1,065
Views
2
CrossRef citations to date
0
Altmetric
Original Articles

Frobenius reciprocity for topological groups

Pages 2102-2117 | Received 08 Jan 2018, Accepted 16 Sep 2018, Published online: 24 Jan 2019

Abstract

We investigate the existence of left and right adjoints to the restriction functor in three categories of continuous representations of a topological group: discrete, linear complete and compact.

1991 MATHEMATICS SUBJECT CLASSIFICATION:

Given a representation of an abstract group G one can always define a representation of a subgroup HG by restricting the group action to the subgroup. This defines a functor ResHG:Rep(G)Rep(H).

We are interested in functors going in the opposite direction, in particular, ones which are adjoint to ResHG. Such functors are called induction functors and the adjointness relation is known as Frobenius reciprocity. A criterion for existence of adjoints to a given functor is given by the Freyd’s Adjoint Functor Theorem. It states that a functor F:CD between two categories C and D, with C locally small, has a left (respectively right) adjoint if the category C is complete (respectively cocomplete), F commutes with small limits (respectively colimits), and an abstract condition (SSC) is satisfied [Citation15]. In the case of F being the restriction functor ResHG,C being the category Rep(G) of representations of an abstract group G, and D the category Rep(H) of representations of a subgroup HG, all these conditions are satisfied and the functor ResHG has both a left and a right adjoint given by ⊗ and Hom(,) respectively. Thus, we have two induction functors: IndHG,CoindHG:Rep(H)Rep(G).

To distinguish between them we call the right adjoint induction and denote it IndHG, and the left adjoint coinduction and denote it CoindHG. Since adjoints are unique up to isomorphism, it is clear that in the case of abstract groups CoindHG is given by the tensor product and IndHG is given by the Hom(,) functor. For finite groups, and more generally for abstract groups of arbitrary cardinality, provided that H is of finite index in G, IndHG and CoindHG are isomorphic. However, in general this is not true. It is worth noting that in some of the literature the opposite convention is adopted - the left adjoint functor is called induction and the right - coinduction. This is always the case when studying the representation theory of rings. However, for p-adic groups the notions are the same as ours which is the reason for our terminology.

In this article, we wish to extend the idea of Frobenius reciprocity to continuous representations of topological groups and to investigate existence of the induction and coinduction functors. In particular, we fix a topological group G and a closed subgroup HG. We look at continuous representations of such groups over an associative unitial ring R. These are topological modules (V,TV) over R, such that the action map G×VV is continuous with respect to the topology TV on V and the product topology on G×V. Varying TV we obtain different categories of continuous representations for G. We are interested in three such: the category of discrete representations Md(G), where V is endowed with the discrete topology, the category of linearly topologized and complete representations Mltc(G), where the topology on V is linear and complete, and the category of compact representations Mc(G), where V is given a linear, complete topology in which all quotients of open submodules are of finite length. In each of these categories we investigate the existence of a left and a right adjoint to the restriction functor ResHG. Our main tool is Freyd’s Adjoint Functor Theorem. As opposed to the abstract groups case, in this setting we do not always have adjoints. Having topological, as well as algebraic conditions, creates some difficulties. As a start we need to construct products and coproducts in the aforementioned categories in a correct way. Furthermore, the functor ResHG has to commute with those, which is not always the case when taking a closed subgroup HG. In the present paper we address all these problems in detail. More precisely, each section contains a main result which is a criterion for the existence of a left and a right adjoint to ResHG, i.e, we give variants of Frobenius reciprocity in each of the three categories described above. Putting those together, we establish the following:

Theorem

(Main). Let G be a topological group and HG a closed subgroup. The restriction functor ResHG:M(G)M(H)

has the following properties:

  1. In Md(G) the functor ResHG always has a right adjoint, given by the induction functor IndHG, and has a left adjoint CoindHG if H is also open.

  2. In Mltc(G) and Mc(G) the functor ResHG always has a left adjoint CoindHG and has a right adjoint IndHG if H is also open.

Now let us give a section-by-section outline of the present paper.

Section 1 is a short introduction to the categories of interest. We set the notation and give a brief description of the objects and morphisms in each category. A more precise definition of a continuous representation of G is given.

In Section 2, we investigate Frobenius reciprocity in the category of discrete representations Md(G) of G. Considering representations over modules endowed with the discrete topology is the standard approach to continuous representations. In particular, if G is a locally compact totally disconnected group, then Md(G) is precisely the category of smooth representations of G [Citation5,16,17]. This category is widely studied as examples of groups with such topology include p-adic groups [Citation5,Citation16], topological Kac-Moody groups and groups of Kac-Moody type [Citation11]. The construction of the induction functor from a closed subgroup HG in the case of smooth representations is well-known. Following Bushnell and Henniart in terminology, we generalize the construction of IndHG to arbitrary topological groups [Citation5]. Next, we move on to coinduction, which is more subtle. In the case of locally compact totally disconnected groups, the coinduction functor is called compact induction or induction with compact support [Citation5,Citation16]. In Theorem 2.3, we establish a sufficient condition for its existence in Md(G). In particular, we claim that HG must also be an open subgroup. It is worth remarking that Bushnell and Henniart work over a field of characteristic zero. This is extended by Vignéras to positive characteristic. She also describes the induction and compact induction functors from a closed subgroup of a locally profinite group for modules over a commutative ring [Citation16]. We, however, keep the generality of an associative ring with identity and arbitrary topological groups.

In Section 3, we move on to the category of linearly topologized and complete continuous G-modules Mltc(G). First, we give a precise formulation of the continuity condition. We then begin our investigation of adjoints to ResHG. We wish to use Freyd’s Theorem to establish whether IndHG and CoindHG exist in Mltc(G). The main condition, which the category needs to satisfy to ensure such existence, is to be complete and cocomplete, i.e., to be closed under taking small limits and small colimits. Since Mltc(G) is abelian, using the product-equalizer construction of limits and the coproduct-coequalizer construction of colimits, it is enough to show that the category is closed under taking products and coproducts. We show how to construct these in Mltc(G) in Lemma 3.1 and Lemma 3.2, respectively. The main result of Section 3 is Theorem 3.3. It establishes the existence of the left adjoint to ResHG, i.e., CoindHG in the case when HG is closed, and the existence of the right adjoint, i.e., IndHG when H is also open.

In Section 4, we describe the category of (linearly) compact continuous representations of G, denoted Mc(G). The notion of linear compactness of vector spaces first appears in Lefschetz’ “Algebraic Topology” [Citation14]. He calls a vector space linearly compact if it is linearly topologized, Hausdorff, and satisfies the finite intersection property on cosets of closed subspaces. Such spaces are complete [Citation14]. This leads to an alternative definition of linearly compact vector spaces given by Drinfeld as linearly topologized complete Hausdorff spaces with the property that open subspaces have finite codimension [Citation8]. Dieudonné unifies these definitions by showing they are all equivalent in the case of vector spaces [Citation7]. Compact vector spaces are also topological duals of discrete ones [Citation8,Citation13]. Taking the duality viewpoint, we begin the section by constructing examples to show that if R is a field, then ResHG is cocontinuous in neither Mc(G), nor Mltc(G), unless H is open. We then move on to the case of modules over an associative ring. In this setting the definitions for compactness given above are not equivalent. A linearly compact topological R-module V is linearly topologized, Hausdorff, and such that every family of closed cosets in V has the finite intersection property [Citation18]. However, this is not equivalent to V being linearly topologized, complete, and such that open submodules have finite colength. We wish to take the point of view of the latter definition as it is closer to the Beilinson-Drinfeld approach to linearly compact topological vector spaces [Citation1]. Modules defined as above are known in the literature as pseudocompact [Citation3,9,12]. These come up in deformation theory, in particular, they are useful when describing lifts and deformations of representations of a profinite group over a perfect field of characteristic p [Citation3].

After we have fixed the definition of the compact topology for a module over a ring, we follow our strategy from Section 3: we construct products (Lemma 4.5), coproducts (Lemma 4.6) and investigate the existence of IndHG and CoindHG in Mc(G). Theorem 4.7 is the main result of the section. It establishes the existence of the coinduction functor for HG closed, and the existence of the induction functor, given that H is also open.

We finish the article with a brief discussion of the category of Tate representations MT(G). Tate spaces, or locally linearly compact spaces as defined by Lefschetz [Citation14], are complete linearly topologized vector spaces, such that the basis at zero is given by mutually commensurable subspaces [Citation2]. Equivalently, a Tate space is a vector space which splits as a topological direct sum of a discrete and a compact space [Citation8]. The latter definition also generalizes to modules over a commutative ring [Citation1]. Hence, for a topological group G one can define the category of Tate representations, as the category with objects Tate spaces, on which G acts continuously. These are an interesting object to study as they appear not only in the phenomenal work of Tate, but also in other areas of mathematics, such as the algebraic geometry of curves, the study of chiral algebras and infinite dimensional Lie algebras, as well as in Conformal Theory [Citation1,Citation2]. We do not fully investigate the analogue of Frobenius reciprocity in MT(G), but we pose some questions about it.

1. Introducing the categories

Throughout let R be an associative ring with 1 and G a topological group. We are interested in studying continuous representations of G over R. Let us explain precisely what we mean by this.

First, recall that (V,TV) is called a topological R-module if TV makes (V,+) into a topological group and the R-action map ·:R×VV,(r,v)r·v is continuous with respect to TV on the right and the product topology on the left (where R is endowed with the discrete topology). With this in mind, we make the following definition:

Definition 1.1.

Let (V,TV) be a topological (left) R-module and π:GAutR(V) a homomorphism. Then the pair (π,V) is a representation of G. It is called continuous if the map ϕ:G×VV defined by (g,v)π(g)v is continuous with respect to the product topology on the left and TV on the right.

Whenever we talk about topological R-modules, we always mean left modules, but of course the results remain true for right R-modules. From the definition above it is clear that the continuity condition depends on the topology we put on V. Hence, by changing this topology we obtain different categories of continuous representations. We are mainly interested in three such:

  • Md(G) - category of discrete representations of G. The objects are continuous representations (π,V) of G, such that (V,TV) is a topological R-module, endowed with the discrete topology. The morphisms between two objects (π1,V1) and (π2,V2) are given by R-linear maps f:V1V2, such that f(π1(g)v)=π2(g)f(v), for gG and vV1.

In the next two categories of interest V is given a linear topology TV. More precisely, we say that a topology TV is linear, or that V is linearly topologized, if the open R-submodules of V form a fundamental system of neighborhoods at zero [Citation18]. This gives rise to the following categories:

  • Mltc(G) - category of linearly topologized complete representations. The objects are pairs (π,V), where V is a continuous representation of G, endowed with a linear topology TV, such that (V,TV) is a complete topological space. The morphisms between two objects (π1,V1) and (π2,V2) are given by continuous R-module homomorphisms f:V1V2, such that f(π1(g)v)=π2(g)f(v), for gG and vV1.

  • Mc(G) - category of compact representations. The objects are pairs (π,V), where (V,TV) is a linearly topologized complete R-module, such that for every open R-submodule WV, V/W is an R-module of finite length. The morphisms are defined in the same way as in Mltc(G).

We give further details on the topologies of the three categories defined above, as well as the explicit meaning of the continuity condition, in the sections to follow.

2. Category of discrete representations

Fix a topological group G and a closed subgroup HG. We study the category Md(G) of discrete representations of G and Md(H) of discrete representations of H. These are connected by the restriction functor: ResHG:Md(G)Md(H),defined by(π,V)(π|H,HV), where π|H:HAutR(V) is the restriction of π:GAutR(V) to H and HV denotes V as an H-module.

Let (π,V) be a representation of G and ϕ:G×VV, given by (g,v)π(g)v, be the map induced by the action of G. A discrete representation (π,V) of G is continuous if for every vV,ϕ1(v) is open in G×V. In other words, for every vV, there exists an open set KvG, such that π(k)v=v, for every kKv.

Note that 1G always satisfies π(1G)v=v. Since the group topology is determined by the fundamental neighborhoods of identity, without loss of generality assume that Kv is an open neighborhood of 1G. We could go even further - for every vV we can construct an open subgroup Kv˜G generated by Kv. Then clearly π(k)v=v, for every kKv˜.

If the topology of G is locally compact and totally disconnected, then Md(G) is the category of smooth representations of G. The smoothness condition there states that StabG(v) is open in G for every vV, which is precisely our continuity condition.

The main goal of this section is to determine when the restriction functor ResHG has a left and a right adjoint in Md(G).

We start by investigating whether a right adjoint to ResHG exists. We claim that it exists and is given by the induction functor IndHG:Md(H)Md(G). We define IndHG generalizing the construction of smooth induction for locally compact totally disconnected groups [Citation5]:

Fix (σ,W)Md(H). Consider the R-module Ŵ of all left H-equivariant functions f:GW, i.e., which satisfy the property

  1. f(hg)=σ(h)f(g), for all hH and gG.

Within Ŵ we find an R-submodule W˜ consisting of “continuous functions”, i.e., functions with the additional property

  • iii. fW˜ if and only if there exists an open neighborhood Kf of 1G, such that f(gk)=f(g), for all gG and kKf.

The homomorphism ρ:GAutR(Ŵ), given by ρ(g)f:xf(xg), for g,xG and fŴ, defines a G-action on both Ŵ and W˜, thus making (ρ,W˜) a continuous representation of G, i.e., (ρ,W˜)Md(G). The pair (ρ,W˜) is called the representation of G continuously induced by σ and is denoted IndHG(σ). Using this construction we define the functor IndHG:Md(H)Md(G). We claim that this is the right adjoint we are looking for.

Lemma 2.1.

For a topological group G and a closed subgroup HG, the functor IndHG:Md(H)Md(G) defined above is right adjoint to the restriction functor.

Proof.

For continuous representations (π,V)Md(G) and (σ,W)Md(H), with notation as above, we want HomG(V,W˜)HomH(HV,W).

We have maps: α:HomG(V,W˜)HomH(HV,W) given by: ψ:VW˜ψ˜:VWψ:vψvψ˜:vψv(1G), and β:HomH(HV,W)HomG(V,W˜) given by: ϕ:VWϕ˜:VW˜

with ϕ˜:vfv:gϕ(g·v).

Clearly ψ˜HomH(HV,W) since ψvW˜. Similarly for ϕ˜HomG(V,W˜). It is routine to check that α and β are inverse to each other. □

Now we move on to the case of the left adjoint. This is more subtle. Let us lay out our conventions first. We use the following standard terminology:

  • A functor F:CD is called:

    • continuous if it preserves small limits,

    • cocontinuous if it preserves small colimits.

  • A category C is called:

    • complete if all small diagrams have limits in C,

    • cocomplete if all small diagrams have colimits in C.

Since limits can be constructed as equalizers of products, a category C is complete if all morphisms in C have equalizers and C is closed under arbitrary products [Citation15]. Hence, to check continuity of a functor F:CD, it is sufficient to check that F preserves those (respectively coproduct and coequalizers for the cocontinuous case).

Recall the following criterion for existence of a left adjoint to a functor [Citation15]:

Theorem

(The Freyd Adjoint Functor Theorem). Given a locally small, complete category C a functor F:CD has a left adjoint if and only if it preserves all small limits and satisfies the following condition:

(SSC) For each dOb(D) there is a small set I and an I-indexed family of morphisms fi:dF(ci) in D, with ciOb(C), such that every morphism h:dF(c) in D, with cOb(C), can be written as a composite h=F(t)°fi, for some index iI and some morphism t:cic in C.

Dualise the statement to obtain a criterion for a right adjoint.

We wish to use Freyd’s Theorem to determine whether the restriction functor has a left adjoint. First note that (SSC) holds in Md(G): it just says that every map in Md(G) can be factored through a quotient. Also note that Md(G) is abelian, so equalizers of all morphisms exist. Since ResHG does not change the morphisms between the objects, it commutes with equalizers.

The next step is to check whether Md(G) is closed under arbitrary products. Take a collection {Vi}iIMd(G), for some arbitrary set I. Let V:=iIVi denote the product of Vi as R-modules. V remains a discrete space with respect to the box topology. It also has an obvious G-module structure - G acts componentwise: g·(v1,v2,,vn,..)=(g·v1,g·v2,,g·vn,..),forgG,viVi.

However, the action is not necessarily continuous:

Fix v:=(v1,v2,,vn,..)V. Since ViMd(G), for every viVi, there exists an open neighborhood Kvi of 1G, such that k·vi=vi, for all kKvi and iI. Thus, Kv=iIKvi has the property that k·v=v, for all kKv. But as I is chosen arbitrarily Kv does not have to be open. Therefore, the representation is not continuous at v and VMd(V). However, consider the continuous part of V, i.e., Vsm:={vV|there existsKvopen inG,such thatk·v=v,for allkKv}.

Clearly Vsm is a continuous representation of G. We claim the following:

Lemma 2.2.

Every collection {Vi}iIMd(G),I an arbitrary set, has a product in Md(G) given by (iIVi)sm. In other words, Md(G) is complete.

Proof.

Let Vsm:=(iIVi)sm. Denote by pi:VsmVi the canonical projections in Md(G). Let AMd(G) and fi:AVi be a family of morphisms in Md(G) indexed by I. As VsmiIViR-mod, there exists a unique R-module homomorphism f:AVsm, such that pi°f=fi, for all iI. It is also a G-map: pi(f(g·a))=fi(g·a)=g·fi(a)=g·(pi(f(a))=pi(g·(f(a)),gG,aA, i.e., f(g·a)=g·f(a).

Thus, f is a morphism in Md(G) and the universal property of the product is satisfied. □

We claim that even though Md(G) is complete, the restriction functor does not always have a left adjoint.

Theorem 2.3.

Let G be a topological group and H a closed subgroup of G. Then ResHG:Md(G)Md(H)

has a left adjoint if H is also open.

Proof.

Taking into consideration the discussion before Lemma 2.2, to establish the existence of a left adjoint, we have to show that ResHG is continuous, i.e.: (iIResHG(Vi))smResHG((iIVi)sm),for some indexing set I.

This is the same as showing that every WMd(H), such that W=(ResHG(Vi))sm, for some ViMd(G), is also an element of Md(G).

Take such WMd(H). Then for every wW there exists an open neighborhood Kw of 1H, such that k·w=w, for all kKw. As H is given the subspace topology, there is an open UG, such that Kw=UH. But since H is open in G, then Kw is an open neighborhood of 1G and thus WMd(G). □

Thus, for an open subgroup HG, we have a functor CoindHG:Md(H)Md(G), which is left adjoint to the restriction functor.

Example 2.4.

Suppose the topology on G is locally compact and totally disconnected. Then CoindHG is given by compact induction of representations, i.e., CoindHG:Md(H)Md(G),(σ,W)(ρ,W˜), where W˜ is the space of all left H-equivariant, continuous functions f:GW, which have compact modulo H support, and ρ(g)f:xf(xg) [Citation5]. It is clear from the constructions that for a locally compact totally disconnected G and H, such that HG is compact, IndHGCoindHG [Citation5,Citation16].

The next example shows that if H is not open in G, then CoindHG is not always defined.

Example 2.5.

Take {Vi}iIMd(G). Fix v=(v1,v2,,vn,..)iIVi with Ki corresponding open neighborhoods of 1G, such that ki·vi=vi, for all kiKi. For each iI construct an open subgroup Ki˜G, generated by Ki. Then k·v=v, for all kK˜:=iIKi˜. However, as I is an arbitrary set K˜ is not necessarily open in G. Moreover, it is closed as every Ki˜ is. Taking H:=K˜ we have v(iIVi)sm, and thus vResHG(iIVi)sm. But as H is open in H, v(iI(ResHG(Vi))sm. Thus, ResHG fails to be continuous and CoindHG is not defined.

3. Linearly topologized and complete G-modules

Let R be an associative ring with 1, V a topological R-module and G a topological group. In this section we investigate the category Mltc(G) of linearly topologized and complete R-modules which admit a continuous action of G.

Let (π,V)Mltc(G). Then the map ϕ:G×VV defined by ϕ:(g,v)π(g)v is continuous. In particular, for an open UV there exists an open neighborhood K of 1G and an open submodule WV, such that if g·xU, then Kg·(x+W)U, for some gG and xV.

Fix a closed subgroup HG. As in Section 2, given a continuous representation (π,V) of G, we obtain a representation of H by restricting the map ϕ:G×VV to H. As a restriction of a continuous map it remains continuous. Thus, once again we have a functor ResHG:Mltc(G)Mltc(H).

We wish to investigate the existence of adjoints to this functor. We start with the left adjoint. Following the same strategy as in Section 2 we begin by constructing arbitrary products in Mltc(G).

Lemma 3.1.

Arbitrary products exist in Mltc(G). More precisely, the product of a collection of objects of Mltc(G) is their product as R-modules, endowed with the product topology.

Proof.

For an arbitrary collection {(πi,Vi)}iI of elements of Mltc(G), let V:=iIVi denote their product in R-mod which is a topological R-module with respect to the product topology [Citation4].

Following Lefschetz we show that the topology on V is linear [Citation14]. Let {Uij},jJi be a base of neighborhoods of 0 in Vi, consisting of open submodules. Then for any finite subset KI,kKUkj×iIKVi is a base of neighborhoods of 0 in V consisting of open submodules, giving the linearity of the product topology.

Let zn=(zn1,zn2,..,zni,),zniVi and nL for an ordinal number L, be a Cauchy net in V. Since all Vi are complete, zni is a convergent Cauchy net in Vi. Let zi:=limnLzni.

Set z=(z1,z2,,zi,). Let Ui be an open neighborhood of zi in Vi and define U=iJUi×iIJVi, for some finite subset JI. Since each net zni is convergent, there exists some li, such that zniUi, for all nli in L. Pick the largest li,iJ, say l. Then for all nl in L,zniUi for all iI. Thus, zU and zn is convergent in V with limnLzn=z.

The G-action on V is componentwise. We want to show it is continuous. Let U:=iJUi×iIJVi with JI finite. Then UV is open.

Since all Vi are continuous G-modules, for every iI there exists an open neighborhood Ni of 1G and an open submodule WiVi, such that if g·xiUi, for gG,xiVi, then Nig·(xi+Wi)Ui. Fix N:=iJNi. Since J is finite, N is an open neighborhood of 1G and furthermore Ng·(xi+Wi)Ui, for all iJ. Let W:=iJWi×iIJVi. This is an open submodule of V. Thus, we found NG and and open submodule WV, such that for g·xU, with x=(x1,..,xi,..),Ng·(x+W)U. Hence, VMltc(G).

Let AMltc(G),pi:VVi be the projections in Mltc(G) and fi:AVi be a family of morphisms in Mltc(G) indexed by I. As V is the product of Vi in R-mod, there exists a unique R-module homomorphism f:AV, making the following diagram commute:

The map f has the following properties:

  1. (pi°f)(g·a)=fi(g·a)=g·fi(a)=g·(pi°f)(a),for gG and aA, i.e., f is G-linear.

  2. For UV open, Ui:=pi(U)Vi is open. By continuity of fi it follows that fi1(Ui)A is open, for every iI. Thus, f1(U)=fi1(pi(U))A is open, showing that f is continuous.

Thus, f is a morphism in Mltc(G), finishing the proof. □

To continue our investigation of adjoint functors, we would also need existence of arbitrary coproducts in Mltc(G). We construct them explicitly. Let {Vi}iI be an arbitrary collection of elements in Mltc(G). Denote by V:=iIVi their coproduct in R-mod and by αi:ViV the canonical injections. In this case they are just inclusion maps. We follow Higgins in defining the topology on V [Citation10]:

Consider pairs (W,τW), such that:

  1. WMltc(G), such that there exists a surjective R-module homomorphism qW:VW, which is also G-linear,

  2. τW is a topology on W in which the maps qWi:ViW that factor through qW are continuous.

All such pairs (W,τW) taken up to isomorphism form a set. Hence, we can form a product (W,τW)W. The map (1) q:V(W,τW)W,given by v(qW(v))(W,τW)(1) is an embedding. We endow (W,τW)W with the product topology and V with the topology induced by q. This is a group topology [Citation10]. The map ϕ:R×(W,τW)W(W,τW)W is continuous, hence, the restriction ϕ|q(V):R×q(V)q(V) is also continuous. Thus, the subspace topology on q(V), and respectively the induced one on V, is an R-module topology. By Lemma 3.1 (W,τW)W lies in Mltc(G). Every subspace of a linearly topologized space is linearly topologized [Citation14]. Thus, as Vq(V), both as an R-module and as a topological space, the topology on it is linear. A priori V is not necessarily complete. However, its closure V¯ is, as it is a closed subspace of a complete space [Citation4].

Lemma 3.2.

V¯ as defined above is the coproduct of {Vi}iI in Mltc(G).

Proof.

By definition V¯ is a linearly topologized and complete space. As V=iIVi and Vi is a G-module for every iI, then clearly so is V. Since (W,τW)WMltc(G), the map G×(W,τW)W(W,τW)W is continuous. Hence, its restriction to a subspace is also continuous. Therefore, V is a continuous G-module and, hence, so is its closure V¯. Thus, V¯Mltc(G) as required.

Let us check that V¯ is indeed the coproduct of {Vi}iI. Let A be any module in Mltc(G) and βi:ViA be morphisms in Mltc(G) indexed by I. Since V is the coproduct of {Vi}iI in R-mod, there exists a unique R-linear homomorphism f:V¯A, such that for every iI the diagram below commutes:

The map f is G-linear: (f°αi)(g·vi)=f(g·(αi(vi)))=βi(g·vi)=g·βi(vi)=g·f(αi(vi)), for viVi,gG.

Lastly, let UA be open. By continuity of βi, βi1(U)Vi is open, for every iI. Since the Vi’s appear amongst the (W,τW), then βi1(U)×(W,τW),WViW is open in (W,τW)W and moreover βi1(U)×WVi(W,τW),W=qi(βi1(U)). Thus, q1(βi1(U)×WVi(W,τW),W)=q1(qi(βi1(U)))=αi(βi1(U))=f1(U), where qi is given componentwise by the qWi defined above. By definition of the topology on V¯, it follows that f is continuous, finishing the proof. □

With notation as before, we have the following diagram:

Since qWi is continuous for each iI, then so is qi [Citation4]. By definition of the topology on V, q is continuous. Hence, αi is continuous for each i. This means that the topology on V is contained in the final topology with respect to αi. However, the continuity of the αi implies that V appears as one of the W, thus, the coproduct topology defined above coincides with the final topology.

Now we would like to give an explicit description of the basis of open neighborhoods of 0 in V. Chasco and Domínguez describe this basis with respect to the final topology for a coproduct of topological abelian groups [Citation6]. We generalize their construction to topological R-modules:

Let {Ui}iI be a sequence of neighborhoods of 0, with Ui a neighborhood of 0 in Vi. Let JI be finite. Then (2) U:=|J|=nnN,|K|=|J|KI,iKαi(Ui)(2) is a sequence of neighborhoods of 0 in V. Hence, the basis is given by U={U|{Ui}iK,with UiVi open neighbourhood of 0}.

This indeed agrees with our description of the topology: If B(W,τW)W is an open neighborhood of 0 in (W,τW)W,q1(B) would be the corresponding open in V and q1(B)V=q1(B)iIVi=iIq1(B)Vi, can be written in the form of (2).

Theorem 3.3.

Let G be a topological group and HG a closed subgroup. Let ResHG:Mltc(G)Mltc(H) be the restriction functor. The following hold:

  1. ResHG has a left adjoint CoindHG:Mltc(H)Mltc(G).

  2. If H is also open, then ResHG has a right adjoint IndHG:Mltc(H)Mltc(G),

given by IndHG:WFunH(G,W), where FunH(G,W) is the space of all functions f:GW, such that f(gh)=h·f(g).

Proof.

We wish to apply Freyd’s Adjoint Functor Theorem to prove the statements. First, Mltc(G) is an abelian category. Hence, equializers and coequalizers exist. By the product-equalizer (coproduct-coequalizer) construction of limits (respectively colimits), Lemma 3.1 and Lemma 3.2 imply that Mltc(G) is both complete and cocomplete. Since the restriction functor does not change morphisms, it commutes with equalizers and coequalizers. Thus, to show existence of a left and a right adjoint to ResHG we only need to check whether it commutes with products and coproducts. Let us start with products, i.e., we want ResHG(iVi)iResHG(Vi),

where ViMltc(G), for all i. By Lemma 3.1 iResHG(Vi)Mltc(H) and iResHG(Vi)iHViH(iVi)ResHG(iVi).

Note that H(iVi) is already a continuous H-module. Thus, the second isomorphism holds because iHVi and H(iVi) have the same structure as H-modules, as well as topological spaces. This completes the proof of (1).

Now we move on to coproducts. To have a right adjoint to ResHG we need to check cocontinuity, i.e.: iResHG(Vi)ResHG(iVi),

where by ⊕ we denote the coproduct in Mltc(H) and Mltc(G) respectively. This amounts to showing that iHViH(iVi).

Since iHVi and HiVi have the same structure as H-modules and the action of H on both is continuous, to obtain the isomorphism, we need to show that the coproduct topologies on both sides are the same. In particular, for every H-module (W,τW) for which there exists a surjective H-map φ:iHViW with the property that the maps qi:ViW are continuous, we have to find a module (W˜,τW˜)Mltc(G), for which there exists a surjective G-map φ˜:HiViW˜, such that the maps qi˜:ViW˜ factoring through φ˜ are continuous. In addition, for every open UW there must be an open U˜W˜, such that φ˜1(U˜)φ(U).

Fix (W,τW) with the above properties. Let FunH(G,W) be the space of all right H-equivariant functions f:GW, i.e., which satisfy the property: f(gh)=h·f(g), for gG,hH.

This is a G-module via g·f:xf(xg). We claim that W˜:=FunH(G,W) satisfies the desired conditions. First, we show that W˜Mltc(G). Let X be a set of right coset representatives of H in G. There is an H-module isomorphism: ψ:FunH(G,W)aXaHW,ψ:f(aif(1))aiX.

Identifying the right hand-side as |X| copies of W, we can put the product topology on it. By Lemma 3.1 aXaHWMltc(H). Endow FunH(G,W) with the topology induced by ψ. Since FunH(G,W)aXaHW as H-modules and they have the same topology, then FunH(G,W)Mltc(H).

Let UFunH(G,W) be open. Suppose g·fU, for gG and fFunH(G,W). The function g·f is H-equivariant and we can rewrite it as g·f:xf(xg)=f(yh)=h·f(y):=h·wyU, for g,x,yG,hH and wyW. Since W is a continuous representation of H, there exist an open neighborhood K of 1H and an open submodule ZW, such that Kh(wy+Z)U. But as HG is open, K is an open neighborhood of 1G.

Set Z˜:=ψ1(1GZ×aX,a1GaW). This is an open submodule of FunH(G,W). Moreover, the pair (K,Z˜) has the property Kg(f+Z˜)U. In particular, FunH(G,W) is a continuous representation of G.

Next, extend every surjective H-map φ:iHViW to a surjective G-map φ˜:HiViW˜ by φ˜:vfv:gφ(g·v).

Fix an open UW. Then bU×aX,abaWaXaW is open. Let U˜=ψ1(bU×aX,abaW). By definition U˜ is open in W˜ and

U˜={fFunH(G,W)|f(1)U}. Then φ˜1(U˜)={vV|fvU˜}={vV|fv(1)U}{vV|φ(v)U}=φ1(U).

4. Category of compact representations

Let us start by defining the category Mc(G) of compact representations of a topological group G. As always the objects are pairs (π,V), where (V,TV) is a topological module over an associative unitial ring R and π:GAutR(V) is a continuous representation of G. We need to describe the topology TV on V.

Firstly, we look at the case when R=F is a (skew-) field. As explained in the introduction, there are a few equivalent definitions of (linear) compactness for topological vector spaces. Following Beilinson-Drinfeld we view them as topological duals V of discrete vector spaces V [Citation1,Citation8]. By a topological dual we mean the space of all continuous linear functionals on V. The topology on V is given by orthogonal complements of finite dimensional subspaces of V with respect to the canonical pairing [Citation1].

Define D:Md(G)Mc(G) by VVandff,wheref:φφ°f.

We claim that this is a contravariant functor, which induces an anti-equivalence of categories. First, let us check that D is indeed a functor.

Lemma 4.1.

Suppose R=F is a field and let VMd(G). Then VMc(G), where V:={f:VF} is the space of all continuous linear functionals on V.

Proof.

We make V into a G-module by defining an action of G by left translation. More precisely, we have an action map ϕ:G×VV, given by ϕ:(g,f)λgf, where λgf:xf(g1x). We claim that ϕ is continuous.

Let MV be open. Suppose that g·fM, for some gG and fV. By definition M:={f:VF|f(m)=0 for all mM}, where MV is finite dimensional. Then M=Fm1,..,mn, for some miV,i=1,,n. Since VMd(G), there exists open neighborhoods Ki of 1G, such that ki·mi=mi, for every kiKi and i=1,..,n. Since G is endowed with a group topology, we can choose Ki to be symmetric. Let K=i=1nKi. This is an open symmetric neighborhood of 1G. Now consider N:=g·M=Fg·m1,,g·mn. This is a finite dimensional subspace of V. Hence, N:={f:VF|f(n)=0 for all nN}V is open. Thus, KG and NV are both open and gK·(f+N)M, finishing the proof. □

Lemma 4.1 shows that D maps objects to objects. Let us check it does the same on morphisms. Let V1,V2Md(G) and f:V1V2 be a morphism. Then f:V2V1 has the following properties:

  1. f(g·φ)=(g·φ)°f:v1φ(g1f(v1))=φ(f(g1v1))=φ°(g·f)=g·f(φ), where gG,v1V1,φV2. Thus f is G-linear.

  2. Let UV1 be open. Then f(U)1={φV2|(φ°f)(U)=0}, for UV1 of finite dimension. But since f is a linear map, then f(U)V2 is also a finite dimensional subspace. By definition of the topology on V2 it follows that f(U)1 is open and f is continuous.

Therefore, D is indeed a functor. It is clear from the definition of a compact vector space that D is bijective. Thus:

Lemma 4.2.

The functor D induces an anti-equivalence between Md(G) and Mc(G). In particular, D maps products in Md(G) to coproducts in Mc(G).

Note that Kohlhase shows a special case of our result: he establishes that if G is a locally profinite group and (π,V) is a discrete representation of G over a field F, then V has the structure of a compact module over the algebra Λ(G), where Λ(G) is a generalization of the Iwasawa algebra of a compact group [Citation13]. In particular, he obtains an anti-equivalence between Md(G) and the category of pseudocompact Λ(G)-modules. However, dropping the restriction of the topology on G, we obtain an anti-equivalence between the category of discrete representations and the category of compact G-modules.

In the following example we use our observations about D to gain information about the cocontinuity of ResHG in Mc(G).

Example 4.3.

Suppose {Vi}iI is an arbitrary collection of discrete vector spaces over a field F, such that G acts continuously on each Vi. By Lemma 4.1 ViMc(G). By Freyd’s Theorem to have a right adjoint to ResHG in Mc(G) we need (3) ResHG(iIVi)iIResHG(Vi),(3) where denotes the coproduct. By Lemma 4.2 (4) ResHG(iIVi)ResHG(((iIVi)sm)).(4)

However, (5) iIResHG(Vi)((iIHVi)sm),(5)

By Theorem 2.3 ResHG((iIVi)sm)iI((HVi)sm) if H is open. Thus, for HG open 3 holds and there is a well-defined functor IndHG:Mc(H)Mc(G).

As explained in the introduction, all linearly compact vector spaces are complete. This means that Mc(G) is a subcategory of Mltc(G). In the next example we consider the coproduct iIVi, for ViMc(G), in the category Mltc(G). We aim to illustrate that in the case of topological vector spaces, if HG is not open, IndHG is not always defined in Mltc(G), too. We first wish to note that our definition of the compact topology on a topological vector space V using the duality viewpoint is equivalent to requiring that V is Hausdorff, linear, complete and such that every open subspace has finite codimension [Citation8]. We are now ready for our example.

Example 4.4.

For a collection {Vi}iIMc(G) defined as in Example 4.3, consider their coproduct iIVi in Mltc(G). Keeping the notation and conventions of Section 3, recall that this is the coproduct in VectF with topology induced by the embedding q:iIVi,(W,τW)W.

Since we are considering the coproduct as an object of Mltc(G), rather than Mc(G), some of the (W,τW) can be linearly topologized and complete, but not compact. Suppose that this is the case. The continuity of the maps qWi:ViW implies that for every open submodule UW and for all iI,qWi(U)1 is open and hence of finite codimension in Vi. Since qWi=qW°αi it follows that qW1(U) is of finite codimension in iIVi. Now take an arbitrary open submodule N(W,τW)W. Then N=jJUj×(W,τW)W, where Uj are open submodules of some of the W’s and J is finite. Applying the same argument as above and using the fact that q=(qW)(W,τW), we conclude that q1(N) is open in iIVi. As all open submodules of iIVi correspond to inverse images of open submodules N(W,τW)W, the topology on iIVi is compact. By the uniqueness of coproducts and Example 4.3 it follows that if H is not open, the restriction functor ResHG:Mltc(G)Mltc(H) is not cocontinuous.

We move on to R-modules, where R is an associative ring with 1. We call an R-module V compact if V admits a linear complete topology with the additional property that if UV is an open submodule, then V/U is of finite length. Such modules are sometimes called pseudocompact [Citation3,9,12]. We call a topological R-module V linearly compact if it is linearly topologized, Hausdorff, and such that every family of closed cosets in V has the finite intersection property [Citation18]. Every compact module is linearly compact [Citation12]. We denote by Mc(G) the category of all compact R-modules which admit a continuous action of the topological group G. We now construct products and coproducts in Mc(G).

Lemma 4.5.

Arbitrary products exist in Mc(G).

Proof.

A product of linearly compact R-modules is linearly compact with respect to the product topology [Citation18]. Since every compact module is linearly compact, then the category of compact modules is closed under products.

By exactly the same argument as in Lemma 3.1 for an arbitrary collection {Vi}iI of elements of Mc(G) the product V:=iIVi in R-mod, endowed with the product topology, is a continuous G-module with respect to the componentwise action of G. □

We now wish to form coproducts in Mc(G). For an arbitrary collection {Vi}iIMc(G), we form the coproduct V:=iIVi in R-mod. To define a topology TV on V we mimic the procedure from Section 3: Let WMc(G). Suppose there exists a surjective R-linear map qW:VW which commutes with the G-action, such that the maps qWi:ViW, factoring through qW, are continuous. The topology TV on V is induced by the embedding (6) q:V(W,τW)W,v(qW(v))(W,τW).(6)

Lemma 4.6.

Let {Vi}iI be an arbitrary collection of elements of Mc(G). Their coproduct is the module (V,TV) described above.

Proof.

By Lemma 4.5 (W,τw)WMltc(G). Since V(W,τw)W, the topology on V is linear [Citation14]. Let U(W,τw)W be a basic open. Then U=Ui×(W,τw)W, where UiW for some W, is an open submodule. By definition q1(U) is open in V. But q1(U)=iIq1(U)Vi. Since each qWi:ViW is continuous, it follows that q1(U)Vi is an open submodule of Vi. Hence, the quotient is of finite length. This implies that V/q1(U) is also of finite length, showing that the topology TV is compact. Every compact space is linearly compact. Thus, V is closed in (W,τW)W [Citation18]. In particular, V is complete [Citation4]. The map G×(W,τW)W(W,τW)W is continuous, and thus so is its restriction to a subspace, i.e., VMc(G). As Mc(G)Mltc(G) and the coproducts in the two categories are constructed in the same way, Lemma 3.2 implies that V satisfies the universal property of the coproduct in Mc(G), finishing the proof. □

Having constructed products and coproducts in Mc(G), in order to establish existence of a left and a right adjoint to the restriction functor ResHG in Mc(G), we need to check whether it is continuous and cocontinuous.

Theorem 4.7.

Let G be a topological group and HG a closed subgroup. The restriction functor ResHG:Mc(G)Mc(H) has a left adjoint. Hence, we have a well-defined CoindHG:Mc(H)Mc(G). It has a right adjoint if H is open.

Proof.

Since Mc(G) is an abelian subcategory of Mltc(G) and the products and coproducts in Mc(G) are the same as in Mltc(G), then the statement is just a corollary to Theorem 3.3. □

Having already studied the categories Mc(G) and Mltc(G), there is another category of topological vector spaces which is interesting to consider. We call a vector space Tate, if it splits as a direct sum of a discrete and a compact vector space. Thus, we can form a category MT(G) of Tate spaces on which G acts continuously. However, if one would like to investigate Frobenius reciprocity in MT(G), one encounters a difficulty straight away - the category is not even closed under products. Therefore, we propose to look at the free product completion MT(G)¯ of MT(G). Since products and coproducts in Mltc(G) exist and MT(G)Mltc(G), then MT(G)¯ would also be a subcategory of Mltc(G). If the products in the two categories coincide, then by Theorem 3.3 the restriction functor in MT(G)¯ would always have a left adjoint. An interesting question in this case would be whether for (σ,W)MT(H)¯, there exists a canonical G-submodule of CoindHG(σ) which is Tate. Coproducts and induction can be approached similarly. Firstly, we construct a free coproduct completion MT(G)̂ of MT(G). If coproducts there coincide with coproducts in Mltc(G), then we expect to have existence of an induction functor in MT(G)̂ in the case when HG is also open.

Acknowledgments

The research is supported by the author’s EPSRC PhD studentship. The author would like to thank Dmitriy Rumynin for proposing the problem and for the numerous valuable suggestions and helpful comments, and also Inna Capdeboscq for the many helpful conversations. The author is also indebted to the anonymous referee for the comments made.

Disclosure statement

No potential conflict of interest was reported by the author.

References

  • Beilinson, A., Drinfeld, V. (1995). Quantization of Hitchin’s integrable system and Hecke eigensheaves. 142–148. Preprint, available at http://www.math.uchicago.edu/benzvi.
  • Beilinson, A., Feigin, B., Mazur, B. (1991). Notes on Conformal Field Theory.
  • Bleher, F. M. (2014). Universal deformation rings of group representations with an application of brauer’s generalized decomposition numbers. Contemp. Math. 607:97–112.
  • Bourbaki, N. (1966). Elements of Mathematics: General Topology (Part 1), Hermann. 1–198.
  • Bushnell, C. J., Henniart, G. (2006). The Local Langlands Conjecture for GL(2). A Series of Comprehensive Studies in Mathematics, Vol. 335. Berlin, Heidelberg, New York: Springer, pp. 1–123.
  • Chasco, M. J., Domínguez, X. (2003). Topologies on the direct sum of topological abelian groups. Topology Appl. 133(3):209–223.
  • Dieudonné, J. (1951). Linearly compact spaces and double vector spaces over sfields. Am. J. Math. 73(1):13–19.
  • Drinfeld, V. (2006). Infinite-dimensional Vector Bundles in Algebraic Geometry. The Unity of Mathematics. Boston: Birkhäuser, pp. 263–304.
  • Gabriel, P. (1962). Des catégories abélienne. Bul. Soc. Math. France. 79:323–448.
  • Higgins, P. J. (1977). Coproducts of topological abelian groups. J. Algebra 44(1):152–159.
  • Hristova, K., Rumynin, D. (2018). Kac-Moody groups and cosheaves on Davis building. J. Algebra 515, 202–235.
  • Iovanov, M. C., Mesyan, Z., Reyes, M. L. (2016). Infinite-dimensional diagonalization and semisimplicity. Isr. J. Math. 215(2):801–855.
  • Kohlhase, J. (2017). Smooth duality in natural characteristic. Adv. Math. 317:1–49.
  • Lefschetz, S. (1991). Algebraic Topology. Providence, Rhode Island: AMS Colloquium Publications, pp. 72–88.
  • Mac Lane, S. (1998). Categories for the Working Mathematician. New York: Springer-Verlag, pp. 55–137.
  • Vignéras, M.-F. (1996). Represéntations-Modulaires D’un Groupe Réductif p-Adique Avec ℓ≠p. Boston: Birkhäuser, pp. 29–48.
  • Vignéras, M.-F. (2016). The Right Adjoint of the Parabolic Induction, Arbeitstagung Bonn 2013. Basel: Birkhäuser / Switzerland: Springer, pp. 405–425.
  • Warner, S. (1993). Topological rings. Northholand Math. Stud. 178:232–241.