960
Views
2
CrossRef citations to date
0
Altmetric
Research Article

Principal series for general linear groups over finite commutative rings

, ORCID Icon &
Pages 4857-4868 | Received 28 Aug 2020, Accepted 12 May 2021, Published online: 05 Jun 2021

Abstract

We construct, for any finite commutative ring R, a family of representations of the general linear group GLn(R) whose intertwining properties mirror those of the principal series for GLn over a finite field.

2020 Mathematics Subject Classification:

1. Introduction

Among the irreducible, complex representations of reductive groups over finite fields, the simplest to construct and to classify are the principal series: those obtained by Harish-Chandra induction from a minimal Levi subgroup; see, for instance, [Citation13]. In this paper we use a generalization of Harish-Chandra induction to construct a “principal series” of representations of the group GLn(R), where R is any finite commutative ring with identity. Our main results assert that the well-known intertwining relations among the principal series for GLn over a finite field also hold for the representations that we construct.

The study of the principal series for reductive groups over finite fields can be viewed as the first step in the program to understand all irreducible complex representations of such groups in terms of what Harish-Chandra called the ‘philosophy of cusp forms’ [Citation10, Citation20]. This program has met with considerable success. The basic ideas appear already in Green’s determination [Citation8] of the irreducible characters of GLn(k), where k is a finite field, and these ideas have since been developed and generalized to a very great extent; see [Citation7] for an overview.

The theory for groups over finite rings is in a far less advanced state. Most efforts so far have been directed toward groups over principal ideal rings: see for instance [Citation21] and references therein. By contrast, the results presented below are valid for all finite rings, with the essential jump in generality being from principal ideal rings to local rings. Moreover, our results depend on the algebraic properties of the base ring in only a very limited way: for instance, we give a uniform construction of a family of irreducible representations of GLn(R) for all finite local rings R, and to our knowledge these are the first results obtained in this degree of generality.

The present paper is part of a project whose aim is to extend the philosophy of cusp forms to reductive groups over finite rings. Our construction, which is a special case of a general induction procedure developed in [Citation3], extends in a natural way to produce more general ‘Harish-Chandra series’. The analysis of the intertwining properties of these more general series seems, however, to be substantially more involved than the results for the principal series presented here. See [Citation3, Section 5] and [Citation4] for some partial results in this more general setting.

1.1. Notation and definitions

Let R be a finite commutative ring with 1. Let G=GLn(R), let L(R×)n be the subgroup of diagonal matrices in G, and let U and V be the upper-unipotent subgroup and the lower-unipotent subgroup, respectively, in G. Let B = LU be the subgroup of upper-triangular matrices. We write G(R), L(R), etc., when it is necessary to specify R.

The ring R decomposes as a direct product of local rings: RR1××Rm, and this decomposition is unique up to permuting the factors [Citation17, Theorem VI.2]. There is a corresponding decomposition G(R)G(R1)××G(Rm), and similarly for L, U, and V. If R is a local ring then we let N(R) be the subgroup of monomial matrices in G(R), that is, products of permutation matrices with diagonal matrices. If R is not local then we define N(R)=N(R1)××N(Rm), where the Ri are the local factors of R as above. Let W(R)=N(R)/L(R). It will be convenient to realize W(R) as a subgroup of G(R), as follows: if R is local, then we identify W(R) with the group of permutation matrices; and in the general case we identify W(R) with the product of the permutation subgroups in G(R)G(R1)××G(Rm). Note that following Lemma 4, we will be able to assume without loss of generality that R is a local ring.

If χ:LGL(X) is a representation of L on a complex vector space X, and if wW, then we let w*χ denote the representation χ°Adw1:LGL(X). We let Wχ={wW|w*χχ}.

For each subgroup HG we let eH denote the idempotent in the complex group ring C[G] corresponding to the trivial character of H: eH=|H|1hHh. Since L normalizes U and V, the idempotents eU and eV commute with C[L] inside C[G].

We consider the functors i:Rep(L)Rep(G)XC[G]eUeVC[L]Xr:Rep(G)Rep(L)YeUeVC[G]C[G]Y, where Rep(G) denotes the category of complex representations, identified in the usual way with the category of left C[G]-modules. This is a special case of the construction defined in [Citation3, Section 2], which generalizes a definition due to Dat [Citation6]. The functors i and r are two-sided adjoints to one another; see [Citation3, Theorem 2.15] for a proof of this and other basic properties.

Definition.

Let us say that an irreducible representation of G is in the principal series if it is isomorphic to a subrepresentation of iχ for some representation χ of L.

Example.

For each representation χ:LGL(X) of L, the representation iχ=C[G]eUeVC[L]X of G is a nonzero quotient of the representation C[G]eUC[L]X, the latter being the representation of G obtained by first extending χ from L to LU by letting U act trivially on X, and then inducing from LU to G. If this representation C[G]eUC[L]X is irreducible, then it must equal iχ.

If R is a field, then the map C[G]eUffeVC[G]eV is known to be an isomorphism of C[G]-C[L] bimodules; see [Citation15, Theorem 2.4]. It follows that in this case the functors i and r are naturally isomorphic to the familiar functors of Harish-Chandra induction and restriction, i.e., the functors of tensor product with the bimodules C[G]eU and eUC[G], respectively. The same is not true if R is not a product of fields, as the following example illustrates.

Example.

Let 1L denote the trivial representation of L. Then we have C[G]eUC[L]1LC[G/LU], with G acting by permutations of G/LU; and likewise C[G]eVC[L]1LC[G/LV]. Let w0G be the permutation matrix that conjugates U into V, and vice versa; then the map gLVgw0LU induces a G-equivariant isomorphism C[G/LV]C[G/LU]. Making these identifications, the map (*) C[G]eUC[L]1Lf1feV1C[G]eVC[L]1L(*) becomes, up to a nonzero scalar multiple, the map C[G/LU]C[G/LU] of multiplication on the right by the characteristic function of the double coset LUw0LU. If R is a field, then the latter map is well-known to be invertible (as are all of the standard generators of the Iwahori-Hecke algebra C[LU\G/LU]; see for instance [Citation5, §67 A]).

By contrast, suppose now that R is not a field. Let m be a maximal ideal of R, and let V0 be the subgroup of V comprising those lower-unipotent matrices over R that reduce, modulo m, to the identity matrix. The product I = LUV0 is a subgroup of G (namely, the group of upper-triangular-modulo-m matrices). Since V0 is a subgroup of V we have eV=eV0eV, and so the map (*) factors through the map C[G]eUC[L]1Lf1feV01C[G]eV0C[L]1L, whose image is isomorphic to the permutation module C[G/I]. The latter has strictly smaller dimension than C[G/LU], and so (*) cannot be an isomorphism.

For general rings, the permutation module C[G/LU] can be quite complicated. For instance, for R=Z/pkZ (with p a prime and k a positive integer), the results of [Citation18] show that the intertwining algebra of this representation depends both on p and on k. By contrast, it follows from Theorem 2 below that for any R the intertwining algebra of i1L is isomorphic to the tensor product C[Sn]m, where G=GLn(R) and where m is the number of maximal ideals in R.

Example.

Suppose that R is a finite discrete valuation ring, with maximal ideal m and residue field k, and let r be the largest integer such that mr0. Reduction modulo mr gives rise to a group extension 0Gr(Mn(k),+)G(R)G(R/mr)0, which one can use to study the representations of G(R) via Clifford theory; see [Citation11], for example. In [Citation12], Hill identified a class of representations that are particularly amenable to this approach: an irreducible representation π of G(R) is called regular if its restriction to Gr contains a character whose stabilizer under the adjoint action of G(k) is an abelian group (see [Citation12, Theorem 3.6] for details and alternative characterizations of regularity). Explicit constructions of all such representations are given in [Citation16, Citation22].

An application of [Citation3, Theorem 3.4] gives the following criterion for regularity of the induced representations iχ: if χ is an irreducible representation of L(R), then iχ is regular if and only if the restriction of χ to the subgroup L(R)Grkn has trivial stabilizer under the permutation action of Sn. Moreover, the representations iχ, for χ satisfying the above condition, account for all of the regular representations associated to the split semisimple classes in Mn(k).

For n = 2, all of the principal series representations of G(R)=GL2(R) can be described in terms of regular representations, as follows. Let χ:LC× be an irreducible representation of L. If iχ is irreducible, then there is a character τ:R×C×, an integer k, and a regular representation π of G(R/mk) associated to a split semisimple class in M2(k) such that iχ is isomorphic to the representation (τ°det)π, where π is pulled back to a representation of G(R). If iχ is not irreducible, then there is a character τ:R×C× such that iχ is isomorphic to the representation (τ°det)(1GSt), where 1G is the trivial representation, and St is the Steinberg representation of G(k) pulled back to G(R).

To prove these assertions, we use the obvious isomorphism LR××R× to write χ as a product χ1χ2. The criterion for regularity given above shows that if iχ is not itself regular, then χ1 and χ2 agree on 1+mr. Supposing this to be the case, we use Lemma 14 (below) to write iχ(χ1°det)i(1χ11χ2), where the character 1χ11χ2 is trivial on LGr and is therefore pulled back from a character χ of L(R/mr). Now [Citation3, Theorem 3.4] implies that i(1χ11χ2) is the pullback to G(R) of the representation iχ of G(R/mr). If iχ is not regular then we can repeat the above procedure, as many times as necessary. In the case where iχ is not irreducible we have χ1=χ2, by Theorem 1 (below), and then Lemma 14 gives iχ(χ1°det)i1L, where i1L is the pullback to G(R) of the representation i1L(k) (by [Citation3, Theorem 3.4]). The latter representation is, as is well known, isomorphic to sum of the trivial representation and the Steinberg representation.

For n3 the relationship between the principal series and the regular representations becomes more complicated.

2. Main results

We will show that the following well-known properties of the Harish-Chandra functors are shared by the functors i and r for R an arbitrary finite commutative ring.

Theorem 1.

There is a natural isomorphism riwWw* of functors on Rep(L). Consequently, if χ and σ are irreducible representations of L, then dimC(HomG(iχ,iσ))=#{wW|w*χ=σ}.

When σ=χ, we have the following more precise statement:

Theorem 2.

For each irreducible representation χ of L one has EndG(iχ)C[Wχ] as algebras.

Theorems

1 and 2 readily imply the following combinatorial formula for the number of principal series representations. Following [Citation1], we let Pk(n) denote the number of multipartitions of n with k parts: i.e., the number of k-tuples (λ(1),,λ(k)), where each λ(i) is a partition of some non-negative integer ni, and ini=n.

Corollary 3.

If R is isomorphic to a product R1××Rm of finite local rings, and for each j we set kj=|Rj×|, then the principal series of GLn(R) contains precisely jPkj(n) distinct isomorphism classes of irreducible representations.

Remarks.

  • In the case where R is a field, Theorems 1 and 2 are essentially due to Green [Citation8]; see [Citation23] for the case χ=1L, and see [Citation19] for an exposition. Both of these results have been generalized to arbitrary Harish-Chandra series for arbitrary reductive groups: see [Citation10] and [Citation14], respectively.

  • Theorems 1 and 2 can be extended, using [Citation3, Theorem 2.15(5)], to the setting of smooth representations of the profinite groups G(O), where O is the ring of integers in a nonarchimedean local field.

  • Some of our results apply beyond the case of GLn. For instance, an analogue of Theorem 1 holds whenever G is a split classical group: indeed, such groups are easily seen to satisfy properties (a)–(f) in Proposition 5 below, and our proof of Theorem 1 relies only on those properties. We have restricted our attention here to GLn, both in order to simplify the exposition, and because that is the case in which we use these results in [Citation4].

  • On the other hand, adapting our proof of Theorem 1 to the case where L is replaced by a larger Levi subgroup does not seem to be so straightforward. For one thing, the failure of Proposition 5(d) in this more general setting greatly complicates matters.

3. Proofs

The first step in the proof of the main results is to reduce to the case of local rings.

Lemma 4.

If Theorems 1 and 2 and Corollary 3 are true for all finite commutative local rings, then they are true for all finite commutative rings.

Proof.

Let R be a finite commutative ring, and write R as a product of local rings R1××Rm. All of the groups and the representation categories in Theorems 1 and 2 and in Corollary 3 then decompose into products accordingly: G(R)G(R1)××G(Rm),Rep(G(R))Rep(G(R1))××Rep(G(Rm)), and so on. The bimodule C[G(R)]eU(R)eV(R) decomposes as the tensor product of the bimodules C[G(Rj)]eU(Rj)eV(Rj), and likewise for eU(R)eV(R)C[G(R)], so the functors i and r are compatible with the above decompositions. By definition, the group W also decomposes compatibly. Thus Theorems 1 and 2 and Corollary 3 over R follow immediately from the corresponding results over the local factors Rj. □

Assume from now on that R is a finite commutative local ring Let m denote the maximal ideal of R, and let k denote the residue field R/m. Recall that WSn is then the group of permutation matrices in G. We write for the word-length function on W with respect to the standard generating set S={(12),,(n1n)}.

The following proposition collects the group-theoretical ingredients of the proof of Theorem 1.

Proposition 5.

  1. The multiplication map U×L×VG is injective.

  2. The reduction-mod-m map G(R)G(k) is surjective.

  3. For each subgroup H of G, let H0 denote the intersection of H with the kernel G0 of the above reduction homomorphism. Then the multiplication map U0×L0×V0G0 is a bijection, and the same is true for any ordering of the three factors.

  4. For each wW the multiplication maps (UUw)×(UVw)Uand(VUw)×(VVw)V

    are bijections, where Uw=w1Uw, etc.

  5. G is the disjoint union G=wWGw, where Gw=VwLUG0.

  6. For each r,tW with (t)(r) and tr one has ULVt1Ur=.

Proof.

Parts (a), (b), and (d) are well-known and easily verified.

For part (c), the map U0×L0×V0G0 is injective by part (a). Now the ideal m is nilpotent, so every matrix of the form 1+x with xMn(m) is invertible, and thus G0={1+x|xMn(m)}, while L0, U0, and V0 are the subgroups in which x is, respectively, diagonal, strictly upper-triangular, or strictly lower-triangular. Counting matrix entries then shows that the finite sets U0×L0×V0 and G0 have equal cardinality, and so the injective multiplication map is bijective.

Part (e) follows immediately from the Bruhat decomposition of G(k) [Citation5, (65.4)].

In part (f) we may assume without loss of generality that R is a field, since ULVt1Ur is empty if its reduction modulo m is empty. This assumption implies that (B,N,W,S) is a BN-pair in G, where we are writing B for the upper-triangular subgroup LU of G; see, e.g., [Citation5, (65.10)]. Let w0 denote the longest element (1,2,,n)(n,,2,1) of W. It follows from [Citation2, Ch. IV §2 Lemme 1] that, under the stated assumptions on t and r, we have tBw0BBrw0B=. Since ULVw0=ULw0U=Bw0B, while t1Urw0t1Brw0B, we conclude that ULVt1Ur=.

We equip C[G] with the Hermitian inner product | for which the group elements gG constitute an orthonormal basis; and with the conjugate-linear involution * defined on basis elements by g*=g1. The two structures are related by the identity abc|d=b|a*dc* for all a,b,c,dC[G]. An element aC[G] is called self-adjoint if a=a*.

Lemma 6.

There is a self-adjoint, invertible element zC[G] that commutes with eU, eV, and C[L], and that satisfies z(eUeV)2=eUeV and z(eVeU)2=eVeU.

Proof.

This follows from a general fact about pairs of orthogonal projections on a finite-dimensional Hilbert space: see [Citation9, Theorem 2], for example. □

Remark.

If R is a field then [Citation15, Theorem 2.4] implies that there is a unique element z as in Lemma 6. This is not the case over a general ring.

Lemma 7.

For each wW we have eVeUweVw=eVeUeVw.

Proof.

It is clear that eV=eVe(VUw) and similarly that eU=e(UUw)eU. Proposition 5(d) gives e(VUw)e(UUw)=eUw, and it follows that eVeU=eVeUweU. The same reasoning gives eUweVw=eUweUeVw, and so eVeUeVw=eVeUweUeVw=eVeUweVw.

Lemma 8.

For each wW the map φw:eUweVwC[G]xeVxeVeUC[G] is an isomorphism of C[L]-C[G] bimodules.

Proof.

The following argument is taken from [Citation6, Lemme 2.9]. The map φw is well-defined, because eVeUweVwC[G]=eVeUeVwC[G]eVeUC[G] by Lemma 7. The map φw is injective, because for each fC[G] we have w1zweUweVw(eVeUweVwf)=zw(eUweVw)2f=eUweVwf where z is as in Lemma 6, and in the first equality we used that V=(VVw)(VUw). The domain and target of φw are isomorphic as vector spaces: indeed, eVeUC[G]=w0weUweVwC[G], where w0 is the longest element of W. Since φw is injective it is thus also an isomorphism. □

For each subset KG, we let C[K] denote the vector subspace of C[G] spanned by K.

Proposition 9.

For each wW the map Φ:C[wL]eUeVC[Gw]eUeVwleUeVwleUeV is an isomorphism of C[L]-bimodules.

Here the sets wL and Gw are invariant under multiplication by L, on either side, and we are using these multiplication actions to view C[wL] and C[Gw] as C[L]-bimodules.

Proof.

Φ is clearly a bimodule map. Let us show that it is injective. For hC[L] we have Φ(wh)=eUeVeUw1eVw1wh.

The maps eUw1eVw1C[G]xeVxeVeUC[G] and eVeUC[G]xeUxeUeVC[G] are isomorphisms by Lemma 8, so we are left to prove that the map wheUw1eVw1wh=weUheV is injective on C[wL]. It is, because Proposition 5(a) implies that the cosets UlV are all disjoint as l ranges over L. Thus Φ is injective.

To prove that Φ is surjective, first note that Gw=VwLG0U because G0 is normal in G. Since eVv=eV and ueU = eU for all vV and uU, we find that eUeVC[Gw]eUeV is spanned by elements of the form eUeVwlgeUeV, where lL and gG0. We will show that each element of this form is in the image of Φ.

For each xVw we have gx=x(x1gx)VwG0=Vw(V0wL0U0w)=VwL0U0w by Proposition 5(c). Let α:VwVw,β:VwL0 and γ:VwU0w be the (unique) functions satisfying gx=α(x)β(x)γ(x) for all xVw. Writing eU=eUVweUUw and eV=eVUweVVw, we then have eVwlgeUeV=eVwlgeUVweUUweVUweVVw=eVwlg(|UVw|1xUVwx)eUweVVw=|UVw|1xUVweVwlα(x)β(x)γ(x)eUweVVw.

Since γ(x)Uw we have γ(x)eUw=eUw for each xUVw. Since α(x)Vw we have wlα(x)l1w1V, and consequently eVwlα(x)=eVwl for each x. Continuing the computation with the space-saving notation h=|UVw|1xUVwlβ(x)C[L], we find that eVwlgeUeV=eVwheUweVVw=eVeUw1VwheUweVVw=eVwheUVweUUweVUweVVw=eVwheUeV, and so eUeVwlveUeV=Φ(wh).

Proposition 10.

The set {eUeVwleUeVC[G]|wW,lL} is linearly independent.

Proof.

We know from Proposition 9 that for each wW the set {eUeVwleUeV|lL} is linearly independent. We must show that for different choices of w these sets are independent from one another.

Suppose we had elements hwC[L], not all zero, with wWeUeVwhweUeV=0. Let tW be an element of minimal length such that ht is nonzero. To compactify the notation we shall write y=t1.

Let z be as in Lemma 6, and write ζ=y1zy. Thus ζ is a self-adjoint, invertible element of C[G] which commutes with eUy and eVy and which satisfies ζ(eUyeVy)2=eUyeVy. For each rW with rt such that hr0 we have ζ2eUy(eUeVthteUeV)|eUeVrhreUeV=ζ2eUyeUeVeUyeVytht|rhreUreVreUeV=ζ2eUyeUeVeUyeVythteVeUeVreUrhr*|r=ζ2eUyeUeVeUyeVyeUyeVryeUryththr*|r=ζ2eUyeUeVyeUyeVyeUyeVyeUryththr*|r=ζ2(eUyeVy)3eUryththr*|r=eUyeVyeUryththr*|r=teUhthr*eVeUr|r=eUhthr*eV|t1eUr=0. 

Here we have repeatedly used the equality abc|d=b|a*dc*; in the fourth step we used Lemma 7 to replace eUeVeUy with eUeVyeUy and to replace eUyeVryeUry with eUyeVyeUry; in the fifth step we used Proposition 5(d) to write eUyeUeVy=eUyeVy; and in the final equality we used Proposition 5(f), which applies because of the minimality of (t), and which implies that the functions eUhthr*eV and t1eUr are supported on disjoint subsets of G and are therefore orthogonal.

It follows from this that 0=ζ2eUyeUeVthteUeVwWeUeVwhweUeV=ζ2eUyeUeVthteUeVeUeVthteUeV=ζeUyeUeVthteUeVζeUyeUeVthteUeV, \ where the last equality holds because ζ is self-adjoint, eUy is a self-adjoint idempotent, and ζ and eUy commute. Thus ζeUyeUeVthteUeV=0. Since ζ is invertible, and left multiplication by eUy is injective on eUeVC[G] (Lemma 8), we conclude that eUeVthteUeV=0. By Proposition 9 this implies that ht = 0, contradicting our choice of t and completing the proof of the proposition. □

Proof of Theorem 1.

The functor ri is naturally isomorphic to the functor of tensor product (over C[L]) with the C[L]-bimodule eUeVC[G]eUeV, while the functor wWw* is naturally isomorphic to the tensor product with the bimodule C[WL]. Since G=Gw we have eUeVC[G]eUeV=wWeUeVC[Gw]eUeV.

Proposition 9

thus implies that the C[L]-bimodule map C[WL]=wWC[wL]heUeVheUeVeUeVC[G]eUeV is surjective. Proposition 10 implies that this map is injective, so it is an isomorphism of bimodules, and induces a natural isomorphism of functors riw*. The formula for the intertwining number follows from this isomorphism and from the fact that i and r are adjoints. □

We now turn to the proof of Theorem 2. Every irreducible representation χ of the abelian group L(R×)n has the form χ1χn:diag(r1,,rn)χ1(r1)χn(rn) where each χi is a linear character R×C×. For each such χ we let eχ=|L|1lLχ(l)1l be the corresponding primitive central idempotent in C[L].

Lemma 11.

The algebra EndG(iχ) is isomorphic to the subalgebra eχeUeVC[G]eUeVeχ of C[G].

Proof.

We have iχC[G]eUeVC[L]C[L]eχC[G]eUeVeχ=C[G]zeUeVeχ where z is as in Lemma 6. Since zeUeV and eχ are commuting idempotents in C[G], their product E=zeUeVeχ is an idempotent and we have EndG(C[G]E)(EC[G]E)opp via the action of EC[G]E on C[G]E by right multiplication. Now EC[G]E is a finite-dimensional complex semisimple algebra, so it is isomorphic to its opposite, and we have EC[G]E=eχeUeVC[G]eUeVeχ.

Lemma 12.

For the trivial representation 1L of L we have EndG(i1L)C[W] as algebras.

Proof.

First suppose that R is a field, so that the functor i is isomorphic to the functor of Harish-Chandra induction. Then, as we noted above, i1L is isomorphic to the permutation representation on C[G/LU], and the isomorphism EndG(i1L)C[W] is a special case of well-known results of Iwahori-Matsumoto and Tits (see [Citation5, §68] for an exposition).

Now let R be a local ring with residue field k. The quotient map Rk induces a surjective map of algebras (13) eL(R)eU(R)eV(R)C[G(R)]eU(R)eV(R)eL(R)eL(k)eU(k)eV(k)C[G(k)]eU(k)eV(k)eL(k).(13)

Theorem 1 implies that the domain of (13) is isomorphic as a vector space to C[W], while we have just seen that the range of (13) is isomorphic as an algebra to C[W]. Since (13) is surjective, it is an algebra isomorphism. □

Remark.

The isomorphism in Lemma 12 is not canonical. One can trace through the various maps appearing in the proof to construct a set of Iwahori-Hecke generators of eLeUeVC[G]eUeVeL, although this will depend on the choice of an element z as in Lemma 6.

Lemma 14.

Let χ=χ1χn be an irreducible representation of L, let τ:R×C× be a character of R×, and let χ=τχ1τχn. Then iχ(τ°det)iχ.

Proof.

The algebra automorphism (15) C[G]C[G],gτ(detg)g(15) sends eχ to eχ, and fixes eU and eV . Thus (15) induces an isomorphism of C[G]-modules C[G]eUeVC[L]Cχ(τ°det)C[G]eUeVC[L]Cχ.

Lemma 16.

If χ=χ1n is a tensor-multiple of a single character of R×, then EndG(iχ)EndG(i1L) as algebras.

Proof.

Lemma 14 ensures that iχ(χ1°det)i1L.

Lemma 17.

For each wW there is a natural isomorphism of functors i°w*i.

Proof.

The functor i°w* is given by tensor product with the C[G]-C[L] bimodule C[G]eUeVw=C[G]eUweVw, while the functor i is given by tensor product with C[G]eUeV. These two bimodules are isomorphic, by Lemma 8. □

Lemma 17

implies that in order to compute the intertwining algebra EndG(iχ) for an arbitrary character χ of L we may permute the factors χi so that χ takes the form (18) χ=χ1n1χ2n2χknkwhereχjχi unless j=i.(18)

(The exponents indicate tensor powers.) We then have WχSn1××Snk.

In the next lemma we shall consider general linear groups of different sizes, and we shall accordingly embellish the notation with subscripts to indicate the size of the matrices involved: so, for example, La denotes the diagonal subgroup in Ga=GLa(R), and ia is a functor from Rep(La) to Rep(Ga).

Lemma 19.

If χ is as in (18) then EndGn(inχ)j=1kEndGnj(inj(χjnj)) as algebras.

Proof.

Let us write L for the block-diagonal subgroup Gn1××GnkGn, which contains as subgroups the groups U=Un1××Unk and V=Vn1××Vnk. Let U be the subgroup of block-upper-unipotent matrices U={[1n1×n101nk×nk]Gn}, and let V=(U)t be the corresponding group of block-lower-unipotent matrices.

Let i:Rep(L)Rep(Gn) be the functor of tensor product with the C[Gn]-C[L] bimodule C[Gn]eUeV. The semidirect product decompositions U=Un=UU and V=Vn=VV give equalities eU=eUeU and eV=eVeV, and hence an isomorphism of C[Gn]-C[Ln] bimodules C[Gn]eUneVnC[Gn]eUeVC[L]C[L]eUeV.

It follows that inχi(j=1kinj(χjnj)).

Since i is a functor, we obtain from this isomorphism a map of algebras (20) i:j=1kEndGnj(inj(χnnj))EndGn(inχ).(20)

Now, the C[L]-bimodule map C[L]C[Gn]eUeV,hheUeV is injective, because the multiplication map L×U×VGn is one-to-one. It follows from this that the identity functor on Rep(L) is a subfunctor of ResLGn°i. Thus i is a faithful functor, and in particular the map (20) is injective. Since the domain and the range of this map have the same dimension as complex vector spaces, by Theorem 1, we conclude that (20) is an algebra isomorphism. □

Proof of Theorem 2.

Lemma 17 allows us to assume that χ has the form (18), and in this case we have algebra isomorphisms EndG(iχ)Lem. 19jEndGnj(inj(χjnj))Lem. 16jEndGnj(inj1Lnj)Lem. 12jC[Snj]C[Wχ]. 

Proof of Corollary 3.

Choose an ordering {χ1,,χk} of the character group R×̂. Lemma 17 and the intertwining number formula in Theorem 1 imply that for each principal series representation π of GLn(R) there is a unique k-tuple of non-negative integers n1,,nk having ini=n, such that π embeds in i(χ1n1χknk).

Theorem 2 implies that the number of distinct irreducible subrepresentations of i(χ1n1χknk) is equal to the number of distinct irreducible representations of Sn1××Snk. The latter number is equal to the number of k-tuples (λ(1),,λ(k)), where each λ(i) is a partition of ni. Allowing the exponents ni to vary shows that the total number of principal series representations is equal to Pk(n), as claimed. □

Additional information

Funding

The first and second authors were partly supported by the Danish National Research Foundation through the Centre for Symmetry and Deformation (DNRF92). The first author was also supported by fellowships from the Max Planck Institute for Mathematics in Bonn, and from the Radboud Excellence Initiative at Radboud University Nijmegen. The second author was also supported by the Research Training Group 1670 “Mathematics Inspired by String Theory and Quantum Field Theory.” The third author acknowledges the support of the Israel Science Foundation [grant number 1862/16] and of the Australian Research Council [grant number FT160100018].

References

  • Andrews, G. E. (2008). A survey of multipartitions: congruences and identities. In: Alladi, K., ed. Surveys in Number Theory, Volume 17 of Developments in Mathematics. New York: Springer, pp. 1–19.
  • Bourbaki, N. (1968). Éléments de athématique. Fasc. XXXIV. Groupes et lgèbres de Lie. Chapitre IV: Groupes de Coxeter et ystèmes de Tits. Chapitre V: Groupes ngendrés ar es éflexions. Chapitre VI: systèmes de acines. Actualités Scientifiques et Industrielles, No. 1337. Paris: Hermann.
  • Crisp, T., Meir, E., Onn, U. (2019). A variant of Harish-Chandra functors. J. Inst. Math. Jussieu. 18(5):993–1049. DOI: 10.1017/S1474748017000305.
  • Crisp, T., Meir, E., Onn, U. (2020). An inductive approach to representations of general linear groups over compact discrete valuation rings. arXiv:2005.05553.
  • Curtis, C. W., Reiner, I. (1987). Methods of Representation Theory. Vol. II. Pure and Applied Mathematics (New York). New York: John Wiley & Sons, Inc.
  • Dat, J.-F. (2009). Finitude pour les représentations lisses de groupes p-adiques. J. Inst. Math. Jussieu. 8(2):261–333. DOI: 10.1017/S1474748008000054.
  • Digne, F., Michel, J. (1991). Representations of Finite Groups of Lie Type, Volume 21 of London Mathematical Society Student Texts. Cambridge: Cambridge University Press.
  • Green, J. A. (1955). The characters of the finite general linear groups. Trans. Amer. Math. Soc. 80(2):402–447. DOI: 10.1090/S0002-9947-1955-0072878-2.
  • Halmos, P. R. (1969). Two subspaces. Trans. Amer. Math. Soc. 144:381–389. DOI: 10.1090/S0002-9947-1969-0251519-5.
  • Harish-Chandra. (1970). Eisenstein series over finite fields. In: Browder, F. E., ed. Functional Analysis and Related Fields (Proc. Conf. M. Stone, Univ. Chicago, Chicago, Ill., 1968) New York: Springer, pp. 76–88.
  • Hill, G. (1993). A Jordan decomposition of representations for GLn(O). Commun. Algebra 21(10):3529–3543.
  • Hill, G. (1995). Regular elements and regular characters of GLn(O). J. Algebra 174(2):610–635.
  • Howlett, R. B., Kilmoyer, R. W. (1980). Principal series representations of finite groups with split BN pairs. Commun. Algebra 8(6):543–583. DOI: 10.1080/00927878008822475.
  • Howlett, R. B., Lehrer, G. I. (1980). Induced cuspidal representations and generalised Hecke rings. Invent. Math. 58(1):37–64. DOI: 10.1007/BF01402273.
  • Howlett, R. B., Lehrer, G. I. (1994). On Harish-Chandra induction and restriction for modules of Levi subgroups. J. Algebra 165(1):172–183. DOI: 10.1006/jabr.1994.1104.
  • Krakovski, R., Onn, U., Singla, P. (2018). Regular characters of groups of type An over discrete valuation rings. J. Algebra 496:116–137. DOI: 10.1016/j.jalgebra.2017.10.018.
  • McDonald, B. R. (1974). Finite Rings with Identity. Pure and Applied Mathematics, Vol. 28. New York: Marcel Dekker Inc.
  • Onn, U., Prasad, A., Vaserstein, L. (2006). A note on Bruhat decomposition of GL(n) over local principal ideal rings. Commun. Algebra 34(11):4119–4130.
  • Springer, T. A. (1970). Characters of special groups. In: Seminar on Algebraic Groups and Related Finite Groups (the Institute for Advanced Study, Princeton, N.J., 1968/69), Lecture Notes in Mathematics, Vol. 131. Berlin: Springer, pp. 121–166.
  • Springer, T. A. (1970). Cusp forms for finite groups. In: Seminar on Algebraic Groups and Related Finite Groups (the Institute for Advanced Study, Princeton, N.J., 1968/69), Lecture Notes in Mathematics, Vol. 131. Berlin: Springer, pp. 97–120.
  • Stasinski, A. (2017). Representations of GLN over finite local principal ideal rings: an overview. In: Around Langlands Correspondences, Volume 691 of Contemp. Math. Providence, RI: Amer. Math. Soc., pp. 337–358.
  • Stasinski, A., Stevens, S. (2017). The regular representations of GLN over finite local principal ideal rings. Bull. Lond. Math. Soc. 49(6):1066–1084.
  • Steinberg, R. (1951). A geometric approach to the representations of the full linear group over a Galois field. Trans. Amer. Math. Soc. 71(2):274–282. DOI: 10.1090/S0002-9947-1951-0043784-0.