This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On the intertwining differential operators from a line bundle to a vector bundle over the real projective space

Toshihisa Kubo  and  Bent Ørsted Faculty of Economics, Ryukoku University, 67 Tsukamoto-cho, Fukakusa, Fushimi-ku, Kyoto 612-8577, Japan toskubo@econ.ryukoku.ac.jp Department of Mathematics, Aarhus University, Ny Munkegade 118 DK-8000 Aarhus C, Denmark orsted@imf.au.dk
Abstract.

We classify and construct SL(n,)SL(n,\mathbb{R})-intertwining differential operators 𝒟\mathcal{D} from a line bundle to a vector bundle over the real projective space n1\mathbb{R}\mathbb{P}^{n-1} by the F-method. This generalizes a classical result of Bol for SL(2,)SL(2,\mathbb{R}). Further, we classify the KK-type formulas for the kernel Ker(𝒟)\mathrm{Ker}(\mathcal{D}) and image Im(𝒟)\mathrm{Im}(\mathcal{D}) of 𝒟\mathcal{D}. The standardness of the homomorphisms φ\varphi corresponding to the differential operators 𝒟\mathcal{D} between generalized Verma modules are also discussed.

Key words and phrases:
Bol operator, intertwining differential operator, generalized Verma module, F-method, KK-type formula, standard map
2020 Mathematics Subject Classification:
22E46, 17B10

1. Introduction

Gerrit van Dijk worked on harmonic analysis on pp-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 PP to a reductive Lie group GG and the explicit structure of such in concrete cases. In particular, he jointly with Molchanov studied the case of G/PG/P 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 GG be a real reductive Lie group with complexified Lie algebra 𝔤{\mathfrak{g}}. Let PP be a parabolic subgroup with Langlands decomposition P=MAN+P=MAN_{+}. We write Irr(M)fin\mathrm{Irr}(M)_{\mathrm{fin}} for the set of equivalence classes of finite-dimensional irreducible representations of MM. Likewise, let Irr(A)\mathrm{Irr}(A) denote the set of characters of AA. Then, for the outer tensor product ϖνtriv\varpi\boxtimes\nu\boxtimes\mathrm{triv} of ϖIrr(M)fin\varpi\in\mathrm{Irr}(M)_{\mathrm{fin}}, νIrr(A)\nu\in\mathrm{Irr}(A), and the trivial representation triv\mathrm{triv} of N+N_{+}, we put

I(ϖ,ν)=IndPG(ϖνtriv)I(\varpi,\nu)=\mathrm{Ind}_{P}^{G}(\varpi\boxtimes\nu\boxtimes\mathrm{triv})

for an (unnormalized) parabolically induced representation of GG. Let DiffG(I(σ,λ),I(ϖ,ν))\mathrm{Diff}_{G}(I(\sigma,\lambda),I(\varpi,\nu)) denote the space of GG-intertwining differential operators 𝒟:I(σ,λ)I(ϖ,ν)\mathcal{D}\colon I(\sigma,\lambda)\to I(\varpi,\nu).

There are two main problems in this paper. The first problem concerns the classification and construction of intertwining differential operators 𝒟DiffG(I(σ,λ),I(ϖ,ν))\mathcal{D}\in\mathrm{Diff}_{G}(I(\sigma,\lambda),I(\varpi,\nu)) as follows.

Problem A.

Do the following.

  1. (A1)

    Classify (σ,λ),(ϖ,ν)Irr(M)fin×Irr(A)(\sigma,\lambda),(\varpi,\nu)\in\mathrm{Irr}(M)_{\mathrm{fin}}\times\mathrm{Irr}(A) such that

    DiffG(I(σ,λ),I(ϖ,ν)){0}.\mathrm{Diff}_{G}(I(\sigma,\lambda),I(\varpi,\nu))\neq\{0\}.
  2. (A2)

    For (σ,λ),(ϖ,ν)Irr(M)fin×Irr(A)(\sigma,\lambda),(\varpi,\nu)\in\mathrm{Irr}(M)_{\mathrm{fin}}\times\mathrm{Irr}(A) classified in (A1), determine the dimension

    dimDiffG(I(σ,λ),I(ϖ,ν)).\dim\mathrm{Diff}_{G}(I(\sigma,\lambda),I(\varpi,\nu)).
  3. (A3)

    For (σ,λ),(ϖ,ν)Irr(M)fin×Irr(A)(\sigma,\lambda),(\varpi,\nu)\in\mathrm{Irr}(M)_{\mathrm{fin}}\times\mathrm{Irr}(A) classified in (A1), construct generators

    𝒟DiffG(I(σ,λ),I(ϖ,ν)).\mathcal{D}\in\mathrm{Diff}_{G}(I(\sigma,\lambda),I(\varpi,\nu)).

Let KK be a maximal compact subgroup of GG. For 𝒟DiffG(I(σ,λ),I(ϖ,ν))\mathcal{D}\in\mathrm{Diff}_{G}(I(\sigma,\lambda),I(\varpi,\nu)), the kernel Ker(𝒟)\mathrm{Ker}(\mathcal{D}) and image Im(𝒟)\mathrm{Im}(\mathcal{D}) are GG-invariant subspaces of I(σ,λ)I(\sigma,\lambda) and I(ϖ,ν)I(\varpi,\nu), respectively. Let Ker(𝒟)K\mathrm{Ker}(\mathcal{D})_{K} and Im(𝒟)K\mathrm{Im}(\mathcal{D})_{K} denote the space of KK-finite vectors of Ker(𝒟)\mathrm{Ker}(\mathcal{D}) and Im(𝒟)\mathrm{Im}(\mathcal{D}), respectively. The following is the other main problem of this paper.

Problem B.

Classify the KK-type formulas of Ker(𝒟)K\mathrm{Ker}(\mathcal{D})_{K} and Im(𝒟)K\mathrm{Im}(\mathcal{D})_{K}.

For instance, if G=SL(3,)G=SL(3,\mathbb{R}), then the maximal compact subgroup KK is K=SO(3)K=SO(3). Problem B then asks how Ker(𝒟)K\mathrm{Ker}(\mathcal{D})_{K} and Im(𝒟)K\mathrm{Im}(\mathcal{D})_{K} decompose as SO(3)SO(3)-modules. We shall answer this question in Corollary 6.7 for SL(n,)SL(n,\mathbb{R}) with n3n\geq 3.

Problem A for the first order intertwining differential operators for general (G,P)(G,P) was done around 2000 independently by Ørsted [63], Slovák–Souček [66], and Johnson–Korányi–Reimann [29], which generalize the work of Fegan [20] for G=SO(n,1)G=SO(n,1). 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 (SL(n,),P1,n1;(±,triv))(SL(n,\mathbb{R}),P_{1,n-1};(\pm,\mathrm{triv}))

In this paper we consider Problems A and B for (G,P;σ)=(SL(n,),P1,n1;(±,triv))(G,P;\sigma)=(SL(n,\mathbb{R}),P_{1,n-1};(\pm,\mathrm{triv})), where P1,n1P_{1,n-1} is the maximal parabolic subgroup of GG corresponding to the partition n=1+(n1)n=1+(n-1) so that G/P1,n1G/P_{1,n-1} is diffeomorphic to the real projective space n1\mathbb{R}\mathbb{P}^{n-1}. The representation (±,triv)Irr(M)fin(\pm,\mathrm{triv})\in\mathrm{Irr}(M)_{\mathrm{fin}} denotes a one-dimensional representation of MSL±(n1,)M\simeq SL^{\pm}(n-1,\mathbb{R}). 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 SL(2,)SL(2,\mathbb{R})-intertwining differential operators for an equivariant line bundle over S1S^{1} are only the powers of the standard derivative dkdxk\frac{d^{k}}{dx^{k}} (cf. [5] and [64, Thm. 2.1.2]). The operator dkdxk\frac{d^{k}}{dx^{k}} 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 S1S^{1}, which is a double cover of 1\mathbb{R}\mathbb{P}^{1}, to SL(n,)SL(n,\mathbb{R}) on n1\mathbb{R}\mathbb{P}^{n-1}.

It is classically known that the Bol operators dkdxk\frac{d^{k}}{dx^{k}} are the residue operators of the Knapp–Stein operator. In contrast to the case of n=2n=2, the Knapp–Stein operator does not exist for (SL(n,),P1,n1)(SL(n,\mathbb{R}),P_{1,n-1}) for n3n\geq 3 (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 n3n\geq 3.

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 DiffG(I(σ,λ),I(ϖ,ν))\mathrm{Diff}_{G}(I(\sigma,\lambda),I(\varpi,\nu)) of intertwining differential operators is isomorphic to the space of (𝔤,P)({\mathfrak{g}},P)-homomorphisms between generalized Verma modules (see also [12, 14, 23, 27, 49]). Schematically, we have

Hom(Verma)DiffG(I(σ,λ),I(ϖ,ν)).\textnormal{Hom}(\text{Verma})\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\mathrm{Diff}_{G}(I(\sigma,\lambda),I(\varpi,\nu)). (1.1)

The precise statement will be given in Theorem 2.5. In general, it is easier to work with the Verma module side “Hom(Verma)\operatorname{Hom}(\text{Verma})” than the differential operator side DiffG(I(σ,λ),I(ϖ,ν))\mathrm{Diff}_{G}(I(\sigma,\lambda),I(\varpi,\nu)). 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 N+N_{+} is abelian, it requires involved combinatorial computations (see, for instance, [30, Chap. 5]).

The novel idea of the F-method is to further identify DiffG(I(σ,λ),I(ϖ,ν))\mathrm{Diff}_{G}(I(\sigma,\lambda),I(\varpi,\nu)) 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 “Sol(PDE)\textnormal{Sol}(\text{PDE})” to (1.1) as follows.

Sol(PDE)\textstyle{\textnormal{Sol}(\text{PDE})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}Hom(Verma)\textstyle{\textnormal{Hom}(\text{Verma})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\hskip 10.0pt\sim}\scriptstyle{\sim}DiffG(I(σ,λ),I(ϖ,ν)).\textstyle{\mathrm{Diff}_{G}(I(\sigma,\lambda),I(\varpi,\nu)).} (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 (𝔤,P)({\mathfrak{g}},P)-homomorphisms in “Hom(Verma)\textnormal{Hom}(\text{Verma})” from the solutions in “Sol(PDE)\textnormal{Sol}(\text{PDE})”. We thus study the algebraic counterpart of Problem A for (𝔤,P)=(𝔰𝔩(n,),P1,n1)({\mathfrak{g}},P)=(\mathfrak{sl}(n,\mathbb{C}),P_{1,n-1}). It is achieved in Theorems 4.6 and 4.8. We further consider a variant of Problem A for 𝔤{\mathfrak{g}}-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 (𝔤,𝔭)=(𝔰𝔩(n,),𝔭1,n1)({\mathfrak{g}},{\mathfrak{p}})=(\mathfrak{sl}(n,\mathbb{C}),{\mathfrak{p}}_{1,n-1}).

A 𝔤{\mathfrak{g}}-homomorphism between generalized Verma modules is called a standard map if it is induced from a 𝔤{\mathfrak{g}}-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 𝔤{\mathfrak{g}}-homomorphisms in Theorem 5.23 are all standard.

1.5. The KK-type formulas for Ker(𝒟)K\mathrm{Ker}(\mathcal{D})_{K} and Im(𝒟)K\mathrm{Im}(\mathcal{D})_{K}

Kable [31] and the authors [58] recently showed a Peter–Weyl type theorem for the kernel Ker(𝒟)\mathrm{Ker}(\mathcal{D}) of an intertwining differential operator 𝒟\mathcal{D}. The theorem allows us to compute the KK-type formula of Ker(𝒟)\mathrm{Ker}(\mathcal{D}) explicitly by solving the hypergeometric/Heun differential equation ([57, 58]). Tamori [68] independently used a similar idea to determine the KK-type formula of Ker(𝒟)\mathrm{Ker}(\mathcal{D}) 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 FF. (See Möllers–Schwarz [62] for recent development of this matter.) Since the KK-type structures of the induced representations are also known, the determination of the KK-type formula of Ker(𝒟)\mathrm{Ker}(\mathcal{D}) is equivalent to showing Ker(𝒟){0}\mathrm{Ker}(\mathcal{D})\neq\{0\}. As the length is two, it simultaneously determines the KK-type formula of Im(𝒟)\mathrm{Im}(\mathcal{D}).

In order to show that Ker(𝒟){0}\mathrm{Ker}(\mathcal{D})\neq\{0\}, we show 𝒟f0=0\mathcal{D}f_{0}=0 for a lowest weight vector f0Ff_{0}\in F. We shall discuss the details in Section 6. We remark that one can also show Ker(𝒟){0}\mathrm{Ker}(\mathcal{D})\neq\{0\} by using the lowest KK-type in [33] for n=3n=3.

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 (G,P)=(SL(n,),P1,n1)(G,P)=(SL(n,\mathbb{R}),P_{1,n-1}). 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 (G,P;σ)=(SL(n,),P1,n1;(±,triv))(G,P;\sigma)=(SL(n,\mathbb{R}),P_{1,n-1};(\pm,\mathrm{triv})). These are accomplished in Theorems 4.2 and 4.5. Further, a variant of Problem A for (𝔤,P)({\mathfrak{g}},P)-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 KK-type formulas of the kernel Ker(𝒟)\mathrm{Ker}(\mathcal{D}) and the image Im(𝒟)\mathrm{Im}(\mathcal{D}) of the intertwining differential operators 𝒟\mathcal{D} classified in Section 4. These are obtained in Corollary 6.7.

Section 7 is an appendix discussing the standardness of the 𝔤{\mathfrak{g}}-homomorphisms φ\varphi 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 φ\varphi 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 𝒟\mathcal{D}. We mainly follow the arguments in [41] and [44] in this section.

2.1. General framework

Let GG be a real reductive Lie group and P=MAN+P=MAN_{+} a Langlands decomposition of a parabolic subgroup PP of GG. We denote by 𝔤(){\mathfrak{g}}(\mathbb{R}) and 𝔭()=𝔪()𝔞()𝔫+(){\mathfrak{p}}(\mathbb{R})={\mathfrak{m}}(\mathbb{R})\oplus{\mathfrak{a}}(\mathbb{R})\oplus{\mathfrak{n}}_{+}(\mathbb{R}) the Lie algebras of GG and P=MAN+P=MAN_{+}, respectively.

For a real Lie algebra 𝔶()\mathfrak{y}(\mathbb{R}), we write 𝔶\mathfrak{y} and 𝒰(𝔶)\mathcal{U}({\mathfrak{y}}) for its complexification and the universal enveloping algebra of 𝔶{\mathfrak{y}}, respectively. For instance, 𝔤,𝔭,𝔪,𝔞{\mathfrak{g}},{\mathfrak{p}},{\mathfrak{m}},{\mathfrak{a}}, and 𝔫+{\mathfrak{n}}_{+} are the complexifications of 𝔤(),𝔭(),𝔪(),𝔞(){\mathfrak{g}}(\mathbb{R}),{\mathfrak{p}}(\mathbb{R}),{\mathfrak{m}}(\mathbb{R}),{\mathfrak{a}}(\mathbb{R}), and 𝔫+(){\mathfrak{n}}_{+}(\mathbb{R}), respectively.

For λ𝔞Hom(𝔞(),)\lambda\in{\mathfrak{a}}^{*}\simeq\operatorname{Hom}_{\mathbb{R}}({\mathfrak{a}}(\mathbb{R}),\mathbb{C}), we denote by λ\mathbb{C}_{\lambda} the one-dimensional representation of AA defined by aaλ:=eλ(loga)a\mapsto a^{\lambda}:=e^{\lambda(\log a)}. For a finite-dimensional irreducible representation (σ,V)(\sigma,V) of MM and λ𝔞\lambda\in{\mathfrak{a}}^{*}, we denote by σλ\sigma_{\lambda} the outer tensor representation σλ\sigma\boxtimes\mathbb{C}_{\lambda}. As a representation on VV, we define σλ:maaλσ(m)\sigma_{\lambda}\colon ma\mapsto a^{\lambda}\sigma(m). By letting N+N_{+} act trivially, we regard σλ\sigma_{\lambda} as a representation of PP. Let 𝒱:=G×PVG/P\mathcal{V}:=G\times_{P}V\to G/P be the GG-equivariant vector bundle over the real flag variety G/PG/P associated with the representation (σλ,V)(\sigma_{\lambda},V) of PP. We identify the Fréchet space C(G/P,𝒱)C^{\infty}(G/P,\mathcal{V}) of smooth sections with

C(G,V)P:={fC(G,V):f(gp)=σλ1(p)f(g)for any pP},C^{\infty}(G,V)^{P}:=\{f\in C^{\infty}(G,V):f(gp)=\sigma_{\lambda}^{-1}(p)f(g)\;\;\text{for any $p\in P$}\},

the space of PP-invariant smooth functions on GG. Then, via the left regular representation LL of GG on C(G)C^{\infty}(G), we realize the parabolically induced representation π(σ,λ)=IndPG(σλ)\pi_{(\sigma,\lambda)}=\mathrm{Ind}_{P}^{G}(\sigma_{\lambda}) on C(G/P,𝒱)C^{\infty}(G/P,\mathcal{V}). We denote by RR the right regular representation of GG on C(G)C^{\infty}(G).

Similarly, for a finite-dimensional irreducible representation (ην,W)(\eta_{\nu},W) of MAMA, we define an induced representation π(η,ν)=IndPG(ην)\pi_{(\eta,\nu)}=\mathrm{Ind}_{P}^{G}(\eta_{\nu}) on the space C(G/P,𝒲)C^{\infty}(G/P,\mathcal{W}) of smooth sections for a GG-equivariant vector bundle 𝒲:=G×PWG/P\mathcal{W}:=G\times_{P}W\to G/P. We write DiffG(𝒱,𝒲)\mathrm{Diff}_{G}(\mathcal{V},\mathcal{W}) for the space of intertwining differential operators 𝒟:C(G/P,𝒱)C(G/P,𝒲)\mathcal{D}\colon C^{\infty}(G/P,\mathcal{V})\to C^{\infty}(G/P,\mathcal{W}).

Let 𝔤()=𝔫()𝔪()𝔞()𝔫+()\mathfrak{g}(\mathbb{R})=\mathfrak{n}_{-}(\mathbb{R})\oplus\mathfrak{m}(\mathbb{R})\oplus\mathfrak{a}(\mathbb{R})\oplus\mathfrak{n}_{+}(\mathbb{R}) be the Gelfand–Naimark decomposition of 𝔤()\mathfrak{g}(\mathbb{R}), and write N=exp(𝔫())N_{-}=\exp({\mathfrak{n}}_{-}(\mathbb{R})). We identify NN_{-} with the open Bruhat cell NPN_{-}P of G/PG/P via the embedding ι:NG/P\iota\colon N_{-}\hookrightarrow G/P, n¯n¯P\bar{n}\mapsto\bar{n}P. Via the restriction of the vector bundle 𝒱G/P\mathcal{V}\to G/P to the open Bruhat cell NιG/PN_{-}\stackrel{{\scriptstyle\iota}}{{\hookrightarrow}}G/P, we regard C(G/P,𝒱)C^{\infty}(G/P,\mathcal{V}) as a subspace of C(N)VC^{\infty}(N_{-})\otimes V.

We view intertwining differential operators 𝒟:C(G/P,𝒱)C(G/P,𝒲)\mathcal{D}\colon C^{\infty}(G/P,\mathcal{V})\to C^{\infty}(G/P,\mathcal{W}) as differential operators 𝒟:C(N)VC(N)W\mathcal{D}^{\prime}\colon C^{\infty}(N_{-})\otimes V\to C^{\infty}(N_{-})\otimes W such that the restriction 𝒟|C(G/P,𝒱)\mathcal{D}^{\prime}|_{C^{\infty}(G/P,\mathcal{V})} to C(G/P,𝒱)C^{\infty}(G/P,\mathcal{V}) is a map 𝒟|C(G/P,𝒱):C(G/P,𝒱)C(G/P,𝒲)\mathcal{D}^{\prime}|_{C^{\infty}(G/P,\mathcal{V})}\colon C^{\infty}(G/P,\mathcal{V})\to C^{\infty}(G/P,\mathcal{W}) (see (2.1) below).

C(N)V\textstyle{C^{\infty}(N_{-})\otimes V\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟\scriptstyle{\mathcal{D}^{\prime}}C(N)W\textstyle{C^{\infty}(N_{-})\otimes W}C(G/P,𝒱)\textstyle{C^{\infty}(G/P,\mathcal{V})\;\;\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟=𝒟|C(G/P,𝒱)\scriptstyle{\stackrel{{\scriptstyle\phantom{a}}}{{\hskip 20.0pt\mathcal{D}=\mathcal{D}^{\prime}|_{\small{C^{\infty}(G/P,\mathcal{V})}}}}}ι\scriptstyle{\iota^{*}}C(G/P,𝒲)\textstyle{\;\;C^{\infty}(G/P,\mathcal{W})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι\scriptstyle{\iota^{*}} (2.1)

In particular, we regard DiffG(𝒱,𝒲)\mathrm{Diff}_{G}(\mathcal{V},\mathcal{W}) as

DiffG(𝒱,𝒲)Diff(C(N)V,C(N)W).\mathrm{Diff}_{G}(\mathcal{V},\mathcal{W})\subset\mathrm{Diff}_{\mathbb{C}}(C^{\infty}(N_{-})\otimes V,C^{\infty}(N_{-})\otimes W). (2.2)

To describe the F-method, one needs to introduce the following:

  1. (1)

    infinitesimal representation dπ(σ,λ)d\pi_{(\sigma,\lambda)} (Section 2.2),

  2. (2)

    duality theorem (Section 2.3),

  3. (3)

    algebraic Fourier transform ^\widehat{\;\;\cdot\;\;} of Weyl algebras (Section 2.4),

  4. (4)

    Fourier transformed representation dπ(σ,λ)^\widehat{d\pi_{(\sigma,\lambda)^{*}}} (Section 2.5),

  5. (5)

    algebraic Fourier transform FcF_{c} of generalized Verma modules (Section 2.6).

After reviewing these objects, we shall discuss the F-method in Section 2.7.

2.2. Infinitesimal representation dπ(σ,λ)d\pi_{(\sigma,\lambda)}

We start with the representation dπ(σ,λ)d\pi_{(\sigma,\lambda)} of 𝔤{\mathfrak{g}} on C(N)VC^{\infty}(N_{-})\otimes V derived from the induced representation π(σ,λ)\pi_{(\sigma,\lambda)} of GG on C(G/P,𝒱)C^{\infty}(G/P,\mathcal{V}).

For a representation η\eta of GG, we denote by dηd\eta the infinitesimal representation of 𝔤(){\mathfrak{g}}(\mathbb{R}). For instance, dLdL and dRdR denote the infinitesimal representations 𝔤(){\mathfrak{g}}(\mathbb{R}) of the left and right regular representations of GG on C(G)C^{\infty}(G). As usual, we naturally extend representations of 𝔤(){\mathfrak{g}}(\mathbb{R}) to ones for its universal enveloping algebra 𝒰(𝔤)\mathcal{U}({\mathfrak{g}}) of its complexification 𝔤{\mathfrak{g}}. The same convention is applied for closed subgroups of GG.

For gNMAN+g\in N_{-}MAN_{+}, we write

g=p(g)p0(g)p+(g),g=p_{-}(g)p_{0}(g)p_{+}(g),

where p±(g)N±p_{\pm}(g)\in N_{\pm} and p0(g)MAp_{0}(g)\in MA. Similarly, for Y𝔤=𝔫𝔩𝔫+Y\in{\mathfrak{g}}={\mathfrak{n}}_{-}\oplus{\mathfrak{l}}\oplus{\mathfrak{n}}_{+} with 𝔩=𝔪𝔞{\mathfrak{l}}={\mathfrak{m}}\oplus{\mathfrak{a}}, we write

Y=Y𝔫+Y𝔩+Y𝔫+,Y=Y_{{\mathfrak{n}}_{-}}+Y_{{\mathfrak{l}}}+Y_{{\mathfrak{n}}_{+}},

where Y𝔫±𝔫±Y_{{\mathfrak{n}}_{\pm}}\in{\mathfrak{n}}_{\pm} and Y𝔩𝔩Y_{\mathfrak{l}}\in{\mathfrak{l}}.

Let X𝔤()X\in{\mathfrak{g}}(\mathbb{R}). Since NMAN+N_{-}MAN_{+} is an open dense subset of GG, we have exp(tX)NMAN+\exp(tX)\in N_{-}MAN_{+} for sufficiently small t>0t>0. Thus, we have

X\displaystyle X =ddt|t=0exp(tX)\displaystyle=\frac{d}{dt}\bigg{|}_{t=0}\exp(tX)
=ddt|t=0p(exp(tX))p0(exp(tX))p+(exp(tX))\displaystyle=\frac{d}{dt}\bigg{|}_{t=0}p_{-}(\exp(tX))p_{0}(\exp(tX))p_{+}(\exp(tX))
=ddt|t=0p(exp(tX))+ddt|t=0p0(exp(tX))+ddt|t=0p+(exp(tX)),\displaystyle=\frac{d}{dt}\bigg{|}_{t=0}p_{-}(\exp(tX))+\frac{d}{dt}\bigg{|}_{t=0}p_{0}(\exp(tX))+\frac{d}{dt}\bigg{|}_{t=0}p_{+}(\exp(tX)),

which implies

ddt|t=0exp(tX𝔫)\displaystyle\frac{d}{dt}\bigg{|}_{t=0}\exp(tX_{{\mathfrak{n}}_{-}}) =X𝔫\displaystyle=X_{{\mathfrak{n}}_{-}} =ddt|t=0p(exp(tX)),\displaystyle=\frac{d}{dt}\bigg{|}_{t=0}p_{-}(\exp(tX)),
ddt|t=0exp(tX𝔩)\displaystyle\frac{d}{dt}\bigg{|}_{t=0}\exp(tX_{{\mathfrak{l}}}) =X𝔩\displaystyle=X_{{\mathfrak{l}}} =ddt|t=0p0(exp(tX)),\displaystyle=\frac{d}{dt}\bigg{|}_{t=0}p_{0}(\exp(tX)),
ddt|t=0exp(tX𝔫+)\displaystyle\frac{d}{dt}\bigg{|}_{t=0}\exp(tX_{{\mathfrak{n}}_{+}}) =X𝔫+\displaystyle=X_{{\mathfrak{n}}_{+}} =ddt|t=0p+(exp(tX)).\displaystyle=\frac{d}{dt}\bigg{|}_{t=0}p_{+}(\exp(tX)).

Further, for n¯N\bar{n}\in N_{-} and sufficiently small t>0t>0, we have

exp(tX)n¯=p(exp(tX)n¯)p0(exp(tX)n¯)p+(exp(tX)n¯).\exp(-tX)\bar{n}=p_{-}(\exp(-tX)\bar{n})p_{0}(\exp(-tX)\bar{n})p_{+}(\exp(-tX)\bar{n}).

Observe that

p(exp(tX)n¯)=p(n¯exp(tAd(n¯1)X))=n¯p(exp(tAd(n¯1)X)),\displaystyle p_{-}(\exp(-tX)\bar{n})=p_{-}(\bar{n}\exp(-t\mathrm{Ad}(\bar{n}^{-1})X))=\bar{n}p_{-}(\exp(-t\mathrm{Ad}(\bar{n}^{-1})X)),
p0(exp(tX)n¯)=p0(n¯exp(tAd(n¯1)X))=p0(exp(tAd(n¯1)X)).\displaystyle p_{0}(\exp(-tX)\bar{n})=p_{0}(\bar{n}\exp(-t\mathrm{Ad}(\bar{n}^{-1})X))=p_{0}(\exp(-t\mathrm{Ad}(\bar{n}^{-1})X)).

Then, for FC(G/P,𝒱)C(G,V)PF\in C^{\infty}(G/P,\mathcal{V})\simeq C^{\infty}(G,V)^{P}, we have

F(exp(tX)n¯)\displaystyle F(\exp(-tX)\bar{n}) =F(p(exp(tX)n¯)p0(exp(tX)n¯))\displaystyle=F(p_{-}(\exp(-tX)\bar{n})p_{0}(\exp(-tX)\bar{n}))
=F(n¯p(exp(tAd(n¯1)X))p0(exp(tAd(n¯1)X)))\displaystyle=F(\bar{n}p_{-}(\exp(-t\mathrm{Ad}(\bar{n}^{-1})X))p_{0}(\exp(-t\mathrm{Ad}(\bar{n}^{-1})X)))
=σλ(p0(exp(tAd(n¯1)X))1)F(n¯p(exp(tAd(n¯1)X))).\displaystyle=\sigma_{\lambda}(p_{0}(\exp(-t\mathrm{Ad}(\bar{n}^{-1})X))^{-1})F(\bar{n}p_{-}(\exp(-t\mathrm{Ad}(\bar{n}^{-1})X))).

We remark that the N+N_{+}-invariance property that F(gn)=F(g)F(gn)=F(g) for nN+n\in N_{+} is applied in the first line. Since

ddt|t=0p0(exp(tAd(n¯1)X))1\displaystyle\frac{d}{dt}\bigg{|}_{t=0}p_{0}(\exp(-t\mathrm{Ad}(\bar{n}^{-1})X))^{-1} =(Ad(n¯1)X)𝔩,\displaystyle=(\mathrm{Ad}(\bar{n}^{-1})X)_{\mathfrak{l}},
ddt|t=0p(exp(tAd(n¯1)X))\displaystyle\frac{d}{dt}\bigg{|}_{t=0}p_{-}(\exp(-t\mathrm{Ad}(\bar{n}^{-1})X)) =ddt|t=0exp(t(Ad(n¯1)X)𝔫),\displaystyle=\frac{d}{dt}\bigg{|}_{t=0}\exp(-t(\mathrm{Ad}(\bar{n}^{-1})X)_{{\mathfrak{n}}_{-}}),

the representation dπ(σ,λ)(X)d\pi_{(\sigma,\lambda)}(X) on the image ι(C(G/P,𝒱))\iota^{*}(C^{\infty}(G/P,\mathcal{V})) of the inclusion ι:C(G/P,𝒱)C(N)V\iota^{*}\colon C^{\infty}(G/P,\mathcal{V})\hookrightarrow C^{\infty}(N_{-})\otimes V is given by

dπ(σ,λ)(X)F(n¯)\displaystyle d\pi_{(\sigma,\lambda)}(X)F(\bar{n}) =ddt|t=0F(exp(tX)n¯)\displaystyle=\frac{d}{dt}\bigg{|}_{t=0}F(\exp(-tX)\bar{n})
=ddt|t=0σλ(p0(exp(tAd(n¯1)X))1)F(n¯p(exp(tAd(n¯1)X)))\displaystyle=\frac{d}{dt}\bigg{|}_{t=0}\sigma_{\lambda}(p_{0}(\exp(-t\mathrm{Ad}(\bar{n}^{-1})X))^{-1})F(\bar{n}p_{-}(\exp(-t\mathrm{Ad}(\bar{n}^{-1})X)))
=ddt|t=0σλ(p0(exp(tAd(n¯1)X))1)F(n¯)+ddt|t=0F(n¯p(exp(tAd(n¯1)X)))\displaystyle=\frac{d}{dt}\bigg{|}_{t=0}\sigma_{\lambda}(p_{0}(\exp(-t\mathrm{Ad}(\bar{n}^{-1})X))^{-1})F(\bar{n})+\frac{d}{dt}\bigg{|}_{t=0}F(\bar{n}p_{-}(\exp(-t\mathrm{Ad}(\bar{n}^{-1})X)))
=dσλ((Ad(n¯1)X)𝔩)F(n¯)(dR((Ad(1)X)𝔫)F)(n¯).\displaystyle=d\sigma_{\lambda}((\mathrm{Ad}(\bar{n}^{-1})X)_{\mathfrak{l}})F(\bar{n})-\left(dR((\mathrm{Ad}(\cdot^{-1})X)_{{\mathfrak{n}}_{-}})F\right)(\bar{n}). (2.3)

Equation (2.2) shows that, for X𝔤()X\in{\mathfrak{g}}(\mathbb{R}), the formula dπ(σ,λ)(X)d\pi_{(\sigma,\lambda)}(X) on ι(C(G/P,𝒱))\iota^{*}(C^{\infty}(G/P,\mathcal{V})) can be extended to the whole space C(N)VC^{\infty}(N_{-})\otimes V. By extending the formula complex linearly to 𝔤{\mathfrak{g}}, we have

dπ(σ,λ)(X)f(n¯)=dσλ((Ad(n¯1)X)𝔩)f(n¯)(dR((Ad(1)X)𝔫)f)(n¯)d\pi_{(\sigma,\lambda)}(X)f(\bar{n})=d\sigma_{\lambda}((\mathrm{Ad}(\bar{n}^{-1})X)_{\mathfrak{l}})f(\bar{n})-\left(dR((\mathrm{Ad}(\cdot^{-1})X)_{{\mathfrak{n}}_{-}})f\right)(\bar{n}) (2.4)

for X𝔤X\in{\mathfrak{g}} and f(n¯)C(N)Vf(\bar{n})\in C^{\infty}(N_{-})\otimes V.

2.3. Duality theorem

For a finite-dimensional irreducible representation (σλ,V)(\sigma_{\lambda},V) of MAMA, we write V=Hom(V,)V^{\vee}=\operatorname{Hom}_{\mathbb{C}}(V,\mathbb{C}) and ((σλ),V)((\sigma_{\lambda})^{\vee},V^{\vee}) for the contragredient representation of (σλ,V)(\sigma_{\lambda},V). By letting 𝔫+{\mathfrak{n}}_{+} act on VV^{\vee} trivially, we regard the infinitesimal representation dσλd\sigma^{\vee}\boxtimes\mathbb{C}_{-\lambda} of (σλ)(\sigma_{\lambda})^{\vee} as a 𝔭{\mathfrak{p}}-module. The induced module

M𝔭(V):=𝒰(𝔤)𝒰(𝔭)VM_{\mathfrak{p}}(V^{\vee}):=\mathcal{U}({\mathfrak{g}})\otimes_{\mathcal{U}({\mathfrak{p}})}V^{\vee}

is called a generalized Verma module. Via the diagonal action of PP on M𝔭(V)M_{\mathfrak{p}}(V^{\vee}), we regard M𝔭(V)M_{\mathfrak{p}}(V^{\vee}) as a (𝔤,P)({\mathfrak{g}},P)-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

𝒟HD:HomP(W,M𝔭(V))DiffG(𝒱,𝒲),\mathcal{D}_{H\to D}\colon\operatorname{Hom}_{P}(W^{\vee},M_{{\mathfrak{p}}}(V^{\vee}))\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\operatorname{Diff}_{G}(\mathcal{V},\mathcal{W}), (2.6)

where, for φHomP(W,M𝔭(V))\varphi\in\operatorname{Hom}_{P}(W^{\vee},M_{\mathfrak{p}}(V^{\vee})) and FC(G/P,𝒱)C(G,V)PF\in C^{\infty}(G/P,\mathcal{V})\simeq C^{\infty}(G,V)^{P}, the element 𝒟HD(φ)FC(G/P,𝒲)C(G,W)P\mathcal{D}_{H\to D}(\varphi)F\in C^{\infty}(G/P,\mathcal{W})\simeq C^{\infty}(G,W)^{P} is given by

𝒟HD(φ)F,w=jdR(uj)F,vjfor wW,\langle\mathcal{D}_{H\to D}(\varphi)F,w^{\vee}\rangle=\sum_{j}\langle dR(u_{j})F,v_{j}^{\vee}\rangle\;\;\text{for $w^{\vee}\in W^{\vee}$}, (2.7)

where φ(w)=jujvjM𝔭(V)\varphi(w^{\vee})=\sum_{j}u_{j}\otimes v_{j}^{\vee}\in M_{\mathfrak{p}}(V^{\vee}).

Remark 2.8.

By the Frobenius reciprocity, (2.6) is equivalent to

𝒟HD:Hom𝔤,P(M𝔭(W),M𝔭(V))DiffG(𝒱,𝒲).\mathcal{D}_{H\to D}\colon\operatorname{Hom}_{{\mathfrak{g}},P}(M_{\mathfrak{p}}(W^{\vee}),M_{{\mathfrak{p}}}(V^{\vee}))\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\operatorname{Diff}_{G}(\mathcal{V},\mathcal{W}). (2.9)
Remark 2.10.

In general PP is not connected. If it is connected, then

Hom𝔤,P(M𝔭(W),M𝔭(V))=Hom𝔤(M𝔭(W),M𝔭(V)).\operatorname{Hom}_{{\mathfrak{g}},P}(M_{\mathfrak{p}}(W^{\vee}),M_{{\mathfrak{p}}}(V^{\vee}))=\operatorname{Hom}_{{\mathfrak{g}}}(M_{\mathfrak{p}}(W^{\vee}),M_{{\mathfrak{p}}}(V^{\vee})).

Thus, in the case, the isomorphism (2.9) is equivalent to

𝒟HD:Hom𝔤(M𝔭(W),M𝔭(V))DiffG(𝒱,𝒲).\mathcal{D}_{H\to D}\colon\operatorname{Hom}_{{\mathfrak{g}}}(M_{\mathfrak{p}}(W^{\vee}),M_{{\mathfrak{p}}}(V^{\vee}))\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\operatorname{Diff}_{G}(\mathcal{V},\mathcal{W}). (2.11)

2.4. Algebraic Fourier transform ^\widehat{\;\;\cdot\;\;} of Weyl algebras

Let UU be a complex finite-dimensional vector space with dimU=n\dim_{\mathbb{C}}U=n. Fix a basis u1,,unu_{1},\ldots,u_{n} of UU and let (z1,,zn)(z_{1},\ldots,z_{n}) denote the coordinates of UU with respect to the basis. Then the algebra

[U;z,z]:=[z1,,zn,z1,,zn]\mathbb{C}[U;z,\tfrac{\partial}{\partial z}]:=\mathbb{C}[z_{1},\ldots,z_{n},\tfrac{\partial}{\partial z_{1}},\ldots,\tfrac{\partial}{\partial z_{n}}]

with relations zizj=zjziz_{i}z_{j}=z_{j}z_{i}, zizj=zjzi\frac{\partial}{\partial z_{i}}\frac{\partial}{\partial z_{j}}=\frac{\partial}{\partial z_{j}}\frac{\partial}{\partial z_{i}}, and zjzi=δi,j+zizj\frac{\partial}{\partial z_{j}}z_{i}=\delta_{i,j}+z_{i}\frac{\partial}{\partial z_{j}} is called the Weyl algebra of UU, where δi,j\delta_{i,j} is the Kronecker delta. Similarly, let (ζ1,,ζn)(\zeta_{1},\ldots,\zeta_{n}) denote the coordinates of the dual space UU^{\vee} of UU with respect to the dual basis of u1,,unu_{1},\ldots,u_{n}. We write [U;ζ,ζ]\mathbb{C}[U^{\vee};\zeta,\tfrac{\partial}{\partial\zeta}] for the Weyl algebra of UU^{\vee}. Then the map determined by

zi^:=ζi,zi^:=ζi\widehat{\frac{\partial}{\partial z_{i}}}:=-\zeta_{i},\quad\widehat{z_{i}}:=\frac{\partial}{\partial\zeta_{i}} (2.12)

gives a Weyl algebra isomorphism

^:[U;z,z][U;ζ,ζ],TT^.\widehat{\;\;\cdot\;\;}\;\colon\mathbb{C}[U;z,\tfrac{\partial}{\partial z}]\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\mathbb{C}[U^{\vee};\zeta,\tfrac{\partial}{\partial\zeta}],\quad T\mapsto\widehat{T}. (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 ζi-\zeta_{i} 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 Ez=i=1nziziE_{z}=\sum_{i=1}^{n}z_{i}\frac{\partial}{\partial z_{i}} for zz and Eζ=i=1nζiζiE_{\zeta}=\sum_{i=1}^{n}\zeta_{i}\frac{\partial}{\partial\zeta_{i}} for ζ\zeta, we have Ez^=(n+Eζ)\widehat{E_{z}}=-(n+E_{\zeta}).

The Weyl algebra [U;z,z]\mathbb{C}[U;z,\tfrac{\partial}{\partial z}] naturally acts on the space 𝒟(U)\mathcal{D}^{\prime}(U) of distributions on UU. In particular, the action of [U;z,z]\mathbb{C}[U;z,\tfrac{\partial}{\partial z}] preserves the subspace 𝒟0(U)\mathcal{D}^{\prime}_{0}(U) of distributions supported at 0. Let 𝒥\mathcal{J} be the annihilator of the Dirac delta function δ0\delta_{0}, that is, the kernel of the homomorphism

Ψ:[U;z,z]𝒟0(U),PPδ0.\Psi\colon\mathbb{C}[U;z,\tfrac{\partial}{\partial z}]\longrightarrow\mathcal{D}^{\prime}_{0}(U),\quad P\mapsto P\delta_{0}.

Then 𝒥\mathcal{J} is the left ideal of [U;z,z]\mathbb{C}[U;z,\tfrac{\partial}{\partial z}] generated by the coordinate functions z1,,znz_{1},\ldots,z_{n}. Since the map Ψ\Psi is surjective (see, for instance, [70, Thm. 5.5]), it induces the isomorphism

Ψ¯:[U;z,z]/𝒥𝒟0(U).\bar{\Psi}\colon\mathbb{C}[U;z,\tfrac{\partial}{\partial z}]/\mathcal{J}\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\mathcal{D}^{\prime}_{0}(U). (2.14)

On the other hand, by applying the algebraic Fourier transform ^\widehat{\;\;\cdot\;\;} in (2.13) to [U;z,z]/𝒥\mathbb{C}[U;z,\tfrac{\partial}{\partial z}]/\mathcal{J}, the space [U;z,z]/𝒥\mathbb{C}[U;z,\tfrac{\partial}{\partial z}]/\mathcal{J} is isomorphic to the space Pol(U)=[U;ζ]\mathrm{Pol}(U^{\vee})=\mathbb{C}[U^{\vee};\zeta] of polynomials on UU^{\vee}. Thus, we have

𝒟0(U)[]Ψ¯1[U;z,z]/𝒥[]^Pol(U).\mathcal{D}^{\prime}_{0}(U)\stackrel{{\scriptstyle[}}{{\sim}}]{\bar{\Psi}^{-1}}{\longrightarrow}\mathbb{C}[U;z,\tfrac{\partial}{\partial z}]/\mathcal{J}\stackrel{{\scriptstyle[}}{{\sim}}]{\widehat{\;\;\cdot\;\;}}{\longrightarrow}\mathrm{Pol}(U^{\vee}). (2.15)

It is remarked that the composition ^Ψ¯1\widehat{\;\;\cdot\;\;}\circ\bar{\Psi}^{-1} maps 𝒟0(U)δ01Pol(U)\mathcal{D}^{\prime}_{0}(U)\ni\delta_{0}\mapsto 1\in\mathrm{Pol}(U^{\vee}).

2.5. Fourier transformed representation dπ(σ,λ)^\widehat{d\pi_{(\sigma,\lambda)^{*}}}

For 2ρ2ρ(𝔫+)=Trace(ad|𝔫+)𝔞2\rho\equiv 2\rho({\mathfrak{n}}_{+})=\mathrm{Trace}(\mathrm{ad}|_{{\mathfrak{n}}_{+}})\in\mathfrak{a}^{*}, we denote by 2ρ\mathbb{C}_{2\rho} the one-dimensional representation of PP defined by pχ2ρ(p)=|det(Ad(p):𝔫+𝔫+)|p\mapsto\chi_{2\rho}(p)=\left|\mathrm{det}(\mathrm{Ad}(p)\colon\mathfrak{n}_{+}\to\mathfrak{n}_{+})\right|. For the contragredient representation ((σλ),V)((\sigma_{\lambda})^{\vee},V^{\vee}) of (σλ,V)(\sigma_{\lambda},V), we put σλ:=σ2ρλ\sigma^{*}_{\lambda}:=\sigma^{\vee}\boxtimes\mathbb{C}_{2\rho-\lambda}. As for σλ\sigma_{\lambda}, we regard σλ\sigma^{*}_{\lambda} as a representation of PP. Define the induced representation π(σ,λ)=IndPG(σλ)\pi_{(\sigma,\lambda)^{*}}=\mathrm{Ind}_{P}^{G}(\sigma^{*}_{\lambda}) on the space C(G/P,𝒱)C^{\infty}(G/P,\mathcal{V}^{*}) of smooth sections for the vector bundle 𝒱=G×P(V2ρ)\mathcal{V}^{*}=G\times_{P}(V^{\vee}\otimes\mathbb{C}_{2\rho}) associated with σλ\sigma^{*}_{\lambda}, which is isomorphic to the tensor bundle of the dual vector bundle 𝒱=G×PV\mathcal{V}^{\vee}=G\times_{P}V^{\vee} and the bundle of densities over G/PG/P. Then the integration on G/PG/P gives a GG-invariant non-degenerate bilinear form IndPG(σλ)×IndPG(σλ)\mathrm{Ind}^{G}_{P}(\sigma_{\lambda})\times\mathrm{Ind}^{G}_{P}(\sigma^{*}_{\lambda})\to\mathbb{C} for IndPG(σλ)\mathrm{Ind}^{G}_{P}(\sigma_{\lambda}) and IndPG(σλ)\mathrm{Ind}^{G}_{P}(\sigma^{*}_{\lambda}).

As for C(G/P,𝒱)C^{\infty}(G/P,\mathcal{V}), the space C(G/P,𝒱)C^{\infty}(G/P,\mathcal{V}^{*}) can be regarded as a subspace of C(N)VC^{\infty}(N_{-})\otimes V^{\vee}. Then, by replacing σλ\sigma_{\lambda} with σλ\sigma_{\lambda}^{*} in (2.4), we have

dπ(σ,λ)(X)f(n¯)=dσλ((Ad(n¯1)X)𝔩)f(n¯)(dR((Ad(1)X)𝔫)f)(n¯)d\pi_{(\sigma,\lambda)^{*}}(X)f(\bar{n})=d\sigma_{\lambda}^{*}((\mathrm{Ad}(\bar{n}^{-1})X)_{\mathfrak{l}})f(\bar{n})-\left(dR((\mathrm{Ad}(\cdot^{-1})X)_{{\mathfrak{n}}_{-}})f\right)(\bar{n}) (2.16)

for X𝔤X\in{\mathfrak{g}} and f(n¯)C(N)Vf(\bar{n})\in C^{\infty}(N_{-})\otimes V^{\vee}. Via the exponential map exp:𝔫()N\exp\colon{\mathfrak{n}}_{-}(\mathbb{R})\simeq N_{-}, one can regard dπ(σ,λ)(X)d\pi_{(\sigma,\lambda)^{*}}(X) as a representation on C(𝔫())VC^{\infty}(\mathfrak{n}_{-}(\mathbb{R}))\otimes V^{\vee}. It then follows from (2.16) that dπ(σ,λ)d\pi_{(\sigma,\lambda)^{*}} gives a Lie algebra homomorphism

dπ(σ,λ):𝔤[𝔫();x,x]End(V),d\pi_{(\sigma,\lambda)^{*}}\colon\mathfrak{g}\longrightarrow\mathbb{C}[{\mathfrak{n}}_{-}(\mathbb{R});x,\tfrac{\partial}{\partial x}]\otimes\mathrm{End}(V^{\vee}),

where (x1,,xn)(x_{1},\ldots,x_{n}) are coordinates of 𝔫(){\mathfrak{n}}_{-}(\mathbb{R}) with n=dim𝔫()n=\dim{\mathfrak{n}}_{-}(\mathbb{R}). We extend the coordinate functions x1,,xnx_{1},\ldots,x_{n} for 𝔫(){\mathfrak{n}}_{-}(\mathbb{R}) holomorphically to the ones z1,,znz_{1},\ldots,z_{n} for 𝔫{\mathfrak{n}}_{-}. Thus we have

dπ(σ,λ):𝔤[𝔫;z,z]End(V).d\pi_{(\sigma,\lambda)^{*}}\colon\mathfrak{g}\longrightarrow\mathbb{C}[{\mathfrak{n}}_{-};z,\tfrac{\partial}{\partial z}]\otimes\mathrm{End}(V^{\vee}).

Now we fix a non-degenerate symmetric bilinear form κ\kappa on 𝔫+{\mathfrak{n}}_{+} and 𝔫{\mathfrak{n}}_{-}. Via κ\kappa, we identify 𝔫+{\mathfrak{n}}_{+} with the dual space 𝔫{\mathfrak{n}}_{-}^{\vee} of 𝔫{\mathfrak{n}}_{-}. Then the algebraic Fourier transform ^\widehat{\;\;\cdot\;\;} of Weyl algebras (2.13) gives a Weyl algebra isomorphism

^:[𝔫;z,z][𝔫+;ζ,ζ].\widehat{\;\;\cdot\;\;}\;\colon\mathbb{C}[{\mathfrak{n}}_{-};z,\tfrac{\partial}{\partial z}]\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\mathbb{C}[{\mathfrak{n}}_{+};\zeta,\tfrac{\partial}{\partial\zeta}].

In particular, it gives a Lie algebra homomorphism

dπ(σ,λ)^:𝔤[𝔫+;ζ,ζ]End(V).\widehat{d\pi_{(\sigma,\lambda)^{*}}}\colon\mathfrak{g}\longrightarrow\mathbb{C}[{\mathfrak{n}}_{+};\zeta,\tfrac{\partial}{\partial\zeta}]\otimes\mathrm{End}(V^{\vee}). (2.17)

2.6. Algebraic Fourier transform FcF_{c} of generalized Verma modules

Our aim is to construct a (𝔤,P)({\mathfrak{g}},P)-module isomorpshism

M𝔭(V)Pol(𝔫+)V,M_{\mathfrak{p}}(V^{\vee})\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\mathrm{Pol}({\mathfrak{n}}_{+})\otimes V^{\vee},

where Pol(𝔫+)\mathrm{Pol}({\mathfrak{n}}_{+}) denotes the space [𝔫+;ζ]\mathbb{C}[{\mathfrak{n}}_{+};\zeta] of polynomial functions on 𝔫+{\mathfrak{n}}_{+}. It is classically known that the map uvdL(u)(δ[o]v)u\otimes v^{\vee}\mapsto dL(u)(\delta_{[o]}\otimes v^{\vee}) gives a (𝔤,P)({\mathfrak{g}},P)-module isomorphism

M𝔭(V)𝒟[o](G/P,𝒱),uvdL(u)(δ[o]v),M_{\mathfrak{p}}(V^{\vee})\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\mathcal{D}^{\prime}_{[o]}(G/P,\mathcal{V}^{*}),\quad u\otimes v^{\vee}\mapsto dL(u)(\delta_{[o]}\otimes v^{\vee}), (2.18)

where [o][o] denotes the identity coset [o]=eP[o]=eP of G/PG/P and δ[o]\delta_{[o]} denotes the Dirac delta function supported at [o][o] (see, for instance, [44, 50]). It thus suffices to construct a (𝔤,P)({\mathfrak{g}},P)-module isomorphism

𝒟[o](G/P,𝒱)Pol(𝔫+)V.\mathcal{D}^{\prime}_{[o]}(G/P,\mathcal{V}^{*})\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\mathrm{Pol}({\mathfrak{n}}_{+})\otimes V^{\vee}.

First, observe that the restriction of the dual vector bundle 𝒱G/P\mathcal{V}^{*}\to G/P to the open Bruhat cell ι:𝔫()NG/P\iota\colon\mathfrak{n}_{-}(\mathbb{R})\simeq N_{-}\hookrightarrow G/P gives an isomorphism

ι:𝒟[o](G/P,𝒱)𝒟0(𝔫(),V).\iota^{*}\colon\mathcal{D}^{\prime}_{[o]}(G/P,\mathcal{V}^{*})\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\mathcal{D}^{\prime}_{0}({\mathfrak{n}}_{-}(\mathbb{R}),V^{\vee}).

Via the isomorphism ι\iota^{*}, one can induce a (𝔤,P)({\mathfrak{g}},P)-module structure on 𝒟0(𝔫(),V)\mathcal{D}^{\prime}_{0}({\mathfrak{n}}_{-}(\mathbb{R}),V^{\vee}) from 𝒟[o](G/P,𝒱)\mathcal{D}^{\prime}_{[o]}(G/P,\mathcal{V}^{*}). In particular, the Lie algebra 𝔤{\mathfrak{g}} acts on 𝒟0(𝔫(),V)\mathcal{D}^{\prime}_{0}({\mathfrak{n}}_{-}(\mathbb{R}),V^{\vee}) via the infinitesimal representation dπ(σ,λ)d\pi_{(\sigma,\lambda)^{*}}. Consequently, by the isomorphism (2.15), one obtains the following chain of (𝔤,P)({\mathfrak{g}},P)-isomorphisms:

M𝔭(V)\displaystyle M_{\mathfrak{p}}(V^{\vee}) 𝒟[o](G/P,𝒱)\displaystyle\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\mathcal{D}^{\prime}_{[o]}(G/P,\mathcal{V}^{*}) 𝒟0(𝔫(),V)\displaystyle\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\mathcal{D}^{\prime}_{0}({\mathfrak{n}}_{-}(\mathbb{R}),V^{\vee}) Pol(𝔫+)V\displaystyle\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\mathrm{Pol}({\mathfrak{n}}_{+})\otimes V^{\vee} (2.19)
uv\displaystyle u\otimes v^{\vee}\hskip 8.0pt dL(u)(δ[o]v)\displaystyle\mapsto dL(u)(\delta_{[o]}\otimes v^{\vee}) dπ(σ,λ)(u)(δ0v)\displaystyle\mapsto d\pi_{(\sigma,\lambda)^{*}}(u)(\delta_{0}\otimes v^{\vee}) dπ(σ,λ)^(u)(1v).\displaystyle\mapsto\widehat{d\pi_{(\sigma,\lambda)^{*}}}(u)(1\otimes v^{\vee}).

We denote by FcF_{c} the composition of the three (𝔤,P)(\mathfrak{g},P)-module isomorphisms in (2.19).

Definition 2.20 ([44, Sect. 3.4]).

We call the (𝔤,P)(\mathfrak{g},P)-module isomorphism

Fc:M𝔭(V)Pol(𝔫+)V,uvdπ(σ,λ)^(u)(1v)F_{c}\colon M_{\mathfrak{p}}(V^{\vee})\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\mathrm{Pol}({\mathfrak{n}}_{+})\otimes V^{\vee},\quad u\otimes v^{\vee}\longmapsto\widehat{d\pi_{(\sigma,\lambda)^{*}}}(u)(1\otimes v^{\vee})\\ (2.21)

the algebraic Fourier transform of the generalized Verma module M𝔭(V)M_{\mathfrak{p}}(V^{\vee}).

2.7. The F-method

Observe that the algebraic Fourier transform FcF_{c} in (2.21) gives an MAMA-representation isomorphism

M𝔭(V)𝔫+(Pol(𝔫+)V)dπ(σ,λ)^(𝔫+),M_{\mathfrak{p}}(V^{\vee})^{{\mathfrak{n}}_{+}}\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}(\mathrm{Pol}({\mathfrak{n}}_{+})\otimes V^{\vee})^{\widehat{d\pi_{(\sigma,\lambda)^{*}}}({\mathfrak{n}}_{+})}, (2.22)

which induces a linear isomorphism

HomMA(W,M𝔭(V)𝔫+)HomMA(W,(Pol(𝔫+)V)dπ(σ,λ)^(𝔫+)).\operatorname{Hom}_{MA}(W^{\vee},M_{\mathfrak{p}}(V^{\vee})^{{\mathfrak{n}}_{+}})\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\operatorname{Hom}_{MA}\left(W^{\vee},(\mathrm{Pol}(\mathfrak{n}_{+})\otimes V^{\vee})^{\widehat{d\pi_{(\sigma,\lambda)^{*}}}({\mathfrak{n}}_{+})}\right). (2.23)

Here MAMA acts on Pol(𝔫+)\mathrm{Pol}({\mathfrak{n}}_{+}) via the action

Ad#(l):p(X)p(Ad(l1)X)for lMA.\mathrm{Ad}_{\#}(l)\colon p(X)\mapsto p(\mathrm{Ad}(l^{-1})X)\;\;\text{for $l\in MA$}. (2.24)

Recall from (2.17) that we have

dπ(σ,λ)^(𝔫+)[𝔫+;ζ,ζ]End(V).\widehat{d\pi_{(\sigma,\lambda)^{*}}}({\mathfrak{n}}_{+})\subset\mathbb{C}[{\mathfrak{n}}_{+};\zeta,\tfrac{\partial}{\partial\zeta}]\otimes\mathrm{End}(V^{\vee}).

Since the nilpotent radical 𝔫+{\mathfrak{n}}_{+} acts trivially on 1vM𝔭(V)1\otimes v^{\vee}\in M_{\mathfrak{p}}(V^{\vee}), and since the algebraic Fourier transform FcF_{c} is a 𝔤{\mathfrak{g}}-module isomorphism, the Fourier image dπ(σ,λ)^(𝔫+)\widehat{d\pi_{(\sigma,\lambda)^{*}}}({\mathfrak{n}}_{+}) must act on 1vPol(𝔫+)V1\otimes v^{\vee}\in\mathrm{Pol}({\mathfrak{n}}_{+})\otimes V^{\vee} trivially. Therefore, the operator dπ(σ,λ)^(C)\widehat{d\pi_{(\sigma,\lambda)^{*}}}(C) for C𝔫+C\in{\mathfrak{n}}_{+} is a differential operator of order at least 1; in particular, the space (Pol(𝔫+)V)dπ(σ,λ)^(𝔫+)(\mathrm{Pol}(\mathfrak{n}_{+})\otimes V^{\vee})^{\widehat{d\pi_{(\sigma,\lambda)^{*}}}({\mathfrak{n}}_{+})} is the space of polynomial solutions to the system of partial differential equations defined by dπ(σ,λ)^(C)\widehat{d\pi_{(\sigma,\lambda)^{*}}}(C) for C𝔫+C\in{\mathfrak{n}}_{+}. It is remarked that the operators dπ(σ,λ)^(C)\widehat{d\pi_{(\sigma,\lambda)^{*}}}(C) are not vector fields; for instance, if 𝔫+\mathfrak{n}_{+} is abelian, then dπ(σ,λ)^(C)\widehat{d\pi_{(\sigma,\lambda)^{*}}}(C) are second order differential operators (cf. Proposition 5.4 and [44, Prop. 3.10]).

To emphasize that (Pol(𝔫+)V)dπ(σ,λ)^(𝔫+)(\mathrm{Pol}(\mathfrak{n}_{+})\otimes V^{\vee})^{\widehat{d\pi_{(\sigma,\lambda)^{*}}}({\mathfrak{n}}_{+})} is the space of polynomial solutions to a system of differential equations, we let

Sol(𝔫+;V,W)=HomMA(W,(Pol(𝔫+)V)dπ(σ,λ)^(𝔫+)).\mathrm{Sol}(\mathfrak{n}_{+};V,W)=\operatorname{Hom}_{MA}(W^{\vee},(\mathrm{Pol}(\mathfrak{n}_{+})\otimes V^{\vee})^{\widehat{d\pi_{(\sigma,\lambda)^{*}}}({\mathfrak{n}}_{+})}). (2.25)

Via the identification HomMA(W,Pol(𝔫+)V)((Pol(𝔫+)V)W)MA\textrm{Hom}_{MA}(W^{\vee},\mathrm{Pol}({\mathfrak{n}}_{+})\otimes V^{\vee})\simeq\big{(}(\mathrm{Pol}({\mathfrak{n}}_{+})\otimes V^{\vee})\otimes W\big{)}^{MA}, we have

Sol(𝔫+;V,W)\displaystyle\mathrm{Sol}({\mathfrak{n}}_{+};V,W)
={ψHomMA(W,Pol(𝔫+)V): ψ satisfies the system of PDEs (2.27) below.}\displaystyle=\{\psi\in\operatorname{Hom}_{MA}(W^{\vee},\mathrm{Pol}({\mathfrak{n}}_{+})\otimes V^{\vee}):\text{ $\psi$ satisfies the system of PDEs \eqref{eqn:Fsys} below.}\} (2.26)
(dπ(σ,λ)^(C)idW)ψ=0for all C𝔫+.(\widehat{d\pi_{(\sigma,\lambda)^{*}}}(C)\otimes\mathrm{id}_{W})\psi=0\,\,\text{for all $C\in{\mathfrak{n}}_{+}$}. (2.27)

We call the system of PDEs (2.27) the F-system ([42, Fact 3.3 (3)]). Since

HomP(W,M𝔭(V))=HomMA(W,M𝔭(V)𝔫+),\operatorname{Hom}_{P}(W^{\vee},M_{\mathfrak{p}}(V^{\vee}))=\operatorname{Hom}_{MA}(W^{\vee},M_{\mathfrak{p}}(V^{\vee})^{{\mathfrak{n}}_{+}}),

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

FcidW:HomP(W,M𝔭(V))Sol(𝔫+;V,W).F_{c}\otimes\mathrm{id}_{W}\colon\operatorname{Hom}_{P}(W^{\vee},M_{\mathfrak{p}}(V^{\vee}))\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\mathrm{Sol}(\mathfrak{n}_{+};V,W).

Equivalently, we have

Hom𝔤,P(M𝔭(W),M𝔭(V))Sol(𝔫+;V,W).\mathrm{Hom}_{\mathfrak{g},P}(M_{\mathfrak{p}}(W^{\vee}),M_{\mathfrak{p}}(V^{\vee}))\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\mathrm{Sol}(\mathfrak{n}_{+};V,W).

The diagram (2.29) below is the refinement of (1.2) in the introduction.

Sol(𝔫+;V,W)\textstyle{\mathrm{Sol}(\mathfrak{n}_{+};V,W)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}Hom𝔤,P(M𝔭(W),M𝔭(V))\textstyle{\operatorname{Hom}_{\mathfrak{g},P}(M_{\mathfrak{p}}(W^{\vee}),M_{\mathfrak{p}}(V^{\vee}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\hskip 25.0pt\sim}𝒟HD\scriptstyle{\hskip 25.0pt\mathcal{D}_{H\to D}}\scriptstyle{\sim}FcidW\scriptstyle{F_{c}\otimes\mathrm{id}_{W}}DiffG(𝒱,𝒲)\textstyle{\operatorname{Diff}_{G}(\mathcal{V},\mathcal{W})}
(2.29)

2.8. The case of abelian nilradical 𝔫+{\mathfrak{n}}_{+}

Now suppose that the nilpotent radical 𝔫+{\mathfrak{n}}_{+} is abelian. In this case intertwining differential operators 𝒟DiffG(𝒱,𝒲)\mathcal{D}\in\mathrm{Diff}_{G}(\mathcal{V},\mathcal{W}) have constant coefficients. Indeed, observe that, as 𝔫+{\mathfrak{n}}_{+} is abelian, so is 𝔫(){\mathfrak{n}}_{-}(\mathbb{R}). Thus, we have

exp(X)exp(Y)=exp(X+Y)for all X,Y𝔫().\exp(X)\exp(Y)=\exp(X+Y)\quad\text{for all $X,Y\in{\mathfrak{n}}_{-}(\mathbb{R})$}.

For a given ordered basis (X1,,Xn)(X_{1},\ldots,X_{n}) with n=dim𝔫()n=\dim{\mathfrak{n}}_{-}(\mathbb{R}), one may identify C(N)C^{\infty}(N_{-}) with C(N)C(n)C^{\infty}(N_{-})\simeq C^{\infty}(\mathbb{R}^{n}) via the exponential map

nN,(x1,,xn)exp(x1X1++xnXn).\mathbb{R}^{n}\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}N_{-},\quad(x_{1},\ldots,x_{n})\mapsto\exp(x_{1}X_{1}+\cdots+x_{n}X_{n}). (2.30)

Then, for f(x1,,xn)C(n)f(x_{1},\ldots,x_{n})\in C^{\infty}(\mathbb{R}^{n}), we have

dR(Xj)f(x1,,xn)\displaystyle dR(X_{j})f(x_{1},\ldots,x_{n}) =ddt|t=0f(exp(x1X1++xnXn)exp(tXj))\displaystyle=\frac{d}{dt}\bigg{|}_{t=0}f\big{(}\exp(x_{1}X_{1}+\cdots+x_{n}X_{n})\exp(tX_{j})\big{)}
=ddt|t=0f(exp(x1X1++(xj+t)Xj++xnXn))\displaystyle=\frac{d}{dt}\bigg{|}_{t=0}f\big{(}\exp(x_{1}X_{1}+\cdots+(x_{j}+t)X_{j}+\cdots+x_{n}X_{n})\big{)}
=ddt|t=0f(x1,,xj1,xj+t,xj+1,,xn)\displaystyle=\frac{d}{dt}\bigg{|}_{t=0}f(x_{1},\ldots,x_{j-1},x_{j}+t,x_{j+1},\ldots,x_{n})
=xjf(x1,,xn).\displaystyle=\frac{\partial}{\partial x_{j}}f(x_{1},\ldots,x_{n}).

Since intertwining differential operators 𝒟\mathcal{D} are given by linear combinations of the differential operators dR(X1)k1dR(Xn)kndR(X_{1})^{k_{1}}\cdots dR(X_{n})^{k_{n}} over \mathbb{C} by Theorem 2.5, this shows that the coefficients of 𝒟\mathcal{D} must be constant.

As in (2.2), one may view DiffG(𝒱,𝒲)\mathrm{Diff}_{G}(\mathcal{V},\mathcal{W}) as

DiffG(𝒱,𝒲)[𝔫;z]Hom(V,W).\mathrm{Diff}_{G}(\mathcal{V},\mathcal{W})\subset\mathbb{C}[{\mathfrak{n}}_{-};\frac{\partial}{\partial z}]\otimes\mathrm{Hom}_{\mathbb{C}}(V,W).

Since 𝔫+{\mathfrak{n}}_{+} is regarded as the dual space of 𝔫{\mathfrak{n}}_{-}, one can define the symbol map

symb:[𝔫;z][𝔫+;ζ],ziζi.\mathrm{symb}\colon\mathbb{C}[{\mathfrak{n}}_{-};\frac{\partial}{\partial z}]\longrightarrow\mathbb{C}[{\mathfrak{n}}_{+};\zeta],\quad\frac{\partial}{\partial z_{i}}\mapsto\zeta_{i}.

As Pol(𝔫+)=[𝔫+;ζ]\mathrm{Pol}({\mathfrak{n}}_{+})=\mathbb{C}[{\mathfrak{n}}_{+};\zeta], this shows that we have

symb:[𝔫;z]Hom(V,W)\displaystyle\mathrm{symb}\colon\mathbb{C}[{\mathfrak{n}}_{-};\frac{\partial}{\partial z}]\otimes\mathrm{Hom}_{\mathbb{C}}(V,W)\stackrel{{\scriptstyle\sim}}{{\longrightarrow}} Pol(𝔫+)Hom(V,W)\displaystyle\;\mathrm{Pol}({\mathfrak{n}}_{+})\otimes\mathrm{Hom}_{\mathbb{C}}(V,W)
\bigcup \bigcup (2.31)
DiffG(𝒱,𝒲)\displaystyle\textnormal{Diff}_{G}(\mathcal{V},\mathcal{W})\hskip 56.9055pt symb(DiffG(𝒱,𝒲))\displaystyle\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\hskip 14.22636pt\mathrm{symb}(\textnormal{Diff}_{G}(\mathcal{V},\mathcal{W}))\hskip 14.22636pt

On the other hand, by (2.25), we also have

Sol(𝔫+;V,W)(Pol(𝔫+)Hom(V,W))MA,dπ(σ,λ)^(𝔫+).\mathrm{Sol}(\mathfrak{n}_{+};V,W)\simeq\big{(}\mathrm{Pol}(\mathfrak{n}_{+})\otimes\operatorname{Hom}_{\mathbb{C}}(V,W)\big{)}^{MA,\,\widehat{d\pi_{(\sigma,\lambda)^{*}}}({\mathfrak{n}}_{+})}.

Hence,

Pol(𝔫+)Hom(V,W)\textstyle{\mathrm{Pol}(\mathfrak{n}_{+})\otimes\operatorname{Hom}_{\mathbb{C}}(V,W)}

\bigcup

  

\bigcup

      
Sol(𝔫+;V,W)\textstyle{\mathrm{Sol}(\mathfrak{n}_{+};V,W)\hskip 10.0pt}symb(DiffG(𝒱,𝒲))\textstyle{\mathrm{symb}(\textnormal{Diff}_{G}(\mathcal{V},\mathcal{W}))}

Theorem 2.32 below shows that, in fact, we have

Sol(𝔫+;V,W)=symb(DiffG(𝒱,𝒲)).\mathrm{Sol}(\mathfrak{n}_{+};V,W)=\mathrm{symb}(\textnormal{Diff}_{G}(\mathcal{V},\mathcal{W})).
Theorem 2.32 ([44, Cor. 4.3]).

Suppose that the nilpotent radical 𝔫+\mathfrak{n}_{+} is abelian. Then the symbol map symb\mathrm{symb} gives a linear isomorphism

symb1:Sol(𝔫+;V,W)DiffG(𝒱,𝒲).\mathrm{symb}^{-1}\colon\mathrm{Sol}(\mathfrak{n}_{+};V,W)\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\operatorname{Diff}_{G}(\mathcal{V},\mathcal{W}).

Further, the diagram (2.33) below commutes.

Sol(𝔫+;V,W)\textstyle{\mathrm{Sol}(\mathfrak{n}_{+};V,W)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}symb1\scriptstyle{\;\;\mathrm{symb}^{-1}}\textstyle{\circlearrowleft{}}Hom𝔤,P(M𝔭(W),M𝔭(V))\textstyle{\operatorname{Hom}_{\mathfrak{g},P}(M_{\mathfrak{p}}(W^{\vee}),M_{\mathfrak{p}}(V^{\vee}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\hskip 25.0pt\sim}𝒟HD\scriptstyle{\hskip 25.0pt\mathcal{D}_{H\to D}}\scriptstyle{\sim}FcidW\scriptstyle{F_{c}\otimes\mathrm{id}_{W}}DiffG(𝒱,𝒲)\textstyle{\operatorname{Diff}_{G}(\mathcal{V},\mathcal{W})}
(2.33)

2.9. A recipe of the F-method for abelian nilradical 𝔫+{\mathfrak{n}}_{+}

By (2.7) and Theorem 2.32, one can classify and construct 𝒟DiffG(𝒱,𝒲)\mathcal{D}\in\operatorname{Diff}_{G}(\mathcal{V},\mathcal{W}) and φHom𝔤,P(M𝔭(W),M𝔭(V))\varphi\in\operatorname{Hom}_{\mathfrak{g},P}(M_{\mathfrak{p}}(W^{\vee}),M_{\mathfrak{p}}(V^{\vee})) by computing ψSol(𝔫+;V,W)\psi\in\mathrm{Sol}(\mathfrak{n}_{+};V,W) as follows.

  1. Step 1

    Compute dπ(σ,λ)(C)d\pi_{(\sigma,\lambda)^{*}}(C) and dπ(σ,λ)^(C)\widehat{d\pi_{(\sigma,\lambda)^{*}}}(C) for C𝔫+C\in{\mathfrak{n}}_{+}.

  2. Step 2

    Classify and construct ψHomMA(W,Pol(𝔫+)V)\psi\in\operatorname{Hom}_{MA}(W^{\vee},\mathrm{Pol}({\mathfrak{n}}_{+})\otimes V^{\vee}).

  3. Step 3

    Solve the F-system (2.27) for ψHomMA(W,Pol(𝔫+)V)\psi\in\operatorname{Hom}_{MA}(W^{\vee},\mathrm{Pol}({\mathfrak{n}}_{+})\otimes V^{\vee}).

  4. Step 4

    For ψSol(𝔫+;V,W)\psi\in\mathrm{Sol}(\mathfrak{n}_{+};V,W) obtained in Step 3, do the following.

    1. Step 4a

      Apply symb1\mathrm{symb}^{-1} to ψSol(𝔫+;V,W)\psi\in\mathrm{Sol}(\mathfrak{n}_{+};V,W) to obtain 𝒟DiffG(𝒱,𝒲)\mathcal{D}\in\operatorname{Diff}_{G}(\mathcal{V},\mathcal{W}).

    2. Step 4b

      Apply Fc1idWF_{c}^{-1}\otimes\mathrm{id}_{W} to ψSol(𝔫+;V,W)\psi\in\mathrm{Sol}(\mathfrak{n}_{+};V,W) to obtain φHom𝔤,P(M𝔭(W),M𝔭(V))\varphi\in\operatorname{Hom}_{\mathfrak{g},P}(M_{\mathfrak{p}}(W^{\vee}),M_{\mathfrak{p}}(V^{\vee})).

In Section 5, we shall carry out the recipe for the case (G,P)=(SL(n,),P1,n1)(G,P)=(SL(n,\mathbb{R}),P_{1,n-1}).

3. Specialization to (SL(n,),P1,n1)(SL(n,\mathbb{R}),P_{1,n-1})

The aim of this short section is to specialize the general framework in Section 2 to the case (G,P)=(SL(n,),P1,n1)(G,P)=(SL(n,\mathbb{R}),P_{1,n-1}), where P1,n1P_{1,n-1} denotes the maximal parabolic subgroup of GG corresponding to the partition n=1+(n1)n=1+(n-1). Throughout this section, we assume n2n\geq 2.

3.1. Notation

Let G=SL(n,)G=SL(n,\mathbb{R}) with Lie algebra 𝔤()=𝔰𝔩(n,){\mathfrak{g}}(\mathbb{R})=\mathfrak{sl}(n,\mathbb{R}) for n2n\geq 2. We put

Nj+:=E1,j+1,Nj:=Ej+1,1for j{1,,n1}N_{j}^{+}:=E_{1,j+1},\quad N_{j}^{-}:=E_{j+1,1}\quad\text{for $j\in\{1,\ldots,n-1\}$}

and

H0:=1n((n1)E1,1r=2nEr,r)=1ndiag(n1,1,1,,1),H_{0}:=\frac{1}{n}((n-1)E_{1,1}-\sum^{n}_{r=2}E_{r,r})=\frac{1}{n}\mathrm{diag}(n-1,-1,-1,\ldots,-1),

where Ei,jE_{i,j} denote the matrix units. We normalize H0H_{0} as H~0:=nn1H0\widetilde{H}_{0}:=\frac{n}{n-1}H_{0}, namely,

H~0=1n1((n1)E1,1r=2nEr,r)=1n1diag(n1,1,1,,1).\displaystyle\widetilde{H}_{0}=\frac{1}{n-1}((n-1)E_{1,1}-\sum^{n}_{r=2}E_{r,r})=\frac{1}{n-1}\mathrm{diag}(n-1,-1,-1,\ldots,-1).

Let

𝔫+()=Ker(ad(H0)id)\displaystyle{\mathfrak{n}}_{+}(\mathbb{R})=\mathrm{Ker}(\mathrm{ad}(H_{0})-\mathrm{id}) =Ker(ad(H~0)nn1id),\displaystyle=\mathrm{Ker}(\mathrm{ad}(\widetilde{H}_{0})-\tfrac{n}{n-1}\mathrm{id}), (3.1)
𝔫()=Ker(ad(H0)+id)\displaystyle{\mathfrak{n}}_{-}(\mathbb{R})=\mathrm{Ker}(\mathrm{ad}(H_{0})+\mathrm{id}) =Ker(ad(H~0)+nn1id).\displaystyle=\mathrm{Ker}(\mathrm{ad}(\widetilde{H}_{0})+\tfrac{n}{n-1}\mathrm{id}).

Then we have

𝔫±()=span{N1±,,Nn1±}.{\mathfrak{n}}_{\pm}(\mathbb{R})=\text{span}_{\mathbb{R}}\{N_{1}^{\pm},\ldots,N_{n-1}^{\pm}\}.

For X,Y𝔤()X,Y\in{\mathfrak{g}}(\mathbb{R}), let Tr(X,Y)=Trace(XY)\text{Tr}(X,Y)=\text{Trace}(XY) denote the trace form of 𝔤(){\mathfrak{g}}(\mathbb{R}). Then Ni+N_{i}^{+} and NjN_{j}^{-} satisfy Tr(Ni+,Nj)=δi,j\text{Tr}(N_{i}^{+},N_{j}^{-})=\delta_{i,j}. In what follows, we identify the dual 𝔫(){\mathfrak{n}}_{-}(\mathbb{R})^{\vee} of 𝔫(){\mathfrak{n}}_{-}(\mathbb{R}) with 𝔫()𝔫+(){\mathfrak{n}}_{-}(\mathbb{R})^{\vee}\simeq{\mathfrak{n}}_{+}(\mathbb{R}) via the trace form Tr(,)\text{Tr}(\cdot,\cdot).

Let 𝔞()=H~0{\mathfrak{a}}(\mathbb{R})=\mathbb{R}\widetilde{H}_{0} and

𝔪()={(0X):X𝔰𝔩(n1,)}𝔰𝔩(n1,).{\mathfrak{m}}(\mathbb{R})=\left\{\begin{pmatrix}0&\\ &X\end{pmatrix}:X\in\mathfrak{sl}(n-1,\mathbb{R})\right\}\simeq\mathfrak{sl}(n-1,\mathbb{R}). (3.2)

Here 𝔰𝔩(1,)\mathfrak{sl}(1,\mathbb{R}) is regarded as 𝔰𝔩(1,)={0}\mathfrak{sl}(1,\mathbb{R})=\{0\}. We have 𝔪()𝔞()=Ker(ad(H0)){\mathfrak{m}}(\mathbb{R})\oplus{\mathfrak{a}}(\mathbb{R})=\mathrm{Ker}(\mathrm{ad}(H_{0})) and the decomposition 𝔤()=𝔫()𝔪()𝔞()𝔫+(){\mathfrak{g}}(\mathbb{R})={\mathfrak{n}}_{-}(\mathbb{R})\oplus{\mathfrak{m}}(\mathbb{R})\oplus{\mathfrak{a}}(\mathbb{R})\oplus{\mathfrak{n}}_{+}(\mathbb{R}) is a Gelfand–Naimark decomposition of 𝔤(){\mathfrak{g}}(\mathbb{R}). The subalgebra 𝔭():=𝔪()𝔞()𝔫+(){\mathfrak{p}}(\mathbb{R}):={\mathfrak{m}}(\mathbb{R})\oplus{\mathfrak{a}}(\mathbb{R})\oplus{\mathfrak{n}}_{+}(\mathbb{R}) is a maximal parabolic subalgebra of 𝔤(){\mathfrak{g}}(\mathbb{R}). It is remarked that 𝔫±(){\mathfrak{n}}_{\pm}(\mathbb{R}) are abelian.

Let PP be the normalizer NG(𝔭())N_{G}({\mathfrak{p}}(\mathbb{R})) of 𝔭(){\mathfrak{p}}(\mathbb{R}) in GG. We write P=MAN+P=MAN_{+} for the Langlands decomposition of PP corresponding to 𝔭()=𝔪()𝔞()𝔫+(){\mathfrak{p}}(\mathbb{R})={\mathfrak{m}}(\mathbb{R})\oplus{\mathfrak{a}}(\mathbb{R})\oplus{\mathfrak{n}}_{+}(\mathbb{R}). Then A=exp(𝔞())=exp(H~0)A=\exp({\mathfrak{a}}(\mathbb{R}))=\exp(\mathbb{R}\widetilde{H}_{0}) and N+=exp(𝔫+())N_{+}=\exp({\mathfrak{n}}_{+}(\mathbb{R})). The group MM is given by

M={(det(g)1g):gSL±(n1,)}SL±(n1,).M=\left\{\begin{pmatrix}\det(g)^{-1}&\\ &g\\ \end{pmatrix}:g\in SL^{\pm}(n-1,\mathbb{R})\right\}\simeq SL^{\pm}(n-1,\mathbb{R}).

Here SL±(1,)SL^{\pm}(1,\mathbb{R}) is regarded as SL±(1,)={±1}SL^{\pm}(1,\mathbb{R})=\{\pm 1\}. As MM is not connected, let M0M_{0} denote the identity component of MM. We write

γ=diag(1,1,,1,1).\gamma=\mathrm{diag}(-1,1,\ldots,1,-1).

Then M0SL(n1,)M_{0}\simeq SL(n-1,\mathbb{R}) and

M/M0={[In],[γ]}/2,M/M_{0}=\{[I_{n}],[\gamma]\}\simeq\mathbb{Z}/2\mathbb{Z},

where InI_{n} is the n×nn\times n identity matrix and [g]=gM0[g]=gM_{0}.

For a closed subgroup JJ of GG, we denote by Irr(J)\mathrm{Irr}(J) and Irr(J)fin\mathrm{Irr}(J)_{\mathrm{fin}} the sets of equivalence classes of irreducible representations of JJ and finite-dimensional irreducible representations of JJ, respectively.

For λ\lambda\in\mathbb{C}, we define a one-dimensional representation λ=(χλ,)\mathbb{C}_{\lambda}=(\chi^{\lambda},\mathbb{C}) of A=exp(H~0)A=\exp(\mathbb{R}\widetilde{H}_{0}) by

χλ:exp(tH~0)exp(λt).\chi^{\lambda}\colon\exp(t\widetilde{H}_{0})\longmapsto\exp(\lambda t). (3.3)

Then Irr(A)\mathrm{Irr}(A) is given by

Irr(A)={λ:λ}.\mathrm{Irr}(A)=\{\mathbb{C}_{\lambda}:\lambda\in\mathbb{C}\}\simeq\mathbb{C}.

For α{±}\alpha\in\{\pm\}, a one-dimensional representation α\mathbb{C}_{\alpha} of MM is defined by

(det(g)1g)sgnα(det(g)),\begin{pmatrix}\det(g)^{-1}&\\ &g\\ \end{pmatrix}\longmapsto\mathrm{sgn}^{\alpha}(\det(g)),

where

sgnα(det(g))={1if α=+,sgn(det(g))if α=.\mathrm{sgn}^{\alpha}(\det(g))=\begin{cases}1&\text{if $\alpha=+$},\\ \mathrm{sgn}(\det(g))&\text{if $\alpha=-$}.\end{cases}

Then Irr(M)fin=Irr(SL±(n1,))fin\mathrm{Irr}(M)_{\mathrm{fin}}=\mathrm{Irr}(SL^{\pm}(n-1,\mathbb{R}))_{\mathrm{fin}} is given by

Irr(M)fin={αϖ:(α,ϖ){±}×Irr(SL(n1,))fin}.\mathrm{Irr}(M)_{\mathrm{fin}}=\{\mathbb{C}_{\alpha}\otimes\varpi:(\alpha,\varpi)\in\{\pm\}\times\mathrm{Irr}(SL(n-1,\mathbb{R}))_{\mathrm{fin}}\}.

Since Irr(P)finIrr(M)fin×Irr(A)\mathrm{Irr}(P)_{\mathrm{fin}}\simeq\mathrm{Irr}(M)_{\mathrm{fin}}\times\mathrm{Irr}(A), the set Irr(P)fin\mathrm{Irr}(P)_{\mathrm{fin}} can be parametrized by

Irr(P)fin{±}×Irr(SL(n1,))fin×.\mathrm{Irr}(P)_{\mathrm{fin}}\simeq\{\pm\}\times\mathrm{Irr}(SL(n-1,\mathbb{R}))_{\mathrm{fin}}\times\mathbb{C}.

For (α,ϖ,λ){±}×Irr(SL(n1,))fin×(\alpha,\varpi,\lambda)\in\{\pm\}\times\mathrm{Irr}(SL(n-1,\mathbb{R}))_{\mathrm{fin}}\times\mathbb{C}, we write

I(ϖ,λ)α=IndPG((αϖ)λ)I(\varpi,\lambda)^{\alpha}=\mathrm{Ind}_{P}^{G}\left((\mathbb{C}_{\alpha}\otimes\varpi)\boxtimes\mathbb{C}_{\lambda}\right) (3.4)

for unnormalized parabolically induced representation IndPG((αϖ)λ)\mathrm{Ind}_{P}^{G}\left((\mathbb{C}_{\alpha}\otimes\varpi)\boxtimes\mathbb{C}_{\lambda}\right) of GG. For instance, the unitary axis of I(triv,λ)αI(\mathrm{triv},\lambda)^{\alpha} is Re(λ)=n2\text{Re}(\lambda)=\frac{n}{2}, where triv\mathrm{triv} denotes the trivial representation of SL(n1,)SL(n-1,\mathbb{R}). Similarly, for the representation space WW of (α,ϖ,λ)Irr(P)fin(\alpha,\varpi,\lambda)\in\mathrm{Irr}(P)_{\mathrm{fin}}, we write

M𝔭(ϖ,λ)α=𝒰(𝔤)𝒰(𝔭)W.M_{\mathfrak{p}}(\varpi,\lambda)^{\alpha}=\mathcal{U}({\mathfrak{g}})\otimes_{\mathcal{U}({\mathfrak{p}})}W. (3.5)

In the next section we classify and construct intertwining differential operators

𝒟DiffG(I(triv,λ)α,I(ϖ,ν)β)\mathcal{D}\in\mathrm{Diff}_{G}(I(\mathrm{triv},\lambda)^{\alpha},I(\varpi,\nu)^{\beta})

and (𝔤,P)({\mathfrak{g}},P)-homomorphisms

φHom𝔤,P(M𝔭(ϖ,ν)α,M𝔭(triv,λ)β).\varphi\in\operatorname{Hom}_{{\mathfrak{g}},P}(M_{\mathfrak{p}}(\varpi,\nu)^{\alpha},M_{\mathfrak{p}}(\mathrm{triv},\lambda)^{\beta}).

4. Classification and construction of 𝒟\mathcal{D} and φ\varphi

The aim of this section is to state the classification and construction results for intertwining differential operators 𝒟DiffG(I(triv,λ)α,I(ϖ,ν)β)\mathcal{D}\in\mathrm{Diff}_{G}(I(\mathrm{triv},\lambda)^{\alpha},I(\varpi,\nu)^{\beta}) as well as (𝔤,P)({\mathfrak{g}},P)-homomorphisms φHom𝔤,P(M𝔭(ϖ,ν)α,M𝔭(triv,λ)β)\varphi\in\operatorname{Hom}_{{\mathfrak{g}},P}(M_{\mathfrak{p}}(\varpi,\nu)^{\alpha},M_{\mathfrak{p}}(\mathrm{triv},\lambda)^{\beta}). These are given in Theorems 4.2 and 4.5 for 𝒟\mathcal{D} and Theorems 4.6 and 4.8 for φ\varphi. The proofs of the theorems will be discussed in Section 5. We remark that the case of n=3n=3 is studied in [56].

4.1. Classification and construction of intertwining differential operators 𝒟\mathcal{D}

We start with the classification of the parameters (α,β;ϖ;λ,ν){±}2×Irr(SL(n1,))fin×2(\alpha,\beta;\varpi;\lambda,\nu)\in\{\pm\}^{2}\times\mathrm{Irr}(SL(n-1,\mathbb{R}))_{\mathrm{fin}}\times\mathbb{C}^{2} such that DiffG(I(triv,λ)α,I(ϖ,ν)β){0}\mathrm{Diff}_{G}(I(\mathrm{triv},\lambda)^{\alpha},I(\varpi,\nu)^{\beta})\neq\{0\}.

For α{±}{±1}\alpha\in\{\pm\}\equiv\{\pm 1\} and k0k\in\mathbb{Z}_{\geq 0}, we mean α+k{±}\alpha+k\in\{\pm\} by

α+k={+if α=(1)k,if α=(1)k+1.\alpha+k=\begin{cases}+&\text{if $\alpha=(-1)^{k}$},\\ -&\text{if $\alpha=(-1)^{k+1}$}.\end{cases}

Then we define ΛGn{±}2×Irr(SL(n1,))fin×2\Lambda^{n}_{G}\subset\{\pm\}^{2}\times\mathrm{Irr}(SL(n-1,\mathbb{R}))_{\mathrm{fin}}\times\mathbb{C}^{2} as

ΛGn={(α,α+k;polyn1k;1k,1+kn1):α{±}andk0},\Lambda^{n}_{G}=\{(\alpha,\alpha+k;\mathrm{poly}_{n-1}^{k};1-k,1+\tfrac{k}{n-1}):\alpha\in\{\pm\}\;\text{and}\;k\in\mathbb{Z}_{\geq 0}\}, (4.1)

where polyn1k\mathrm{poly}_{n-1}^{k} denotes the irreducible representation on Polk(n1)=Sk((n1))\mathrm{Pol}^{k}(\mathbb{C}^{n-1})=S^{k}((\mathbb{C}^{n-1})^{\vee}) of SL(n1,)SL(n-1,\mathbb{R}) induced by the standard action on n1\mathbb{C}^{n-1}. We regard poly1k\mathrm{poly}_{1}^{k} as poly1k=triv\mathrm{poly}_{1}^{k}=\mathrm{triv} for all k0k\in\mathbb{Z}_{\geq 0}.

Theorem 4.2.

The following three conditions on (α,β;ϖ;λ,ν){±}2×Irr(SL(n1,))fin×2(\alpha,\beta;\varpi;\lambda,\nu)\in\{\pm\}^{2}\times\mathrm{Irr}(SL(n-1,\mathbb{R}))_{\mathrm{fin}}\times\mathbb{C}^{2} are equivalent.

  1. (i)

    DiffG(I(triv,λ)α,I(ϖ,ν)β){0}\mathrm{Diff}_{G}(I(\mathrm{triv},\lambda)^{\alpha},I(\varpi,\nu)^{\beta})\neq\{0\}.

  2. (ii)

    dimDiffG(I(triv,λ)α,I(ϖ,ν)β)=1\dim\mathrm{Diff}_{G}(I(\mathrm{triv},\lambda)^{\alpha},I(\varpi,\nu)^{\beta})=1.

  3. (iii)

    One of the following two conditions holds:

    1. (iii-a)

      (β,ϖ,ν)=(α,triv,λ)(\beta,\varpi,\nu)=(\alpha,\mathrm{triv},\lambda).

    2. (iii-b)

      (α,β;ϖ;λ,ν)ΛGn(\alpha,\beta;\varpi;\lambda,\nu)\in\Lambda^{n}_{G}.

We next consider the explicit formula of 𝒟DiffG(I(triv,λ)α,I(ϖ,ν)β)\mathcal{D}\in\mathrm{Diff}_{G}(I(\mathrm{triv},\lambda)^{\alpha},I(\varpi,\nu)^{\beta}) for (α,β;ϖ;λ,ν)(\alpha,\beta;\varpi;\lambda,\nu) in (iii) of Theorem 4.2. We write

Polk(n1)=k[y1,,yn1].\mathrm{Pol}^{k}(\mathbb{C}^{n-1})=\mathbb{C}^{k}[y_{1},\ldots,y_{n-1}].

Then, as in (2.2) and (2.30), we understand 𝒟DiffG(I(triv,λ)α,I(ϖ,ν)β)\mathcal{D}\in\mathrm{Diff}_{G}(I(\mathrm{triv},\lambda)^{\alpha},I(\varpi,\nu)^{\beta}) for (α,β;ϖ;λ,ν)ΛGn(\alpha,\beta;\varpi;\lambda,\nu)\in\Lambda^{n}_{G} as a map

𝒟:C(n1)C(n1)k[y1,,yn1]\mathcal{D}\colon C^{\infty}(\mathbb{R}^{n-1})\longrightarrow C^{\infty}(\mathbb{R}^{n-1})\otimes\mathbb{C}^{k}[y_{1},\ldots,y_{n-1}]

via the diffeomorphism

n1N,(x1,,xn1)exp(x1X1++xn1Xn1).\mathbb{R}^{n-1}\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}N_{-},\quad(x_{1},\ldots,x_{n-1})\mapsto\exp(x_{1}X_{1}+\cdots+x_{n-1}X_{n-1}). (4.3)

For k0k\in\mathbb{Z}_{\geq 0}, we put

Ξk:={(k1,,kn1)(0)n1:j=1n1kj=k}.\Xi_{k}:=\{(k_{1},\ldots,k_{n-1})\in(\mathbb{Z}_{\geq 0})^{n-1}:\sum_{j=1}^{n-1}k_{j}=k\}.

For 𝐤=(k1,,kn1)Ξk\mathbf{k}=(k_{1},\ldots,k_{n-1})\in\Xi_{k}, we write

y~𝐤\displaystyle\widetilde{y}_{\mathbf{k}} =1k1!kn1!y1k1yn1kn1,\displaystyle=\frac{1}{k_{1}!\cdots k_{n-1}!}\cdot y_{1}^{k_{1}}\cdots y_{n-1}^{k_{n-1}},
kx𝐤\displaystyle\frac{\partial^{k}}{\partial x^{\mathbf{k}}}\ =kx1k1xn1kn1.\displaystyle=\frac{\partial^{k}}{\partial x_{1}^{k_{1}}\cdots\partial x_{n-1}^{k_{n-1}}}.

We define 𝒟kDiff(C(n1),C(n1)k[y1,,yn1])\mathcal{D}_{k}\in\mathrm{Diff}_{\mathbb{C}}(C^{\infty}(\mathbb{R}^{n-1}),C^{\infty}(\mathbb{R}^{n-1})\otimes\mathbb{C}^{k}[y_{1},\ldots,y_{n-1}]) for k0k\in\mathbb{Z}_{\geq 0} by

𝒟k=𝐤Ξkkx𝐤y~𝐤.\mathcal{D}_{k}=\sum_{\mathbf{k}\in\Xi_{k}}\frac{\partial^{k}}{\partial x^{\mathbf{k}}}\otimes\widetilde{y}_{\mathbf{k}}. (4.4)

For k=0k=0, we understand 𝒟0\mathcal{D}_{0} as the identity operator 𝒟0=id\mathcal{D}_{0}=\mathrm{id}.

Theorem 4.5.

We have

DiffG(I(triv,λ)α,I(ϖ,ν)β)={idif (β,ϖ,ν)=(α,triv,λ),𝒟kif (α,β;ϖ;λ,ν)ΛGn,{0}otherwise.\mathrm{Diff}_{G}(I(\mathrm{triv},\lambda)^{\alpha},I(\varpi,\nu)^{\beta})=\begin{cases}\mathbb{C}\mathrm{id}&\text{if $(\beta,\varpi,\nu)=(\alpha,\mathrm{triv},\lambda)$,}\\ \mathbb{C}\mathcal{D}_{k}&\text{if $(\alpha,\beta;\varpi;\lambda,\nu)\in\Lambda^{n}_{G}$,}\\ \{0\}&\text{otherwise.}\end{cases}

4.2. Classification and construction of (𝔤,P)({\mathfrak{g}},P)-homomorphisms φ\varphi

Let 𝔤=𝔤()=𝔰𝔩(n,){\mathfrak{g}}={\mathfrak{g}}(\mathbb{R})\otimes_{\mathbb{R}}\mathbb{C}=\mathfrak{sl}(n,\mathbb{C}). We regard 𝔰𝔩(1,)\mathfrak{sl}(1,\mathbb{C}) as 𝔰𝔩(1,)={0}\mathfrak{sl}(1,\mathbb{C})=\{0\}.

Define Λ(𝔤,P)n{±}2×Irr(SL(n1,))fin×2\Lambda^{n}_{({\mathfrak{g}},P)}\subset\{\pm\}^{2}\times\mathrm{Irr}(SL(n-1,\mathbb{R}))_{\mathrm{fin}}\times\mathbb{C}^{2} as

Λ(𝔤,P)n={(α,α+k;symn1k;k1,(1+kn1)):α{±}andk0},\Lambda^{n}_{({\mathfrak{g}},P)}=\{(\alpha,\alpha+k;\mathrm{sym}_{n-1}^{k};k-1,-(1+\tfrac{k}{n-1})):\alpha\in\{\pm\}\;\text{and}\;k\in\mathbb{Z}_{\geq 0}\},

where symn1k\mathrm{sym}_{n-1}^{k} denotes the irreducible representation on Sk(n1)S^{k}(\mathbb{C}^{n-1}) of SL(n1,)SL(n-1,\mathbb{R}). As for poly1k\mathrm{poly}_{1}^{k}, we regard sym1k\mathrm{sym}_{1}^{k} as sym1k=triv\mathrm{sym}_{1}^{k}=\mathrm{triv} for all k0k\in\mathbb{Z}_{\geq 0}.

The classification of the parameters (α,β;σ;s,r){±}2×Irr(SL(n1,))fin×2(\alpha,\beta;\sigma;s,r)\in\{\pm\}^{2}\times\mathrm{Irr}(SL(n-1,\mathbb{R}))_{\mathrm{fin}}\times\mathbb{C}^{2} such that Hom𝔤,P(M𝔭(σ,r)β,M𝔭(triv,s)α){0}\operatorname{Hom}_{{\mathfrak{g}},P}(M_{\mathfrak{p}}(\sigma,r)^{\beta},M_{\mathfrak{p}}(\mathrm{triv},s)^{\alpha})\neq\{0\} is given as follows.

Theorem 4.6.

The following three conditions on (α,β;σ;s,r){±}2×Irr(SL(n1,))fin×2(\alpha,\beta;\sigma;s,r)\in\{\pm\}^{2}\times\mathrm{Irr}(SL(n-1,\mathbb{R}))_{\mathrm{fin}}\times\mathbb{C}^{2} are equivalent.

  1. (i)

    Hom𝔤,P(M𝔭(σ,r)β,M𝔭(triv,s)α){0}\operatorname{Hom}_{{\mathfrak{g}},P}(M_{\mathfrak{p}}(\sigma,r)^{\beta},M_{\mathfrak{p}}(\mathrm{triv},s)^{\alpha})\neq\{0\}.

  2. (ii)

    dimHom𝔤,P(M𝔭(σ,r)β,M𝔭(triv,s)α)=1\dim\operatorname{Hom}_{{\mathfrak{g}},P}(M_{\mathfrak{p}}(\sigma,r)^{\beta},M_{\mathfrak{p}}(\mathrm{triv},s)^{\alpha})=1.

  3. (iii)

    One of the following two conditions holds:

    1. (iii-a)

      (β,σ,r)=(α,triv,s)(\beta,\sigma,r)=(\alpha,\mathrm{triv},s).

    2. (iii-b)

      (α,β;σ;s,r)Λ(𝔤,P)n(\alpha,\beta;\sigma;s,r)\in\Lambda^{n}_{({\mathfrak{g}},P)}.

To give the explicit formula of φHom𝔤,P(M𝔭(σ,r)β,M𝔭(triv,s)α)\varphi\in\operatorname{Hom}_{{\mathfrak{g}},P}(M_{\mathfrak{p}}(\sigma,r)^{\beta},M_{\mathfrak{p}}(\mathrm{triv},s)^{\alpha}), we write

Sk(n1)=k[e1,,en1],S^{k}(\mathbb{C}^{n-1})=\mathbb{C}^{k}[e_{1},\ldots,e_{n-1}],

where eje_{j} are the standard basis elements of n1\mathbb{C}^{n-1}.

For 𝐤=(k1,,kn1)Ξk\mathbf{k}=(k_{1},\ldots,k_{n-1})\in\Xi_{k}, we write

e𝐤\displaystyle e_{\mathbf{k}} =e1k1en1kn1\displaystyle=e_{1}^{k_{1}}\cdots e_{n-1}^{k_{n-1}} Sk(n1),\displaystyle\in S^{k}(\mathbb{C}^{n-1}),
N𝐤\displaystyle N_{\mathbf{k}}^{-} =(N1)k1(Nn1)kn1\displaystyle=(N_{1}^{-})^{k_{1}}\cdots(N_{n-1}^{-})^{k_{n-1}} Sk(𝔫).\displaystyle\in S^{k}({\mathfrak{n}}_{-}).

Observe that we have

k[y1,,yn1]=Polk(n1)=Sk((n1))Sk(n1).\mathbb{C}^{k}[y_{1},\ldots,y_{n-1}]=\mathrm{Pol}^{k}(\mathbb{C}^{n-1})=S^{k}((\mathbb{C}^{n-1})^{\vee})\simeq S^{k}(\mathbb{C}^{n-1})^{\vee}.

We then define yj(n1)y_{j}\in(\mathbb{C}^{n-1})^{\vee} in such a way that yi(ej)=δi,jy_{i}(e_{j})=\delta_{i,j}, which gives y~𝐤(e𝐤)=δ𝐤,𝐤\widetilde{y}_{\mathbf{k}}(e_{\mathbf{k}^{\prime}})=\delta_{\mathbf{k},\mathbf{k}^{\prime}} for 𝐤,𝐤Ξk\mathbf{k},\mathbf{k}^{\prime}\in\Xi_{k}.

We define φkHom(Sk(n1),Sk(𝔫))\varphi_{k}\in\operatorname{Hom}_{\mathbb{C}}(S^{k}(\mathbb{C}^{n-1}),S^{k}({\mathfrak{n}}_{-})) by means of

φk=𝐤ΞkN𝐤(e𝐤)=𝐤ΞkN𝐤y~𝐤.\varphi_{k}=\sum_{\mathbf{k}\in\Xi_{k}}N_{\mathbf{k}}^{-}\otimes(e_{\mathbf{k}})^{\vee}=\sum_{\mathbf{k}\in\Xi_{k}}N_{\mathbf{k}}^{-}\otimes\widetilde{y}_{\mathbf{k}}. (4.7)

Since M𝔭(triv,s)αS(𝔫)M_{\mathfrak{p}}(\mathrm{triv},s)^{\alpha}\simeq S({\mathfrak{n}}_{-}) as linear spaces, we have

φkHom(Sk(n1),M𝔭(triv,s)α).\varphi_{k}\in\operatorname{Hom}_{\mathbb{C}}(S^{k}(\mathbb{C}^{n-1}),M_{\mathfrak{p}}(\mathrm{triv},s)^{\alpha}).

Further, the following holds.

Theorem 4.8.

We have

Hom𝔤,P(M𝔭(σ,r)β,M𝔭(triv,s)α)={idif (β,σ,r)=(α,triv,s),φkif (α,β,σ;s,r)Λ(𝔤,P)n,{0}otherwise.\operatorname{Hom}_{{\mathfrak{g}},P}(M_{\mathfrak{p}}(\sigma,r)^{\beta},M_{\mathfrak{p}}(\mathrm{triv},s)^{\alpha})=\begin{cases}\mathbb{C}\mathrm{id}&\text{if $(\beta,\sigma,r)=(\alpha,\mathrm{triv},s)$,}\\ \mathbb{C}\varphi_{k}&\text{if $(\alpha,\beta,\sigma;s,r)\in\Lambda^{n}_{({\mathfrak{g}},P)}$,}\\ \{0\}&\text{otherwise.}\end{cases}

Here, by abuse of notation, we regard φk\varphi_{k} as a map

φkHom𝔤,P(M𝔭(symn1k,(1+kn1))α+k,M𝔭(triv,k1)α)\varphi_{k}\in\operatorname{Hom}_{{\mathfrak{g}},P}(M_{\mathfrak{p}}(\mathrm{sym}^{k}_{n-1},-(1+\tfrac{k}{n-1}))^{\alpha+k},M_{\mathfrak{p}}(\mathrm{triv},k-1)^{\alpha})

defined by

φk(uw):=uφk(w)for u𝒰(𝔤) and wSk(n1).\varphi_{k}(u\otimes w):=u\varphi_{k}(w)\quad\text{for $u\in\mathcal{U}({\mathfrak{g}})$ and $w\in S^{k}(\mathbb{C}^{n-1})$.}
Remark 4.9.

If Hom𝔤,P(M𝔭(σ,r)β,M𝔭(triv,s)α){0}\operatorname{Hom}_{{\mathfrak{g}},P}(M_{\mathfrak{p}}(\sigma,r)^{\beta},M_{\mathfrak{p}}(\mathrm{triv},s)^{\alpha})\neq\{0\}, then the infinitesimal character of M𝔭(triv,s)αM_{\mathfrak{p}}(\mathrm{triv},s)^{\alpha} agrees with that of M𝔭(σ,r)βM_{\mathfrak{p}}(\sigma,r)^{\beta}. It follows from the arguments on the standardness of the homomorphism φk\varphi_{k} in Section 7 that the infinitesimal characters are indeed equal for (α,β;σ;s,r)Λ(𝔤,P)n(\alpha,\beta;\sigma;s,r)\in\Lambda^{n}_{({\mathfrak{g}},P)}. (See the proof of Theorem 7.4.)

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 𝔫+{\mathfrak{n}}_{+} 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 dπ(σ,λ)(C)d\pi_{(\sigma,\lambda)^{*}}(C) and dπ(σ,λ)^(C)\widehat{d\pi_{(\sigma,\lambda)^{*}}}(C) for C𝔫+C\in{\mathfrak{n}}_{+}

For σ=αtriv\sigma=\alpha\otimes\mathrm{triv} and λχλ\lambda\equiv\chi^{\lambda}, we simply write

dπλ=dπ(αtriv,χλ)d\pi_{\lambda^{*}}=d\pi_{(\alpha\otimes\mathrm{triv},\chi^{\lambda})^{*}}

with λ=2ρ(𝔫+)λdχ\lambda^{*}=2\rho({\mathfrak{n}}_{+})-\lambda d\chi. We put Ex=j=1n1xjxjE_{x}=\sum_{j=1}^{n-1}x_{j}\frac{\partial}{\partial x_{j}} for the Euler homogeneity operator for xx.

Proposition 5.1.

For Nj+𝔫+N_{j}^{+}\in{\mathfrak{n}}_{+}, we have

dπλ(Nj+)=xj{(nλ)+Ex}.d\pi_{\lambda^{*}}(N_{j}^{+})=x_{j}\{(n-\lambda)+E_{x}\}. (5.2)
Proof.

It follows from (2.16) that dπλ(Nj+)d\pi_{\lambda^{*}}(N_{j}^{+}) is given by

dπλ(Nj+)f(n¯)=λ((Ad(n¯1)Nj+)𝔩)f(n¯)(dR((Ad(1)Nj+)𝔫)f)(n¯).d\pi_{\lambda^{*}}(N_{j}^{+})f(\bar{n})=\lambda^{*}((\mathrm{Ad}(\bar{n}^{-1})N_{j}^{+})_{\mathfrak{l}})f(\bar{n})-\big{(}dR((\mathrm{Ad}(\cdot^{-1})N_{j}^{+})_{{\mathfrak{n}}_{-}})f\big{)}(\bar{n}). (5.3)

A direct computation shows that

(Ad(n¯1)Nj+)𝔩\displaystyle(\mathrm{Ad}(\bar{n}^{-1})N_{j}^{+})_{\mathfrak{l}} =xj(E1,1Ej+1,j+1)r=1rjn1xrNr,\displaystyle=x_{j}(E_{1,1}-E_{j+1,j+1})-\sum_{\begin{subarray}{c}r=1\\ r\neq j\end{subarray}}^{n-1}x_{r}N_{r}^{-},
(Ad(n¯1)Nj+)𝔫\displaystyle-(\mathrm{Ad}(\bar{n}^{-1})N_{j}^{+})_{{\mathfrak{n}}_{-}} =xjr=1n1xrNr.\displaystyle=x_{j}\sum_{r=1}^{n-1}x_{r}N_{r}^{-}.

Since λ(E1,1Ej+1,j+1)=nλ\lambda^{*}(E_{1,1}-E_{j+1,j+1})=n-\lambda and dR(Nr)=xrdR(N_{r}^{-})=\frac{\partial}{\partial x_{r}} via the diffeomorphism (4.3), this shows the proposition. ∎

For later convenience, we next give the formula for ζjdπλ^(Nj+)-\zeta_{j}\widehat{d\pi_{\lambda^{*}}}(N_{j}^{+}) instead of dπλ^(Nj+)\widehat{d\pi_{\lambda^{*}}}(N_{j}^{+}). In the following, we silently extend the coordinate functions x1,,xn1x_{1},\ldots,x_{n-1} on 𝔫(){\mathfrak{n}}_{-}(\mathbb{R}) in (5.2) holomorphically to the ones z1,,zn1z_{1},\ldots,z_{n-1} on 𝔫{\mathfrak{n}}_{-} as in Section 2.5.

For j{1,,n1}j\in\{1,\ldots,n-1\}, we write ϑj=ζjζj\vartheta_{j}=\zeta_{j}\frac{\partial}{\partial\zeta_{j}} for the Euler operator for ζj\zeta_{j}. We also write Eζ=j=1n1ϑjE_{\zeta}=\sum_{j=1}^{n-1}\vartheta_{j} for the Euler homogeneity operator for ζ\zeta. Observe that we have Ez^=((n1)+Eζ)\widehat{E_{z}}=-((n-1)+E_{\zeta}).

Proposition 5.4.

For Nj+𝔫+N_{j}^{+}\in{\mathfrak{n}}_{+}, we have

ζjdπλ^(Nj+)=ϑj(λ1+Eζ).-\zeta_{j}\widehat{d\pi_{\lambda^{*}}}(N_{j}^{+})=\vartheta_{j}(\lambda-1+E_{\zeta}). (5.5)
Proof.

This follows from a direct application of the algebraic Fourier transform (2.12) of Weyl algebras to (5.2). ∎

As Eζ|Polk(𝔫+)=kidE_{\zeta}|_{\mathrm{Pol}^{k}({\mathfrak{n}}_{+})}=k\cdot\mathrm{id}, the operator ζjdπλ^(Nj+)|Polk(𝔫+)-\zeta_{j}\widehat{d\pi_{\lambda^{*}}}(N_{j}^{+})|_{\mathrm{Pol}^{k}({\mathfrak{n}}_{+})} restricted to Polk(𝔫+)\mathrm{Pol}^{k}({\mathfrak{n}}_{+}) is given by

ζjdπλ^(Nj+)|Polk(𝔫+)=(λ1+k)ϑj.-\zeta_{j}\widehat{d\pi_{\lambda^{*}}}(N_{j}^{+})|_{\mathrm{Pol}^{k}({\mathfrak{n}}_{+})}=(\lambda-1+k)\vartheta_{j}. (5.6)

In Step 3 (Section 5.3), we use (5.6) to solve the F-system in concern.

Remark 5.7.

In general, the Fourier transformed operator dπλ^(Nj+)\widehat{d\pi_{\lambda^{*}}}(N_{j}^{+}) is not necessarily reduced to a first order operator as in (5.6). In fact, in [45, 47], such operators give rise to the Jacobi differential equation and Gegenbauer differential equation.

5.2. Step 2: Classify and construct ψHomMA(W,Pol(𝔫+)V)\psi\in\operatorname{Hom}_{MA}(W^{\vee},\mathrm{Pol}({\mathfrak{n}}_{+})\otimes V^{\vee})

For (ϖ,W)Irr(M0)fin(\varpi,W)\in\mathrm{Irr}(M_{0})_{\mathrm{fin}}, we write

Wβ=βWW_{\beta}=\mathbb{C}_{\beta}\otimes W

for the representation space of (β,ϖ)Irr(M)fin(\beta,\varpi)\in\mathrm{Irr}(M)_{\mathrm{fin}}. Then, in this step, we wish to classify and construct

ψHomMA(Wβν,Polk(𝔫+)αλ).\psi\in\operatorname{Hom}_{MA}(W_{\beta}^{\vee}\boxtimes\mathbb{C}_{-\nu},\mathrm{Pol}^{k}({\mathfrak{n}}_{+})_{\alpha}\otimes\mathbb{C}_{-\lambda}).

5.2.1. Notation

We start by introducing some notation. For 𝐤Ξk\mathbf{k}\in\Xi_{k}, we write

ζ𝐤=ζ1k1ζn1kn1Polk(𝔫+).\zeta_{\mathbf{k}}=\zeta_{1}^{k_{1}}\cdots\zeta_{n-1}^{k_{n-1}}\in\mathrm{Pol}^{k}({\mathfrak{n}}_{+}).

We then define ψkHom(Sk(n1),Polk(𝔫+))\psi_{k}\in\operatorname{Hom}_{\mathbb{C}}(S^{k}(\mathbb{C}^{n-1}),\mathrm{Pol}^{k}({\mathfrak{n}}_{+})) by

ψk=𝐤Ξkζ𝐤y~𝐤,\psi_{k}=\sum_{\mathbf{k}\in\Xi_{k}}\zeta_{\mathbf{k}}\otimes\widetilde{y}_{\mathbf{k}}, (5.8)

where y~𝐤Polk(n1)Sk(n1)\widetilde{y}_{\mathbf{k}}\in\mathrm{Pol}^{k}(\mathbb{C}^{n-1})\simeq S^{k}(\mathbb{C}^{n-1})^{\vee} are regarded as the dual basis of e𝐤Sk(n1)e_{\mathbf{k}}\in S^{k}(\mathbb{C}^{n-1}).

Recall from (2.24) that MAMA acts on Pol(𝔫+)\mathrm{Pol}({\mathfrak{n}}_{+}) via the action Ad#\mathrm{Ad}_{\#}. In the present case, (Ad#,Polk(𝔫+))(\mathrm{Ad}_{\#},\mathrm{Pol}^{k}({\mathfrak{n}}_{+})) is an irreducible representation of M0SL(n1,)M_{0}\simeq SL(n-1,\mathbb{R}), which is equivalent to

(M0,Ad#,Polk(𝔫+))(M0,Ad𝔫k,Sk(𝔫))(SL(n1,),symn1k,Sk(n1)).(M_{0},\mathrm{Ad}_{\#},\mathrm{Pol}^{k}({\mathfrak{n}}_{+}))\simeq(M_{0},\mathrm{Ad}^{k}_{{\mathfrak{n}}_{-}},S^{k}({\mathfrak{n}}_{-}))\simeq(SL(n-1,\mathbb{R}),\mathrm{sym}_{n-1}^{k},S^{k}(\mathbb{C}^{n-1})). (5.9)

Since ψk\psi_{k} maps ψk:e𝐤ζ𝐤\psi_{k}\colon e_{\mathbf{k}}\mapsto\zeta_{\mathbf{k}}, we have

ψkHomM0(Sk(n1),Polk(𝔫+)).\psi_{k}\in\operatorname{Hom}_{M_{0}}(S^{k}(\mathbb{C}^{n-1}),\mathrm{Pol}^{k}({\mathfrak{n}}_{+})). (5.10)

5.2.2. Classification and construction of ψHomMA(W,Pol(𝔫+)V)\psi\in\operatorname{Hom}_{MA}(W^{\vee},\mathrm{Pol}({\mathfrak{n}}_{+})\otimes V^{\vee})

As

HomMA(Wβν,Polk(𝔫+)αλ)HomM0(W,Polk(𝔫+)),\operatorname{Hom}_{MA}(W_{\beta}^{\vee}\boxtimes\mathbb{C}_{-\nu},\mathrm{Pol}^{k}({\mathfrak{n}}_{+})_{\alpha}\otimes\mathbb{C}_{-\lambda})\subset\operatorname{Hom}_{M_{0}}(W^{\vee},\mathrm{Pol}^{k}({\mathfrak{n}}_{+})),

we first consider HomM0(W,Polk(𝔫+))\operatorname{Hom}_{M_{0}}(W^{\vee},\mathrm{Pol}^{k}({\mathfrak{n}}_{+})).

Since (symn1k,Sk(n1))(polyn1k,Polk(n1))(\mathrm{sym}_{n-1}^{k},S^{k}(\mathbb{C}^{n-1}))^{\vee}\simeq(\mathrm{poly}_{n-1}^{k},\mathrm{Pol}^{k}(\mathbb{C}^{n-1})), it follows from (5.9) that the following two conditions on a representation WW of M0SL(n1,)M_{0}\simeq SL(n-1,\mathbb{R}) are equivalent.

  1. (i)

    W(Ad#,Polk(𝔫+))W^{\vee}\simeq(\mathrm{Ad}_{\#},\mathrm{Pol}^{k}({\mathfrak{n}}_{+})).

  2. (ii)

    W(polyn1k,Polk(n1))W\simeq(\mathrm{poly}_{n-1}^{k},\mathrm{Pol}^{k}(\mathbb{C}^{n-1})).

Now HomM0(W,Polk(𝔫+))\operatorname{Hom}_{M_{0}}(W^{\vee},\mathrm{Pol}^{k}({\mathfrak{n}}_{+})) is given as follows.

Proposition 5.11.

The following three conditions on a representation WW of M0SL(n1,)M_{0}\simeq SL(n-1,\mathbb{R}) are equivalent.

  1. (i)

    HomM0(W,Polk(𝔫+)){0}\operatorname{Hom}_{M_{0}}(W^{\vee},\mathrm{Pol}^{k}({\mathfrak{n}}_{+}))\neq\{0\}.

  2. (ii)

    dimHomM0(W,Polk(𝔫+))=1\dim\operatorname{Hom}_{M_{0}}(W^{\vee},\mathrm{Pol}^{k}({\mathfrak{n}}_{+}))=1.

  3. (iii)

    W(polyn1k,Polk(n1))W\simeq(\mathrm{poly}_{n-1}^{k},\mathrm{Pol}^{k}(\mathbb{C}^{n-1})).

Consequently, we have

HomM0(W,Polk(𝔫+))={ψkif (ϖ,W)(polyn1k,Polk(n1)),{0}otherwise.\operatorname{Hom}_{M_{0}}(W^{\vee},\mathrm{Pol}^{k}({\mathfrak{n}}_{+}))=\begin{cases}\mathbb{C}\psi_{k}&\text{if $(\varpi,W)\simeq(\mathrm{poly}_{n-1}^{k},\mathrm{Pol}^{k}(\mathbb{C}^{n-1}))$,}\\ \{0\}&\text{otherwise.}\end{cases} (5.12)
Proof.

As the M0M_{0}-representation (Ad#,Polk(𝔫+))(\mathrm{Ad}_{\#},\mathrm{Pol}^{k}({\mathfrak{n}}_{+})) is irreducible, the first assertion follows from Schur’s lemma and the preceding argument. The second assertion (5.12) is a simple consequence of the first and (5.10). ∎

A direct computation shows that we have

(M,Ad𝔫k,Sk(𝔫))(SL±(n1,),sgnksymn1k,Sk(n1)).(M,\mathrm{Ad}^{k}_{{\mathfrak{n}}_{-}},S^{k}({\mathfrak{n}}_{-}))\simeq(SL^{\pm}(n-1,\mathbb{R}),\mathrm{sgn}^{k}\otimes\mathrm{sym}_{n-1}^{k},S^{k}(\mathbb{C}^{n-1})).

Here we mean sgnk\mathrm{sgn}^{k} by sgnk=triv\mathrm{sgn}^{k}=\mathrm{triv} if k0(mod 2)k\equiv 0\ (\mathrm{mod}\ 2); sgn\mathrm{sgn} if k1(mod 2)k\equiv 1\ (\mathrm{mod}\ 2). Since (M,Ad#,Polk(𝔫+))(M,Ad𝔫k,Sk(𝔫))(M,\mathrm{Ad}_{\#},\mathrm{Pol}^{k}({\mathfrak{n}}_{+}))\simeq(M,\mathrm{Ad}^{k}_{{\mathfrak{n}}_{-}},S^{k}({\mathfrak{n}}_{-})), it implies

(M,Ad#,Polk(𝔫+))(SL±(n1,),sgnksymn1k,Sk(n1)).(M,\mathrm{Ad}_{\#},\mathrm{Pol}^{k}({\mathfrak{n}}_{+}))\simeq(SL^{\pm}(n-1,\mathbb{R}),\mathrm{sgn}^{k}\otimes\mathrm{sym}_{n-1}^{k},S^{k}(\mathbb{C}^{n-1})). (5.13)
Proposition 5.14.

We have

HomMA(Wβν,Polk(𝔫+)αλ)={ψkif (α,β;ϖ;λ,ν)=(α,α+k;polyn1k;λ,λ+nn1k),{0}otherwise.\operatorname{Hom}_{MA}(W_{\beta}^{\vee}\boxtimes\mathbb{C}_{-\nu},\mathrm{Pol}^{k}({\mathfrak{n}}_{+})_{\alpha}\otimes\mathbb{C}_{-\lambda})=\begin{cases}\mathbb{C}\psi_{k}&\text{if $(\alpha,\beta;\varpi;\lambda,\nu)=(\alpha,\alpha+k;\mathrm{poly}_{n-1}^{k};\lambda,\lambda+\frac{n}{n-1}k)$,}\\ \{0\}&\text{otherwise.}\end{cases}
Proof.

It follows from (3.1) that, via the action Ad#\mathrm{Ad}_{\#}, the group A=exp(H~0)A=\exp(\mathbb{R}\widetilde{H}_{0}) acts on Polk(𝔫+)\mathrm{Pol}^{k}({\mathfrak{n}}_{+}) by a character χnn1k\chi^{-\frac{n}{n-1}k}. Now Proposition 5.11 and (5.13) conclude the assertion. ∎

In Proposition 5.14, the value of λ\lambda\in\mathbb{C} is still arbitrary. One approach to determine λ\lambda\in\mathbb{C} for which DiffG(I(triv,λ)α,I(ϖ,ν)β){0}\mathrm{Diff}_{G}(I(\mathrm{triv},\lambda)^{\alpha},I(\varpi,\nu)^{\beta})\neq\{0\} is to solve the square norm of the infinitesimal characters in question. (See, for instance, [3, Sect. 7].) In the following, we instead compute the 𝔫+{\mathfrak{n}}_{+}-invariance via the F-method to obtain such λ\lambda\in\mathbb{C} more directly.

5.3. Step 3: Solve the F-system for ψHomMA(W,Pol(𝔫+)V)\psi\in\operatorname{Hom}_{MA}(W^{\vee},\mathrm{Pol}({\mathfrak{n}}_{+})\otimes V^{\vee})

For (α,β;ϖ;λ,ν){±}2×Irr(SL(n1,))fin×2(\alpha,\beta;\varpi;\lambda,\nu)\in\{\pm\}^{2}\times\mathrm{Irr}(SL(n-1,\mathbb{R}))_{\mathrm{fin}}\times\mathbb{C}^{2} and k0k\in\mathbb{Z}_{\geq 0}, we put

Solk(𝔫+;trivα,λ,ϖβ,ν)\displaystyle\mathrm{Sol}^{k}({\mathfrak{n}}_{+};\mathrm{triv}_{\alpha,\lambda},\varpi_{\beta,\nu})
:={ψHomMA(Wβν,Polk(𝔫+)αλ): ψ solves the F-system (5.15) below.}.\displaystyle:=\{\psi\in\operatorname{Hom}_{MA}(W_{\beta}^{\vee}\boxtimes\mathbb{C}_{-\nu},\mathrm{Pol}^{k}({\mathfrak{n}}_{+})_{\alpha}\otimes\mathbb{C}_{-\lambda}):\text{ $\psi$ solves the F-system \eqref{eqn:Fsys2} below.}\}.
(dπλ^(Nj+)idW)ψ=0for all j{1,,n1}.(\widehat{d\pi_{\lambda^{*}}}(N_{j}^{+})\otimes\mathrm{id}_{W})\psi=0\quad\text{for all $j\in\{1,\ldots,n-1\}$}. (5.15)

Since

Solk(𝔫+;trivα,λ,ϖβ,ν)HomMA(Wβν,Polk(𝔫+)αλ),\mathrm{Sol}^{k}({\mathfrak{n}}_{+};\mathrm{triv}_{\alpha,\lambda},\varpi_{\beta,\nu})\subset\operatorname{Hom}_{MA}(W_{\beta}^{\vee}\boxtimes\mathbb{C}_{-\nu},\mathrm{Pol}^{k}({\mathfrak{n}}_{+})_{\alpha}\otimes\mathbb{C}_{-\lambda}),

Proposition 5.14 shows that if Solk(𝔫+;trivα,λ,ϖβ,ν){0}\mathrm{Sol}^{k}({\mathfrak{n}}_{+};\mathrm{triv}_{\alpha,\lambda},\varpi_{\beta,\nu})\neq\{0\}, then (α,β;ϖ;λ,ν)(\alpha,\beta;\varpi;\lambda,\nu) must satisfy

(α,β;ϖ;λ,ν)=(α,α+k;polyn1k;λ,λ+nn1k).(\alpha,\beta;\varpi;\lambda,\nu)=(\alpha,\alpha+k;\mathrm{poly}_{n-1}^{k};\lambda,\lambda+\frac{n}{n-1}k). (5.16)
Theorem 5.17.

Let (α,β;ϖ;λ,ν){±}2×Irr(SL(n1,))fin×2(\alpha,\beta;\varpi;\lambda,\nu)\in\{\pm\}^{2}\times\mathrm{Irr}(SL(n-1,\mathbb{R}))_{\mathrm{fin}}\times\mathbb{C}^{2}. The following conditions on (α,β;ϖ;λ,ν)(\alpha,\beta;\varpi;\lambda,\nu) are equivalent.

  1. (i)

    Solk(𝔫+;trivα,λ,ϖβ,ν){0}\mathrm{Sol}^{k}({\mathfrak{n}}_{+};\mathrm{triv}_{\alpha,\lambda},\varpi_{\beta,\nu})\neq\{0\}.

  2. (ii)

    One of the following two conditions holds.

    1. (ii-a)

      (β,ϖ,ν)=(α,triv,λ)(\beta,\varpi,\nu)=(\alpha,\mathrm{triv},\lambda).

    2. (ii-b)

      (α,β;ϖ;λ,ν)=(α,α+k;polyn1k;1k,1+kn1)(\alpha,\beta;\varpi;\lambda,\nu)=(\alpha,\alpha+k;\mathrm{poly}_{n-1}^{k};1-k,1+\tfrac{k}{n-1}).

Further, the space Solk(𝔫+;trivα,λ,ϖβ,ν)\mathrm{Sol}^{k}({\mathfrak{n}}_{+};\mathrm{triv}_{\alpha,\lambda},\varpi_{\beta,\nu}) is given as follows.

  1. (1)

    (β,ϖ,ν)=(α,triv,λ)(\beta,\varpi,\nu)=(\alpha,\mathrm{triv},\lambda):

    Sol0(𝔫+;trivα,λ,trivα,λ)=.\mathrm{Sol}^{0}({\mathfrak{n}}_{+};\mathrm{triv}_{\alpha,\lambda},\mathrm{triv}_{\alpha,\lambda})=\mathbb{C}.
  2. (2)

    (α,β;ϖ;λ,ν)=(α,α+k;polyn1k;1k,1+kn1)(\alpha,\beta;\varpi;\lambda,\nu)=(\alpha,\alpha+k;\mathrm{poly}_{n-1}^{k};1-k,1+\tfrac{k}{n-1}):

    Solk(𝔫+;trivα,λ,ϖβ,ν)=ψk,\mathrm{Sol}^{k}({\mathfrak{n}}_{+};\mathrm{triv}_{\alpha,\lambda},\varpi_{\beta,\nu})=\mathbb{C}\psi_{k},

    where ψk\psi_{k} is the map defined in (5.8).

Proof.

Observe that ψHomMA(Wβν,Polk(𝔫+)αλ)\psi\in\operatorname{Hom}_{MA}(W_{\beta}^{\vee}\boxtimes\mathbb{C}_{-\nu},\mathrm{Pol}^{k}({\mathfrak{n}}_{+})_{\alpha}\otimes\mathbb{C}_{-\lambda}) solves (5.15) if and only if it satisfies a system of PDEs

(ζjdπλ^(Nj+)idW)ψk=0for all j{1,,n1}.(-\zeta_{j}\widehat{d\pi_{\lambda^{*}}}(N_{j}^{+})\otimes\mathrm{id}_{W})\psi_{k}=0\quad\text{for all $j\in\{1,\ldots,n-1\}$.}

By (5.8), we have

(ζjdπλ^(Nj+)idW)ψk=𝐤Ξkζjdπλ^(Nj+)(ζ𝐤)y~𝐤.(-\zeta_{j}\widehat{d\pi_{\lambda^{*}}}(N_{j}^{+})\otimes\mathrm{id}_{W})\psi_{k}=\sum_{\mathbf{k}\in\Xi_{k}}-\zeta_{j}\widehat{d\pi_{\lambda^{*}}}(N_{j}^{+})(\zeta_{\mathbf{k}})\otimes\widetilde{y}_{\mathbf{k}}.

So, one wishes to solve

ζjdπλ^(Nj+)(ζ𝐤)=0for all j{1,,n1} and 𝐤Ξk.-\zeta_{j}\widehat{d\pi_{\lambda^{*}}}(N_{j}^{+})(\zeta_{\mathbf{k}})=0\quad\text{for all $j\in\{1,\ldots,n-1\}$ and $\mathbf{k}\in\Xi_{k}$}. (5.18)

It follows from (5.6) that (5.18) can be simplified to

(λ1+k)ϑj(ζ𝐤)=0for all j{1,,n1} and 𝐤Ξk.(\lambda-1+k)\vartheta_{j}(\zeta_{\mathbf{k}})=0\quad\text{for all $j\in\{1,\ldots,n-1\}$ and $\mathbf{k}\in\Xi_{k}$}. (5.19)

We have (λ1+k)ϑj(ζ𝐤)=(λ1+k)kjζ𝐤(\lambda-1+k)\vartheta_{j}(\zeta_{\mathbf{k}})=(\lambda-1+k)k_{j}\zeta_{\mathbf{k}}. Thus, (5.19) holds if and only if k=0k=0 or λ=1k\lambda=1-k. Now the theorem follows from (5.16). ∎

As for the MAMA-decomposition Pol(𝔫+)=k0Polk(𝔫+)\mathrm{Pol}({\mathfrak{n}}_{+})=\bigoplus_{k\in\mathbb{Z}_{\geq 0}}\mathrm{Pol}^{k}({\mathfrak{n}}_{+}), we put

Sol(𝔫+;trivα,λ,ϖβ,ν):=k0Solk(𝔫+;trivα,λ,ϖβ,ν).\mathrm{Sol}({\mathfrak{n}}_{+};\mathrm{triv}_{\alpha,\lambda},\varpi_{\beta,\nu}):=\bigoplus_{k\in\mathbb{Z}_{\geq 0}}\mathrm{Sol}^{k}({\mathfrak{n}}_{+};\mathrm{triv}_{\alpha,\lambda},\varpi_{\beta,\nu}). (5.20)

Recall from (4.1) that we have

ΛGn={(α,α+k;polyn1k;1k,1+kn1):α{±}andk0}.\Lambda^{n}_{G}=\{(\alpha,\alpha+k;\mathrm{poly}_{n-1}^{k};1-k,1+\tfrac{k}{n-1}):\alpha\in\{\pm\}\;\text{and}\;k\in\mathbb{Z}_{\geq 0}\}.

Corollary 5.21 below is then a direct consequence of Theorem 5.17.

Corollary 5.21.

For (α,β;ϖ;λ,ν){±}2×Irr(SL(n1,))fin×2(\alpha,\beta;\varpi;\lambda,\nu)\in\{\pm\}^{2}\times\mathrm{Irr}(SL(n-1,\mathbb{R}))_{\mathrm{fin}}\times\mathbb{C}^{2}, we have

Sol(𝔫+;trivα,λ,ϖβ,ν)={if (β,ϖ,ν)=(α,triv,λ),ψkif (α,β;ϖ;λ,ν)ΛGn,{0}otherwise.\mathrm{Sol}({\mathfrak{n}}_{+};\mathrm{triv}_{\alpha,\lambda},\varpi_{\beta,\nu})=\begin{cases}\mathbb{C}&\text{if $(\beta,\varpi,\nu)=(\alpha,\mathrm{triv},\lambda)$,}\\ \mathbb{C}\psi_{k}&\text{if $(\alpha,\beta;\varpi;\lambda,\nu)\in\Lambda^{n}_{G}$,}\\ \{0\}&\text{otherwise.}\end{cases}

Here ψk\psi_{k} is the map defined in (5.8).

5.4. Step 4ab: Apply symb1\mathrm{symb}^{-1} and Fc1idWF_{c}^{-1}\otimes\mathrm{id}_{W} to the solutions ψ\psi to the F-system

Observe that 𝒟k\mathcal{D}_{k} in (4.4) and φk\varphi_{k} in (4.7) are given by

𝒟k\displaystyle\mathcal{D}_{k} =𝐤Ξkkx𝐤y~𝐤\displaystyle=\sum_{\mathbf{k}\in\Xi_{k}}\frac{\partial^{k}}{\partial x^{\mathbf{k}}}\otimes\widetilde{y}_{\mathbf{k}} =𝐤Ξksymb1(ζ𝐤)y~𝐤\displaystyle=\sum_{\mathbf{k}\in\Xi_{k}}\mathrm{symb}^{-1}(\zeta_{\mathbf{k}})\otimes\widetilde{y}_{\mathbf{k}} =symb1(ψk),\displaystyle=\mathrm{symb}^{-1}(\psi_{k}),
φk\displaystyle\varphi_{k} =𝐤ΞkN𝐤y~𝐤\displaystyle=\sum_{\mathbf{k}\in\Xi_{k}}N_{\mathbf{k}}^{-}\otimes\widetilde{y}_{\mathbf{k}} =𝐤ΞkFc1(ζ𝐤)y~𝐤\displaystyle=\sum_{\mathbf{k}\in\Xi_{k}}F_{c}^{-1}(\zeta_{\mathbf{k}})\otimes\widetilde{y}_{\mathbf{k}} =(Fc1idW)(ψk).\displaystyle=(F_{c}^{-1}\otimes\mathrm{id}_{W})(\psi_{k}).

Now we are ready to prove Theorems 4.2 and 4.5 and Theorems 4.6 and 4.8.

Proofs of Theorems 4.2, 4.5, 4.6, and 4.8.

By Theorem 2.32, we have

DiffG(I(triv,λ)α,I(ϖ,ν)β)\displaystyle\mathrm{Diff}_{G}(I(\mathrm{triv},\lambda)^{\alpha},I(\varpi,\nu)^{\beta}) =symb1(Sol(𝔫+;trivα,λ,ϖβ,ν)),\displaystyle=\mathrm{symb}^{-1}(\mathrm{Sol}({\mathfrak{n}}_{+};\mathrm{triv}_{\alpha,\lambda},\varpi_{\beta,\nu})),
Hom𝔤,P(M𝔭(ϖ,ν)α,M𝔭(triv,λ)β)\displaystyle\operatorname{Hom}_{{\mathfrak{g}},P}(M_{\mathfrak{p}}(\varpi^{\vee},-\nu)^{\alpha},M_{\mathfrak{p}}(\mathrm{triv},-\lambda)^{\beta}) =(Fc1idW)(Sol(𝔫+;trivα,λ,ϖβ,ν)).\displaystyle=(F_{c}^{-1}\otimes\mathrm{id}_{W})(\mathrm{Sol}({\mathfrak{n}}_{+};\mathrm{triv}_{\alpha,\lambda},\varpi_{\beta,\nu})).

Since (polyn1k)symn1k(\mathrm{poly}_{n-1}^{k})^{\vee}\simeq\mathrm{sym}_{n-1}^{k}, the theorems in consideration follow from Corollary 5.21. ∎

5.5. Classification and construction of 𝔤{\mathfrak{g}}-homomorphisms

Theorem 4.8 concerns (𝔤,P)({\mathfrak{g}},P)-homomorphisms between generalized Verma modules. Then we finish this section by showing the classification and construction of 𝔤{\mathfrak{g}}-homomorphisms.

Let P0P_{0} be the identity component of the parabolic subgroup PP. Then we have P0=M0AN+P_{0}=M_{0}AN_{+}. Thus, Irr(P0)fin\mathrm{Irr}(P_{0})_{\mathrm{fin}} is given by

Irr(P0)finIrr(M0)fin×Irr(A)Irr(𝔰𝔩(n1,))fin×Irr(𝔭)fin.\mathrm{Irr}(P_{0})_{\mathrm{fin}}\simeq\mathrm{Irr}(M_{0})_{\mathrm{fin}}\times\mathrm{Irr}(A)\simeq\mathrm{Irr}(\mathfrak{sl}(n-1,\mathbb{C}))_{\mathrm{fin}}\times\mathbb{C}\simeq\mathrm{Irr}({\mathfrak{p}})_{\mathrm{fin}}.

For (σ,s)Irr(𝔰𝔩(n1,))fin×(\sigma,s)\in\mathrm{Irr}(\mathfrak{sl}(n-1,\mathbb{C}))_{\mathrm{fin}}\times\mathbb{C}, we define a generalized Verma module M𝔭(σ,s)M_{\mathfrak{p}}(\sigma,s) as in (3.5) as a 𝔤{\mathfrak{g}}-module. For (ϖ;λ,ν)Irr(𝔰𝔩(n1,))fin×2(\varpi;\lambda,\nu)\in\mathrm{Irr}(\mathfrak{sl}(n-1,\mathbb{C}))_{\mathrm{fin}}\times\mathbb{C}^{2}, we let

Sol(𝔫+;trivλ,ϖν)={ψHomM0A(Wν,Pol(𝔫+)λ): ψ solves the F-system (5.15).}.\mathrm{Sol}({\mathfrak{n}}_{+};\mathrm{triv}_{\lambda},\varpi_{\nu})=\{\psi\in\operatorname{Hom}_{M_{0}A}(W^{\vee}\boxtimes\mathbb{C}_{-\nu},\mathrm{Pol}({\mathfrak{n}}_{+})\otimes\mathbb{C}_{-\lambda}):\text{ $\psi$ solves the F-system \eqref{eqn:Fsys2}.}\}.

Then, by Remark 2.10, we have

FcidW:Hom𝔤(M𝔭(ϖ,ν),M𝔭(triv,λ))Sol(𝔫+;trivλ,ϖν).F_{c}\otimes\mathrm{id}_{W}\colon\operatorname{Hom}_{\mathfrak{g}}(M_{\mathfrak{p}}(\varpi^{\vee},-\nu),M_{\mathfrak{p}}(\mathrm{triv},-\lambda))\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\mathrm{Sol}({\mathfrak{n}}_{+};\mathrm{triv}_{\lambda},\varpi_{\nu}). (5.22)

Define Λ𝔤nIrr(𝔰𝔩(n1,))fin×2\Lambda^{n}_{{\mathfrak{g}}}\subset\mathrm{Irr}(\mathfrak{sl}(n-1,\mathbb{C}))_{\mathrm{fin}}\times\mathbb{C}^{2} as

Λ𝔤n={(symn1k;k1,(1+kn1)):k0}.\Lambda^{n}_{{\mathfrak{g}}}=\{(\mathrm{sym}_{n-1}^{k};k-1,-(1+\tfrac{k}{n-1})):k\in\mathbb{Z}_{\geq 0}\}.
Theorem 5.23.

We have

Hom𝔤(M𝔭(σ,r),M𝔭(triv,s))={idif (σ,r)=(triv,s),φkif (σ;s,r)Λ𝔤n,{0}otherwise.\operatorname{Hom}_{{\mathfrak{g}}}(M_{\mathfrak{p}}(\sigma,r),M_{\mathfrak{p}}(\mathrm{triv},s))=\begin{cases}\mathbb{C}\mathrm{id}&\text{if $(\sigma,r)=(\mathrm{triv},s)$,}\\ \mathbb{C}\varphi_{k}&\text{if $(\sigma;s,r)\in\Lambda_{{\mathfrak{g}}}^{n}$,}\\ \{0\}&\text{otherwise.}\end{cases}
Proof.

By (5.22), it suffices to compute Sol(𝔫+;trivλ,ϖν)\mathrm{Sol}({\mathfrak{n}}_{+};\mathrm{triv}_{\lambda},\varpi_{\nu}). Then the assertion simply follows from the same arguments in Step 1 (Section 5.1) – Step 4 (Section 5.4). ∎

Corollary 5.24.

The following are equivalent on ss\in\mathbb{C}.

  1. (i)

    M𝔭(triv,s)M_{\mathfrak{p}}(\mathrm{triv},s) is reducible.

  2. (ii)

    s0s\in\mathbb{Z}_{\geq 0}.

Proof.

Observe that M𝔭(triv,s)M_{\mathfrak{p}}(\mathrm{triv},s) is reducible if and only if there exists (σ,r)Irr(𝔭)fin\{(triv,s)}(\sigma,r)\in\mathrm{Irr}({\mathfrak{p}})_{\mathrm{fin}}\backslash\{(\mathrm{triv},s)\} such that Hom𝔤(M𝔭(σ,r),M𝔭(triv,s)){0}\operatorname{Hom}_{{\mathfrak{g}}}(M_{\mathfrak{p}}(\sigma,r),M_{\mathfrak{p}}(\mathrm{triv},s))\neq\{0\}. Now the assertion follows from Theorem 5.23. ∎

Remark 5.25.

The results of Corollary 5.24 is already known in the literature. See, for instance, [1, Ex. 5.2], [24, Thm. 1.1], and [25, p. 794, Table 1].

6. KK-type formulas for Ker(𝒟k)α\mathrm{Ker}(\mathcal{D}_{k})^{\alpha} and Im(𝒟k)α\mathrm{Im}(\mathcal{D}_{k})^{\alpha}

The aim of this section is to classify the KK-type formulas of the kernel Ker(𝒟k)α\mathrm{Ker}(\mathcal{D}_{k})^{\alpha} and image Im(𝒟k)α\mathrm{Im}(\mathcal{D}_{k})^{\alpha} of the non-zero intertwining differential operator

𝒟k:I(triv,1k)αI(polyn1k,1+kn1)α+k.\mathcal{D}_{k}\colon I(\mathrm{triv},1-k)^{\alpha}\longrightarrow I(\mathrm{poly}_{n-1}^{k},1+\tfrac{k}{n-1})^{\alpha+k}.

The KK-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 n=2n=2, to avoid the complication of the exposition, we constraint ourselves to the case n3n\geq 3.

6.1. Composition factors and KK-type structure of I(triv,λ)αI(\mathrm{triv},\lambda)^{\alpha}

Let K=SO(n)K=SO(n) be a maximal compact subgroup of G=SL(n,)G=SL(n,\mathbb{R}). Then we have

KM=\displaystyle K\cap M= S(O(1)×O(n1))\displaystyle S(O(1)\times O(n-1))
={(det(g)1g):gO(n1)}\displaystyle=\left\{\begin{pmatrix}\det(g)^{-1}&\\ &g\end{pmatrix}:g\in O(n-1)\right\}
O(n1).\displaystyle\simeq O(n-1).

We remark that KM/(KM)0M/M0/2K\cap M/(K\cap M)_{0}\simeq M/M_{0}\simeq\mathbb{Z}/2\mathbb{Z}.

For a representation VV of GG, we denote by VKV_{K} the space of KK-finite vectors of VV. Since G=KPG=KP, as KK-representations, we have

I(triv,λ)KαIndKMK(α)K.I(\mathrm{triv},\lambda)^{\alpha}_{K}\simeq\mathrm{Ind}_{K\cap M}^{K}(\alpha)_{K}.

Let m(n)\mathcal{H}^{m}(\mathbb{R}^{n}) be the irreducible representation of KK consisting of spherical harmonics on Sn1nS^{n-1}\subset\mathbb{R}^{n} of homogeneous degree mm (cf. [43, Sect. 7.5]). It is known that the KK-type decomposition of I(triv,λ)KαIndKMK(α)KI(\mathrm{triv},\lambda)^{\alpha}_{K}\simeq\mathrm{Ind}_{K\cap M}^{K}(\alpha)_{K} is given as follows (see, for instance, [26, p. 286]).

I(triv,λ)Kα\displaystyle I(\mathrm{triv},\lambda)^{\alpha}_{K} α=(1)mm0m(n)\displaystyle\simeq\bigoplus_{\stackrel{{\scriptstyle m\in\mathbb{Z}_{\geq 0}}}{{\alpha=(-1)^{m}}}}\mathcal{H}^{m}(\mathbb{R}^{n})
={02(n)if α=+,02+1(n)if α=.\displaystyle=\begin{cases}\bigoplus_{\ell\in\mathbb{Z}_{\geq 0}}\mathcal{H}^{2\ell}(\mathbb{R}^{n})&\text{if $\alpha=+$,}\\[5.0pt] \bigoplus_{\ell\in\mathbb{Z}_{\geq 0}}\mathcal{H}^{2\ell+1}(\mathbb{R}^{n})&\text{if $\alpha=-$.}\end{cases} (6.1)

Theorem 6.2 below states well-known facts on the irreducibility and composition series of I(triv,λ)αI(\mathrm{triv},\lambda)^{\alpha}. For the proof, see, for instance, [26, 71] for α=±\alpha=\pm and [62] for α=+\alpha=+.

Theorem 6.2.

Let n3n\geq 3. For α{±}{±1}\alpha\in\{\pm\}\equiv\{\pm 1\} and λ\lambda\in\mathbb{C}, I(triv,λ)αI(\mathrm{triv},\lambda)^{\alpha} enjoys the following.

  1. (1)

    The induced representation I(triv,λ)αI(\mathrm{triv},\lambda)^{\alpha} is irreducible except the following two cases.

    1. (A)

      λ0\lambda\in-\mathbb{Z}_{\geq 0} and α=(1)λ\alpha=(-1)^{\lambda}.

    2. (B)

      λn+0\lambda\in n+\mathbb{Z}_{\geq 0} and α=(1)λ+n\alpha=(-1)^{\lambda+n}.

  2. (2)

    For Case (A) with λ=m\lambda=-m, there exists a finite-dimensional irreducible subrepresentation F(m)αI(triv,m)αF(-m)^{\alpha}\subset I(\mathrm{triv},-m)^{\alpha} such that I(triv,m)α/F(m)αI(\mathrm{triv},-m)^{\alpha}/F(-m)^{\alpha} is irreducible and infinite-dimensional. The KK-type formulas of F(m)Kα=F(m)αF(-m)^{\alpha}_{K}=F(-m)^{\alpha} are given as follows.

    F(m)K+=0m/22(n)andF(m)K=0(m1)/22+1(n)F(-m)^{+}_{K}\simeq\bigoplus_{\ell=0}^{m/2}\mathcal{H}^{2\ell}(\mathbb{R}^{n})\quad\text{and}\quad F(-m)^{-}_{K}\simeq\bigoplus_{\ell=0}^{(m-1)/2}\mathcal{H}^{2\ell+1}(\mathbb{R}^{n})
  3. (3)

    For Case (B) with λ=n+m\lambda=n+m, there exists an infinite-dimensional irreducible subrepresentation T(n+m)αI(triv,n+m)αT(n+m)^{\alpha}\subset I(\mathrm{triv},n+m)^{\alpha} such that I(triv,n+m)α/T(n+m)αI(\mathrm{triv},n+m)^{\alpha}/T(n+m)^{\alpha} is irreducible and finite-dimensional. The KK-type formulas of T(n+m)KαT(n+m)^{\alpha}_{K} are given as follows.

    T(n+m)K+m+222(n)andT(n+m)Km+122+1(n)T(n+m)^{+}_{K}\simeq\bigoplus_{\ell\geq\frac{m+2}{2}}\mathcal{H}^{2\ell}(\mathbb{R}^{n})\quad\text{and}\quad T(n+m)^{-}_{K}\simeq\bigoplus_{\ell\geq\frac{m+1}{2}}\mathcal{H}^{2\ell+1}(\mathbb{R}^{n})
  4. (4)

    For m0m\in\mathbb{Z}_{\geq 0}, we have the non-split exact sequences of Fréchet GG-modules:

    {0}\displaystyle\{0\} F(m)α\displaystyle\longrightarrow F(-m)^{\alpha} I(triv,m)α\displaystyle\longrightarrow I(\mathrm{triv},-m)^{\alpha} T(n+m)α\displaystyle\longrightarrow T(n+m)^{\alpha} {0},\displaystyle\longrightarrow\{0\},
    {0}\displaystyle\{0\} T(n+m)α\displaystyle\longrightarrow T(n+m)^{\alpha} I(triv,n+m)α\displaystyle\longrightarrow I(\mathrm{triv},n+m)^{\alpha} F(m)α\displaystyle\longrightarrow F(-m)^{\alpha} {0}.\displaystyle\longrightarrow\{0\}.
Remark 6.3.

In [71, Thm. 1.1], the condition “μ<2n\mu<2-n” in b) should be read as “μ<1n\mu<1-n”. For μ=1n\mu=1-n, the induced representation π1n,ν±\pi^{\pm}_{1-n,\nu} in the cited paper is irreducible, as it is dual to the case of μ=1\mu=-1.

Remark 6.4.

The parabolic subgroup Pn1,1P_{n-1,1} corresponding to the partition n=(n1)+1n=(n-1)+1 is considered in [26] and [71], while we consider the one corresponding to n=1+(n1)n=1+(n-1) in this paper. Thus, the complex parameters “μ\mu” in [71], “α\alpha” in [26], and “λ\lambda” in this paper are related as μ=α=λ\mu=\alpha=-\lambda.

Now, for k0k\in\mathbb{Z}_{\geq 0}, we consider the KK-type formulas of the kernel Ker(𝒟k)Kα\mathrm{Ker}(\mathcal{D}_{k})^{\alpha}_{K} and image Im(Dk)Kα\mathrm{Im}(D_{k})^{\alpha}_{K} of the intertwining differential operator

𝒟k:I(triv,1k)αI(polyn1k,1+kn1)α+k.\mathcal{D}_{k}\colon I(\mathrm{triv},1-k)^{\alpha}\longrightarrow I(\mathrm{poly}^{k}_{n-1},1+\tfrac{k}{n-1})^{\alpha+k}.

If k=0k=0, then I(triv,1k)α=I(polyn1k,1+kn1)α=I(triv,1)αI(\mathrm{triv},1-k)^{\alpha}=I(\mathrm{poly}^{k}_{n-1},1+\tfrac{k}{n-1})^{\alpha}=I(\mathrm{triv},1)^{\alpha} and 𝒟0=id\mathcal{D}_{0}=\mathrm{id}. Thus, in this case, we have

Ker(𝒟0)α={0}andIm(𝒟0)α=I(triv,1)α.\mathrm{Ker}(\mathcal{D}_{0})^{\alpha}=\{0\}\quad\text{and}\quad\mathrm{Im}(\mathcal{D}_{0})^{\alpha}=I(\mathrm{triv},1)^{\alpha}.

Therefore, the KK-type formula Im(𝒟0)Kα\mathrm{Im}(\mathcal{D}_{0})^{\alpha}_{K} is given as in (6.1). For k1+0k\in 1+\mathbb{Z}_{\geq 0}, we have the following.

Theorem 6.5.

For α{±}{±1}\alpha\in\{\pm\}\equiv\{\pm 1\} and k1+0k\in 1+\mathbb{Z}_{\geq 0}, the kernel Ker(𝒟k)α\mathrm{Ker}(\mathcal{D}_{k})^{\alpha} is given as follows.

Ker(𝒟k)α={F(1k)αif α=(1)1k,{0}otherwise.\mathrm{Ker}(\mathcal{D}_{k})^{\alpha}=\begin{cases}F(1-k)^{\alpha}&\text{if $\alpha=(-1)^{1-k}$,}\\[3.0pt] \{0\}&\text{otherwise}.\end{cases}

In particular, the composition factors F(1k)αF(1-k)^{\alpha} and T(n+k1)αI(triv,1k)α/F(1k)αT(n+k-1)^{\alpha}\simeq I(\mathrm{triv},1-k)^{\alpha}/F(1-k)^{\alpha} of I(triv,1k)αI(\mathrm{triv},1-k)^{\alpha} can be realized as

F(1k)α=Ker(𝒟k)αandT(n+k1)αIm(𝒟k)α.F(1-k)^{\alpha}=\mathrm{Ker}(\mathcal{D}_{k})^{\alpha}\quad\text{and}\quad T(n+k-1)^{\alpha}\simeq\mathrm{Im}(\mathcal{D}_{k})^{\alpha}.
Proof.

Since the second assertion follows from the first and Theorem 6.2, it suffices to consider the first statement. As 𝒟k0\mathcal{D}_{k}\not\equiv 0 for k1+0k\in 1+\mathbb{Z}_{\geq 0}, by Theorem 6.2, there are only two possibilities on Ker(𝒟k)α\mathrm{Ker}(\mathcal{D}_{k})^{\alpha}, namely, Ker(𝒟k)α={0}\mathrm{Ker}(\mathcal{D}_{k})^{\alpha}=\{0\} or F(1k)αF(1-k)^{\alpha}. It then suffices to show that Ker(𝒟k)α{0}\mathrm{Ker}(\mathcal{D}_{k})^{\alpha}\neq\{0\} as far as I(triv,1k)αI(\mathrm{triv},1-k)^{\alpha} is reducible. In what follows, we understand that I(triv,1k)αI(\mathrm{triv},1-k)^{\alpha} is a subspace of C(n1)C^{\infty}(\mathbb{R}^{n-1}) via the isomorphism (4.3) (see (2.1)). In particular, the Lie algebra 𝔤{\mathfrak{g}} acts on I(triv,1k)αI(\mathrm{triv},1-k)^{\alpha} via the representation dπ1kdπ(αtriv,χ1k)d\pi_{1-k}\equiv d\pi_{(\alpha\otimes\mathrm{triv},\chi^{1-k})} (see (2.4)).

Suppose that I(triv,1k)αI(\mathrm{triv},1-k)^{\alpha} is reducible. Since F(1k)αF(1-k)^{\alpha} is irreducible and finite-dimensional, there exists a lowest weight vector f0(x1,,xn1)F(1k)αf_{0}(x_{1},\ldots,x_{n-1})\in F(1-k)^{\alpha}. We shall show that 𝒟kf0=0\mathcal{D}_{k}f_{0}=0.

As f0f_{0} being a lowest weight vector of F(1k)αI(triv,1k)αF(1-k)^{\alpha}\subset I(\mathrm{triv},1-k)^{\alpha}, we have

dπ1k(Nj)f0(x1,,xn1)=0for all j{1,,n1}.d\pi_{1-k}(N_{j}^{-})f_{0}(x_{1},\ldots,x_{n-1})=0\quad\text{for all $j\in\{1,\ldots,n-1\}$}. (6.6)

A direct computation shows that dπ1k(Nj)=xjd\pi_{1-k}(N_{j}^{-})=-\frac{\partial}{\partial x_{j}}. Thus, (6.6) is equivalent to

xjf0(x1,,xn1)=0for all j{1,,n1},\frac{\partial}{\partial x_{j}}f_{0}(x_{1},\ldots,x_{n-1})=0\quad\text{for all $j\in\{1,\ldots,n-1\}$},

which shows that f0f_{0} is a constant function. Therefore, we have

𝒟kf0(x1,,xn1)=𝐤Ξkkx𝐤f0(x1,,xn1)y~𝐤=0.\mathcal{D}_{k}f_{0}(x_{1},\ldots,x_{n-1})=\sum_{\mathbf{k}\in\Xi_{k}}\frac{\partial^{k}}{\partial x^{\mathbf{k}}}f_{0}(x_{1},\ldots,x_{n-1})\otimes\widetilde{y}_{\mathbf{k}}=0.\qed

Corollary 6.7 below is an immediate consequence of Theorems 6.2 and 6.5.

Corollary 6.7.

For (α,k){±}×(1+0)(\alpha,k)\in\{\pm\}\times(1+\mathbb{Z}_{\geq 0}), the KK-type formulas of Ker(𝒟k)Kα\mathrm{Ker}(\mathcal{D}_{k})^{\alpha}_{K} and Im(𝒟k)Kα\mathrm{Im}(\mathcal{D}_{k})^{\alpha}_{K} are given as follows.

  1. (1)

    α=+:\alpha=+\colon

    1. (1-a)

      k1+20:k\in 1+2\mathbb{Z}_{\geq 0}\colon

      Ker(𝒟k)K+\displaystyle\mathrm{Ker}(\mathcal{D}_{k})^{+}_{K} =F(1k)K+\displaystyle=F(1-k)^{+}_{K} =0(k1)/22(n),\displaystyle\simeq\bigoplus_{\ell=0}^{(k-1)/2}\mathcal{H}^{2\ell}(\mathbb{R}^{n}),
      Im(𝒟k)K+\displaystyle\mathrm{Im}(\mathcal{D}_{k})^{+}_{K} T(n+k1)K+\displaystyle\simeq T(n+k-1)^{+}_{K} k+122(n).\displaystyle\simeq\bigoplus_{\ell\geq\frac{k+1}{2}}\mathcal{H}^{2\ell}(\mathbb{R}^{n}).
    2. (1-b)

      k2(1+0):k\in 2(1+\mathbb{Z}_{\geq 0})\colon

      Ker(𝒟k)K+\displaystyle\mathrm{Ker}(\mathcal{D}_{k})^{+}_{K} ={0},\displaystyle=\{0\},
      Im(𝒟k)K+\displaystyle\mathrm{Im}(\mathcal{D}_{k})^{+}_{K} I(triv,1k)K+\displaystyle\simeq I(\mathrm{triv},1-k)^{+}_{K} 02(n).\displaystyle\simeq\bigoplus_{\ell\in\mathbb{Z}_{\geq 0}}\mathcal{H}^{2\ell}(\mathbb{R}^{n}).
  2. (2)

    α=:\alpha=-\colon

    1. (2-a)

      k1+20:k\in 1+2\mathbb{Z}_{\geq 0}\colon

      Ker(𝒟k)K\displaystyle\mathrm{Ker}(\mathcal{D}_{k})^{-}_{K} ={0},\displaystyle=\{0\},
      Im(𝒟k)K\displaystyle\mathrm{Im}(\mathcal{D}_{k})^{-}_{K} I(triv,1k)K\displaystyle\simeq I(\mathrm{triv},1-k)^{-}_{K} 02+1(n).\displaystyle\simeq\bigoplus_{\ell\in\mathbb{Z}_{\geq 0}}\mathcal{H}^{2\ell+1}(\mathbb{R}^{n}).
    2. (2-b)

      k2(1+0):k\in 2(1+\mathbb{Z}_{\geq 0})\colon

      Ker(𝒟k)K\displaystyle\mathrm{Ker}(\mathcal{D}_{k})^{-}_{K} =F(1k)K\displaystyle=F(1-k)^{-}_{K} =0(k2)/22+1(n),\displaystyle\simeq\bigoplus_{\ell=0}^{(k-2)/2}\mathcal{H}^{2\ell+1}(\mathbb{R}^{n}),
      Im(𝒟k)K\displaystyle\mathrm{Im}(\mathcal{D}_{k})^{-}_{K} T(n+k1)K\displaystyle\simeq T(n+k-1)^{-}_{K} k22+1(n).\displaystyle\simeq\bigoplus_{\ell\geq\frac{k}{2}}\mathcal{H}^{2\ell+1}(\mathbb{R}^{n}).

7. Appendix: the standardness of the homomorphism φk\varphi_{k}

The aim of this appendix is to show that the homomorphisms φk\varphi_{k} 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 𝔤{\mathfrak{g}} be a complex simple Lie algebra. Fix a Cartan subalgebra 𝔥{\mathfrak{h}} and write ΔΔ(𝔤,𝔥)\Delta\equiv\Delta({\mathfrak{g}},{\mathfrak{h}}) for the set of roots of 𝔤{\mathfrak{g}} with respect to 𝔥{\mathfrak{h}}. Choose a positive system Δ+\Delta^{+} and denote by Π\Pi the set of simple roots of Δ\Delta. Let 𝔟{\mathfrak{b}} denote the Borel subalgebra of 𝔤{\mathfrak{g}} associated with Δ+\Delta^{+}, namely, 𝔟=𝔥αΔ+𝔤α{\mathfrak{b}}={\mathfrak{h}}\oplus\bigoplus_{\alpha\in\Delta^{+}}{\mathfrak{g}}_{\alpha}, where 𝔤α{\mathfrak{g}}_{\alpha} is the root space for αΔ+\alpha\in\Delta^{+}.

Let ,\langle\cdot,\cdot\rangle denote the inner product on 𝔥{\mathfrak{h}}^{*} induced from a non-degenerate symmetric bilinear form of 𝔤{\mathfrak{g}}. For αΔ\alpha\in\Delta, we write α=2α/α,α\alpha^{\vee}=2\alpha/\langle\alpha,\alpha\rangle. Also, write sαs_{\alpha} for the reflection with respect to αΔ\alpha\in\Delta. As usual, we let ρ=(1/2)αΔ+α\rho=(1/2)\sum_{\alpha\in\Delta^{+}}\alpha be half the sum of the positive roots.

Let 𝔭𝔟{\mathfrak{p}}\supset{\mathfrak{b}} be a standard parabolic subalgebra of 𝔤{\mathfrak{g}}. Write 𝔭=𝔩𝔫+{\mathfrak{p}}={\mathfrak{l}}\oplus{\mathfrak{n}}_{+} for the Levi decomposition of 𝔭{\mathfrak{p}}. We let Π(𝔩)={αΠ:𝔤α𝔩}\Pi({\mathfrak{l}})=\{\alpha\in\Pi:{\mathfrak{g}}_{\alpha}\subset{\mathfrak{l}}\}.

Now we put

𝐏𝔩+:={μ𝔥:μ,α1+0for all αΠ(𝔩)}.\mathbf{P}^{+}_{{\mathfrak{l}}}:=\{\mu\in{\mathfrak{h}}^{*}:\langle\mu,\alpha^{\vee}\rangle\in 1+\mathbb{Z}_{\geq 0}\;\;\text{for all $\alpha\in\Pi({\mathfrak{l}})$}\}.

For μ𝐏𝔩+\mu\in\mathbf{P}^{+}_{\mathfrak{l}}, let E(μρ)E(\mu-\rho) be the finite-dimensional simple 𝒰(𝔩)\mathcal{U}({\mathfrak{l}})-module with highest weight μρ\mu-\rho. By letting 𝔫+{\mathfrak{n}}_{+} act trivially, we regard E(μρ)E(\mu-\rho) as a 𝒰(𝔭)\mathcal{U}({\mathfrak{p}})-module. Then the induced module

N𝔭(μ):=𝒰(𝔤)𝒰(𝔭)E(μρ)N_{\mathfrak{p}}(\mu):=\mathcal{U}({\mathfrak{g}})\otimes_{\mathcal{U}({\mathfrak{p}})}E(\mu-\rho)

is the generalized Verma module with highest weight μρ\mu-\rho. If 𝔭=𝔟{\mathfrak{p}}={\mathfrak{b}}, then N(μ)N𝔟(μ)N(\mu)\equiv N_{{\mathfrak{b}}}(\mu) is the (ordinary) Verma module with highest weight μρ\mu-\rho.

Let μ,η𝐏𝔩+\mu,\eta\in\mathbf{P}^{+}_{\mathfrak{l}}. It follows from a theorem by BGG-Verma (cf. [13, Thm. 7.6.23] and [28, Thm. 5.1]) that if Hom𝔤(N𝔭(μ),N𝔭(η)){0}\operatorname{Hom}_{{\mathfrak{g}}}(N_{\mathfrak{p}}(\mu),N_{\mathfrak{p}}(\eta))\neq\{0\}, then Hom𝔤(N(μ),N(η)){0}\operatorname{Hom}_{{\mathfrak{g}}}(N(\mu),N(\eta))\neq\{0\}.

Conversely, suppose that there exists a non-zero 𝔤{\mathfrak{g}}-homomorphism φ:N(μ)N(η)\varphi\colon N(\mu)\to N(\eta). Let prμ:N(μ)N𝔭(μ)\mathrm{pr}_{\mu}\colon N(\mu)\to N_{\mathfrak{p}}(\mu) denote the canonical projection map. Then we have φ(Ker(prμ))Ker(prη)\varphi(\mathrm{Ker}(\mathrm{pr}_{\mu}))\subset\mathrm{Ker}(\mathrm{pr}_{\eta}) ([60, Prop. 3.1]). Thus, the map φ\varphi induces a 𝔤{\mathfrak{g}}-homomorphism φstd:N𝔭(μ)N𝔭(η)\varphi_{\mathrm{std}}\colon N_{\mathfrak{p}}(\mu)\to N_{\mathfrak{p}}(\eta) such that the following diagram commutes.

N(μ)\textstyle{N(\mu)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ\scriptstyle{\varphi}prμ\scriptstyle{\mathrm{pr}_{\mu}}N(η)\textstyle{N(\eta\ignorespaces\ignorespaces\ignorespaces\ignorespaces)}prη\scriptstyle{\mathrm{pr}_{\eta}}\textstyle{\circlearrowleft}N𝔭(μ)\textstyle{N_{\mathfrak{p}}(\mu)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φstd\scriptstyle{\varphi_{\mathrm{std}}}N𝔭(η)\textstyle{N_{\mathfrak{p}}(\eta)}

The map φstd\varphi_{\mathrm{std}} is called the standard map from N𝔭(μ)N_{\mathfrak{p}}(\mu) to N𝔭(η)N_{\mathfrak{p}}(\eta) ([60]). As dimHom𝔤(N(μ),N(η))1\dim\operatorname{Hom}_{{\mathfrak{g}}}(N(\mu),N(\eta))\leq 1, the standard map φstd\varphi_{\mathrm{std}} is unique up to scalar. It is known that the standard map φstd\varphi_{\mathrm{std}} could be zero, and even if φstd=0\varphi_{\mathrm{std}}=0, there could be another non-zero map from N𝔭(μ)N_{\mathfrak{p}}(\mu) to N𝔭(η)N_{\mathfrak{p}}(\eta). Any homomorphism that is not standard is called a non-standard map.

It is known when the standard map φstd\varphi_{\mathrm{std}} is zero. To state the criterion, we first introduce the notion of a link between two weights.

Definition 7.1 (Bernstein–Gelfand–Gelfand).

Let μ,η𝔥\mu,\eta\in{\mathfrak{h}}^{*} and β1,,βtΔ+\beta_{1},\ldots,\beta_{t}\in\Delta^{+}. Set η0:=η\eta_{0}:=\eta and ηi:=sβisβ1η\eta_{i}:=s_{\beta_{i}}\cdots s_{\beta_{1}}\eta for 1it1\leq i\leq t. We say that the sequence (β1,,βt)(\beta_{1},\ldots,\beta_{t}) links η\eta to μ\mu if it satisfies the following two conditions.

  1. (1)

    ηt=μ\eta_{t}=\mu.

  2. (2)

    ηi1,βi0\langle\eta_{i-1},\beta^{\vee}_{i}\rangle\in\mathbb{Z}_{\geq 0} for all i{1,,t}i\in\{1,\ldots,t\}.

The criterion on the vanishing of the standard map φstd\varphi_{\mathrm{std}} 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 μ,η𝔥\mu,\eta\in{\mathfrak{h}}^{*} and suppose that Hom𝔤(N(μ),N(η)){0}\operatorname{Hom}_{{\mathfrak{g}}}(N(\mu),N(\eta))\neq\{0\}. Then the following two conditions on (μ,η)(\mu,\eta) are equivalent.

  1. (i)

    The standard map φstd:N𝔭(μ)N𝔭(η)\varphi_{\mathrm{std}}\colon N_{\mathfrak{p}}(\mu)\to N_{\mathfrak{p}}(\eta) is non-zero.

  2. (ii)

    For all sequences (β1,,βt)(\beta_{1},\ldots,\beta_{t}) linking η\eta to μ\mu, we have η1𝐏𝔩+\eta_{1}\in\mathbf{P}^{+}_{\mathfrak{l}}.

7.2. The standardness of the homomorphism φk\varphi_{k}

Now we specialize the situation to the one considered in Section 3, that is, 𝔤=𝔰𝔩(n,){\mathfrak{g}}=\mathfrak{sl}(n,\mathbb{C}) and 𝔭{\mathfrak{p}} is the maximal parabolic subalgebra corresponding to the partition n=1+(n1)n=1+(n-1). Observe that if 𝔤=𝔰𝔩(2,){\mathfrak{g}}=\mathfrak{sl}(2,\mathbb{C}), then N𝔭(μ)=N𝔟(μ)N_{\mathfrak{p}}(\mu)=N_{\mathfrak{b}}(\mu). Thus, we assume that n3n\geq 3.

As usual, a Cartan subalgebra 𝔥𝔤{\mathfrak{h}}\subset{\mathfrak{g}} is taken to be

𝔥={diag(a1,,an):i=1nai=0}.{\mathfrak{h}}=\{\mathrm{diag}(a_{1},\ldots,a_{n}):\sum_{i=1}^{n}a_{i}=0\}.

We take a set of positive roots Δ+\Delta^{+} and simple roots Π\Pi as

Δ+={εiεj:1i<jn}\Delta^{+}=\{\varepsilon_{i}-\varepsilon_{j}:1\leq i<j\leq n\}

and

Π={εiεi+1:1in1}.\Pi=\{\varepsilon_{i}-\varepsilon_{i+1}:1\leq i\leq n-1\}.

We realize 𝔥{\mathfrak{h}}^{*} as a subspace of n\mathbb{R}^{n} and write elements in 𝔥{\mathfrak{h}}^{*} in coordinates. For instance, we write ε1ε2=(1,1,0,,0)\varepsilon_{1}-\varepsilon_{2}=(1,-1,0,\ldots,0).

Recall from Theorem 5.23 that we have

Hom𝔤(M𝔭(symn1k,(1+kn1)),M𝔭(triv,k1))=φk\operatorname{Hom}_{{\mathfrak{g}}}(M_{\mathfrak{p}}(\mathrm{sym}^{k}_{n-1},-(1+\tfrac{k}{n-1})),M_{\mathfrak{p}}(\mathrm{triv},k-1))=\mathbb{C}\varphi_{k} (7.3)

for k0k\in\mathbb{Z}_{\geq 0}, where φk\varphi_{k} is given in (4.7). Let dχ𝔥d\chi\in{\mathfrak{h}}^{*} be the differential of the character χ\chi of AA defined in (3.3). Then

dχ=1n(n1,1,,1).d\chi=\frac{1}{n}(n-1,-1,\ldots,-1).

Define ω1𝔥\omega_{1}\in{\mathfrak{h}}^{*} by

ω1=1n1(0,n2,1,,1).\omega_{1}=\frac{1}{n-1}(0,n-2,-1,\ldots,-1).

For ρ=(1/2)αΔ+α𝔥\rho=(1/2)\sum_{\alpha\in\Delta^{+}}\alpha\in{\mathfrak{h}}^{*}, we have

ρ=12(n1,n3,,(n3),(n1)).\rho=\frac{1}{2}(n-1,n-3,\ldots,-(n-3),-(n-1)).

For simplicity we let

μk\displaystyle\mu^{k} =kω1(1+kn1)dχ+ρ,\displaystyle=k\omega_{1}-(1+\tfrac{k}{n-1})d\chi+\rho,
ηk\displaystyle\eta^{k} =(k1)dχ+ρ.\displaystyle=(k-1)d\chi+\rho.

We have

N𝔭(μk)\displaystyle N_{\mathfrak{p}}(\mu^{k}) =M𝔭(symn1k,(1+kn1)),\displaystyle=M_{\mathfrak{p}}(\mathrm{sym}^{k}_{n-1},-(1+\tfrac{k}{n-1})),
N𝔭(ηk)\displaystyle N_{\mathfrak{p}}(\eta^{k}) =M𝔭(triv,k1).\displaystyle=M_{\mathfrak{p}}(\mathrm{triv},k-1).

Equation (7.3) is then given as

Hom𝔤(N𝔭(μk),N𝔭(ηk))=φk.\operatorname{Hom}_{\mathfrak{g}}\left(N_{\mathfrak{p}}(\mu^{k}),N_{\mathfrak{p}}(\eta^{k})\right)=\mathbb{C}\varphi_{k}.
Theorem 7.4.

Let n3n\geq 3. Then the homomorphism φk\varphi_{k} is standard for all k0k\in\mathbb{Z}_{\geq 0}.

Proof.

Since dimHom𝔤(N𝔭(μk),N𝔭(ηk))=1\dim\operatorname{Hom}_{\mathfrak{g}}(N_{\mathfrak{p}}(\mu^{k}),N_{\mathfrak{p}}(\eta^{k}))=1, it suffices to show the standard map φstdk:N𝔭(μk)N𝔭(ηk)\varphi_{\mathrm{std}}^{k}\colon N_{\mathfrak{p}}(\mu^{k})\to N_{\mathfrak{p}}(\eta^{k}) is non-zero. We remark that Hom𝔤(N(μk),N(ηk)){0}\operatorname{Hom}_{\mathfrak{g}}\left(N(\mu^{k}),N(\eta^{k})\right)\neq\{0\} as Hom𝔤(N𝔭(μk),N𝔭(ηk)){0}\operatorname{Hom}_{\mathfrak{g}}\left(N_{\mathfrak{p}}(\mu^{k}),N_{\mathfrak{p}}(\eta^{k})\right)\neq\{0\}.

First, suppose that k=0k=0. In this case we have μ0=η0=dχ+ρ\mu_{0}=\eta_{0}=-d\chi+\rho. Thus, the standard map φstd0\varphi_{\mathrm{std}}^{0} is φstd0=id\varphi_{\mathrm{std}}^{0}=\mathrm{id}; in particular, φstd0\varphi_{\mathrm{std}}^{0} is non-zero.

Next, suppose that k1+0k\in 1+\mathbb{Z}_{\geq 0}. In this case we have

μk\displaystyle\mu^{k} =12n(2(1k)+n(n3),(n1)(n2+2k),2(1k)+n(n5),,2(1k)n(n1)),\displaystyle=\frac{1}{2n}(2(1-k)+n(n-3),(n-1)(n-2+2k),2(1-k)+n(n-5),\ldots,2(1-k)-n(n-1)),
ηk\displaystyle\eta^{k} =12n((n1)(n2+2k),2(1k)+n(n3),2(1k)+n(n5),,2(1k)n(n1)).\displaystyle=\frac{1}{2n}((n-1)(n-2+2k),2(1-k)+n(n-3),2(1-k)+n(n-5),\ldots,2(1-k)-n(n-1)).

Then ε1ε2\varepsilon_{1}-\varepsilon_{2} is the only element that links ηk\eta^{k} to μk\mu^{k}. Since η1k=sε1ε2ηk=μk𝐏𝔩+\eta_{1}^{k}=s_{\varepsilon_{1}-\varepsilon_{2}}\eta^{k}=\mu^{k}\in\mathbf{P}^{+}_{{\mathfrak{l}}}, Theorem 7.2 shows that the standard map φstdk\varphi_{\mathrm{std}}^{k} 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 𝔭𝔤𝔩(a+1|b)𝔳𝔢𝔠𝔱(a|b)\mathfrak{pgl}(a+1|b)\subset\mathfrak{vect}(a|b), 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 GLn()GL_{n}(\mathbb{C}) and GLn()GL_{n}(\mathbb{\mathbb{R}}), 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 𝒪\mathcal{O}, 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, QQ-Curvature and Holography, Progr. Math., vol. 275, Birkhäuser, Basel, 2009.
  • [31] A.C. Kable, KK-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 G2G_{2} and their K-finite solutions, J. Lie Theory 22 (2012), no. 1, 93–136.
  • [33] by same author, The Heisenberg ultrahyperbolic equation: KK-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 O(p,q)O(p,q), 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 O(p,q)O(p,q) 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: 𝖠n\mathsf{A}_{n}-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 KK-type formulas for the Heisenberg ultrahyperbolic operator s\square_{s} for SL~(3,)\widetilde{SL}(3,\mathbb{R}) and tridiagonal determinants for local Heun functions, to appear in Festschrift in honour of Toshiyuki Kobayashi.
  • [58] by same author, On the space of KK-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 (LieG2,so(7))(\mathrm{Lie}G_{2},so(7)), Arch. Math. (Brno) 49 (2013), 317–332.
  • [62] J. Möllers and B. Schwarz, Structure of the degenerate principal series on symmetric RR-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 S3S^{3}, 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 (𝔤,𝔨)(\mathfrak{g},\mathfrak{k})-modules associated to the ideals of minimal nilpotent orbits for simple Lie groups of type AA, 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 SL(n,)SL(n,\mathbb{R}), 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.