
For , we write
for the representation space of . Then, in this step, we wish to classify and construct
5.2.1. Notation
We start by introducing some notation. For , we write
We then define by
(5.8) |
where are regarded as the dual basis of .
Recall from (2.24) that acts on via the action . In the present case, is an irreducible representation of , which is equivalent to
(5.9) |
Since maps , we have
(5.10) |
5.2.2. Classification and construction of
As
we first consider .
Since , it follows from (5.9) that the following two conditions on a representation of are equivalent.
-
(i)
.
-
(ii)
.
Now is given as follows.
Proposition 5.11.
The following three conditions on a representation of are equivalent.
-
(i)
.
-
(ii)
.
-
(iii)
.
Consequently, we have
(5.12) |
Proof.
A direct computation shows that we have
Here we mean by if ; if . Since , it implies
(5.13) |
Proposition 5.14.
We have
Proof.
5.3. Step 3: Solve the F-system for
For and , we put
(5.15) |
Theorem 5.17.
Let . The following conditions on are equivalent.
-
(i)
.
-
(ii)
One of the following two conditions holds.
-
(ii-a)
.
-
(ii-b)
.
-
(ii-a)
Further, the space is given as follows.
-
(1)
:
- (2)
Proof.
As for the -decomposition , we put
(5.20) |
Recall from (4.1) that we have
Corollary 5.21 below is then a direct consequence of Theorem 5.17.
Corollary 5.21.
5.4. Step 4ab: Apply and to the solutions to the F-system
5.5. Classification and construction of -homomorphisms
Theorem 4.8 concerns -homomorphisms between generalized Verma modules. Then we finish this section by showing the classification and construction of -homomorphisms.
Let be the identity component of the parabolic subgroup . Then we have . Thus, is given by
For , we define a generalized Verma module as in (3.5) as a -module. For , we let
Then, by Remark 2.10, we have
(5.22) |
Define as
Theorem 5.23.
We have
Proof.
Corollary 5.24.
The following are equivalent on .
-
(i)
is reducible.
-
(ii)
.
Proof.
Observe that is reducible if and only if there exists such that . Now the assertion follows from Theorem 5.23. ∎
6. -type formulas for and
The aim of this section is to classify the -type formulas of the kernel and image of the non-zero intertwining differential operator
The -type formulas are obtained in Corollary 6.7. We continue the notation and normalization from Section 3, unless otherwise specified.
Although the main idea in this section works for the case , to avoid the complication of the exposition, we constraint ourselves to the case .
6.1. Composition factors and -type structure of
Let be a maximal compact subgroup of . Then we have
We remark that .
For a representation of , we denote by the space of -finite vectors of . Since , as -representations, we have
Let be the irreducible representation of consisting of spherical harmonics on of homogeneous degree (cf. [43, Sect. 7.5]). It is known that the -type decomposition of is given as follows (see, for instance, [26, p. 286]).
(6.1) |
Theorem 6.2 below states well-known facts on the irreducibility and composition series of . For the proof, see, for instance, [26, 71] for and [62] for .
Theorem 6.2.
Let . For and , enjoys the following.
-
(1)
The induced representation is irreducible except the following two cases.
-
(A)
and .
-
(B)
and .
-
(A)
-
(2)
For Case (A) with , there exists a finite-dimensional irreducible subrepresentation such that is irreducible and infinite-dimensional. The -type formulas of are given as follows.
-
(3)
For Case (B) with , there exists an infinite-dimensional irreducible subrepresentation such that is irreducible and finite-dimensional. The -type formulas of are given as follows.
-
(4)
For , we have the non-split exact sequences of Fréchet -modules:
Remark 6.3.
In [71, Thm. 1.1], the condition “” in b) should be read as “”. For , the induced representation in the cited paper is irreducible, as it is dual to the case of .
Remark 6.4.
Now, for , we consider the -type formulas of the kernel and image of the intertwining differential operator
If , then and . Thus, in this case, we have
Therefore, the -type formula is given as in (6.1). For , we have the following.
Theorem 6.5.
For and , the kernel is given as follows.
In particular, the composition factors and of can be realized as
Proof.
Since the second assertion follows from the first and Theorem 6.2, it suffices to consider the first statement. As for , by Theorem 6.2, there are only two possibilities on , namely, or . It then suffices to show that as far as is reducible. In what follows, we understand that is a subspace of via the isomorphism (4.3) (see (2.1)). In particular, the Lie algebra acts on via the representation (see (2.4)).
Suppose that is reducible. Since is irreducible and finite-dimensional, there exists a lowest weight vector . We shall show that .
As being a lowest weight vector of , we have
(6.6) |
A direct computation shows that . Thus, (6.6) is equivalent to
which shows that is a constant function. Therefore, we have
Corollary 6.7.
For , the -type formulas of and are given as follows.
-
(1)
-
(1-a)
-
(1-b)
-
(1-a)
-
(2)
-
(2-a)
-
(2-b)
-
(2-a)
7. Appendix: the standardness of the homomorphism
The aim of this appendix is to show that the homomorphisms in (4.7) are all standard maps. We achieve this in Theorem 7.4.
7.1. Standard map
We start by introducing the definition of the standard map. Let be a complex simple Lie algebra. Fix a Cartan subalgebra and write for the set of roots of with respect to . Choose a positive system and denote by the set of simple roots of . Let denote the Borel subalgebra of associated with , namely, , where is the root space for .
Let denote the inner product on induced from a non-degenerate symmetric bilinear form of . For , we write . Also, write for the reflection with respect to . As usual, we let be half the sum of the positive roots.
Let be a standard parabolic subalgebra of . Write for the Levi decomposition of . We let .
Now we put
For , let be the finite-dimensional simple -module with highest weight . By letting act trivially, we regard as a -module. Then the induced module
is the generalized Verma module with highest weight . If , then is the (ordinary) Verma module with highest weight .
Let . It follows from a theorem by BGG-Verma (cf. [13, Thm. 7.6.23] and [28, Thm. 5.1]) that if , then .
Conversely, suppose that there exists a non-zero -homomorphism . Let denote the canonical projection map. Then we have ([60, Prop. 3.1]). Thus, the map induces a -homomorphism such that the following diagram commutes.
The map is called the standard map from to ([60]). As , the standard map is unique up to scalar. It is known that the standard map could be zero, and even if , there could be another non-zero map from to . Any homomorphism that is not standard is called a non-standard map.
It is known when the standard map is zero. To state the criterion, we first introduce the notion of a link between two weights.
Definition 7.1 (Bernstein–Gelfand–Gelfand).
Let and . Set and for . We say that the sequence links to if it satisfies the following two conditions.
-
(1)
.
-
(2)
for all .
The criterion on the vanishing of the standard map is first studied by Lepowsky ([60]) and then Boe refined Lepowsky’s criterion ([4]). Theorem 7.2 below is a version of Boe’s criterion [4, Thm. 3.3].
Theorem 7.2.
Let and suppose that . Then the following two conditions on are equivalent.
-
(i)
The standard map is non-zero.
-
(ii)
For all sequences linking to , we have .
7.2. The standardness of the homomorphism
Now we specialize the situation to the one considered in Section 3, that is, and is the maximal parabolic subalgebra corresponding to the partition . Observe that if , then . Thus, we assume that .
As usual, a Cartan subalgebra is taken to be
We take a set of positive roots and simple roots as
and
We realize as a subspace of and write elements in in coordinates. For instance, we write .
Recall from Theorem 5.23 that we have
(7.3) |
for , where is given in (4.7). Let be the differential of the character of defined in (3.3). Then
Define by
For , we have
For simplicity we let
We have
Equation (7.3) is then given as
Theorem 7.4.
Let . Then the homomorphism is standard for all .
Proof.
Since , it suffices to show the standard map is non-zero. We remark that as .
First, suppose that . In this case we have . Thus, the standard map is ; in particular, is non-zero.
Next, suppose that . In this case we have
Then is the only element that links to . Since , Theorem 7.2 shows that the standard map is non-zero. ∎
Acknowledgements. The authors are grateful to Dr. Ryosuke Nakahama and Dr. Masatoshi Kitagawa, and Prof. Toshiyuki Kobayashi for fruitful communication on this paper. They would also like to show their gratitude to the anonymous referees to review the article carefully. The first author was partially supported by JSPS Grant-in-Aid for Scientific Research(C) (JP22K03362).
References
- [1] Z. Bai and W. Xiao, Reducibility of generalized Verma modules for Hermitian symmetric pairs, J. Pure Appl. Algebra 225 (2021), Paper No. 106561, 21pp.
- [2] L. Barchini, A.C. Kable, and R. Zierau, Conformally invariant systems of differential equations and prehomogeneous vector spaces of Heisenberg parabolic type, Publ. Res. Inst. Math. Sci. 44 (2008), no. 3, 749–835.
- [3] by same author, Conformally invariant systems of differential operators, Adv. Math. 221 (2009), no. 3, 788–811.
- [4] B.D. Boe, Homomorphisms between generalized Verma modules, Trans. Amer. Math. Soc. 288 (1985), 791–799.
- [5] G. Bol, Invarianten linearer differentialgleichungen, Abh. Math. Sem. Univ. Hamburg 16 (1949), 1–28.
- [6] S. Bouarroudj, D. Leites, and I. Shchepochkina, Analogs of Bol operators for , Internat. J. Algebra Comput. 32 (2022), 1345–1368.
- [7] by same author, Analogs of Bol operators on superstrings, Internat. J. Algebra Comput. 32 (2022), 807–835.
- [8] D.M.J. Calderbank and T. Diemer, Differential invariants and curved Bernstein-Gelfand-Gelfand sequences, J. Reine Angew. Math. 537 (2001), 67–103.
- [9] A. Čap, A.R. Gover, and M. Hammerl, Normal BGG solutions and polynomials, Internat. J. Math. 23 (2012), 1250117, 29 pp.
- [10] by same author, Projective BGG equations, algebraic sets, and compactifications of Einstein geometries, J. Lond. Math. Soc. (2) 86 (2012), 433–454.
- [11] A. Čap, J. Slovák, and V. Souček, Bernstein-Gelfand-Gelfand sequences, Ann. of Math. (2) 154 (2001), 97–113.
- [12] D.H. Collingwood and B. Shelton, A duality theorem for extensions of induced highest weight modules, Pacific J. Math. 146 (1990), no. 2, 227–237.
- [13] J. Dixmier, Algèbres enveloppantes, Gauthier-Villars, Paris, 1974, reprint of English translation, Enveloping Algebras, American Mathematical Society, Providence, RI, 1996.
- [14] V.K. Dobrev, Canonical construction of differential operators intertwining representations of real semisimple Lie groups, Rep. Math. Phys. 25 (1988), no. 2, 159–181, first as ICTP Trieste preprint IC/86/393 (1986).
- [15] by same author, Invariant Differential Operators for Non-Compact Lie Algebras Parabolically Related to Conformal Lie Algebras, J. High Energy Phys. (2013), 42 pp, 015, front matter + 40 pp.
- [16] by same author, Invariant differential operators. Vol. 1. Noncompact semisimple Lie algebras and groups, De Gruyter Studies in Mathematical Physics, no. 35, De Gruyter, Berlin, 2016.
- [17] by same author, Invariant differential operators. Vol. 2. Quantum groups, De Gruyter Studies in Mathematical Physics, no. 39, De Gruyter, Berlin, 2017.
- [18] by same author, Invariant differential operators. Vol. 3. Supersymmetry, De Gruyter Studies in Mathematical Physics, no. 49, De Gruyter, Berlin, 2018.
- [19] by same author, Invariant differential operators. Vol. 4. AdS/CFT, (super-)Virasoro, affine (super-)algebras, De Gruyter Studies in Mathematical Physics, no. 53, De Gruyter, Berlin, 2019.
- [20] H.D. Fegan, Conformally invariant first order differential operators, Quart. J. Math. Oxford Ser. (2) 27 (1976), 371–378.
- [21] M. Fischmann, A. Juhl, and P. Somberg, Conformal Symmetry Breaking Differential Operators on Differential Forms, Mem. Amer. Math. Soc., vol. 268, Amer. Math. Soc., Providence, RI, 2020.
- [22] J. Gregorovič and L. Zalabová, First BGG operators via homogeneous examples, J. Geom. Phys. 192 (2023), Paper No. 104953, 33 pp.
- [23] M. Harris and H.P. Jakobsen, Singular holomorphic representations and singular modular forms, Math. Ann. 259 (1982), 227–244.
- [24] H. He, On the reducibility of scalar generalized Verma modules of abelian type, Algebr. Represent. Theory 19 (2016), 147–170.
- [25] H. He, T. Kubo, and R. Zierau, On the reducibility of scalar generalized Verma modules associated to maximal parabolic subalgebras, Kyoto J. Math. 59 (2019), 787–813.
- [26] R. Howe and S.T. Lee, Degenerate Principal Series Representations of and , J. Funct. Anal. 166 (2019), 244–309.
- [27] J.-S. Huang, Intertwining differential operators and reducibility of generalized Verma modules, Math. Ann. 297 (1993), 309–324.
- [28] J.E. Humphreys, Representations of semisimple Lie algebras in the BGG category , Graduate Studies in Mathematics, vol. 94, American Mathematical Society, Rhode Island, 2008.
- [29] K.D. Johnson, A. Korányi, and H.M. Reimann, Equivariant first order differential operators for parabolic geometries, Indag. Math. (N.S.) 14 (2003), 385–393.
- [30] A. Juhl, Families of Conformally Covariant Differential Operators, -Curvature and Holography, Progr. Math., vol. 275, Birkhäuser, Basel, 2009.
- [31] A.C. Kable, -finite solutions to conformally invariant systems of differential equations, Tohoku Math. J. 63 (2011), no. 4, 539–559, Centennial Issue.
- [32] by same author, Conformally invariant systems of differential equations on flag manifolds for and their K-finite solutions, J. Lie Theory 22 (2012), no. 1, 93–136.
- [33] by same author, The Heisenberg ultrahyperbolic equation: -finite and polynomial solutions, Kyoto J. Math 52 (2012), no. 4, 839–894.
- [34] by same author, On certain conformally invariant systems of differential equations, New York J. Math 19 (2013), 189–251.
- [35] by same author, On certain conformally invariant systems of differential equations II: Further study of type A systems, Tsukuba J. Math. 39 (2015), 39–81.
- [36] by same author, A note on the construction of second-order conformally invariant systems on generalized flag manifolds, J. Lie Theory 28 (2018), 969–985.
- [37] by same author, The structure of the space of polynomial solutions to the canonical central systems of differential equations on the block Heisenberg groups: A generalization of a theorem of Korányi , Tohoku Math. J. 70 (2018), 523–545.
- [38] T. Kobayashi, F-method for constructing equivariant differential operators, Geometric analysis and integral geometry, Contemp. Math., vol. 598, Amer. Math. Soc., 2013, pp. 139–146.
- [39] by same author, F-method for symmetry breaking operators, Diff. Geometry and its Appl. 33 (2014), 272–289.
- [40] by same author, A program for branching problems in the representation theory of real reductive groups, Representations of reductive groups, Progr. Math., vol. 312, Birkhäuser/Springer, 2015, pp. 277–322.
- [41] by same author, Residue formula for regular symmetry breaking operators, Representation theory and harmonic analysis on symmetric spaces, Contemp. Math., vol. 714, Amer. Math. Soc., 2018, pp. 175–197.
- [42] T. Kobayashi, T. Kubo, and M. Pevzner, Conformal symmetry breaking operators for differential forms on spheres, Lecture Notes in Math., vol. 2170, Springer-Nature, Singapore, 2016.
- [43] T. Kobayashi and G. Mano, The Schrödinger model for the minimal representation of the indefinite orthogonal group , Mem. Amer. Math. Soc. 213 (2011), no. 1000, vi+132.
- [44] T. Kobayashi and M. Pevzner, Differential symmetry breaking operators. I. General theory and F-method, Selecta Math. (N.S.) 22 (2016), 801–845.
- [45] by same author, Differential symmetry breaking operators. II. Rankin–Cohen operators for symmetric pairs, Selecta Math. (N.S.) 22 (2016), 847–911.
- [46] T. Kobayashi and B. Ørsted, Analysis on the minimal representation of I. Realization via conformal geometry, Adv. Math. 180 (2003), no. 2, 486–512.
- [47] T. Kobayashi, B. Ørsted, P. Somberg, and V. Souček, Branching laws for Verma modules and applications in parabolic geometry. I, Adv. Math. 285 (2015), 1796–1852.
- [48] T. Kobayashi and B. Speh, Symmetry breaking for representations of rank one orthogonal groups II, Lecture Notes in Math., vol. 2234, Springer-Nature, Singapore, 2018.
- [49] A. Korányi and H.M. Reimann, Equivariant first order differential operators on boundaries of symmetric spaces, Invent. Math. 139 (2000), no. 2, 371–390.
- [50] B. Kostant, Verma modules and the existence of quasi-invariant differential operators, Noncommutative harmonic analysis, Lecture Notes in Math., vol. 466, Springer, 1975, pp. 101–128.
- [51] L. Křižka and P. Somberg, Algebraic analysis of scalar generalized Verma modules of Heisenberg parabolic type I: -series, Transform. Groups 22 (2017), no. 2, 403–451.
- [52] by same author, Differential invariants on symplectic spinors in contact projective geometry, J. Math. Phys. 58 (2017), 091701, 22 pp.
- [53] by same author, Equivariant differential operators on spinors in conformal geometry, Complex Var. Elliptic Equ. 62 (2017), 583–599.
- [54] T. Kubo, Special values for conformally invariant systems associated to maximal parabolics of quasi-Heisenberg type, Trans. Amer. Math. Soc. 366 (2014), 4649–4696.
- [55] by same author, Systems of differential operators and generalized Verma modules, SIGMA Symmetry Integrability Geom. Methods Appl. 10 (2014), Paper 008, 35pp.
- [56] by same author, On the classification and construction of intertwining differential operators by the F-method, Proceedings of Symposium on Representation Theory 2022, 2022, in Japanese, pp. 41–65.
- [57] T. Kubo and B. Ørsted, Classification of -type formulas for the Heisenberg ultrahyperbolic operator for and tridiagonal determinants for local Heun functions, to appear in Festschrift in honour of Toshiyuki Kobayashi.
- [58] by same author, On the space of -finite solutions to intertwining differential operators, Represent. Theory 23 (2019), 213–248.
- [59] J.M. Lee and T.H. Parker, The Yamabe problem, Bull. Amer. Math. Soc. (N.S.) 17 (1987), 37–91.
- [60] J. Lepowsky, A generalization of the Bernstein–Gelfand–Gelfand resolution, J. Algebra 49 (1977), 496–511.
- [61] T. Milev and P. Somberg, The F-method and a branching problem for generalized Verma modules associated to , Arch. Math. (Brno) 49 (2013), 317–332.
- [62] J. Möllers and B. Schwarz, Structure of the degenerate principal series on symmetric -spaces and small representations, J. Funct. Anal. 266 (2014), 3508–3542.
- [63] B. Ørsted, Generalized gradients and Poisson transforms, Global analysis and harmonic analysis (Marseille-Luminy, 1999), Semin. Congr., 4, Soc. Math. France, Paris, 2000, pp. 235–249.
- [64] V. Ovsienko and S. Tabachnikov, Projective differential geometry old and new, Cambridge Tracts in Mathematics, vol. 165, Cambridge University Press, Cambridge, 2005.
- [65] V. Pérez-Valdés, Conformally covariant differential symmetry breaking operators for a vector bundle of rank 3 over , Internat. J. Math. 34 (2023), Paper No. 2350072, 58 pp.
- [66] J. Slovák and V. Souček, Invariant operators of the first order on manifolds with a given parabolic structure, Global analysis and harmonic analysis (Marseille-Luminy, 1999), Semin. Congr., 4, Soc. Math. France, Paris, 2000, pp. 251–276.
- [67] P. Somberg, Rankin–Cohen brackets for orthogonal Lie algebras and bilinear conformally invariant differential operators, arXiv:1301.2687, preprint (2013).
- [68] H. Tamori, Classification of irreducible -modules associated to the ideals of minimal nilpotent orbits for simple Lie groups of type , Int. Math. Res. Not. IMRN, rnab356 (2021).
- [69] Y. Tanimura, An algebraic approach to Duflo’s polynomial conjecture in the nilpotent case, J. Lie Theory 29 (2019), 839–879.
- [70] G. van Dijk, Distribution theory. Convolution, Fourier transform, and Laplace transform, De Gruyter Grad. Lect., De Gruyter, Berlin, 2013.
- [71] G. van Dijk and V.F. Molchanov, Tensor products of maximal degenerate series representations of the group , J. Math. Pures Appl. 78 (1999), 99–119.
- [72] W. Xiao, Leading weight vectors and homomorphisms between Verma modules, J. Algebra 430 (2015), 62–93.