• 検索結果がありません。

Analysis on the minimal representation of O(p, q) – II. Branching laws

N/A
N/A
Protected

Academic year: 2022

シェア "Analysis on the minimal representation of O(p, q) – II. Branching laws"

Copied!
39
0
0

読み込み中.... (全文を見る)

全文

(1)

Analysis on the minimal representation of O(p, q)

– II. Branching laws

Toshiyuki KOBAYASHI

RIMS, Kyoto University, Sakyo-ku, Kyoto, 606-8502, Japan

Bent ØRSTED

Department of Mathematics and Computer Science, SDU - Odense University, Campusvej 55, DK-5230, Odense M, Denmark

Abstract

This is a second paper in a series devoted to the minimal unitary representation of O(p, q). By explicit methods from conformal geometry of pseudo Riemannian man- ifolds, we find the branching law corresponding to restricting the minimal unitary representation to natural symmetric subgroups. In the case of purely discrete spec- trum we obtain the full spectrum and give an explicit Parseval-Plancherel formula, and in the general case we construct an infinite discrete spectrum.

Contents

Introduction

§4. Criterion for discrete decomposable branching laws

§5. Minimal elliptic representations of O(p, q)

§6. Conformal embedding of the hyperboloid

§7. Explicit branching formula (discretely decomposable case)

§8. Inner product on $p,q and the Parseval-Plancherel formula

§9. Construction of discrete spectra in the branching laws

Email addresses: toshi@kurims.kyoto-u.ac.jp (Toshiyuki KOBAYASHI), orsted@imada.sdu.dk (Bent ØRSTED).

(2)

Introduction

This is the second in a series of papers devoted to the analysis of the mini- mal representation$p,q of O(p, q). We refer to [24] for a general introduction;

also the numbering of the sections is continued from that paper, and we shall refer back to sections there. However, the present paper may be read inde- pendently from [24], and its main object is to study the branching law for the minimal unitary representation $p,q from analytic and geometric point of view. Namely, we shall find by explicit means, coming from conformal geom- etry, the restriction of $p,q with respect to the symmetric pair

(G, G0) = (O(p, q), O(p0, q0)×O(p00, q00)).

If one of the factors inG0 is compact, then the spectrum is discrete (see The- orem 4.2 also for an opposite implication), and we find the explicit branching law; when both factors are non-compact, there will still (generically) be an infinite discrete spectrum, which we also construct (conjecturally almost all of it; see §9.8). We shall see that the (algebraic) situation is similar to the theta-correspondence, where the metaplectic representation is restricted to analogous subgroups.

Let us here state the main results in a little more precise form, referring to sections 8 and 9 for further notation and details.

Theorem A (the branching law for O(p, q) ↓ O(p, q0)× O(q00); see Theo- rem 7.1) If q00 ≥ 1 and q0 +q00 = q, then the twisted pull-back Φf1 of the local conformal map Φ1 between spheres and hyperboloids gives an explicit ir- reducible decomposition of the unitary representation $p,q when restricted to O(p, q0)×O(q00):

(Φ]1) :$p,q|O(p,q0)×O(q00)

X l=0

πp,q0

+,l+q002 −1Hl(Rq00).

The representations appearing in the decompositions are in addition to usual spherical harmonicsHl(Rq) for compact orthogonal groupsO(q), also the rep- resentations π+,λp,q for non-compact orthogonal groups O(p, q). The latter ones may be thought of as discrete series representations on hyperboloids

X(p, q) :={x= (x0, x00)∈Rp+q :|x0|2− |x00|2 = 1}

for λ >0 or their analytic continuation for λ ≤0; they may be also thought of as cohomologically induced representations from characters of certain θ- stable parabolic subalgebras. The fact that they occur in this branching law gives a different proof of the unitarizability of these modules π+,λp,q for λ >

(3)

−1, once we know $p,q is unitarizable (cf. Part I, Theorem 3.6.1). It might be interesting to remark that the unitarizability for λ < 0 (especially, λ =

12 in our setting) does not follow from a general unitarizability theorem on Zuckerman-Vogan’s derived functor modules [35], neither from a general theory of harmonic analysis on semisimple symmetric spaces.

Our intertwining operator (Φ]1) in Theorem A is derived from a conformal change of coordinate (see §6 for its explanation) and is explicitly written.

Therefore, it makes sense to ask also about the relation of unitary inner prod- ucts between the left-hand and the right-hand side in the branching formula.

Here is an answer (see Theorem 8.6): We normalize the inner product k kπp,q+,λ

(see (8.4.2)) such that for λ >0,

kfk2π+,λp,q =λkfk2L2(X(p,q)), for any f ∈(πp,q+,λ)K.

Theorem B (the Parseval-Plancherel formula forO(p, q)↓O(p, q0)×O(q00)) 1) If we develop F ∈Ker∆eM asF =Pl Fl(1)Fl(2) according to the irreducible decomposition in Theorem A, then we have

kFk2$p,q =

X

l=0

kFl(1)k2πp,q0

+,l+q00 2 1

k Fl(2)k2L2(Sq00−1).

2) In particular, if q00 ≥ 3, then all of πp,q0

+,l+q002 −1 are discrete series for the hyperboloid X(p, q0) and the above formula amounts to

kFk2$p,q =

X

l=0

(l+q00

2 −1) kFl(1)k2L2(X(p,q0))k Fl(2)k2L2(Sq00−1).

The formula may be also regarded as an explicit unitarization of the minimal representation$p,qon the “hyperbolic space model” by means of the right side (for an abstract unitarization of$p,q, it suffices to choose a single pair (q0, q00)).

We note that the formula was previously known in the case where (q0, q00) = (0, q) (namely, when each summand in the right side is finite dimensional) by Kostant, Binegar-Zierau by a different approach. The formula is new and seems to be particularly interesting even in the special case q00 = 1, where the minimal representation $p,q splits into two irreducible summands when restricted toO(p, q−1)×O(1).

In Theorem 9.1, we consider a more general setting and prove:

Theorem C (discrete spectrum in the restrictionO(p, q)↓O(p0, q0)×O(p00, q00)) The twisted pull-back of the locally conformal diffeomorphism also constructs

X λ∈A0(p0,q0)∩A0(q00,p00)

π+,λp0,q0 πp−,λ00,q00X

λ∈A0(q0,p0)∩A0(p00,q00)

πp−,λ0,q0 π+,λp00,q00

(4)

as a discrete spectrum in the branching law for the non-compact case.

Even in the special case (p00, q00) = (0,1), our branching formula includes a new and mysterious construction of the minimal representation on the hyperboloid as below (see Corollary 7.2.1): Let Wp,r be the set of K-finite vectors (K = O(p)×O(r)) of the kernel of the Yamabe operator

Ker∆eX(p,r) ={f ∈C(X(p, r)) : ∆X(p,r)f = 1

4(p+r−1)(p+r−3)f}, on which the isometry groupO(p, r) and the Lie algebra of the conformal group O(p, r+ 1) act. The following Proposition is a consequence of Theorem 7.2.2 by an elementary linear algebra.

Proposition D Let m >3 be odd. There is a long exact sequence 0→W1,m−1 −→ϕ1 W2,m−2 −→ϕ2 W3,m−3 −→ · · ·ϕ3 ϕ−→m2 Wm−1,1 ϕ−→m1 0 such thatKerϕp is isomorphic to($p,q)K for any(p, q)such thatp+q=m+1.

We note that each representation spaceWp,q−1 is realized on a different space X(p, q−1) whose isometry groupO(p, q−1) varies according top(1≤p≤m).

So, one may expect that only the intersections of adjacent groups can act (infinitesimally) on Kerϕp. Nevertheless, a larger group O(p, q) can act on a suitable completion of Kerϕp, giving rise to another construction of the minimal representation on the hyperboloid X(p, q−1) = O(p, q −1)/O(p− 1, q−1) ! We note that Kerϕp is roughly half the kernel of the Yamabe operator on the hyperboloid (see§7.2 for details).

We briefly indicate the contents of the paper: In section 4 we recall the relevant facts about discretely decomposable restrictions from [17] and [18], and apply the criteria to our present situation. In particular, we calculate the associated variety of$p,q as well as its asymptoticK-support introduced by Kashiwara- Vergne. Theorem 4.2 and Corollary 4.3 clarify the reason why we start with the subgroupG0 =O(p, q0)×O(q00) (i.e.p00= 0). Section 5 contains the identi- fication of the representationsπp,q+,λandπp,q ofO(p, q) in several ways, namely as: derived functor modules, Dolbeault cohomologies, eigenspaces on hyper- boloids, and quotients or subrepresentations of parabolically induced modules.

In section 6 we give the main construction of embedding conformally a direct product of hyperboloids into a product of spheres; this gives rise to a canoni- cal intertwining operator between solutions to the so-called Yamabe equation, studied in connection with conformal differential geometry, on conformally re- lated spaces. Applying this principle in section 7 we obtain the branching law in the case where one factor inG0is compact, and in particular when one factor is justO(1). In this case we have Corollary 7.2.1, stating that$p,qrestricted to O(p, q−1) is the direct sum of two representations, realized in even respectively odd functions on the hyperboloid forO(p, q−1). Note here the analogy with the

(5)

metaplectic representation. Also note here Theorem 7.2.2, which gives a mys- terious extention of$p,qby$p+1,q−1 - both inside the space of solutions to the Yamabe equation on the hyperboloidX(p, q−1) = O(p, q−1)/O(p−1, q−1).

We also point out that the representationsπ+,λp,q forλ= 0,−12 are rather excep- tional; they are unitary, but outside the usual “fair range” for derived functor modules, see the remarks in section 8.4. Section 8 contains a proof of Theo- rem 3.9.1 of [24] on the spectra of the Knapp-Stein intertwining operators and gives the explicit Parseval-Plancherel formulas for the branching laws. Finally, in section 9 we use certain Sobolev estimates to construct infinitely many dis- crete spectra when both factors in G0 are non-compact. We also conjecture the form of the full discrete spectrum (true in the case of a compact factor). It should be interesting to calculate the full Parseval-Plancherel formula in the case of both discrete and continuous spectrum.

The first author expresses his sincere gratitude to SDU - Odense University for the warm hospitality.

4 Criterion for discrete decomposable branching laws

4.1 Our object of study is the discrete spectra of the branching law of the restriction$p,qwith respect to a symmetric pair (G, G0) = (O(p, q), O(p0, q0)× O(p00, q00)). The aim of this section is to give a necessary and sufficient condition onp0, q0, p00 and q00 for the branching law to be discretely decomposable.

We start with general notation. LetGbe a linear reductive Lie group, and G0 its subgroup which is reductive in G. We take a maximal compact subgroup K of G such that K0 := K ∩G0 is also a maximal compact subgroup. Let g0 = k0+p0 be a Cartan decomposition, and g = k+p its complexification.

Accordingly, we have a direct decompositiong =k+p of the dual spaces.

Let π ∈ G. We say that the restrictionb π|G0 is G0-admissible if π|G0 splits into a direct Hilbert sum of irreducible unitary representations of G0 with each multiplicity finite (see [16]). As an algebraic analogue of this notion, we say the underlying (g, K)-module πK isdiscretely decomposable as a (g0, K0)- module, ifπK is decomposed into an algebraic direct sum of irreducible (g0, K0)- modules (see [18]). We note that if the restriction π|K0 isK0-admissible, then the restrictionπ|G0 is alsoG0-admissible ([16], Theorem 1.2) and the underlying (g, K)-moduleπK is discretely decomposable (see [18], Proposition 1.6). Here are criteria forK0-admissibility and discrete decomosability:

Fact 4.1 (see [17], Theorem 2.9 for (1); [18], Corollary 3.4 for (2))

1)IfASK(π)∩Ad(K)(k0)= 0, thenπisK0-admissible and alsoG0-admissible.

2)IfπKis discretely decomposable as a(g0, K0)-module, thenprg→g0(VgK))⊂

(6)

Ng0.

Here, ASK(π) is the asymptotic cone of

SuppK(π) := {highest weight of τ ∈Kd0 : [π|K0 :τ]6= 0}

whereK0 is the identity component of K, and (k0)⊂k is the annihilator of k0. Let Np (⊂p) be the nilpotent cone for p. VgK) denotes the associated variety of πK, which is an Ad(KC)-invariant closed subset of Np. We write the projection prp→p0 p →p0∗ dual to the inclusion p0 ,→p.

4.2 Let us consider our setting whereπ=$p,qand (G, G0) = (O(p, q), O(p0, q0)× O(p00, q00)).

Theorem 4.2 Suppose p0+p00=p (≥ 2), q0+q00 =q (≥2) and p+q∈ 2N. Then the following three conditions on p0, q0, p00, q00 are equivalent:

i) $p,q is K0-admissible.

ii) $Kp,q is discretely decomposable as a (g0, K0)-module.

iii) min(p0, q0, p00, q00) = 0.

The implication (i) ⇒ (ii) holds by a general theory as we explained ([18], Proposition 1.6); (ii)⇒ (iii) will be proved in §4.4, and (iii) ⇒ (i) in§4.5, by an explicit computation of the asymptotic cone ASK($p,q) and the associated variety Vg($Kp,q) which are used in Fact 4.1.

Remark Analogous results to the equivalence (i)⇔(ii) in Theorem 4.2 were first proved in [18], Theorem 4.2 in the setting where (G, G0) is any reductive symmetric pair and the representation is any Aq(λ) module in the sense of Zuckerman-Vogan, which may be regarded as “representations attached to elliptic orbits”. We note that our representations $p,q are supposed to be attached to nilpotent orbits. We refer [21], Conjecture A to relevant topics.

4.3 The following corollary is a direct consequence of Theorem 4.2, which will be an algebraic background for the proof of the explicit branching law (Theorem 7.1).

Corollary 4.3 Suppose that min(p0, q0, p00, q00) = 0.

1) The restriction of the unitary representation $p,q|G0 is also G0-admissible.

2) The space of K0-finite vectors $p,qK0 coincides with that of K-finite vectors

$Kp,q.

PROOF. See [16], Theorem 1.2 for (1); and [18], Proposition 1.6 for (2). 2

(7)

A geometric counterpart of Corollary 4.3 (2) is reflected as the removal of singularities of matrix coefficients for the discrete spectra in the analysis that we study in §6; namely, any analytic function defined on an open subset M+ (see §6 for notation) of M which is a K0-finite vector of a discrete spectrum, extends analytically onM ifp00 = 0. The reason for this is not only the decay of matrix coefficients but a matching condition of the leading terms fort→ ±∞. This is not the case for min(p0, q0, p00, q00)>0 (see §9).

4.4 Proof of (ii) ⇒ (iii) in Theorem 4.2.

We identify p with p via the Killing form, which is in turn identified with M(p, q;C) by

M(p, q;C)→ p, X 7→

O X

tX O

.

Then the nilpotent cone Np corresponds to the following variety:

{X ∈M(p, q;C) : both XtX and tXX are nilpotent matrices}. (4.4.1) We put

M0,0(p, q;C) :={X ∈M(p, q;C) :XtX=O,tXX =O}.

ThenM0,0(p, q;C)\ {O}is the uniqueKC 'O(p,C)×O(q,C)-orbit of dimen- sionp+q−3. The associated varietyVg($p,qK ) of$p,q is of dimensionp+q−3, which follows easily from theK-type formula of$p,q(see [24], Theorem 3.6.1).

Thus, we have proved:

Lemma 4.4 The associated variety Vg($p,qK ) equals M0,0(p, q;C).

The projection prp→p0 :p →p0∗ is identified with the map

prp→p0 :M(p, q;C)→M(p0, q0;C)⊕M(p00, q00;C),

X1 X2 X3 X4

7→(X1, X4).

Suppose p0p00q0q00 6= 0. If we take X :=E1,1−Ep0+1,q0+1+√

−1Ep0+1,1+√

−1E1,q0+1 ∈M0,0(p, q;C), then prpp0(X) = (E1,1,−Ep0+1,q0+1). But E1,1 6∈ No(p 0,q0) and −Ep0+1,q0+1 6∈

No(p 00,q00). Thus, prg→g0(X)6∈ Ng0. It follows from Fact 4.1 (2) that $Kp,q is not discrete decomposable as a (g0, K0)-module. Hence (ii)⇒ (iii) in Theorem 4.2 is proved. 2

(8)

4.5 Proof of (iii) ⇒ (i) in Theorem 4.2.

We take an orthogonal complementary subspace k000 of k00 in k0 'o(p) +o(q).

Lettc0 be a Cartan subalgebra of k0 such thatt000 :=tc0∩k000 is a maximal abelian subspace ink000. We choose a positive system ∆+(k,tc) which is compatible with a positive system of the restricted root system Σ(k,t00). Then we can find a basis {fi : 1≤i≤[p2] + [q2]}on√

−1t0 such that a positive root system of kis given by

4+(k,tc) ={fi±fj : 1≤i < j ≤[p 2]}

∪{fi±fj : [p

2] + 1≤i < j ≤[p 2] + [q

2]}

{fl : 1≤l≤[p

2]} (p:odd)

{fl : [p

2] + 1 ≤l ≤[p 2] + [q

2]} (q:odd)

, and such that

√−1(t000) =

min(p0,p00)

X

j=1

Rfi+

min(q0,q00)

X

j=1

Rf[p

2]+j (4.5.1)

if we regard (t000) as a subspace of (tc0) by the Killing form.

Supposep0q0p00q00 = 0. Without loss of generality we may and do assumep00= 0, namely,G0 =O(p, q0)×O(q00) withq0+q00=q.

Let us first consider the case p6= 2. Then the irreducible O(p)-representation Ha(Rp) remains irreducible when restricted toSO(p). The corresponding high- est weight is given by af1. It follows from the K-type formula of $p,q (Theo- rem 3.6.1) that

SuppK($p,q) ={af1+bf[p

2]+1:a, b∈N, a+p

2 =b+q 2}. Therefore we have proved:

ASK($p,q) =R+(f1+f[p

2]+1). (4.5.2)

Then ASK($p,q)∩√

−1(t000) ={0} from (4.5.1) and (4.5.2), which implies ASK($p,q)∩√

−1 Ad(K)(k00)={0}

because (t000)meets any Ad(K)-orbit through (k00). Therefore, the restriction

$p,q|K0 isK0-admissible by Fact 4.1 (1).

Ifp= 2, then$p,q splits into two representations (see Remark 3.7.3), say$+2,q and $2,q, when restricted to the connected component SO0(2, q). Likewise,

(9)

Ha(Rp) is a direct sum of two one-dimensional representations when restricted toSO(2) ifa ≥1. Then we have

ASK($±2,q) =R+(±f1+fp+1).

Applying Fact 4.1 (1) to the identity components (G0, G00) of groups (G, G0), we conclude that the restriction$±p,q|K00 isK00-admissible. Hence the restriction

$p,q|K0 is also K0-admissible. Thus, (iii) ⇒ (i) in Theorem 4.2 is proved.

Now the proof of Theorem 4.2 is completed. 2

5 Minimal elliptic representations of O(p, q)

5.1 In this section, we introduce a family of irreducible representations of G = O(p, q), denoted by πp,q+,λ, πp,q, for λ ∈ A0(p, q), in three different realizations. These representations are supposed to be attached to minimal elliptic orbits, for λ > 0 in the sense of the Kirillov-Kostant orbit method.

Here, we set

A0(p, q) :=

{λ∈Z+p+q2 :λ >−1} (p > 1, q6= 0), {λ∈Z+p+q2 :λ≥ p2 −1} (p > 1, q= 0),

∅ (p= 1, q6= 0) or (p= 0),

{−12,12} (p= 1, q= 0).

(5.1.1)

It seems natural to include the parameterλ= 0,−12 in the definition ofA0(p, q) as above, although λ =−12 is outside the weakly fair range parameter in the sense of Vogan [37]. Cohomologically induced representations forλ=−12 and λ = −1 will be discussed in details in a subsequent paper. In particular, the caseλ=−1 is of importance in another geometric construction of the minimal representation via Dolbeault cohomology groups (see Part I, Introduction, Theorem B (4)).

5.2 Let Rp,q be the Euclidean space Rp+q equipped with the flat pseudo- Riemannian metric:

gRp,q =dv02+· · ·+dvp12−dvp2− · · · −dvp+q12. We define a hyperboloid by

X(p, q) := {(x, y)∈Rp,q :|x|2− |y|2 = 1}.

We note X(p,0) ' Sp−1 and X(0, q) = ∅. If p = 1, then X(p, q) has two connected components. The groupGacts transitively onX(p, q) with isotropy

(10)

subgroup O(p−1, q) at

xo :=t(1,0,· · · ,0). (5.2.1) ThusX(p, q) is realized as a homogeneous manifold:

X(p, q)'O(p, q)/O(p−1, q).

We induce a pseudo-Riemannian metric gX(p,q) onX(p, q) from Rp,q (see [24],

§3.2), and write ∆X(p,q) for the Laplace-Beltrami operator on X(p, q). As in [24], Example 2.2, the Yamabe operator is given by

eX(p,q) = ∆X(p,q)− 1

4(p+q−1)(p+q−3). (5.2.2) Forλ ∈C, we set

Cλ(X(p, q)) := {f ∈C(X(p, q)) : ∆X(p,q)f = (−λ2+ 1

4(p+q−2)2)f}

={f ∈C(X(p, q)) :∆eX(p,q)f = (−λ2+ 1

4)f}. (5.2.3) Furthermore, for=±, we write

Cλ,(X(p, q)) :={f ∈Cλ(X(p, q)) :f(−z) = f(z), z∈X(p, q)}. Then we have a direct sum decomposition

Cλ(X(p, q)) =Cλ,+ (X(p, q)) +Cλ,(X(p, q)) (5.2.4) and each space is invariant under left translations of the isometry group G becauseGcommutes with ∆X(p,q). With the notation in§3.5, we note ifq= 0, then Cλ,sgn(−1) k(X(p,0)) is finite dimensional and isomorphic to the space of spherical harmonics:

Hk(Rp)'Cλ(X(p,0)) =Cλ,sgn(−1) k(X(p,0)) (k :=λ+ p−2 2 ).

5.3 Let G = O(p, q) where p, q ≥ 1 and let θ be the Cartan involution corresponding to K = O(p)×O(q). We extend a Cartan subalgebra tc0 of k0 (given in §4.5) to that of g0, denoted by hc0. If both p and q are odd, then dimhc0 = dimtc0 + 1; otherwise hc0 = tc0. The complexification of hc0 is denoted byhc.

We can take a basis {fi : 1≤i≤[p+q2 ]} of (hc) (see §4.5; by a little abuse of notation if both pand q are odd) such that the root system ofg is given by

(11)

4(g,hc) = {±(fi±fj) : 1 ≤i < j ≤[p+q 2 ]}

{±fl : 1≤l ≤[p+q

2 ]} (p+q:odd)

.

Let {Hi} ⊂ hc be the dual basis for {fi} ⊂ (hc). Set t := CH1 (⊂ tc ⊂ hc).

Then the centralizer L of t in G is isomorphic to SO(2)×O(p−2, q). Let q=l+ube aθ-stable parabolic subalgebra ofgwith nilpotent radicalu given by

4(u,hc) := {f1 ±fj : 2≤j ≤[p+q

2 ]} ∪({f1} (p+q:odd) ), and with a Levi partl=l0⊗C given by

l0 ≡Lie(L)'o(2) +o(p−2, q).

Any character of the Lie algebra l0 (or any complex character of l) is deter- mined by its restriction to hc0. So, we shall write Cν for the character of the Lie algebra l0 whose restriction to hc is ν ∈ (hc). With this notation, the character ofL acting on ∧dimuu is written as C2ρ(u) where

ρ(u) := (p+q

2 −1)f1. (5.3.1)

The homogeneous manifoldG/Lcarries aG-invariant complex structure with canonical bundle ∧topTG/L ' G ×L C2ρ(u). As an algebraic analogue of a Dolbeault cohomology of a G-equivariant holomorphic vector bundle over a complex manifold G/L, Zuckerman introduced the cohomological parabolic inductionRjqRgq

j

(j ∈N), which is a covariant functor from the category of metaplectic (l,(L∩K))-modules to that of (g, K)-modules. Here, Le is a metaplectic covering of L defined by the character of L acting on ∧dimuu ' C2ρ(u). In this paper, we follow the normalization in [36], Definition 6.20 which is different from the one in [34] by a ‘ρ-shift’.

The character Cλf1 of l0 lifts to a metaplectic (l,(L∩ K))-module if and only if λ∈Z+p+q2 . In particular, we can define (g, K)-modules Rjq(Cλf1) for λ∈A0(p, q). The Z(g)-infinitesimal character of Rjq(Cλf

1) is given by (λ,p+q

2 −2,p+q

2 −3, . . . ,p+q

2 −[p+q

2 ])∈(hc)

in the Harish-Chandra parametrization if it is non-zero. In the sense of Vogan [37], we have

Cλf1 is in the good range ⇔ λ > p+q 2 −2, Cλf1 is in the weakly fair range ⇔ λ≥0.

(12)

We note that Rjq(Cλf1) = 0 if j 6=p−2 and if λ∈A0(p, q). This follows from a general result in [35] forλ ≥0; and [15] forλ=−12.

5.4 For b ∈ Z, we define an algebraic direct sum of K = O(p)×O(q)- modules by

Ξ(K :b)≡Ξ(O(p)×O(q) :b) := M

m,nN m−n≥b m−n≡b mod 2

Hm(Rp)Hn(Rq). (5.4.1)

Forλ ∈A0(p, q), we put

b≡b(λ, p, q) :=λ− p 2 +q

2 + 1∈Z, (5.4.2)

≡(λ, p, q) := (−1)b. (5.4.3)

We define the line bundle Ln over G/Lby the character nf1 of L (see §5.3).

Here is a summary for different realizations of the representation π+,λp,q: Fact 5.4 Let p, q ∈N (p > 1).

1) For any λ∈A0(p, q), each of the following 5 conditions defines uniquely a (g, K)-module, which are mutually isomorphic. We shall denote it by (πp,q+,λ)K. The (g, K)-module(π+,λp,q)K is non-zero and irreducible.

i) A subrepresentation of the degenerate principal representationIndGPmax(⊗Cλ) (see §3.7) with K-type Ξ(K :b).

i)0 A quotient of IndGPmax(⊗Cλ) with K-type Ξ(K :b).

ii) A subrepresentation of Cλ(X(p, q))K with K-type Ξ(K :b).

iii) The underlying (g, K)-module of the Dolbeault cohomology group Hp−2¯ (G/L,L(λ+p+q2

2 ))K.

iii)0 The Zuckerman-Vogan derived functor module Rp−2q (Cλf

1).

2) In the realization of (ii), if f ∈(πp,q+,λ)K, then there exists an analytic func- tion a∈C(Sp−1×Sq−1) such that

f(ωcosht, ηsinht) =a(ω, η)e(λ+ρ)t(1 +te2tO(1)) as t→ ∞. Here, we put ρ= p+q−22 .

For details, we refer, for example, to [12] for (i) and (i)0; to [31] for (ii) and also for a relation with (i) (under some parity assumption on eigenspaces); to [16], §6 (see also [15]) for (iii)0 ⇔(ii); and to [38] for (iii) ⇔(iii)0. The second statement follows from a general theory of the boundary value problem with regular singularities; or also follows from a classical asymptotic formula of hypergeometric functions (see (8.3.1)) in our specific setting.

(13)

Remark 1) By definition, (i) and (i)0 make sense forp > 1 and q > 0; and others forp > 1 and q≥0.

2) Each of the realization (i), (i)0, (ii), and (iii) also gives a globalization of πp,q+,λ, namely, a continuous representation of G on a topological vector space.

Because all of (π+,λp,q)K (λ∈A0(p, q)) are unitarizable we may and do take the globalizationπ+,λp,q to be the unitary representation of G.

3) If λ > 0 and λ∈ A0(p, q), then the realization (ii) of π+,λp,q gives a discrete series representation forX(p, q). Conversely,

p,q+,λ:λ∈A0(p, q), λ >0}

exhausts the set of discrete series representations forX(p, q).

If (p, q) = (1,0), then O(p, q)'O(1) and it is convenient to define represen- tations of O(1) by

π1,0+,λ=

1 (λ=−12), sgn (λ= 12), 0 (otherwise).

As we defined π+,λp,q in Fact 5.4, we can also define an irreducible unitary representation, denoted by π−,λp,q, for λ ∈ A0(q, p) such that the underlying (g, K)-module has the following K-type

M

m,n∈N mn≤−λ+q2p21 mn≡−λ+q2p21 mod 2

Hm(Rp)Hn(Rq).

Similarly to πp,q+,λ, the representations πp,q are realized in function spaces on another hyperboloid O(p, q)/O(p, q−1).

In order to understand the notation here, we remark:

i) π−,λp,q ∈O(p, q) corresponds to the representation\ π+,λq,p ∈O(q, p) if we identify\ O(p, q) with O(q, p).

ii) π+,λp,0 ' Hk(Rp),where k =λ−p−22 and p≥1, k ∈N.

5.5 The case λ=±12 is delicate, which happens when p+q∈2N+ 1.

First, we assume p+q ∈2N+ 1. By using the equivalent realizations of π+,λp,q in Fact 5.4 and by the classification of the composition series of the most degenerate principal series representation IndGPmax(⊗Cλ) (see [12]), we have

(14)

non-splitting short exact sequences of (g, K)-modules:

0→(πp,q

,12)K →IndGPmax((−1)p−q+12 ⊗C1

2)→(πp,q+,1 2

)K →0, (5.5.1) 0→(πp,q+,

12)K →IndGPmax((−1)p2q1 ⊗C1

2)→(πp,q

,12)K →0. (5.5.2) Because π+,λp,q (λ ∈ A0(p, q)) is self-dual, the dual (g, K)-modules of (5.5.1) and (5.5.2) give the following non-splitting short exact sequences of (g, K)- modules:

0→(πp,q+,1 2

)K →IndGPmax((−1)p−q+12 ⊗C1

2)→(πp,q

,12)K →0, (5.5.3) 0→(πp,q

,12)K →IndGPmax((−1)p2q1 ⊗C1

2)→(π+,p,q

12)K →0. (5.5.4) Next, we assume p+q ∈2N. Then, $p,q is realized as a subrepresentation of some degenerate principal series (see [24], Lemma 3.7.2). More precisely, we have non-splitting short exact sequences of (g, K)-modules

0→$Kp,q→IndGPmax((−1)p2q ⊗C−1)→p,q,1)K⊕(πp,q+,1)K→0, (5.5.5) 0→p,q,1)K⊕(πp,q+,1)K→IndGPmax((−1)p−q2 ⊗C1)→$Kp,q→0, (5.5.6) and an isomorphism of (g, K)-modules:

IndGPmax((−1)p−q+22 ⊗C0)'(πp,q,0)K⊕(πp,q+,0)K. (5.5.7) These results will be used in another realization of the unipotent representa- tion $p,q, namely, as a submodule of the Dolbeault cohomology group in a subsequent paper (cf. Part 1, Introduction, Theorem B (4)).

6 Conformal embedding of the hyperboloid

This section prepares the geometric setup which will be used in §7 and§9 for the branching problem of $p,q|G0. Throughout this section, we shall use the following notation:

|x|2 :=|x0|2+|x00|2 =

p0

X

i=1

(x0i)2+

p00

X

j=1

(x00j)2, for x:= (x0, x00)∈Rp0+p00 =Rp,

|y|2 :=|y0|2+|y00|2 =

q0

X

i=1

(yi0)2+

q00

X

j=1

(yj00)2, for y:= (y0, y00)∈Rq0+q00 =Rq.

(15)

6.1 We define two open subsets of Rp+q by Rp

0+p00,q0+q00

+ :={(x, y) = ((x0, x00),(y0, y00))∈Rp0+p00,q0+q00 :|x0|>|y0|}, Rp

0+p00,q0+q00

:={(x, y) = ((x0, x00),(y0, y00))∈Rp0+p00,q0+q00 :|x0|<|y0|}. Then the disjoint unionRp

0+p00,q0+q00

+ ∪Rp

0+p00,q0+q00

is open dense inRp+q. Let us consider the intersection ofRp

0+p00,q0+q00

± with the submanifoldsM and Ξ given in§3.2:

M ⊂Ξ⊂Rp,q.

Then, we define two open subsets of M 'Sp−1 ×Sq−1 by M±:=M ∩Rp

0+p00,q0+q00

± . (6.1.1)

Likewise, we define two open subsets of the cone Ξ by Ξ± := Ξ∩Rp

0+p00,q0+q00

± . (6.1.2)

We notice that if (x, y) = ((x0, x00),(y0, y00))∈Ξ then

|x0|>|y0| ⇐⇒ |x00|<|y00|

because|x0|2+|x00|2 =|y0|2+|y00|2. The following statement is immediate from definition:

Ξ+ =∅ ⇔ M+ =∅ ⇔ p0q00 = 0. (6.1.3) Ξ =∅ ⇔ M =∅ ⇔ p00q0 = 0. (6.1.4) 6.2 We embed the direct product of hyperboloids

X(p0, q0)×X(q00, p00) = {((x0, y0),(y00, x00)) :|x0|2 − |y0|2 =|y00|2− |x00|2 = 1}. into Ξ+ (⊂Rp,q) by the map

X(p0, q0)×X(q00, p00),→Ξ+, ((x0, y0),(y00, x00))7→(x0, x00, y0, y00). (6.2.1) The image is transversal to rays (see [24],§3.3 for definition) and the induced pseudo-Riemannian metricgX(p0,q0)×X(q00,p00) onX(p0, q0)×X(q00, p00) has signa- ture (p0 −1, q0) + (p00, q00−1) = (p−1, q−1). With the notation in §5.2, we have

gX(p0,q0)×X(q00,p00) =gX(p0,q0)⊕(−gX(q00,p00)).

We note that if p00 = q0 = 0, then X(p0, q0)×X(q00, p00) is diffeomorphic to Sp−1×Sq−1, andgX(p0,0)×X(q00,0) is nothing but the pseudo-Riemannian metric gSp−1×Sq−1 of signature (p−1, q−1) (see [24], §3.3).

(16)

By the same computation as in (3.4.1), we have the relationship among the Yamabe operators on hyperboloids (see also (5.2.2)) by

eX(p0,q0)×X(q00,p00) =∆eX(p0,q0)−∆eX(q00,p00). (6.2.2) We denote by Φ1 the composition of (6.2.1) and the projection Φ : Ξ → M (see [24], (3.2.4)), namely,

Φ1 :X(p0, q0)×X(q00, p00),→M, ((x0, y0),(y00, x00))7→ (x0, x00)

|x| ,(y0, y00)

|y|

!

. (6.2.3) Lemma 6.2 1) The map Φ1 :X(p0, q0)×X(q00, p00)→M is a diffeomorphism onto M+. The inverse map Φ−11 : M+→X(p0, q0)×X(q00, p00) is given by the formula:

((u0, u00),(v0, v00))7→

(u0, v0)

q|u0|2− |v0|2, (v00, u00)

q|v00|2− |u00|2

. (6.2.4)

2) Φ1 is a conformal map with conformal factor |x|1 = |y|1, where x = (x0, x00)∈Rp0+p00 and y= (y0, y00)∈Rq0+q00. Namely, we have

Φ1(gSp−1×Sq−1) = 1

|x|2gX(p0,q0)×X(q00,p00).

PROOF. The first statement is straightforward in light of the formula

|u0|2− |v0|2 =|v00|2− |u00|2 >0 for (u, v) = ((u0, u00),(v0, v00))∈M+ ⊂Sp−1×Sq−1.

The second statement is a special case of Lemma 3.3. 2

6.3 Now, the conformal diffeomorphism Φ1 : X(p0, q0)×X(q00, p00)→ M+ establishes a bijection of the kernels of the Yamabe operators owing to Propo- sition 2.6:

Lemma 6.3 Φf1 gives a bijection from Ker∆eM+ onto Ker∆eX(p0,q0)×X(q00,p00). Here, the twisted pull-backs Φf1 and ^

−11 ) (see Definition 2.3), namely, Φf1 : C(M+) →C(X(p0, q0)×X(q00, p00)), (6.3.1) (Φ^−11 ) : C(X(p0, q0)×X(q00, p00)) → C(M+), (6.3.2)

(17)

are given by the formulae

f1F)(x0, y0, y00, x00) := (|x0|2+|x00|2)p+q−44 F

(x0, x00)

q|x0|2+|x00|2, (y0, y00)

q|y0|2+|y00|2

,

(^

11)f)(u0, u00, v0, v00) := (|u0|2− |v0|2)p+q−44 f

(u0, v0)

q|u0|2− |v0|2, (u00, v00)

q|v00|2− |u00|2

,

respectively. We remark that(Φ^−11 ) = (Φf1)−1.

6.4 Similarly to §6.2, we consider another embedding

X(q0, p0)×X(p00, q00),→Ξ, ((y0, x0),(x00, y00))7→(x0, x00, y0, y00). (6.4.1) The composition of (6.4.1) and the projection Φ : Ξ→M is denoted by

Φ2 :X(q0, p0)×X(p00, q00),→M, ((y0, x0),(x00, y00))7→ (x0, x00)

|x| ,(y0, y00)

|y|

!

. (6.4.2) Obviously, results analogous to Lemma 6.2 and Lemma 6.3 hold for Φ2. For example, here is a lemma parallel to Lemma 6.2:

Lemma 6.4 The map Φ2 : X(q0, p0)×X(p00, q00) → M is a conformal dif- feomorphism onto M. The inverse map Φ21 :M →X(q0, p0)×X(p00, q00) is given by

((u0, u00),(v0, v00))7→

(v0, u0)

q|v0|2− |u0|2, (u00, v00)

q|u00|2− |v00|2

.

7 Explicit branching formula (discrete decomposable case)

If one of p0, q0, p00 or q00 is zero, then the restriction $p,q|G0 is decomposed discretely into irreducible representations of G0 = O(p0, q0)×O(p00, q00) as we saw in §4. In this case, we can determine the branching laws of $p,q|G0 as follows:

Theorem 7.1 Let p+q ∈ 2N. If q00 ≥ 1 and q0 +q00 = q, then we have an irreducible decomposition of the unitary representation$p,q when restricted to

参照

関連したドキュメント

This concept of generalized sign is then used to characterize the entropy condition for discontinuous solutions of scalar conservation laws.. Keywords: Colombeau algebra,

Classical definitions of locally complete intersection (l.c.i.) homomor- phisms of commutative rings are limited to maps that are essentially of finite type, or flat.. The

Yin, “Global existence and blow-up phenomena for an integrable two-component Camassa-Holm shallow water system,” Journal of Differential Equations, vol.. Yin, “Global weak

This paper presents an investigation into the mechanics of this specific problem and develops an analytical approach that accounts for the effects of geometrical and material data on

We study the classical invariant theory of the B´ ezoutiant R(A, B) of a pair of binary forms A, B.. We also describe a ‘generic reduc- tion formula’ which recovers B from R(A, B)

While conducting an experiment regarding fetal move- ments as a result of Pulsed Wave Doppler (PWD) ultrasound, [8] we encountered the severe artifacts in the acquired image2.

To be specic, let us henceforth suppose that the quasifuchsian surface S con- tains two boundary components, the case of a single boundary component hav- ing been dealt with in [5]

For X-valued vector functions the Dinculeanu integral with respect to a σ-additive scalar measure on P (see Note 1) is the same as the Bochner integral and hence the Dinculeanu