by Roger Howe
Group extensions and cohomology
Although he went on to deal with many other subjects, including ergodic theory and harmonic analysis on nilpotent and solvable groups (both discussed further below), Moore also continued to reflect on issues around cohomology theories for topological groups, and published two further papers (or four, depending on how you count [6], [11], [16]) relating to this subject, the last one fairly recently.
In back-to-back papers [11], he revisited and substantially reformulated and generalized the work of his thesis. A major aspect of the new treatment was to expand the class of groups under consideration. In the 1964 papers, the main object of consideration was a separable locally compact group G, and a separable abelian locally compact group \( A, \) on which \( G \) acts by automorphisms (a locally compact \( G \)-module). In the 1976 papers, although \( G \) remains locally compact, the group \( A \) is allowed to be more general. It can be any topological group whose topology is equivalent to a separable complete metric topology (the class of polonais groups). So for example, \( A \) could be the additive group of the Banach space underlying a (strongly continuous) representation of \( G \) (in particular, of a unitary representation of \( G \)).
Enlarging the class of allowable \( A \) permitted several important constructions that made the resulting cohomology theory more flexible, more applicable, and also better defined, in the sense that it could be characterized by a set of three easily stated axioms. The first two axioms ask for standard expected features of cohomology. The first requires the expected functorial properties of the cohomology groups \( H^n (G, A), \) including coboundary operators \( \partial_{n} : H^n (G, A^{\prime\prime}) \rightarrow H^{n+1} (G, A^{\prime}) \) that create a long exact sequence of cohomology groups out of a short exact sequence \[ 1 \rightarrow A^{\prime} \rightarrow A \rightarrow A^{\prime\prime} \rightarrow 1 \] of admissible \( G \)-modules. The second is the normalization property that \( H^0 (G, A) = A^G, \) the subgroup of \( G \)-invariant elements of \( A. \)
The third property is another normalization axiom, requiring that \( H^n (G, I(A)) = 0 \) for \( n > 0, \) where \( A \) is a polonais abelian group, and \( I(A) \) is the “regular representation” of \( G \) on Borel (or measurable, modulo the usual condition of equivalence up to measure zero — either choice leads to a satisfactory theory) functions from \( G \) to \( A. \) (Note that if \( G \) is infinite, then even if \( A \) is locally compact, \( I(A) \) will not be, but it will be polonais.) This axiom allows dimension shifting, and for \( n > 0, \) allows one to determine the \( H^n (G, A) \) in terms of \( H^m (G, B) \) for other \( B \) and \( m < n, \) and so transfers uniqueness to higher cohomology. In the first of the two papers comprising [11], Moore considers the two different cases of \( I(A) \) as described above, and shows that they both satisfy the axioms, and thus give rise to the same cohomology theory. The third axiom is the challenging issue, and requires an argument closely related to the Mackey Imprimitivity Theorem [e2].
Central extensions and number theory
These results about group cohomology for locally compact groups, with their spaces of Borel or measurable functions, have an air of general topology about them, and may seem rather abstract and removed from the examples that have been of greatest interest in representation theory, such as algebraic groups, especially reductive groups, over local or global fields. However, in [6], a work with a different flavor, Moore investigated central extensions of such groups, and found remarkable connections with number theory, particularly the Artin reciprocity map of class field theory. Along with [3], this research showcases Moore’s pathbreaking work in the setting of adelic groups.
The main object of study in [6] is the notion of “fundamental group” for Chevalley (simple, split algebraic) groups over local fields. For connected topological groups, there is a natural notion of fundamental group in the sense of covering spaces. For an abstract group \( G \) that is “perfect”, i.e., equal to its own commutator subgroup, Schur [e1] had shown that a good theory of central extensions exists; in particular, there is a “universal central extension” \( \tilde{G}, \) that can map onto any given central extension. The kernel \( \pi_1 (G) \) of the covering map \[ 1 \rightarrow \pi_1 (G) \rightarrow \tilde{G} \rightarrow G \rightarrow 1 \] is the “Schur fundamental group” of \( G. \) Given some abelian group \( A, \) the central extensions of \( G \) by \( A \) are then classified by \[ H^2 (G, A) = \operatorname{Hom}(\pi_1 (G), A). \]
For topological groups, one should consider central extensions \[ 1 \rightarrow A \rightarrow E \rightarrow G \rightarrow 1 \] with all members of the sequence locally compact, and the mappings continuous. These extensions are classified by \( H^2_{\mathrm{top}} (G, A), \) with these groups being as defined in Moore’s first papers. It may not be the case that \( \pi_1^{\text{top}} (G), \) a universal topological central extension, exists, but if it does, and if \( G \) is perfect, then one can compare \( \pi^{\text{top}}_1 (G) \) with \( \pi_1 (G), \) the Schur fundamental group. The former will be a quotient of the latter.
There are also notions of “connected” and “simply connected” in the context of algebraic groups, which equate to the like-named topological notions when the base field is \( C. \) Suppose that \( G \) is an algebraic group that is connected and simply connected in the sense of algebraic groups. Moore first shows that in almost all cases, \( G_k, \) the group of \( k \)-rational points of \( G, \) is perfect, i.e., connected in the algebraic sense.
He then considers Chevalley groups — simple, split algebraic groups \( G, \) defined over a field \( k, \) which he restricts to be a local field. He asks: is the group \( G_k \) of \( k \)-rational points of \( G, \) simply connected as an abstract group or topological group? And if not, what is \( \pi_1 (G_k), \) and what is \( \pi_1^{\text{top}} (G_k) \)?
Steinberg [e6] had given an explicit construction by generators and relations, of the universal central extension \( \tilde{G}_k, \) using the Bruhat decomposition of \( G, \) relative to a selected \( k \)-split torus \( H \) in \( G. \) Implicit in Steinberg’s work is the construction of a cocycle \[ c : k^{\ast} \times k^{\ast} \rightarrow \pi_1 (G k ), \] such that the values \( c(x, y) \) for \( x \) and \( y \) in \( k^{\ast}, \) generate \( \pi_1 (G_k). \) (Strictly speaking, there is one cocycle \( c_{\alpha} \) for each long root in the root system of \( G \); but these are almost independent of \( \alpha. \))
The question thus becomes, what relations must be satisfied by the generators \( c(x, y) \)? Moore gives a list of four types of identities. (One of the identities is closely related to bilinearity, but somewhat more complicated. However, Moore shows by careful investigation of root systems that for all simple groups except symplectic groups, it can be replaced by bilinearity.) He calls continuous functions \( c \) with values in a given (locally compact, separable) abelian group \( A \) and satisfying his four identities, “Steinberg cocycles” with values in \( A, \) denoted \( S^G (k, A). \) The subgroup of Steinberg cocycles that are continuous as functions on \( k^{\ast} \times k^{\ast} \) are called “topological”, and denoted \( S_{\text{top}}^G (k,A). \) Moore constructs embeddings \begin{align*} & \phi : H^2 (G_k, A)\rightarrow S^G(k,A) \quad \text{and}\\ & \phi_{\text{top}}: H^2_{\text{top}} (G_k , A) \rightarrow S_{\text{top}}^G (k, A). \end{align*} He shows that for \( G = \operatorname{SL}_2, \) these maps are isomorphisms, and conjectures the same for general groups. This was established by H. Matsumoto [e12].
Steinberg cocycles in fact had been encountered before, in a seemingly different context, that of class field theory [e22]. The norm residue symbol \( \nu_k : k^{\ast} \times k^{\ast} \rightarrow \mu_k , \) where \( \mu_k \) is the group of roots of unity of \( k, \) provides a continuous Steinberg cocycle. Moore proves that it is essentially the only one: he shows that for a nonarchimedean local field with roots of unity \( \mu_k, \) the map \( f \rightarrow f \circ \nu_k \) defines an isomorphism \( \operatorname{Hom}(\mu_k , A) \simeq S_{\text{top}}^G (k, A). \) It follows that \( \pi_1^{\text{top}} (G_k )\simeq \mu_k. \) In particular, \( \pi_1^{\text{top}} (G_k) \) exists.
Although the notation above refers to the group \( G, \) in fact, the relations are almost independent of \( G. \) Only because bilinearity of the Steinberg cocycles fails for symplectic groups is there any dependence on the group. Moore notes that this exception permits the existence of the “metaplectic cover” of \( \operatorname{Sp}_{2n}, \) which intervenes in the “Weil representation” [e7] (aka “oscillator representation”), which links representation theory with both quantum mechanics and automorphic forms.
Moore goes on to consider the case of a global field \( k, \) and its ring \( A \) of adèles. General considerations on restricted products imply that \( \pi^{\text{top}}_1 (G_A) \) exists, and is the restricted product \( \prod\pi^{\text{top}}_1 (G_\nu), \) where \( G_\nu = G_{k_\nu}, \) and \( \nu \) runs over all places of \( k. \) In this context, Moore proves a second basic theorem: the sequence \[ \pi_1(G_k) \rightarrow \pi^{\text{top}}_1 (G_A)\rightarrow \mu_k \rightarrow 1 \] is exact. This in particular proves the essential uniqueness of the Artin reciprocity law of abelian class field theory. It also classifies topological central extensions of \( G_A \) that split over \( G_k. \) The primary example of this is the metaplectic cover of \( \operatorname{Sp}_{2n}, \) mentioned above as intervening in the construction of the Weil/oscillator representation.
Ergodic theory
Moore was an early exponent of using representation theory to study ergodic theory. The possibility for doing this arises from the fact that ergodic properties of a transformation \( T \) can be reinterpreted in terms of spectral theory. This equivalence is explained, for example, in Halmos’ well-known text [e3] on ergodic theory.
Recall the basic framework. Let \( X \) denote a measure space with total measure equal to 1 (aka, a probability space). For a measurable subset \( A \subset X, \) denote the measure of \( A \) by \( \mu(A). \) Let \( T : X \rightarrow X \) be an invertible transformation of \( X \) that preserves the measure, meaning that \( \mu(T (A)) = \mu(A) \) for any measurable subset \( A \) of \( X. \) The operators \( T^n, \) for \( n \in \mathbb{Z} \) then form a group of transformations on \( X, \) and ergodic theory is concerned with the “long term” behavior of this action, that is, what happens as \( n \rightarrow \infty. \) There is also interest in one-parameter groups \( T_t \) of measure-preserving transformations, for \( t \in \mathbb{R}, \) with \( T_s \circ T_t = T_{s+t}. \) The definitions and properties below can be adapted to this case, and indeed, to the action of any group on \( X \) by means of measure-preserving transformations.
The transformation \( T \) is called ergodic provided that there is no (measurable) \( T \)-invariant subset \( A \) in \( X \) with measure strictly between 0 and \( 1. \) If \( T \) is ergodic, it is called weakly mixing if the transformation \( T \times T \) on \( X \times X \) is also ergodic. The transformation \( T \) is called strongly mixing if a subset \( A \) of \( X \) gets spread out in a uniform way under successive applications of \( T. \) Formally, this is expressed by the requirement that the measure \( \mu(T^n (A) \cap B), \) of the intersection of a multiply transformed set with some other set, approaches the expectation dictated by randomness: \[ \lim_{\mu\to\infty} \mu(T^n (A) \cap B) = \mu(A) \cdot \mu(B). \] As Halmos explains, these properties can all be reformulated in terms of operator theory. From the transformation \( T \) of \( X, \) define an operator \( U_T \) on the Hilbert space \( L^2 (X) \) of square integrable measurable functions on \( X, \) by the recipe \[ U_T (f)(x) = f (T^{-1} (x)), \] for \( f \) in \( L^2 (X) \) and \( x \in X. \) It is straightforward to verify that \( U_T \) is a unitary operator. Similar definitions apply to a one-parameter group \( T_s \) of transformations. Using these correspondences, one can check that ergodicity of \( T \) is equivalent to the statement that the eigenspace of \( U_T \) for the eigenvalue 1 is one-dimensional, and consists of the constant functions. Weak mixing corresponds to \( U_T \) not having any eigenvectors for any eigenvalue, other than the constant functions. Strong mixing corresponds to the requirement that matrix coefficients of functions orthogonal to the constant functions (i.e., with integral zero), should “vanish at infinity”: \[ \lim_{n\to\infty} (\xi_{\phi,\psi} (n) = ((U_T)^n (\phi), \psi)) = 0, \] for any two functions \( \phi \) and \( \psi \) in \( L^2 (X) \) that are orthogonal to the constant functions. Again, there are analogous definitions for one-parameter groups, and in fact, for any locally compact group of measure preserving transformations.
Guided by these connections, in [4], Moore considers the following properties of a (closed) subgroup \( H \) of a (locally compact) group \( G \):
- \( H \) has property \( (E) \) if, in every unitary representation of \( G, \) any vector that is fixed by all the operators from \( H, \) is also fixed by all the operators from \( G. \)
- \( H \) has property \( (W M ) \) if, in every unitary representation of \( G, \) any finite-dimensional subspace that is invariant by all the operators from \( H, \) is also invariant by all the operators from \( G. \)
Given these definitions, he shows that, if \( X\simeq G/\Gamma \) is a homogeneous space for \( G \) with finite \( G \)-invariant measure, and \( H \) is a subgroup of \( G, \) then property \( (E) \) for \( H \) implies that the action of \( H \) on \( X \) is ergodic, and property \( (W M ) \) for \( H \) implies that the action of \( H \) on \( X \) is weakly mixing. In case \( H \) is a singly generated subgroup or a one-parameter subgroup, he proves an analogous statement about strong mixing.
Focusing on the situation when the large group \( G \) is a semisimple Lie group, which provides a large family of interesting examples, especially by allowing \( \Gamma \) to be an arithmetic subgroup, Moore proves an essentially optimal result about ergodicity of one-parameter group actions.
Note that, if a one parameter group \( H \subset G \) is in fact compact, in other words, is a copy of the unit circle \( \mathbb{T} \) in the complex plane, then one knows that any representation of \( H \) will decompose into a sum of one-dimensional representations. A bit more subtly, if \( H \) is a winding line in a compact torus, i.e., if \( \overline{H}, \) the closure of \( H \) in \( G, \) is compact, then a similar conclusion holds. Moore formulates a general condition on \( H \) that precludes these situations: the notion of total noncompactness.
A general connected semisimple Lie group \( G \) is almost a product of almost simple groups. Precisely, the adjoint group of \( G, \) i.e., the quotient \( G/Z(G), \) where \( Z(G) \) is the center of \( G, \) is a product of simple groups. In other words, given a semisimple Lie group \( G, \) there \( Q \) is a finite number of simple Lie groups \( G_j, \) for \( 1 \leq j \leq r, \) such that \( G \) maps onto the product group \( G \rightarrow \prod_{j=1}^r G_j , \) with discrete, central kernel. Given a subgroup \( H \subset G, \) we can look at the image \( H_j \) of \( H \) in each one of the simple factors \( G_j. \) We say that \( H \) is totally noncompact in \( G \) if the closure of each \( H_j \) in \( G_j \) is noncompact. Moore gives in [4] the following sharp description of the spectral behavior of a totally noncompact one-parameter subgroup.
Let \( L^+ = L^2 (\mathbb{R}^+) \) be the \( L^2 \) space of the positive half-line, and let \( L^{-} = L^2 (R^{-}) \) be the \( L^2 \) space of the negative half line. The one parameter group of multiplication operators \( U_t \) on \( L^2 (\mathbb{R}), \) defined by \[ U_t (f )(x) = e^{2\pi itx} f (x), \] for \( f \) in \( L^2 (\mathbb{R}), \) and \( x \in \mathbb{R}, \) preserves both of \( L^{\pm}, \) and its infinitesimal generator \( 2\pi ix \) has purely positive imaginary spectrum on \( L^+, \) and purely negative imaginary spectrum on \( L^-. \)
Moore shows:
Let \( H \) be a totally noncompact one-parameter subgroup of the semisimple Lie group \( G, \) and let \( \pi \) be a unitary representation of \( G. \) Then the restriction of \( \pi \) to \( H \) is unitarily equivalent to multiplication by \( e^{2\pi ix} \) on \[ m^+ L^+ \oplus m^{-} L^{-}, \] where \( m^{\pm} \) are nonnegative integers, or \( +\infty. \) He also shows by example that all values of \( m^{\pm} \) can occur. (In some sense, the typical case is that both of \( m^{\pm} \) equal \( +\infty. \)) This result implies that the action of \( H \) on any homogeneous probability space for \( G \) will be strongly mixing, but it is considerably more refined. As noted, strong mixing of a one-parameter group of transformations is equivalent to the condition that the matrix coefficients of the group vanish at infinity. Somewhat later, in [15] Moore proved that, for the geodesic flow on certain homogeneous spaces, appropriate matrix coefficients would actually decay exponentially as the group parameter goes to infinity. In between, I had the pleasure of collaborating with Moore on a result that proves ergodicity, and indeed, strong mixing, for a large class of actions of fairly general groups. Matrix coefficients can be studied for a wide class of groups. Given a locally compact group \( G \) and a unitary representation \( \pi \) of \( G \) on the Hilbert space \( H \) with scalar product \( (\phi, \psi) \) for vectors \( \phi \) and \( \psi \) in \( H, \) we can consider the matrix coefficients \[ \xi_{\phi,\psi} (g) = (\pi(g)(\phi), \psi) \] as functions of \( g \in G. \) We will say that the matrix coefficients of \( \pi \) vanish at infinity if, for fixed \( \phi \) and \( \psi, \) and any \( \epsilon > 0, \) there is a compact set \( J \subset G \) such that \( |\xi_{\phi,\psi} (g)| < \epsilon \) for \( g \) not in \( J. \) The paper [13] establishes that, barring some obvious conditions that would prevent it, for a very large class of representations of a large class of Lie groups, the matrix coefficients vanish at infinity. This result has been widely used by workers studying dynamical systems on homogeneous spaces, and even beyond (see, e.g., [e25], [e21], [e24], [e23]).
Harmonic analysis on nilpotent and solvable groups
Moore’s work on representation theory dealt with a number of issues for general groups (see [1], [7], [9], [10], [12]), and also dealt in detail with aspects of representation theory of solvable groups (see [5], [3], [8]). We briefly discuss [5], and then look at [3] in more detail.
Representation theory of solvable groups has been built around the orbit method, first enunciated by A. A. Kirillov for nilpotent groups [e5]. Kirillov’s main result is that the unitary dual of a (connected, simply connected) Lie group \( N \) is naturally parametrized by the set \( \{n^{\ast} /Ad^{\ast} N \} \) of orbits the for action of \( N \) on the dual space \( n^{\ast} \) of linear functionals on the Lie algebra \( n \) of \( N, \) by the dual of the standard action \( AdN \) on \( n \) by conjugation. This is known as the coadjoint action of \( N, \) and is denoted \( Ad^{\ast} N \); its orbits are often referred to as coadjoint orbits.
Kirillov also gave a reasonably explicit recipe for realizing the representation \( \rho_{\lambda} \) corresponding to the orbit \( O_{\lambda} \) of an element \( \lambda \) of \( n^{\ast}. \) Given \( \lambda, \) one can find Lie subalgebras \( m \subset n, \) such that
- \( m \) contains the Lie algebra \( n_\lambda \) of the stabilizer of \( \lambda \) under \( Ad^\ast N \); and
- \( m \) is subordinate to \( \lambda, \) meaning that \( \lambda_{|[m,m]} = 0 \) (which implies, since for a group acting by unipotent operators, the stabilizer of any point is connected, that \( \lambda \) can be exponentiated to define a one-dimensional unitary character \( \chi_\lambda \) of the group \( M = \exp m \)), and
- \( m \) is of maximal dimension subject to (ii).
Condition (iii) implies that \begin{align*} \dim n - \dim m & = \dim m - \dim n_{\lambda} \\ & = \tfrac{1}{2} (\dim n - \dim n_{\lambda}) = \tfrac{1}{2} \dim O_{\lambda}. \end{align*} (In fact, condition (iii) makes condition (i) redundant; but it is perhaps desirable to make (i) explicit.)
Such a subalgebra m is called a maximal subordinate subalgebra, or a polarizing subalgebra, or simply a polarization, for \( \lambda. \) Given a polarization \( m \) for \( \lambda, \) the representation \( \rho_{\lambda} \) of \( N \) is realizable as the induced representation \[ \operatorname{ind}^N_M \chi_{\lambda} . \] Implicit in this statement, of course, is the fact that this induced representation is independent of the choice of polarization \( m \) of \( \lambda, \) and proving this was an important part of Kirillov’s work.
Kirillov’s theory was gradually adapted to larger and larger classes of solvable groups, until the representation theory of the general connected solvable Lie group was described by the orbit method. The first extension was to exponential solvable groups, by a French team led by P. Bernat [e8], [e13]. An exponential solvable group is a connected group \( S, \) with Lie algebra \( s, \) for which the exponential map \[ \exp: s \to S \] is a diffeomorphism. Here the theory went through almost unchanged, although a slight refinement was needed in the definition of polarization.
The next major step beyond exponential solvable groups was taken by Moore and Louis Auslander [5]. They studied solvable groups of “type R”, which are in some sense the opposite of exponential. One definition of type R is that the roots of the adjoint action for the Lie algebra on itself are all pure imaginary. Here there were several new conceptual issues to overcome, including the fact that polarizations might not exist, and even if they did, the linear functional might not exponentiate to a character of the polarizing group. Also, the key regularity condition of being type I, in the sense of von Neumann algebras generated by unitary representations of the group [e19], which always holds for exponential groups [e4], can fail for type R groups.
However, [5] establishes several notable results, including a criterion for when such a group is type I, and that, when this happens, a description of the construction of irreducible unitary representations. They also investigate the stronger regularity condition of being CCR (see [5], Chapter 5), and show that a solvable group of type R that is type I is also CCR. By contrast, nonnilpotent exponential groups, although always of type 1, are never CCR.
After [5], the theory was pushed further by J. Brezin [e11], who found a criterion for a general solvable group to be type I. Then Auslander and Kostant [e10], [e16] showed how to introduce symplectic geometry ideas, and also a more general notion of polarization, to create an account that, while noticeably more complex, retains the essential flavor of Kirillov’s account of nilpotent groups. Over roughly the following decade, the theory was tied up nicely in several follow-on papers of Pukanszky; see, for example, [e9], [e17].
Classifying irreducible representations is only the start of harmonic analysis. It is also of considerable interest to describe explicitly the unitary representations that appear “in nature”. In particular, if \( H \subset G \) is a subgroup, the question of describing the representations that appear in the natural action of \( G \) on the homogeneous space \( G/H \) presents itself in many situations. Particularly interesting examples are the cases when \( H \) is a discrete subgroup and \( G/H \) is compact. (We say, “\( H \) is cocompact.”) Then the representation of \( G \) on \( L^2 (G/H) \) decomposes into a direct sum of irreducible representations, each of which appears with finite multiplicity. These multiplicities may have arithmetic significance. When \( G \) is a reductive algebraic group and \( H \) is an arithmetic subgroup, we are in the domain of automorphic forms, which have been important for number theory for several hundred years, and a subject of particularly intense study since the 1960s.
Moore considered coset spaces \( G/H \) with \( G = N, \) a nilpotent Lie group (taken to be connected and simply connected), and \( H = \Gamma \) a discrete cocompact subgroup. In this situation, it can be shown that \( N \) can be identified with the completion \( N_{\mathbb{R}} \) over \( \mathbb{R} \) of a unipotent algebraic group \( N_{\mathbb{Q}} \) over \( \mathbb{Q}, \) the rational numbers. Thus, one can consider at the same time the completion \( N_p \) of \( N_{\mathbb{Q}} \) over the field of \( p \)-adic numbers for any prime \( p \); and further one can construct \( N_{\mathbb{A}}, \) the adelic points of \( N_{\mathbb{Q}}, \) as the restricted direct product of the \( N_p. \) Moore did this; he was thus a pioneer in investigating the adelic viewpoint, which since has become standard in the theory of automorphic forms.
Associated to \( N_{\mathbb{Q}}, \) there is a Lie algebra \( n_{\mathbb{Q}}, \) with all the appropriate functorial properties, including an exponential map \( \exp : n_{\mathbb{Q}}\to N_{\mathbb{Q}}, \) that is in fact a polynomial mapping, with polynomial inverse, and which defines exponential maps for all the local completions of \( n_{\mathbb{Q}} \) and \( N_{\mathbb{Q}}, \) and for the adelic groups as well. Further, the dual space \( n^{\ast}_{\mathbb{Q}} \) localizes to the dual spaces over all completions of \( Q, \) and these can be assembled into a global dual \( n^{\ast}_{\mathbb{A}}. \)
A standard part of the construction of adeles is that the natural embeddings \( n_{\mathbb{Q}}\hookrightarrow n_{\mathbb{Q}_p} \) can be assembled into a global embedding \( n_{\mathbb{Q}}\hookrightarrow n_{\mathbb{A}}, \) with image that is discrete and cocompact. The parallel facts hold also for the group \( N_{\mathbb{Q}} \) and the dual space \( n^{\ast}_{\mathbb{Q}}. \)
Moore observes that the orbit method applies more or less without change to all the local groups \( N_{\mathbb{Q}_p}, \) and that these can be assembled into an orbital description of the unitary dual of \( N_{\mathbb{A}}. \) In \( n^{\ast}_{\mathbb{A}}, \) one has a collection of rational points, \( n^{\ast}_{\mathbb{Q}} \); their \( Ad^{\ast} N_{\mathbb{A}} \) orbits provide a collection of “rational orbits” in \( n^{\ast}_{\mathbb{A}}. \) A lovely primary result of [3] is that the Hilbert space \( L^2 (N_{\mathbb{A}} /N_{\mathbb{Q}}) \) decomposes into a multiplicity-free sum of representations parametrized by the rational orbits in \( n^{\ast}_{\mathbb{A}}. \)
The more naive question, to describe the multiplicities of representations in \( L^2 (N_{\mathbb{R}}/\Gamma), \) turns out to be trickier to answer. One might hope for something analogous to the situation when \( N \) is actually commutative, so that \( \Gamma \) is a lattice in \( N, \) and the (one-dimensional) representations that appear are given by the points in the dual lattice \( \Gamma^{\ast} \subset N^{\ast}. \) This is approximately the right idea, but there are a number of complications to be addressed before the correct analog can be formulated.
As noted, the exponential map \( \exp : n \to N \) is a global diffeomorphism. Its inverse will be denoted by \( \log. \) Given a discrete subgroup \( \Gamma \subset N, \) the set \( \log \Gamma \subset n \) need not be closed under addition in \( n \); and likewise, given a lattice \( L \subset n, \) its image \( \exp(L) \) need not be a subgroup of \( N. \) There are examples of both failures already in the three-dimensional Heisenberg group.
On the other hand, it may happen that \( \log\Gamma \) is a lattice. Moore then calls \( \Gamma \) a lattice subgroup. He also shows that, for any \( \Gamma, \) there are lattice subgroups \( \Gamma^{\prime} \subset \Gamma \subset \Gamma^{\prime\prime}, \) such that the index \( \Gamma^{\prime\prime} /\Gamma^{\prime} \) is bounded independent of \( \Gamma \) (for given \( N \)).
For lattice subgroups, the hoped-for parallel with the abelian case holds, in the sense that a representation \( \rho_O \) corresponding to a coadjoint orbit \( O \subset n^{\ast} \) appears with positive multiplicity in \( L^2 (N_{\mathbb{R}} /\Gamma) \) if and only if \( O \cap (\log\Gamma)^{\bot} \) is nonempty, where \( (\log\Gamma)^{\bot} \subset n^{\ast} \) is the dual lattice to \( \log\Gamma. \) However, the multiplicity with which such a representation \( \rho_O \) appears in \( L^2 (N_{\mathbb{ R}} /\Gamma) \) is somewhat tricky to compute. It is not something relatively straightforward, such as the cardinality of the intersection \( O \cap (\log\Gamma)^{\bot}, \) or the number of \( Ad^{\ast} \Gamma \) orbits in this set. The exact description of the multiplicity was given in [e15] and [e14]. It requires selecting an appropriate polarization. The paper [e14] also related this multiplicity to the adelic representation on \( L^2 (N_{\mathbb{ A}} /N_{\mathbb{Q}} ), \) underscoring the appropriateness of Moore’s consideration of the global situation.