Home Smooth representations of unit groups of split basic algebras over non-Archimedean local fields
Article Publicly Available

Smooth representations of unit groups of split basic algebras over non-Archimedean local fields

  • Carlos A. M. André EMAIL logo and João Dias
Published/Copyright: May 26, 2021

Abstract

We consider smooth representations of the unit group G = A × of a finite-dimensional split basic algebra 𝒜 over a non-Archimedean local field. In particular, we prove a version of Gutkin’s conjecture, namely, we prove that every irreducible smooth representation of 𝐺 is compactly induced by a one-dimensional representation of the unit group of some subalgebra of 𝒜. We also discuss admissibility and unitarisability of smooth representations of 𝐺.

1 Introduction

Let 𝕜 be a non-Archimedean local field (such as a finite extension of the 𝑝-adic field Q p , or a field F q ( ( t ) ) of formal Laurent series in one variable over a finite field F q ); we equip 𝕜 with the natural topology making it a locally compact totally disconnected topological field. We fix a non-trivial unitary character ϑ : 𝕜 + C × of the additive group of 𝕜, and for each a 𝕜 , we define ϑ a : 𝕜 + C × by

ϑ a ( b ) = ϑ ( a b ) , b 𝕜 ;

then the mapping a ϑ a defines a topological isomorphism between 𝕜 + and its Pontryagin dual (see, for example, [4, Proposition 1.7]).

Let 𝒜 be a finite-dimensional associative 𝕜-algebra (with identity), and let J ( A ) denote the Jacobson radical of 𝒜. We say that 𝒜 is a split basic 𝕜-algebra if J ( A ) is equal to the set of all nilpotent elements of 𝒜 (in [15], B. Szegedy refers to 𝒜 as an N-algebra over 𝕜), and the semisimple 𝕜-algebra A / J ( A ) is isomorphic to a (finite) direct sum of isomorphic copies of the base field 𝕜 (in the terminology of [15], 𝒜 is referred to as a DN-algebra over 𝕜). It follows from Wedderburn’s theorem (and from the usual process of “lifting idempotents”) that a finite-dimensional 𝕜-algebra 𝒜 is a split basic 𝕜-algebra if and only if there are non-zero orthogonal idempotents e 1 , , e n A such that

A = ( 𝕜 e 1 𝕜 e n ) J ( A ) ;

we refer to D = 𝕜 e 1 𝕜 e n as a diagonal subalgebra of 𝒜. The following easy observation is crucial for inductive arguments; a proof in the case where 𝕜 is a finite field can be found in [15, Lemmas 2.2 and 2.3].

Lemma 1

Let 𝒜 be a finite-dimensional split basic 𝕜-algebra. Then every subalgebra of 𝒜 is also a split basic 𝕜-algebra.

Proof

Let ℬ be a subalgebra of 𝒜, and let J ( B ) denote the Jacobson radical of ℬ. As J ( B ) = B J ( A ) , it is clear that ℬ is a basic 𝕜-algebra; moreover, the 𝕜-algebra B / J ( B ) is naturally isomorphic to the subalgebra ( B + J ( A ) ) / J ( A ) of the semisimple 𝕜-algebra A / J ( A ) . Therefore, without loss of generality, we may assume that 𝒜 is a split basic semisimple 𝕜-algebra. Since ℬ is a basic semisimple 𝕜-algebra, there are non-zero orthogonal idempotents e 1 , , e m B such that B = 𝕜 1 e 1 𝕜 n e m , where 𝕜 1 , , 𝕜 n are finite field extensions of 𝕜. On the other hand, let e 1 , , e n A be non-zero orthogonal idempotents such that A = 𝕜 e 1 𝕜 e n . It is straightforward to check that there exists a subset partition I 1 , , I m of { 1 , , n } such that e j = i I j e i , and thus

e j e i = e i , i I j ,  1 j m .

It follows that 𝕜 j e j e i = 𝕜 j e i A e i = 𝕜 e i for all i I j and all 1 j m , which clearly implies that 𝕜 j = 𝕜 for all 1 j m . ∎

In the following result, we list some elementary properties which will be used repeatedly throughout the paper; for a detailed proof (which does not depend on the finiteness of the field 𝕜), we refer to [10, Lemma 2.1].

Lemma 2

Let 𝒜 be a finite-dimensional split basic 𝕜-algebra, let 𝒟 be a diagonal subalgebra of 𝒜, and let e 1 , , e n A be non-zero orthogonal idempotents such that D = 𝕜 e 1 𝕜 e n . The following properties hold.

  1. If 𝒱 is an (arbitrary) 𝒟-bimodule, then 𝒱 decomposes as a direct sum of the (non-zero) homogeneous sub-bimodules e i V e j for 1 i , j n .

  2. For every sub-bimodule V 1 of a 𝒟-bimodule 𝒱, there exists a sub-bimodule V 2 of 𝒱 such that V = V 1 V 2 .

  3. Every 𝒟-bimodule decomposes as a direct sum of one-dimensional sub-bimodules.

  4. If 𝒱 is a one-dimensional 𝒟-bimodule and v V is such that V = 𝕜 v , then there exist uniquely determined 1 i , j n such that v = e i v e j .

Let G = A × denote the unit group of a split basic 𝕜-algebra 𝒜. For any (nilpotent) subalgebra 𝒰 of J ( A ) , the set 1 + U is a subgroup of 𝐺 to which we refer as an algebra subgroup of 𝐺 (as defined in [11]); similarly, if I J ( A ) is an ideal of 𝒜, we refer to 1 + I as an ideal subgroup of 𝐺. In the particular case where I = J ( A ) , it is clear that P = 1 + J ( A ) is a normal subgroup of 𝐺; furthermore, 𝐺 is the semidirect product G = T P , where T G is isomorphic to the unit group of A / J ( A ) . Since 𝒜 is a split basic 𝕜-algebra, 𝑇 is isomorphic to a direct product 𝕜 × × × 𝕜 × of n = dim A / J ( A ) copies of the multiplicative group 𝕜 × of 𝕜. In fact, T = D × is the unit group of a diagonal subalgebra 𝒟 of 𝒜; we will refer to 𝑇 as a diagonal subgroup of 𝐺.

Example 1

Let M n ( 𝕜 ) denote the 𝕜-algebra consisting of all n × n matrices with entries in 𝕜, and let A = B n ( 𝕜 ) denote the Borel subalgebra of M n ( 𝕜 ) consisting of all upper-triangular matrices; hence G = A × is the standard Borel subgroup B n ( 𝕜 ) of the general linear group GL n ( 𝕜 ) (consisting of all invertible matrices in M n ( 𝕜 ) ). In this case, T G is the standard torus consisting of all diagonal matrices, and P = 1 + J ( A ) is the standard unitriangular group; we note that J ( A ) is the nilpotent ideal of 𝒜 consisting of all upper-triangular matrices with zeroes on the main diagonal. In this case, for each 1 i n , the idempotent e i A can be chosen to be the elementary matrices e i = e i , i having a unique non-zero entry (equal to 1) in the position ( i , i ) , and thus D = 𝕜 e 1 𝕜 e n is the subalgebra of 𝒜 consisting of all diagonal matrices.

The topology of 𝕜 naturally induces a topology in G = A × with respect to which 𝐺 becomes a locally compact totally disconnected topological group (that is, a topological group such that every open neighbourhood of the identity contains a compact open subgroup); for simplicity, we follow the terminology of [3] and refer to such a group as an ℓ-group. (For the definition and main properties of ℓ-groups, we refer mainly to [4, Chapter I].) We note that 𝐺 is second countable. On the other hand, any algebra subgroup 𝑄 of 𝐺 is an c -group (see [3]) which means that 𝑄 is the filtered union of its compact open subgroups; in other words, every element of 𝑄 is contained in a compact open subgroup, and any two such subgroups are contained in a third such subgroup.

We recall that a (complex) representation of 𝐺 is a pair ( π , V ) consisting of a complex vector space 𝑉 (not necessarily of finite dimension) and group homomorphism π : G GL ( V ) , where GL ( V ) denotes the group consisting of all linear automorphisms of 𝑉; for simplicity, we will refer to 𝑉 as a (left) 𝐺-module with (linear) 𝐺-action defined by g v = π ( g ) v for all g G and all v V . A 𝐺-module 𝑉 is said to be smooth if, for every v V , the centraliser

G v = { g G : g v = v }

is an open subgroup of 𝐺; in the particular case where the vector space 𝑉 is one-dimensional, we naturally obtain a group homomorphism ϑ : G C × with open kernel. We will refer to such a homomorphism as a smooth character of 𝐺, and denote by G the set of all smooth characters of 𝐺. A smooth character ϑ G can be naturally viewed as a smooth representation of 𝐺; indeed, an arbitrary one-dimensional 𝐺-module is smooth if and only if it affords a smooth character of 𝐺. Throughout the paper, for every smooth character ϑ G , we will denote by C ϑ the one-dimensional smooth 𝐺-module whose underlying vector space is ℂ and where the (linear) 𝐺-action is given by g α = ϑ ( g ) α for all g G and all α C .

It is well known that G is a group with respect to the usual multiplication of characters; it should not be confused with the Pontryagin dual of 𝐺 which consists of all unitary characters ϑ : G C × of 𝐺. By definition, a unitary character of 𝐺 is a continuous group homomorphism ϑ : G C × whose image ϑ ( G ) lies inside the unit circle S 1 in ℂ. We observe that every unitary character of 𝐺 is a smooth character, but the converse is not necessarily true; however, for an arbitrary c -group 𝑄, it is well known that Q is equal to the Pontryagin dual of 𝑄 (see [4, Proposition 1.6]; see also [3, Lemma 4.9]).

A morphism between smooth 𝐺-modules is defined as a morphism of abstract 𝐺-modules; we refer to such a morphism as a homomorphism of 𝐺-modules (or simply a 𝐺-homomorphism) and denote by Hom G ( V , V ) the (complex) vector space consisting of all homomorphisms between 𝐺-modules 𝑉 to V . We say that two smooth representations ( π , V ) and ( π , V ) of 𝐺 are equivalent if there is an isomorphism φ Hom G ( V , V ) (that is, a linear isomorphism which commutes with the 𝐺-actions); if this is the case, then the smooth 𝐺-modules 𝑉 and V are said to be isomorphic, and we write V V . A smooth 𝐺-module 𝑉 is said to be irreducible if V { 0 } and { 0 } and 𝑉 are the only 𝐺-invariant subspaces of 𝑉. Since 𝐺 is an ℓ-group (hence 𝐺 is second countable), there is a natural linear isomorphism Hom G ( V , V ) C for every irreducible smooth 𝐺-module 𝑉; the proof of this version of Schur’s lemma is due to H. Jacquet [12] (it can be found in [4, p. 21]; see also [5]). As a consequence, we deduce that, if 𝐺 is abelian, then every irreducible smooth 𝐺-module is one-dimensional (and hence affords a smooth character of 𝐺).

If 𝐻 is a subgroup of 𝐺 and 𝑉 is a 𝐺-module, we denote by V H the vector subspace of 𝑉 consisting of all v V such that h v = v for all h H ; then 𝑉 is a smooth 𝐺-module if and only if V = K V K , where 𝐾 runs over all compact open subgroups of 𝐺. A smooth 𝐺-module 𝑉 is said to be admissible if, for every compact open subgroup 𝐾 of 𝐺, the subspace V K is finite-dimensional; on the other hand, 𝑉 is said to be unitarisable if 𝑉 has a positive definite Hermitian inner product invariant under the action of 𝐺. We recall the (usual notion) of unitary representations (which are not necessarily smooth) of topological groups. By a unitary representation of a topological group 𝐺, we mean a pair ( π , H ) where ℋ is an Hilbert space over ℂ and π : G U ( H ) is a continuous group homomorphism from 𝐺 to the group of unitary linear automorphisms of ℋ equipped with the strong operator topology; in this case, the representation ( π , H ) is said to be irreducible if H { 0 } and { 0 } and ℋ are the only π ( G ) -invariant closed subspaces of ℋ.

A major difference between ℓ-groups and c -groups (hence, in particular, between G = A × and its algebra subgroups) concerns the admissibility and unitarisability of smooth modules. Indeed, [3, Theorem 1.3] asserts that, if 𝑃 is an algebra group over a non-Archimedean local field 𝕜, then every smooth irreducible 𝑃-module is admissible and unitarisable; in general, by an algebra group over an arbitrary field 𝕜, we mean a group of the form 1 + J , where 𝒥 is a finite-dimensional nilpotent 𝕜-algebra – with product formally defined by

( 1 + a ) ( 1 + b ) = 1 + a + b + a b , a , b J .

In particular, the unitarisability of a smooth character ϑ P is equivalent to the statement that the image ϑ ( P ) lies inside the unit circle S 1 in ℂ (by [4, Proposition 1.6]). However, the absolute value | | of a non-Archimedean local field 𝕜 defines a smooth homomorphism | | : 𝕜 × C × whose image does not lie inside S 1 .

The main goal of this paper is to study smooth modules for the unit group G = A × of an arbitrary finite-dimensional split basic 𝕜-algebra 𝒜; in particular, we aim to establish that every irreducible smooth 𝐺-module may be obtained by induction (with compact supports) from a one-dimensional smooth module for the unit group H = B × of some subalgebra ℬ of 𝒜.

Let 𝐺 be an arbitrary ℓ-group, and let 𝐻 be a closed subgroup of 𝐺; hence 𝐻 is also an ℓ-group. Let 𝑊 be an arbitrary smooth 𝐻-module, and define Ind H G ( W ) to be the (complex) vector space consisting of all functions ϕ : G W which satisfy the following two conditions:

  1. ϕ ( h g ) = h ϕ ( g ) for all h H and all g G ;

  2. there is an compact open subgroup 𝐾 of 𝐺 such that ϕ ( g k ) = ϕ ( g ) for all g G and all k K .

We define a (linear) 𝐺-action on Ind H G ( W ) by the rule

( g ϕ ) ( g ) = ϕ ( g g ) , g , g G , ϕ Ind H G ( W ) ;

then Ind H G ( W ) becomes a smooth 𝐺-module, to which we refer as the smooth 𝐺-module smoothly induced by 𝑊. If ρ : H GL ( W ) is the smooth representation of 𝐻 which is canonically determined by the smooth 𝐻-module 𝑊, then we will denote by Ind H G ( ρ ) the smooth representation of 𝐺 which is canonically determined by the smooth 𝐺-module Ind H G ( W ) . In particular, for every smooth character ϑ H , we obtain a smooth representation Ind H G ( ϑ ) of 𝐺 on the vector space Ind H G ( C ϑ ) consisting of all complex-valued functions ϕ : G C which satisfy the two conditions as above; note that condition (i) reads as ϕ ( h g ) = ϑ ( h ) ϕ ( g ) for all h H and all g G .

On the other hand, we define c - Ind H G ( W ) to be the vector subspace of Ind H G ( W ) consisting of all functions ϕ : G W which are compactly supported modulo 𝐻, which means that supp ( ϕ ) H C for some compact subset C G ; as usual, supp ( ϕ ) denotes the support of 𝜙. It is clear that c - Ind H G ( W ) is a 𝐺-invariant vector subspace of Ind H G ( W ) , and thus c - Ind H G ( W ) becomes a smooth 𝐺-module to which we refer as the smooth 𝐺-module compactly induced (or simply c-induced) by 𝑊. As in the case of smooth induction, if ρ : H GL ( W ) is the smooth representation of 𝐻 which is canonically determined by the smooth 𝐻-module 𝑊, then we will denote by c - Ind H G ( ρ ) the smooth representation of 𝐺 which is canonically determined by the smooth 𝐺-module c - Ind H G ( W ) . In the general situation, it is obvious that c - Ind H G ( W ) = Ind H G ( W ) whenever the coset space H G is compact; however, the canonical inclusion map c - Ind H G ( W ) Ind H G ( W ) may be an isomorphism even when H G is not compact (this may occur, for example, in the case where 𝐺 is an algebra group and 𝐻 is an algebra subgroup of 𝐺; see [3, Theorem 1.3]).

The main goal of this paper is to prove the following result.

Theorem 1

Let 𝒜 be a finite-dimensional split basic algebra over a perfect non-Archimedean local field 𝕜, let G = A × be the unit group of 𝒜, and let 𝑉 be an irreducible smooth 𝐺-module. Then there exists a subalgebra ℬ of 𝒜 and a smooth character of the unit group H = B × such that V c - Ind H G ( C ϑ ) .

We should mention that the analogous result has been proved by Z. Halasi in the case where 𝕜 is a finite field (see [10, Theorem 1.3]), and this result extends a previous result (see [9, Theorem 1.2]) for arbitrary algebra groups over finite fields. More generally, E. Gutkin in [8] claimed that every unitary irreducible representation of an algebra group over a locally compact self-dual field is induced (in the sense of unitary representations) by a unitary character of some algebra subgroup; this statement obviously includes the case of an algebra group over a finite field, and the proof of it was shown to be defective by I. M. Isaacs in [11, Section 10]. However, the theorem by Z. Halasi has been successfully generalised by M. Boyarchenko in the paper [3]; in particular, [3, Theorem 1.3]) asserts that every irreducible smooth representation of an algebra group over any non-Archimedean local field 𝕜 is admissible and unitarisable. In general, this is not true for the unit group of an arbitrary split basic 𝕜-algebra. By way of example, the multiplicative group 𝕜 × of a non-Archimedean local field has smooth characters which are not unitarisable; as we mentioned above, the absolute value of 𝕜 defines a smooth character of 𝕜 × whose image does not lie in the unit circle S 1 . On the other hand, the following example shows that there are irreducible smooth representations which are not admissible.

Example 2

Let 𝕜 be an arbitrary non-Archimedean local field, and let A = B 2 ( 𝕜 ) be the standard Borel algebra of M 2 ( 𝕜 ) ; hence G = A × is the standard Borel subgroup B 2 ( 𝕜 ) of GL 2 ( 𝕜 ) consisting of all upper-triangular matrices. It is obvious that 𝐺 is the semidirect product G = T P , where T 𝕜 × × 𝕜 × is the subgroup of 𝐺 consisting of all diagonal matrices and P 𝕜 + is the abelian normal subgroup of 𝐺 consisting of all unipotent matrices.

We consider the conjugation action of 𝐺 on P : for every g G and every ϑ P , we define ϑ g P by ϑ g ( x ) = ϑ ( g x g - 1 ) for all x P . It is clear that there are exactly two 𝐺-orbits on P , namely, the singleton { 1 P } consisting of the trivial character of 𝑃, and its complement P { 1 P } . Both these 𝐺-orbits are locally closed, and thus by [14, Corollaire 2 au Théorème 3], there are two distinct families of irreducible smooth 𝐺-modules each one corresponding to one of these two 𝐺-orbits. On the one hand, one has the family consisting of one-dimensional 𝐺-modules corresponding to the smooth characters ϑ G which satisfy P ker ( ϑ ) ; indeed, 𝑃 lies in the kernel of every smooth character of 𝐺.

On the other hand, there is a family corresponding to a fixed non-trivial smooth character ϑ P . In this case, the centraliser[1] G ϑ is the (internal) direct product G ϑ = Z P , where Z = Z ( G ) is the centre of 𝐺; note that Z 𝕜 × , whereas P 𝕜 + (hence G ϑ 𝕜 × × 𝕜 + ). Since G ϑ is abelian, every irreducible smooth G ϑ -module is one-dimensional. By Rodier’s result, for every smooth character τ ( G ϑ ) satisfying τ P = ϑ , the c-induced smooth G ϑ -module c - Ind G ϑ G ( C τ ) is irreducible (and clearly of dimension at least 2), and moreover, the mapping τ c - Ind G ϑ G ( C τ ) defines a one-to-one correspondence between smooth characters of G ϑ satisfying τ P = ϑ and irreducible smooth 𝐺-modules with dimension at least 2.

Now, let 𝒪 denote the ring of algebraic integers of 𝕜, let

T 0 = { [ α 0 0 β ] : α , β O × } ,

where O × denotes the unit group of 𝒪, and consider the subgroup G 0 = T 0 G ϑ of 𝐺; note that G 0 is an open (hence also a closed) subgroup of 𝐺. Let τ ( G ϑ ) be an arbitrary smooth character of G ϑ satisfying τ P = ϑ , and consider the c-induced 𝐺-module V = c - Ind G ϑ G ( C τ ) . By the transitivity of c-induction (see [1, Proposition 2.25 (b)]), V = c - Ind G 0 G ( V 0 ) , where V 0 = c - Ind G ϑ G 0 ( C τ ) . Since 𝑉 is irreducible, it is obvious that V 0 is an irreducible smooth G 0 -module. Since G 0 is an open subgroup of 𝐺, it follows from [4, Lemma 2.5] that V 0 is naturally embedded as a vector subspace of 𝑉 and that we have a direct sum decomposition

V = g G g V 0 ,

where 𝒢 is a complete set of representatives of the coset space G / G 0 .

Let 𝐾 be a sufficiently small compact open subgroup of G 0 such that ( V 0 ) K 0 (hence V K 0 ); note that we can choose 𝐾 to be the compact open subgroup consisting of the matrices x G 0 such that x 1 , 1 , x 2 , 2 O × and x 1 , 2 ϖ r O for a suitable r Z (here ϖ O is the uniformiser of 𝕜 so that ϖ O is the unique maximal ideal of 𝒪). Let

g m = [ ϖ m 0 0 ϖ - m ] , m N ;

then, for every m N , we have g m G 0 and g m K g m - 1 K . It follows that ( g m V 0 ) K = g m V 0 K { 0 } for every m N , and thus

V K = g G ( g V 0 ) K

has infinite dimension. Therefore, the smooth 𝐺-module 𝑉 is not admissible, and hence it follows from Rodier’s theorem that an irreducible smooth 𝐺-module is admissible if and only if it is one-dimensional.

2 Proof of Theorem 1

The proof of Theorem 1 relies on a refinement of the general techniques used by M. Boyarchenko in the paper [3].

Notation

Henceforth, we fix the following notation which we will use repeatedly, without always recalling their meaning.

  • 𝕜 is a perfect non-Archimedean local field.

  • 𝒜 is a finite-dimensional split basic 𝕜-algebra.

  • G = A × is the unit group of 𝒜.

  • J = J ( A ) is the Jacobson radical of 𝒜, and P = 1 + J .

  • 𝒟 is a diagonal subalgebra of 𝒜, and T = D × is a diagonal subgroup of 𝐺.

We recall that 𝐺 is a second countable ℓ-group and that 𝑃 is a normal c -subgroup of 𝐺; moreover, 𝐺 is the semidirect product G = T P . We also recall that, by an algebra subgroup of 𝐺, we mean a subgroup of 𝐺 of the form Q = 1 + J , where 𝒥 is a (nilpotent) subalgebra of J ( A ) (hence Q P ), whereas an ideal subgroup of 𝐺 is an algebra subgroup Q = 1 + J , where J J ( A ) is an ideal of 𝒜 (in particular, every ideal subgroup of 𝐺 is a normal subgroup of 𝐺).

Let 𝑄 be an arbitrary algebra subgroup of 𝐺 (hence 𝑄 is a subgroup of 𝑃), and let 𝑊 be an arbitrary smooth 𝑄-module. For every smooth character ϑ Q , denote by W ( ϑ ) the vector subspace of 𝑊 linearly spanned by vectors x w - ϑ ( x ) w for x Q and w W , and consider the quotient W ϑ = W / W ( ϑ ) ; note that W ϑ is the largest quotient of 𝑊 where 𝑄 acts via the character 𝜗. We define the spectral support of 𝑊 to be the subset

Spec Q ( W ) = { ϑ Q : W ϑ { 0 } }

of Q ; we recall that, since 𝑄 is an c -group, Q is equal to the Pontryagin dual of 𝑄. If 𝑉 is an arbitrary smooth 𝐺-module, then 𝑉 is also a smooth 𝑄-module, and thus we may define the quotient V ϑ = V / V ( ϑ ) as above; in this case, we refer to Spec Q ( V ) as the spectral support of 𝑉 with respect to 𝑄. In the case where 𝑄 is an ideal subgroup of 𝐺, then 𝑄 is a normal subgroup, and hence 𝐺 acts by conjugation on Q : for every g G and every ϑ Q , we define ϑ g Q by ϑ g ( x ) = ϑ ( g x g - 1 ) for all x Q . We also observe that V ϑ is a smooth G ϑ -module which satisfies

x v ¯ = ϑ ( x ) v ¯ , x Q , v ¯ V ϑ

(that is, the restriction Res Q G ϑ ( V ϑ ) of V ϑ to 𝑄 is isotypic of type 𝜗). In this situation, the following auxiliary result will be important for us; for any subgroup 𝐻 of 𝐺, we will denote by [ H , H ] ¯ the closure of the commutator subgroup [ H , H ] of 𝐻.

Lemma 3

Let 𝑉 be an arbitrary smooth 𝐺-module, and let 𝑄 be an ideal subgroup of 𝐺. Then Spec Q ( V ) is a 𝐺-invariant subset of Q and

V 0 = ϑ Spec Q ( V ) V ( ϑ )

is a 𝐺-submodule of 𝑉 (that is, a 𝐺-invariant vector subspace of 𝑉). In particular, if 𝑉 is irreducible and Spec Q ( V ) is non-empty, then [ Q , Q ] ¯ acts trivially on 𝑉, and hence 𝑉 becomes an irreducible smooth ( G / [ Q , Q ] ¯ ) -module in a natural way.

Proof

For the first assertion, it is enough to observe that V ( ϑ g ) = g - 1 V ( ϑ ) for all ϑ Q and all g G . For the second assertion, we start by observing that V 0 V (otherwise, V ( ϑ ) = V for all ϑ Spec Q ( V ) ), and thus V 0 = { 0 } (because 𝑉 is irreducible). It follows that the natural linear map

V ϑ Spec Q ( V ) V ϑ

is injective, and thus [ Q , Q ] ¯ acts trivially on 𝑉 (because [ Q , Q ] ¯ ker ( ϑ ) for all ϑ Q ). ∎

By [14, Corollaire 1 au Théorème 3], we conclude that Spec Q ( V ) = ϑ G whenever 𝑄 is an ideal subgroup of 𝐺 and 𝑉 is an irreducible smooth 𝐺-module such that Spec Q ( V ) ; indeed, we have the following result.

Lemma 4

Let 𝑄 be an ideal subgroup of 𝐺, and let ϑ Q . Then the 𝐺-orbit ϑ G = { ϑ g : g G } is a locally closed subset of Q .

Proof

We claim the Q has the structure of an affine algebraic 𝕜-variety. To show this, we consider the affine algebraic group Q / [ Q , Q ] ¯ ; since 𝕜 is perfect, both 𝑄 and [ Q , Q ] ¯ are 𝕜-split, and hence Q / [ Q , Q ] ¯ is also 𝕜-split (see [2, Corollary 15.5]). On the one hand, if 𝕜 has characteristic zero, then the usual exponential map shows that Q is naturally isomorphic to the vector space over 𝕜 which is dual to the Lie algebra of Q / [ Q , Q ] ¯ , and hence it has a natural structure of affine algebraic 𝕜-variety. On the other hand, if 𝕜 has prime characteristic, then Q / [ Q , Q ] ¯ is isomorphic as a topological group to its dual Q (see [13, Corollary 5.16]), and thus Q also adquires the structure of an affine algebraic 𝕜-variety, as required. Since 𝐺 acts 𝕜-morphically on Q , it follows by [2, Proposition 6.7] that the 𝐺-orbit ϑ G is a locally closed subset of Q (with respect to the Zariski topology), and the result follows because the Zariski topology is coarser than the topology of Q as an c -group. ∎

Lemma 5

Let 𝑉 be an irreducible smooth 𝐺-module, let 𝑄 be an ideal subgroup of 𝐺, and let ϑ Q be such that V ϑ { 0 } . Then Spec Q ( V ) = ϑ G ; moreover, V ϑ is an irreducible smooth G ϑ -module and V c - Ind G ϑ G ( V ϑ ) .

Proof

We consider the closure [ Q , Q ] ¯ of the commutator subgroup of 𝑄 and the quotient group G ¯ = G / [ Q , Q ] ¯ ; then G ¯ is an ℓ-group, and Q ¯ = Q / [ Q , Q ] ¯ is an abelian normal subgroup of G ¯ . By Lemma 3, 𝑉 is an irreducible smooth G ¯ -module; moreover, Lemma 4 clearly implies that every G ¯ -orbit is a locally closed subset of Q ¯ Q . Therefore, it follows from [14, Corollaire 1 au Théorème 3] that Spec Q ¯ ( V ) = ϑ G ¯ and that V ϑ is an irreducible smooth G ¯ ϑ -module; moreover, [14, Corollaire 2 au Théorème 3] implies that

V c - Ind G ¯ ϑ G ¯ ( V ϑ ) .

The result follows because ϑ G = ϑ G ¯ , [ Q , Q ] ¯ G ϑ and G ¯ ϑ = G ϑ / [ Q , Q ] ¯ . ∎

Our next aim is to prove that, for every irreducible smooth 𝐺-module 𝑉, there exists an ideal subgroup 𝑄 of 𝐺 (depending on 𝑉) such that Spec Q ( V ) ; the existence of the ideal subgroup 𝑄 will be established using a slightly modified key construction due to M. Boyarchenko (see [3, Section 5.2]). We start by proving the following auxiliary result.

Lemma 6

Let 𝑉 be an irreducible smooth 𝐺-module, and let 𝑄 be an arbitrary algebra subgroup of 𝐺. Then the smooth 𝑄-module Res Q G ( V ) has an irreducible quotient.

Proof

By [3, Theorem 1.3], every irreducible smooth 𝑄-module is admissible. Since 𝑄 is also an c -group, Res Q G ( V ) has an irreducible quotient by [3, Corollary 4.8]. ∎

Next, we deal with the case where Res P G ( V ) has a one-dimensional quotient.

Lemma 7

Let 𝑉 be an irreducible smooth 𝐺-module, and let 𝑊 be an irreducible quotient of Res P G ( V ) . Suppose that 𝑊 is one-dimensional, and let ϑ P be the character afforded by 𝑊. Then 𝜗 is 𝐺-invariant if and only if 𝑉 is one-dimensional.

Proof

It is clear that 𝜗 is 𝐺-invariant whenever 𝑉 is one-dimensional. Conversely, suppose that 𝜗 is 𝐺-invariant. Then V ( ϑ ) is a 𝐺-invariant vector subspace of 𝑉, and hence either V ( ϑ ) = { 0 } or V ( ϑ ) = V (because 𝑉 is irreducible). Since 𝑃 is an c -group, [1, Proposition 2.35] implies that W ϑ = W / W ( ϑ ) is an epimorphic image of V ϑ = V / V ( ϑ ) , and thus V ϑ 0 (because W ( ϑ ) = { 0 } , and hence W W ϑ ). Therefore, we must have V ( ϑ ) = { 0 } , and thus

x v = ϑ ( x ) v , x P , v V .

It follows that the restriction Res T G ( V ) of 𝑉 to the diagonal subgroup 𝑇 of 𝐺 is irreducible; indeed, if V is 𝑇-submodule of Res T G ( V ) , then V is also a 𝐺-submodule of 𝑉 (because every vector subspace of 𝑉 is 𝑃-invariant), and so either V = { 0 } or V = V . Since 𝑇 is an abelian ℓ-group, Schur’s lemma implies that 𝑉 is one-dimensional, and this completes the proof. ∎

In the following, we let 𝑉 be an irreducible smooth 𝐺-module and assume that dim V 2 . Let 𝑊 be an irreducible quotient of Res P G ( V ) . On the one hand, suppose that 𝑊 is one-dimensional, and let ϑ P be the character afforded by 𝑊. Since dim V 2 , we have G ϑ G (by the previous lemma); moreover, by Lemma 5, V ϑ is an irreducible smooth G ϑ -module and V c - Ind G ϑ G ( V ϑ ) . As in the proof of Lemma 7, we conclude that V ϑ is one-dimensional; note that W ϑ is a one-dimensional irreducible quotient of Res P G ϑ ( V ϑ ) . Therefore, Theorem 1 holds in this situation once we prove that G ϑ is the unit group of some subalgebra of 𝒜; we observe that G ϑ is the semidirect product G ϑ = T ϑ P .

Proposition 1

For every ideal subgroup 𝑄 of 𝐺 and every ϑ Q , the centraliser T ϑ is the unit group of a subalgebra of 𝒟.

Proof

If J J ( A ) is an ideal of 𝒜 and ϑ ( 1 + J ) , then it is straightforward to check that

D ϑ = { d D : ϑ ( 1 + a d ) = ϑ ( 1 + d a ) for all a J }

is a subalgebra of 𝒟 with ( D ϑ ) × = T ϑ . By way of example, let d , d D ϑ , and let a J be arbitrary. Then, since a ( 1 + d a ) - 1 = ( 1 + a d ) - 1 a , we deduce that

ϑ ( 1 + d a + d a ) = ϑ ( 1 + d a ( 1 + d a ) - 1 ) ϑ ( 1 + d a )
= ϑ ( 1 + a ( 1 + d a ) - 1 d ) ϑ ( 1 + a d )
= ϑ ( 1 + a d ) ϑ ( 1 + ( 1 + a d ) - 1 a d )
= ϑ ( 1 + a d + a d ) ,
and thus d + d D ϑ . ∎

As we mentioned above, this completes the proof of the following particular case of Theorem 1.

Proposition 2

Let 𝑉 be an irreducible smooth 𝐺-module, and let 𝑊 be an irreducible quotient of Res P G ( V ) . Suppose that 𝑊 is one-dimensional, and let ϑ P be the character afforded by 𝑊. Then G ϑ is the unit group of some subalgebra of 𝒜, and V ϑ is a one-dimensional smooth G ϑ -module such that

V c - Ind G ϑ G ( V ϑ ) .

Proof

We have already proved that the smooth G ϑ -module V ϑ is one-dimensional and that V = c - Ind G ϑ G ( V ϑ ) . If 𝒟 is a diagonal subalgebra of 𝒜 and T = D × , then the previous proposition ensures that T ϑ is the unit group of some subalgebra D ϑ of 𝒟. Since G ϑ = T ϑ P and since J ( A ) is an ideal of 𝒜, it follows that A ϑ = D ϑ J ( A ) is a subalgebra of 𝒜 and that G ϑ = ( A ϑ ) × is the unit group of A ϑ . ∎

The proof of Theorem 1 will proceed by induction on dim A . By the results above, the inductive step depends on the existence of an ideal subgroup 𝑄 of 𝐺 such that Spec Q ( V ) , where 𝑉 is an arbitrary irreducible smooth 𝐺-module; moreover, we are reduced to the case where dim W 2 for every irreducible quotient 𝑊 of Res P G ( V ) (which obviously implies that dim V 2 ).

In what follows, we fix the following notation which we will use repeatedly in the subsequent results (without always recalling their meaning). Let n 2 be an integer such that J n { 0 } , and write N = 1 + J n . Since J n J n - 1 are ideals of 𝒜, it follows from Lemma 2 that there exists an ideal ℒ of 𝒜 such that J n L J n - 1 and dim L = dim J n + 1 . We fix such an ideal and set Q = 1 + L . (As usual, if 𝑔 and ℎ are elements of a group, then we define g h = h - 1 g h and [ g , h ] = g - 1 h - 1 g h = g - 1 g h .)

Lemma 8

Let ς N be 𝑃-invariant, and define

J ς = { a J : ς ( [ 1 + a , 1 + u ] ) = 1 for all u L } .

Then J ς is a subalgebra of 𝒥 satisfying J 2 J ς and dim J ς dim J - 1 . Furthermore, if we define the map φ ς : P Q by the rule

φ ς ( g ) ( h ) = ς ( [ g , h ] ) , g P , h Q ,

then φ ς is a group homomorphism with ker ( φ ς ) = 1 + J ς and φ ς ( P ) N , where N = { τ Q : N ker ( τ ) } is the orthogonal subgroup of 𝑁 in Q ; hence φ ς naturally defines a group homomorphism φ ¯ ς : P ( Q / N ) .

Proof

We first observe that the map φ ς is a well-defined group homomorphism. On the one hand, we have [ P , Q ] [ 1 + J , 1 + J n - 1 ] 1 + J n = N . On the other hand, since [ g , h k ] = [ g , k ] [ g , h ] k , we deduce that

φ ς ( g ) ( h k ) = ς ( [ g , k ] ) ς ( [ g , h ] ) = φ ς ( g ) ( h ) φ ς ( g ) ( k )

for all g P and all h , k Q ; we recall that 𝜍 is 𝑃-invariant. It follows that, for every g P , the map φ ς ( g ) : Q C × is indeed a (smooth) character of 𝑄. Similarly, since [ g h , k ] = [ g , k ] h [ h , k ] , we have

φ ς ( g h ) ( k ) = ς ( [ g , k ] ) ς ( [ h , k ] ) = φ ς ( g ) ( k ) φ ς ( h ) ( k )

for all g , h P and all k Q , and so φ ς is a group homomorphism.

Now, since [ P , N ] ker ( ς ) (because 𝜍 is 𝑃-invariant), the image φ ς ( P ) clearly lies in N ; moreover, it is obvious (by the definition) that ker ( φ ς ) = 1 + J ς . Let a J and α 𝕜 be arbitrary. As in the proof of [3, Proposition 3.1 (b), p. 547], we deduce that

[ 1 + α a , 1 + u ] [ 1 + a , 1 + α u ] - 1 [ 1 + J , 1 + J n ]

for all u J n - 1 . Since 𝜍 is 𝑃-invariant (and [ 1 + J , 1 + J n ] = [ P , N ] ), we conclude that

( ) ς ( [ 1 + α a , 1 + u ] ) = ς ( [ 1 + a , 1 + α u ] )

for all u J n - 1 , and this clearly implies that α a J ς for all α 𝕜 and all a J ς . On the other hand, [9, Theorem 1.4] implies that

[ 1 + J 2 , 1 + L ] [ 1 + J 2 , 1 + J n - 1 ] [ 1 + J , 1 + J n ] = [ P , N ] ,

and thus J 2 J ς . Since

( 1 + u + v ) - 1 ( 1 + u ) ( 1 + v ) = 1 + ( 1 + u + v ) - 1 u v 1 + J 2 ,

we see that u + v J ς for all u , v J ς . It follows that J ς is an ideal of 𝒥 with J 2 J ς .

Finally, note that N ( Q / N ) and that

Q / N = ( 1 + L ) / ( 1 + J n ) 1 + ( L / J n ) 𝕜 + .

Therefore, we get an isomorphism of abelian groups N ( 𝕜 + ) , and hence N acquires the structure of a vector space over 𝕜, where the scalar multiplication is defined by

( α τ ) ( 1 + u ) = τ ( 1 + α u ) , α 𝕜 , τ N , u L .

On the other hand, since 𝕜 is a self-dual field, there is a 𝕜-linear isomorphism 𝕜 ( 𝕜 + ) , and thus N is one-dimensional. Furthermore, the argument used above can be repeated to show that the mapping a φ ς ( 1 + a ) defines a 𝕜-linear homomorphism φ ^ ς : J N with kernel J ς ; we note that implies that

α φ ^ ς ( a ) = φ ^ ς ( α a ) , α 𝕜 , a J .

It follows that dim J - dim J ς 1 , and this completes the proof. ∎

The following result is essentially a particular case of [10, Lemma 3.4]; for convenience of the reader, we include a proof which avoids the finiteness of the base field 𝕜.

Lemma 9

Let ς N be 𝐺-invariant, let ℐ be an ideal of 𝒜 with J 2 I J , and let J ς J be defined as in Lemma 8. Then I ς = I J ς is an ideal of 𝒜.

Proof

The result is obvious in the case where J ς = J ; hence we assume that J ς J so that dim J ς = dim J - 1 (by the previous lemma). The result is also clearly true in the case where I J ς ; thus we may assume that J ς + I = J , which implies that dim I ς = dim I - 1 . We now proceed by induction on dim I , the result being obvious if dim I = dim J 2 + 1 ; indeed, since J 2 I J ς I , either

I J ς = J 2 or I J ς = I .

Therefore, we may assume that dim I dim J 2 + 2 and that the result is true whenever I is an ideal of 𝒜 with J 2 I J and dim I < dim I .

Let I ς be the unique ideal of 𝒜 which is maximal with respect to the condition I ς I ς ; hence we must prove that I ς = I ς . Since I ς is clearly a 𝒟-bimodule, Lemma 2 ensures that I = I ς V for some sub-bimodule 𝒱 of ℐ. Let V ς = V J ς , and note that I ς = I ς V ς ; hence I ς = I ς if and only if V ς = { 0 } . By way of contradiction, we assume that V ς { 0 } ; note that V ς V (otherwise, I ς = I ). Since 𝑄 is a 𝑇-invariant subgroup of 𝐺 (because it is an ideal subgroup of 𝐺), we deduce that

ς ( [ 1 + t - 1 a t , h ] ) = ς ( [ 1 + a , t h t - 1 ] t ) = ς ( [ 1 + a , t h t - 1 ] ) = 1

for all a V ς , all t T and all h Q . Since V ς J ς (and since 𝒱 is clearly 𝑇-invariant), it follows that V ς is a 𝑇-invariant vector subspace of 𝒱.

On the other hand, suppose that V { 0 } is a proper sub-bimodule of 𝒱, and let I = I ς + V . Then I is an ideal of 𝒜 with I I , and thus I J ς is an ideal of 𝒜 (by the inductive hypothesis). Since I ς I J ς I J ς = I ς , we conclude that I J ς = I ς (by the maximality of I ς ), and thus

V ς V = ( J ς I ) V = I ς V = { 0 } .

Therefore, the vector subspaces 𝒱 and V ς of ℐ satisfy the assumptions of [10, Lemma 2.2] (we note that the proof of this result holds for an arbitrary field). In particular, if e 1 , , e n D are non-zero orthogonal idempotents such that D = 𝕜 e 1 𝕜 e n , then dim e r V 1 and dim V e r 1 for all 1 r n .

Next, we consider the ideal subgroup Q = 1 + L of 𝐺, and the group homomorphism φ ς : P Q (as defined in the previous lemma); we recall that we have ker ( φ ς ) = 1 + J ς . Since ℒ is an ideal of 𝒜 with dim L = dim J n + 1 , we have L = J n 𝕜 u , where u = e i u e j for some 1 i , j n ; hence Q = ( 1 + 𝕜 u ) N .

Firstly, suppose that e j V = V e i = { 0 } , and let v V be arbitrary. Then we have u v = u e j v = 0 and v u = v e i u = 0 , so [ 1 + v , 1 + α u ] = 1 for all α 𝕜 . It follows that 1 + v ker ( φ ς ) = 1 + J ς , and thus V J ς . Therefore, in this case, we have V I J ς = I ς , and hence I = I ς + V I ς , which implies that I ς = I is an ideal of 𝒜.

Now, suppose that e j V { 0 } , and let v V be such that e j V = 𝕜 v ; since 𝒱 has a 𝕜-basis consisting of vectors w V satisfying D w = w D = 𝕜 w , it is clear that v = v e k for some 1 k n (see Lemma 2). Then V = V 𝕜 v for some sub-bimodule V of 𝒱; in particular, we have e j V = V e k = { 0 } . On the one hand, suppose that V e i = { 0 } . Then the argument above shows that V I ς , and so V V I ς = V ς . It follows that V = { 0 } (because V V ), and thus V = 𝕜 v . By the definition of 𝒱, we conclude that I = I ς 𝕜 v , and hence we have dim I = dim I ς + 1 . Since I ς I ς and dim I ς = dim I - 1 , we must have I ς = I ς , and hence I ς is an ideal of 𝒜. If k = i , then V e i = V e i 𝕜 v . Since dim ( V e i ) 1 , we get V e i = 0 , which is the previously handled case.

Since k i (otherwise, v = v e i V e i = 𝕜 w ), we have v u = 0 , and thus

[ 1 + v , 1 + u ] = 1 - u v ,

where u J is such that ( 1 + u ) - 1 = 1 - u . Since u v u A v , we see that ( u v ) 2 ( u A v ) 2 = { 0 } , and thus S = 1 + 𝕜 ( u v ) is a 𝑇-invariant algebra subgroup of 𝑁; indeed, we have D ( u v ) D = 𝕜 u v (because u = e i u and v = v e k ). Let α 𝕜 × be arbitrary, and choose t T such that t - 1 u = α u and v t = v ; note that, since i k , it is enough to choose t T satisfying t e i = α - 1 e i and t e k = e k . It follows that

[ 1 + v , 1 + u ] t = ( 1 - u v ) t = 1 - t - 1 u v t = 1 - α u v ,

and thus the restriction ς S of 𝜍 to 𝑆 is a (smooth) character of 𝑆 which is constant on S { 1 } (because 𝜍 is 𝑇-invariant). Therefore, ς S must be the trivial character, and thus ς ( [ 1 + v , 1 + u ] ) = 1 . It follows that 1 + v 1 + J ς , and so v V J ς = V ς . Since 𝕜 v V and D v = v D = 𝕜 v , we must have 𝕜 v = { 0 } , a contradiction.

The proof is complete. ∎

We are now able to prove the following crucial result.

Proposition 3

Let ς N be 𝑃-invariant, and let J ς J be as in Lemma 8. Then [ Q , Q ] ker ( ς ) , and there exists ϑ Q such that ϑ N = ς ; moreover, the following properties hold.

  1. P ϑ = 1 + J ς for all ϑ Q such that ϑ N = ς .

  2. If P ϑ P and if ϑ Q is such that ϑ N = ς , then there exists g P such that ϑ = ϑ g .

Proof

We start by observing that [ Q , Q ] ker ( ς ) . Indeed, since dim L = J n + 1 , there exists a L such that L = J n 𝕜 a , and hence Q = ( 1 + 𝕜 a ) N . Since [ 1 + α a , 1 + β a ] = 1 for all α , β 𝕜 , we see that ς ( [ 1 + 𝕜 a , 1 + 𝕜 a ] ) = { 1 } , and this clearly implies that ς ( [ Q , Q ] ) = { 1 } (because 𝜍 is 𝑃-invariant). Let 𝑉 be an irreducible quotient of the smoothly induced 𝑄-module Ind N Q ( C ς ) ; as in the proof of Lemma 6, the existence of 𝑉 follows from [3, Theorem 1.3] and [3, Corollary 4.8]. Since 𝑁 is a normal subgroup of 𝑄 and 𝜍 is 𝑄-invariant, we have x ϕ = ς ( x ) ϕ for all x N and all ϕ Ind N Q ( C ς ) , and thus x v = ς ( x ) v for all x N and all v V . Since [ Q , Q ] ker ( ς ) , it follows from Schur’s lemma that dim V = 1 , and thus 𝑉 affords a character ϑ Q which clearly satisfies ϑ N = ς .

In order to prove (i) and (ii), we consider the group homomorphism φ ς : P Q as defined in Lemma 8; we recall that φ ς ( P ) N and that ker ( φ ς ) = 1 + J ς , where J ς is an ideal of 𝒥 satisfying J 2 J ς and dim J ς dim J - 1 . On the one hand, (i) follows because P ϑ = ker ( φ ς ) = 1 + J ς for all ϑ Q such that ϑ N = ς . On the other hand, let us assume that P ϑ P (hence ker ( φ ς ) P and J ς J ), and let x P be such that φ ς ( x ) Q is not identically equal to 1. Let a J be such that x = 1 + a , and note that implies that φ ς ( 1 + α a ) φ ς ( P ) N for all α 𝕜 . Moreover, since J / J ς is one-dimensional, we conclude that the 𝕜-linear map φ ^ ς : J N (as defined in the proof of Lemma 8) is surjective, and thus the group homomorphism φ ς : P N is also surjective and induces an isomorphism P / P ϑ N ; in particular, it follows that the mapping α φ ς ( 1 + α a ) defines a group isomorphism 𝕜 + N and that N = { φ ς ( 1 + α a ) : α 𝕜 } .

To conclude the proof of (ii), let ϑ Q be such that ϑ N = ς , and consider the character ϑ ϑ - 1 Q . It is obvious that ϑ ϑ - 1 N , and thus there exists α 𝕜 such that ϑ ϑ - 1 = φ ς ( 1 + α a ) . If we set g = 1 + α a , then

ϑ ( x ) ϑ ( x ) - 1 = ς ( [ g , x ] ) = ς ( g - 1 x - 1 g x ) = ς ( g x g - 1 x - 1 ) = ϑ ( g x g - 1 x - 1 ) = ϑ ( g x g - 1 ) ϑ ( x ) - 1

(where the third equation follows from the 𝑃-invariance of 𝜍), and hence we get ϑ ( x ) = ϑ ( g x g - 1 ) for all x Q , as required. ∎

Proposition 4

Let ς N be 𝑃-invariant, and let ϑ Q be such that ϑ N = ς . Then G ϑ is the unit group of some subalgebra of 𝒜.

Proof

In the case where 𝜗 is 𝑃-invariant, the result follows by Proposition 1; hence we assume that P ϑ P . Let 𝒟 be a diagonal subalgebra of 𝒜, and let T = D × . By Proposition 1, we know that T ς is the unit group of some subalgebra D ς of 𝒟; similarly, T ϑ is the unit group of some subalgebra D ϑ of 𝒟.

Since 𝜍 is 𝑃-invariant, we see that the centraliser G ς of 𝜍 is the unit group of the subalgebra A ς = D ς J of 𝒜; indeed, we have G ς = T ς P . Let J ς be the ideal of 𝒥 defined as in Lemma 8, and note that 1 + J ς = P ϑ is the centraliser of 𝜗 in 𝑃 (by the previous proposition, because ϑ N = ς ). Since 𝜍 is G ς -invariant, Lemma 9 implies that J ς is an ideal of A ς , and thus T ς P ϑ is the unit group of the subalgebra B ς = D ς J ς of A ς (and hence of 𝒜). We also observe that G ϑ G ς (because ϑ N = ς ).

Since J ς is an ideal of 𝒥 with dim J ς = dim J - 1 , there exists a J such that J = J ς 𝕜 a and D a = a D = 𝕜 a (see Lemma 2). On the one hand, suppose that [ D ς , J ] = 0 (so that d a = a d for all d D ); in particular, we see that [ T ς , P ] = 1 (because T ς = ( D ς ) × ). Let y G ϑ be arbitrary, and write y = t x for t T ς , x P (note that this decomposition exists because G ϑ G ς = T ς P ). For every g Q , we deduce that

ϑ ( g ) = ϑ ( y g y - 1 ) = ϑ ( t x g x - 1 t - 1 ) = ϑ ( x g x - 1 )

(where the last equality holds because [ T ς , P ] = 1 , and hence t x g x - 1 = x g x - 1 t ). It follows that x G ϑ , and thus t = y x - 1 T ϑ . It follows that y = t x T ϑ P ϑ , and so G ϑ = T ϑ P ϑ is the unit group of the subalgebra B ϑ = D ϑ J ς of 𝒜. On the other hand, suppose that [ D ς , J ] 0 , and observe that the above element a J can be chosen such that [ D ς , a ] 0 ; indeed, if [ D ς , a ] = 0 , then we may replace 𝑎 by an element a + b , where b J ς satisfies D b = b D and is such that [ D ς , b ] 0 .

Since G ϑ G ς = T ς P and P ϑ G ϑ , for every element g G ϑ , there exist unique elements t T ς and x 1 + 𝕜 a such that g t x P ϑ . In fact, for every t T ς , there is a unique element x ( t ) 1 + 𝕜 a such that t x ( t ) G ϑ . To see this, let t T ς be arbitrary. Then ϑ t Q satisfies ( ϑ t ) N = ς t = ς , and thus Proposition 3 implies that ϑ t = ϑ x for some x P . Therefore, ϑ t x - 1 = ϑ , and hence t x - 1 G ϑ . Since P = ( 1 + 𝕜 a ) ( 1 + J ς ) = ( 1 + 𝕜 a ) P ϑ and P ϑ G ϑ , we have x - 1 x ( t ) P ϑ for some x ( t ) 1 + 𝕜 a , and so ϑ t x ( t ) = ϑ x x ( t ) = ϑ ; note that x ( t ) is uniquely determined by t T ς .

Suppose that T ς = T ϑ . If this is the case, then ϑ x ( t ) = ϑ t x ( t ) = ϑ , and thus x ( t ) G ϑ P = P ϑ for all t T ς . By the above, we conclude that G ϑ = T ς P ϑ is the unit group of the subalgebra B ς of 𝒜. Therefore, we henceforth assume that T ς T ϑ .

For every t T ς , let α ( t ) 𝕜 be such that x ( t ) = 1 + α ( t ) a . It is straightforward to check that the mapping t T ϑ α ( t ) defines an injective map

α : T ς / T ϑ 𝕜 .

Since T ς = ( D ς ) × of 𝒟, the stabiliser

( T ς ) a = { t T ς : t - 1 a t = a }

is the unit group of the subalgebra ( D ς ) a = { d D ς : d a = a d } of D ς . Moreover, since [ D ς , a ] 0 , the mapping d d a - a d defines a surjective 𝕜-linear map D ς 𝕜 a with kernel ( D ς ) a , and so dim ( D ς ) a = dim D ς - 1 . On the other hand, it is straightforward to check that 𝛼 induces (by restriction) a group homomorphism

α ~ : ( ( T ς ) a T ϑ ) / T ϑ 𝕜 + .

Since ( ( T ς ) a T ϑ ) / T ϑ ( T ς ) a / ( T ϑ ( T ς ) a ) is either the trivial group or isomorphic to the direct product of a finite number of copies of the multiplicative group 𝕜 × (because ( T ς ) a and T ϑ ( T ς ) a are unit groups of subalgebras of 𝒟), we must have ( T ς ) a = T ϑ ( T ς ) a ; indeed, if we choose a root of unity ζ 𝕜 × of order coprime to the characteristic of the residue field of 𝕜, then we must have α ~ ( ζ ) = 0 . Therefore, we conclude that ( T ς ) a T ϑ , and so ( D ς ) a D ϑ D ς . Since dim ( D ς ) a = dim D ς - 1 , it follows that D ϑ = ( D ς ) a , and thus T ϑ = ( T ς ) a and T ς / T ϑ 𝕜 × .

Since P ϑ is an ideal subgroup of G ς and P ϑ G ϑ G ς , it is also a normal subgroup of G ϑ , and thus the mapping t ( t x ( t ) ) P ϑ defines a bijection

β : T ς G ϑ / P ϑ .

Since 𝑃 is a normal subgroup of 𝐺, we have

( t t x ( t t ) ) ( t x ( t ) ) - 1 ( t x ( t ) ) - 1 P G ϑ = P ϑ ,

so β ( t t ) = β ( t ) β ( t ) for all t , t T ς . It follows that 𝛽 is a group isomorphism, and hence G ϑ / P ϑ is an abelian group. Therefore, we conclude that ( T ϑ P ϑ ) / P ϑ is a normal subgroup of G ϑ / P ϑ , and thus T ϑ P ϑ is a normal subgroup of G ϑ . Since β ( T ϑ ) = ( T ϑ P ϑ ) / P ϑ , we see that 𝛽 naturally induces a group isomorphism

β ~ : T ς / T ϑ G ϑ / ( T ϑ P ϑ ) .

Now, for every t T ς , we have t a t - 1 𝕜 a , and hence there is λ ( t ) 𝕜 × such that t a t - 1 = λ ( t ) a . The mapping t λ ( t ) defines a group homomorphism λ : T ς 𝕜 × with ker ( λ ) = ( T ς ) a = T ϑ . On the other hand, as J = J ς 𝕜 a , for every x P , there exists μ ( x ) 𝕜 such that x ( 1 + μ ( x ) a ) P ϑ , and the mapping x μ ( x ) defines a group homomorphism μ : P 𝕜 + with ker ( μ ) = P ϑ . Since every element g T ς P is uniquely written as a product g = x t for t T ς and x P , we may define a map ψ : T ς P GL 2 ( 𝕜 ) by the rule

ψ ( x t ) = [ λ ( t ) μ ( x ) 0 1 ] , t T ς , x P .

(Since T ς P = P T ς , this is a well-defined map.) Since ( x t ) ( x t ) = ( x t x t - 1 ) ( t t ) and μ ( x t x t - 1 ) = λ ( t ) μ ( x ) + μ ( x ) , we see that ψ ( ( x t ) ( x t ) ) = ψ ( x t ) ψ ( x t ) for all t , t T ς and all x , x P , which means that 𝜓 is a group homomorphism. It is clear that ker ( ψ ) = T ϑ P ϑ , and so 𝜓 induces a group isomorphism ψ ~ : ( T ς P ) / ( T ϑ P ϑ ) M 2 , where M 2 denotes the mirabolic subgroup

M 2 = { [ r s 0 1 ] : r 𝕜 × , s 𝕜 }

of GL 2 ( 𝕜 ) .

Finally, consider the image M 2 = ψ ~ ( G ϑ / ( T ϑ P ϑ ) ) ; recall that G ϑ is a subgroup of T ς P . Since there are group isomorphisms G ϑ / ( T ϑ P ϑ ) T ς / T ϑ 𝕜 × , we conclude that M 2 𝕜 × is a commutative subgroup of M 2 . For every t T ς , we have t x ( t ) G ϑ , and

ψ ( x ( t ) t ) = [ λ ( t ) μ ( x ( t ) ) 0 1 ] ;

moreover, note that the matrix ψ ( t x ( t ) ) is semisimple for all t T ς T ϑ . Therefore, since M 2 is commutative and consists of semisimple matrices, there exists y GL 2 ( 𝕜 ) such that

y [ λ ( t ) μ ( x ( t ) ) 0 1 ] y - 1 = [ λ ( t ) 0 0 1 ]

for all t T ς ; in fact, we may choose y M 2 . Let g T ς P be such that ψ ( g ) = y . Then

ψ ( g G ϑ g - 1 ) = x M 2 x - 1 = { [ λ ( t ) 0 0 1 ] : t T ς } = ψ ( T ς ) ,

and thus g G ϑ g - 1 = T ς P ϑ is the unit group of the subalgebra B ς of 𝒜. It follows that G ϑ is the unit group of the subalgebra g - 1 B ς g of 𝒜, and this completes the proof. ∎

We are now able to prove our main result.

Proof of Theorem 1

We proceed by induction on dim A , the result being obvious if dim A = 1 . Therefore, we assume that dim A 2 and that the result is true whenever A is a subalgebra 𝒜 with dim A dim A .

Let 𝑉 be an arbitrary irreducible smooth 𝐺-module, and let V be an irreducible quotient of Res P G ( V ) (the existence of which is guaranteed by Lemma 6). In spite of Proposition 2, we may assume that dim V 2 . In this situation, there is an integer m 2 such that J m { 0 } and J m + 1 = { 0 } ; note that J 2 { 0 } (otherwise, P = 1 + J is abelian, and hence V must be one-dimensional). Since 1 + J m lies in the centre of 𝑃, Schur’s lemma implies that 1 + J m acts on V by scalar multiplications, and thus we may choose the smallest positive integer 𝑛 for which there exists ς ( 1 + J n ) such that g v = ς ( g ) v for all g 1 + J n and all v V ; recall that, by construction, 𝜍 is 𝑃-invariant. We note that, since V is an irreducible smooth 𝑃-module with dim V 2 , we must have n 2 ; furthermore, since [ 1 + J , 1 + J n - 1 ] 1 + J n , the minimal choice of 𝑛 implies that 𝜍 is not identically equal to 1 (otherwise, Schur’s lemma would imply that the subgroup 1 + J n - 1 acts on V by scalar multiplications). Since J n - 1 and J n are ideals of 𝒜, Lemma 2 implies that J n - 1 = L 1 + + L t for some ideals L 1 , , L t of 𝒜 satisfying J n L i J n - 1 and dim ( L i / J n ) = 1 for all 1 i t . By the minimal choice of 𝑛, we must have ς [ 1 + J , 1 + L i ] { 1 } for some 1 i t (otherwise, we would have [ 1 + J , 1 + J n - 1 ] ker ( ς ) , and hence 1 + J n - 1 would act on V by scalar multiplications).

Let N = 1 + J n , and let Q = 1 + L , where we set L = L i . The argument used in the proof of Lemma 6 shows the smooth 𝑄-module Res Q P ( V ) has an irreducible quotient V ′′ . Since [ Q , Q ] ker ( ς ) (by Proposition 3), Schur’s lemma implies that V ′′ is one-dimensional, and thus it affords a character ϑ Q . (Notice that the extreme case where n = 2 and dim J = dim J 2 + 1 cannot occur; indeed, in this situation, we must have Q = P , and hence V ′′ = V , which contradicts the assumption dim V 2 .) In particular, we have V ϑ { 0 } , and thus V ϑ { 0 } (by [1, Proposition 2.35] because 𝑃 is an c -group). By Lemma 5, V ϑ is an irreducible G ϑ -module and we have V = c - Ind G ϑ G ( V ϑ ) . Since 𝑁 acts on V (hence on V ′′ ) via the character 𝜍, we must have ϑ N = ς , and thus G ϑ is the unit group of some subalgebra A of 𝒜 (by Proposition 4). Since ϑ ( [ P , Q ] ) = ς ( [ P , Q ] ) { 1 } , we must have P ϑ P , and thus G ϑ G . Therefore, we have dim A dim A , and thus it follows by induction that there exists a subalgebra ℬ of A such that V ϑ c - Ind H G ϑ ( W ) , where H = B × is the unit group of ℬ and 𝑊 is a one-dimensional 𝐻-module. By transitivity of c-induction [1, Proposition 2.25 (b)], we conclude that

V c - Ind G ϑ G ( c - Ind H G ϑ ( W ) ) c - Ind H G ( W ) ,

as required. ∎

3 Admissibility and unitarisability

We conclude the present work with two remarks concerning the admissibility and unitarisability of an arbitrary irreducible smooth 𝐺-module. On the one hand, we establish the following result. (We let the notation be as in the previous section.)

Theorem 2

Let 𝑉 be an irreducible smooth 𝐺-module, and let ℬ be a subalgebra of 𝒜 such that V c - Ind H G ( W ) , where H = B × is the unit group of ℬ and 𝑊 is a one-dimensional smooth 𝐻-module. Then the following are equivalent.

  1. 𝐻 contains a diagonal subgroup 𝑇 of 𝐺.

  2. The smooth 𝐺-module 𝑉 is admissible.

  3. There is an isomorphism of 𝐺-modules V Ind H G ( W ) .

Proof

For simplicity of reading, we consider several (independent) steps; in the first step, we establish that both (ii) and (iii) imply (i).

Step 1

Assume that either 𝑉 is admissible or V Ind H G ( W ) . Then 𝐻 contains a diagonal subgroup 𝑇 of 𝐺.

Proof

Let D be a diagonal subalgebra of ℬ, and let T = ( D ) × be the unit group of D (hence T is a diagonal subgroup of 𝐻). On the other hand, let 𝒟 be a diagonal subalgebra of 𝒜 that contains D ; to see that such a diagonal subalgebra exists, it is enough to consider the inclusion map ( D + J ) / J ( D + J ) / J , where 𝒟 is a diagonal algebra of 𝒜, and use it to extend the basis of orthogonal idempotents of D to a basis of orthogonal idempotents of 𝒟. Note that 𝒟 is clearly a D -bimodule, and thus Lemma 2 implies that D = D D ′′ for some sub-bimodule D ′′ of 𝒟. Then T = T T ′′ , where T ′′ = ( D ′′ ) × is the unit group of D ′′ , and 𝐺 decomposes as the semidirect product G = G T ′′ , where G = T P ; moreover, we have B = D J ( B ) , where J ( B ) denotes the Jacobson radical of ℬ, and thus H G . Let V = c - Ind H G ( W ) so that

V c - Ind H G ( W ) c - Ind G G ( V ) .

On the one hand, assume that V Ind H G ( W ) . Then c - Ind H G ( W ) = Ind H G ( W ) (because 𝑉 is irreducible), and so Ind H G ( W ) is irreducible. This implies that

V = Ind H G ( W ) , and thus Ind H G ( W ) Ind G G ( V ) ;

moreover, we conclude that Ind G G ( V ) = Ind G G ( V ) . Now, since G = G T ′′ is a semidirect product, every element of 𝐺 is uniquely factorised as a product g t for g G and t T ′′ , and hence every function ϕ Ind G G ( W ) is uniquely determined by the rule

ϕ ( g t ) = g ϕ ( t ) , g G , t T ′′ ;

in particular, a function ϕ Ind G G ( W ) lies in c - Ind G G ( W ) ) if and only if its restriction to T ′′ has compact support. Since T ′′ ( 𝕜 × ) r for some non-negative integer r 0 , we conclude that Ind G G ( V ) = c - Ind G G ( V ) if and only if T ′′ = { 1 } (that is, if and only if r = 0 ); in other words, we have Ind G G ( V ) = c - Ind G G ( V ) if and only if G = G , which occurs if and only if T H . Finally, since

Ind G G ( V ) = c - Ind G G ( V ) ,

we conclude that T H , as required.

On the other hand, suppose that V c - Ind H G ( W ) is admissible. Let

δ G : G R + × and δ H : H R + ×

be the modular characters of 𝐺 and 𝐻, respectively; for the definition, we refer to [6, Section 1.4] (see also [4, Section 3.3]). If V denotes the smooth dual of 𝑉 (see [4, Section 2.8]), then [4, Theorem 3.5] implies that there is a natural isomorphism

V ( c - Ind H G ( W ) ) Ind H G ( δ G / H W ) ,

where δ G / H = ( δ H ) - 1 ( δ G ) H and where the smooth 𝐻-module δ G / H W has underlying vector space equal to W and 𝐻-action defined by

h w = δ G / H ( h ) ( h w ) , h H , w W .

Since 𝑉 is admissible, the smooth dual V is irreducible [4, Proposition 2.10], and so the smooth 𝐺-module Ind H G ( δ G / H W ) is also irreducible. By the above, we conclude that T H , and this completes the proof. ∎

We next prove that (i) implies (ii) in the particular situation where the subalgebra ℬ has codimension one in 𝒜.

Step 2

Let J 0 = J ( B ) denote the Jacobson radical of ℬ, and suppose that

dim J 0 = dim J - 1 , where J = J ( A ) .

Moreover, assume that 𝐻 contains a diagonal subgroup 𝑇 of 𝐺. Then the smooth 𝐺-module 𝑉 is admissible.

Proof

Firstly, we note that J 2 J 0 ; otherwise, since dim J = dim J 0 + 1 , we must have J 0 + J 2 = J , and so [11, Lemma 3.1] implies that J 0 = J . Now, let P 0 = 1 + J 0 , and note that P 0 is an ideal subgroup of 𝐺 with H = P 0 T (because T H ). Let τ ( P 0 ) be the character of P 0 afforded by the one-dimensional smooth P 0 -module Res P 0 H ( W ) , and consider the irreducible smooth G τ -module V τ ; note that V c - Ind G τ G ( V τ ) . Since 𝐻 is clearly a subgroup of G τ , there are isomorphisms

V c - Ind H G ( W ) c - Ind G τ G ( c - Ind H G τ ( W ) ) ;

in particular, we conclude that the smooth G τ -module c - Ind H G τ ( W ) is irreducible. Since x ϕ = τ ( x ) ϕ for all x P 0 and all ϕ c - Ind H G τ ( W ) , it follows from [14, Corollaire 2 au Théorème 3] that c - Ind H G τ ( W ) V τ ; note that, as in the proof of Lemma 5, we should consider the closure [ P , P ] ¯ of the commutator subgroup of 𝑃 and replace 𝐺 by the quotient group G / [ P , P ] ¯ and 𝑃 by P / [ P , P ] ¯ (see also [3, Theorem 4.11]). Since x v = τ ( x ) v for all x P 0 and all v V τ , it follows that Res T G τ ( V τ ) is irreducible, and hence V τ is one-dimensional (by Schur’s lemma); in particular, we conclude that H = G τ and that W V τ .

Since dim V 2 , the proof of Theorem 1 guarantees that there exists an ideal ℒ of 𝒜 satisfying J 2 L J and dim ( L / J 2 ) = 1 , and such that V c - Ind G ϑ G ( V ϑ ) for some smooth character ϑ ( 1 + L ) . Since τ ( [ 1 + J , 1 + J 0 ] ) 1 (because otherwise P = 1 + J would be contained in G τ = H ), we see from the construction that we may choose L J 0 and ϑ = τ 1 + L ; furthermore, it follows from Proposition 3 that ς = ϑ 1 + J 2 is a 𝑃-invariant smooth character of 1 + J 2 such that P ϑ = 1 + J ς , where J ς = { a J : ς ( [ 1 + a , 1 + u ] ) = 1 for all u L } is an ideal of 𝒜 with dim J ς = dim J - 1 (see also Lemma 8). Since ς = τ 1 + J 2 and since 1 + J 0 P τ , we must have

J 0 J ς , and thus J 0 = J ς

(because dim J 0 = dim J - 1 = dim J ς ). On the other hand, we recall from Proposition 3 that the 𝑃-orbit ϑ P ( 1 + L ) of 𝜗 consists of all ϑ ( 1 + L ) which satisfy ( ϑ ) 1 + J 2 = ς ; in particular, ϑ P is a closed subset of ( 1 + L ) . Since T H = G τ , we conclude that G ϑ = T ( 1 + J 0 ) = H = G τ , and so the 𝐺-orbit ϑ G = ϑ P is a closed subset of ( 1 + L ) . Finally, we note that V ϑ = V τ (by the choice of ϑ = τ 1 + L ) and that V τ is an admissible smooth G τ -module (because it is one-dimensional). Therefore, it follows from [14, Théorème 4] that the smooth 𝐺-module V c - Ind G ϑ G ( V ϑ ) c - Ind G τ G ( V τ ) is also admissible, and this completes the proof. ∎

In the next step, we establish that (i) implies (ii).

Step 3

Assume that 𝐻 contains a diagonal subgroup 𝑇 of 𝐺. Then the smooth 𝐺-module 𝑉 is admissible.

Proof

As in the previous proof, we argue by induction on the dimension of 𝒜, the result being obvious in the case where 𝒜 is one-dimensional; indeed, the result is obvious in the case where 𝑉 is one-dimensional, which clearly includes the case where the 𝕜-algebra 𝒜 is semisimple (because, if this is the case, then 𝒜 must be commutative). Therefore, we may assume that dim V 2 ; in particular, the Jacobson radical J = J ( A ) of 𝒜 must be non-zero.

Let J ( B ) denote the Jacobson radical of ℬ. If J ( B ) + J 2 = J , then J ( B ) = J (by [11, Lemma 3.1]), and thus H = T ( 1 + J ) = G and V = W is one-dimensional. It follows that J ( B ) + J 2 J , and so there exists an ideal J of 𝒜 such that J ( B ) + J 2 J J and dim J = dim J - 1 . Let P = 1 + J , and let G = T P ; note that P is an ideal subgroup of 𝐺 and that G = ( A ) × is the unit group of the subalgebra A = D J of 𝒜, where 𝒟 denotes the diagonal subalgebra of 𝒜 such that T = D × . Since H T ( 1 + J ( B ) ) G , it is obvious that V = c - Ind H G ( W ) is an irreducible smooth G -module; indeed, by the transitivity of compact induction, we see that V c - Ind G G ( V ) . Since A is a proper subalgebra of 𝒜, we know by induction that V is admissible.

If V is one-dimensional, then it follows from Step 2 that 𝑉 is admissible; thus we may assume that dim V 2 . In this situation, we repeat step by step the proof of Theorem 1 (but applied to the sequence of ideals J J 2 J 3 and to the commutators [ 1 + J , 1 + J m ] for m N ) to construct an ideal subgroup Q = 1 + L , where L J is an ideal of 𝒜 satisfying J n L J n - 1 and dim ( L / J n ) = 1 for a suitable integer n 2 and such that

V c - Ind G ϑ G ( V ϑ )

for some smooth character ϑ Q . Indeed, the proof of Theorem 1 shows that

V c - Ind G ϑ G ( V ϑ ) ;

we claim that V ϑ = c - Ind G ϑ G ϑ ( V ϑ ) . On the one hand, we note that there are isomorphisms

V c - Ind G G ( V ) c - Ind G G ( c - Ind G ϑ G ( V ϑ ) ) c - Ind G ϑ G ( V ϑ ) c - Ind G ϑ G ( c - Ind G ϑ G ϑ ( V ϑ ) ) ,

and hence c - Ind G ϑ G ϑ ( V ϑ ) is an irreducible smooth G ϑ -module. On the other hand, since x ϕ = ϑ ( x ) ϕ for all x Q and all ϕ c - Ind G ϑ G ϑ ( V ϑ ) , it follows from [14, Corollaire 2 au Théorème 3] that

c - Ind G ϑ G ϑ ( V ϑ ) V ϑ ,

which is precisely what we claimed.

Now, we know from Proposition 4 that G ϑ is the unit group of some subalgebra A ϑ of A ; moreover, by Theorem 1, we know that there is a subalgebra B of A ϑ such that

V ϑ c - Ind H G ϑ ( W ) ,

where H = ( B ) × is the unit group of B , and W is a one-dimensional smooth H -module. (Notice that H G ϑ ; however, in the general situation, we are not assuming that H G ϑ .) Since V is admissible, the smooth G ϑ -module V ϑ is also admissible, and hence H contains a diagonal subgroup T of 𝐺 (by Step 1). Since V ϑ c - Ind H G ϑ ( W ) (by the transitivity of c-induction), we conclude that the smooth G ϑ -module V ϑ is admissible (by induction, because G ϑ is the unit group of a proper subalgebra of 𝒜). Finally, we recall from Proposition 3 that ς = ϑ 1 + J n is a 𝑃-invariant smooth character of 1 + J n and that the 𝑃-orbit ϑ P Q of 𝜗 consists of all ϑ Q which satisfy ( ϑ ) 1 + J n = ς ; in particular, ϑ P is a closed subset of Q . Since T H G ϑ and since G = T P (because T is a diagonal subgroup of 𝐺), it follows that ϑ G = ϑ P is a closed subset of Q , and thus we conclude that the smooth 𝐺-module V c - Ind G ϑ G ( V ϑ ) is admissible (by [14, Théorème 4]), as required. ∎

Finally, we prove that (i) implies (iii).

Step 4

Assume that 𝐻 contains a diagonal subgroup 𝑇 of 𝐺. Then there is an isomorphism of smooth 𝐺-modules V Ind H G ( W ) .

Proof

Let δ G : G R + × and δ H : H R + × be the modular characters of 𝐺 and 𝐻, respectively; we claim that ( δ G ) H = δ H . To see this, we observe that there is a chain of closed subgroups H = H 0 H 1 H n = G such that H i - 1 is normal in H i for all 1 i n ; for each 1 i n , let δ i denote the modular character of H i . By [6, Corollary 1.5.5 (a)], we have ( δ i ) H i - 1 = δ i - 1 for all 1 i n ; in particular, we conclude that ( δ G ) H = δ H , as claimed. It follows that the map δ G / H = ( δ H ) - 1 ( δ G ) H is identically equal to 1, and thus [4, Theorem 3.5] implies that

V ( c - Ind H G ( W ) ) Ind H G ( W ) .

On the other hand, since T H , the smooth 𝐺-module 𝑉 is admissible by Step 3, and thus its smooth dual V is also irreducible by [4, Proposition 2.10]. Since c - Ind H G ( W ) is a submodule of Ind H G ( W ) , we conclude that V c - Ind H G ( W ) , and thus

( V ) ( c - Ind H G ( W ) ) Ind H G ( ( W ) )

(again by [4, Theorem 3.5]). Since ( W ) W (because 𝑊 is one-dimensional) and since ( V ) V (by [4, Proposition 2.9] because 𝑉 is admissible), we conclude that V Ind H G ( W ) , as required. ∎

The proof of Theorem 2 is now complete. ∎

We finish the paper with the following result concerning the unitarisability of an arbitrary irreducible smooth 𝐺-module.

Theorem 3

Let 𝑉 be an irreducible smooth 𝐺-module. Then there is an isomorphism of 𝐺-modules V V V ′′ , where V is a unitarisable irreducible smooth 𝐺-module and V ′′ is a one-dimensional smooth 𝐺-module and where the 𝐺-action on V V ′′ is given by g ( v v ′′ ) = ( g v ) ( g v ′′ ) for all g G , all v V and all v ′′ V ′′ .

Proof

Let ℬ be a subalgebra of 𝒜 such that V c - Ind H G ( W ) , where H = B × is the unit group of ℬ and 𝑊 is a one-dimensional smooth 𝐻-module; moreover, let τ H be the smooth character of 𝐻 afforded by 𝑊. Since ℬ is a split basic 𝕜-algebra, the group 𝐻 decomposes as the semidirect product H = S Q , where 𝑆 is the unit group of a diagonal algebra of ℬ and where 𝑄 is the ideal subgroup of 𝐻 which corresponds to the Jacobson radical of ℬ; note that 𝑄 is a normal subgroup of 𝐻 and that 𝑆 is isomorphic to a finite direct product of copies of 𝕜 × .

Let τ H be defined by τ ( s x ) = τ - 1 ( s ) for all s S and all x Q ; we note that

( τ τ ) ( s x ) = ( τ - 1 τ S ) ( s ) τ ( x ) = τ ( x ) , s S , x Q ,

and thus τ τ is a unitary character of 𝐻 (because 𝑄 is an c -group, and hence τ Q is a unitary character of 𝑄; see [3, Lemma 4.9]). Let W C τ be a one-dimensional smooth 𝐻-module which affords the character τ , and consider the tensor product W W . We note that W W is one-dimensional and affords the unitary character τ τ H ; hence the smooth 𝐻-module W W is unitarisable. We define V to be the c-induced smooth 𝐺-module V = c - Ind H G ( W W ) and claim that V is unitarisable. To see this, as in the proof of Step 4 above, we see that ( δ G ) H = δ H , where δ G and δ H are the modular characters of 𝐺 and 𝐻, respectively, and thus it follows from [6, Theorem 1.5.3] that there exists a non-zero Radon measure 𝜇 on the quotient space G / H which is invariant for the action of 𝐺. We denote by , the 𝐺-invariant positive definite Hermitian inner product on W W and define

ϕ , ψ = G / H ϕ ( g ) , ψ ( g ) d μ , ϕ , ψ c - Ind H G ( W W ) ;

note that, if X G is a complete set of representatives of the cosets of G / H , then every function ϕ c - Ind H G ( W W ) is completely determined by its values in 𝑋 and hence defines a function on G / H which clearly has compact support (note also that the 𝐺-invariance of the inner product on W W ensures that the values of this function do not depend on the particular choice of the set 𝑋). It is straightforward to check that the formula above defines a positive definite Hermitian inner product on c - Ind H G ( W W ) which is 𝐺-invariant (because the measure 𝜇 is 𝐺-invariant). Therefore, we conclude that V c - Ind H G ( W W ) is unitarisable, as claimed.

Finally, as in the proof of Theorem 2 (Step 1), we see that there exists a subgroup T of 𝑇 such that 𝑇 decomposes as a direct product T = S T , and thus the smooth character σ = ( τ ) S = ( τ - 1 ) S S can be extended to a smooth character σ T of 𝑇; moreover, since 𝐺 is the semidirect product G = T P , where P = 1 + J ( A ) , there is a smooth character ϑ G such that ϑ ( t x ) = σ ( t ) for all t T and all x P . Let W ′′ C ϑ be a one-dimensional smooth 𝐺-module which affords the character 𝜗, and consider the tensor product W ′′ c - Ind H G ( W ) . For every w ′′ W ′′ and every ϕ c - Ind H G ( W ) , we define the function

ψ ( w ′′ ϕ ) : G W ′′ W

by the rule

ψ ( w ′′ ϕ ) ( g ) = ( g w ′′ ) ϕ ( g ) ) = ϑ ( g ) ( w ′′ ϕ ( g ) ) , g G ;

it is easy to check that the mapping w ′′ ϕ ψ ( w ′′ ϕ ) defines an isomorphism of 𝐺-modules

W ′′ c - Ind H G ( W ) c - Ind H G ( Res H G ( W ′′ ) W ) .

Since we clearly have Res H G ( W ′′ ) W (as 𝐻-modules), we conclude that

W ′′ c - Ind H G ( W ) c - Ind H G ( W W ) ,

and thus

V c - Ind H G ( W ) V ′′ c - Ind H G ( W W ) V V ′′ ,

where V ′′ = ( W ′′ ) C ϑ - 1 . The result follows. ∎

Award Identifier / Grant number: UID/MAT/04721/2013

Funding statement: This research was made within the activities of the Group for Linear, Algebraic and Combinatorial Structures of the Center for Functional Analysis, Linear Structures and Applications (University of Lisbon, Portugal), and was partially supported by the Portuguese Science Foundation (FCT) through the Strategic Project UID/MAT/04721/2013. The second author was partially supported by the Lisbon Mathematics PhD program (funded by the Portuguese Science Foundation). This work is part of the second author’s PhD thesis.

  1. Communicated by: Robert Guralnick

References

[1] I. N. Bernšteĭn and A. V. Zelevinskiĭ, Representations of the group GL ( n , F ) , where 𝐹 is a local non-Archimedean field, Uspehi Mat. Nauk 31 (1976), no. 3 (189), 5–70. 10.1070/RM1976v031n03ABEH001532Search in Google Scholar

[2] A. Borel, Linear Algebraic Groups, 2nd ed., Grad. Texts Math. 126, Springer, New York, 1991. 10.1007/978-1-4612-0941-6Search in Google Scholar

[3] M. Boyarchenko, Representations of unipotent groups over local fields and Gutkin’s conjecture, Math. Res. Lett. 18 (2011), no. 3, 539–557. 10.4310/MRL.2011.v18.n3.a14Search in Google Scholar

[4] C. J. Bushnell and G. Henniart, The Local Langlands Conjecture for GL ( 2 ) , Grundlehren Math. Wiss. 335, Springer, Berlin, 2006. 10.1007/3-540-31511-XSearch in Google Scholar

[5] P. Cartier, Representations of 𝑝-adic groups: A survey, Automorphic Forms, Representations and 𝐿-Functions, Proc. Sympos. Pure Math. 33, American Mathematical Society, Providence (1979), 111–155. 10.1090/pspum/033.1/546593Search in Google Scholar

[6] A. Deitmar and S. Echterhoff, Principles of Harmonic Analysis, 2nd ed., Universitext, Springer, Cham, 2014. 10.1007/978-3-319-05792-7Search in Google Scholar

[7] G. B. Folland, A Course in Abstract Harmonic Analysis, 2nd ed., Textb. Math., CRC Press, Boca Raton, 2016. 10.1201/b19172Search in Google Scholar

[8] E. A. Gutkin, Representations of algebraic unipotent groups over a selfdual field, Funct. Anal. Appl. 7 (1974), 322–323. 10.1007/BF01075739Search in Google Scholar

[9] Z. Halasi, On the characters and commutators of finite algebra groups, J. Algebra 275 (2004), no. 2, 481–487. 10.1016/j.jalgebra.2004.01.021Search in Google Scholar

[10] Z. Halasi, On the characters of the unit group of DN-algebras, J. Algebra 302 (2006), no. 2, 678–685. 10.1016/j.jalgebra.2006.02.036Search in Google Scholar

[11] I. M. Isaacs, Characters of groups associated with finite algebras, J. Algebra 177 (1995), no. 3, 708–730. 10.1006/jabr.1995.1325Search in Google Scholar

[12] H. Jacquet, Sur les représentations des groupes réductifs 𝑝-adiques, C. R. Acad. Sci. Paris Sér. A-B 280 (1975), 1271–1272. Search in Google Scholar

[13] H. Klüver, The unitary character group of abelian unipotent groups, Münster J. Math. 1 (2008), 181–219. Search in Google Scholar

[14] F. Rodier, Décomposition spectrale des représentations lisses, Non-Commutative Harmonic Analysis, Lecture Notes in Math. 587, Springer, Berlin (1977), 177–195. 10.1007/BFb0087921Search in Google Scholar

[15] B. Szegedy, On the characters of the group of upper-triangular matrices, J. Algebra 186 (1996), no. 1, 113–119. 10.1006/jabr.1996.0365Search in Google Scholar

Received: 2019-06-19
Revised: 2020-01-22
Published Online: 2021-05-26
Published in Print: 2021-11-01

© 2021 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 16.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/jgth-2019-0084/html
Scroll to top button