Startseite Convexity theorems for semisimple symmetric spaces
Artikel
Lizenziert
Nicht lizenziert Erfordert eine Authentifizierung

Convexity theorems for semisimple symmetric spaces

  • Dana Bălibanu und Erik P. van den Ban EMAIL logo
Veröffentlicht/Copyright: 10. Januar 2016

Abstract

We prove a convexity theorem for semisimple symmetric spaces G/H which generalizes an earlier theorem of the second named author to a setting without restrictions on the minimal parabolic subgroup involved. The new more general result specializes to Kostant’s non-linear convexity theorem for a real semisimple Lie group G in two ways, firstly by taking H maximal compact and secondly by viewing G as a symmetric space for G×G.


Communicated by Karl-Hermann Neeb


A Proof of Lemma 2.11

Finally, we prove Lemma 2.11.

We begin by showing that the result holds for G a complex semisimple Lie group, connected with trivial center. That proof will be based on the following general lemma, inspired by [21, Proposition 1].

Let 𝔥 be a complex abelian Lie algebra and let 𝒩 be the class of complex finite dimensional nilpotent Lie algebras 𝔫, equipped with a representation of 𝔥 by derivations, such that the following conditions are fulfilled:

  1. the representation of 𝔥 in 𝔫 is semisimple,

  2. all weight spaces of 𝔥 in 𝔫 have complex dimension one.

If 𝔫 belongs to the class 𝒩, we write Λ(𝔫) for the set of 𝔥-weights in 𝔫. If λΛ(𝔫), then the associated weight space is denoted by 𝔫λ.

Lemma A.1

Let nN and let N be the connected, simply-connected Lie group with Lie algebra n. Let λ1,,λm be the distinct weights of h in n. Then the map

ψ:(X1,,Xm)expX1expXm

defines a diffeomorphism

𝔫λ1××𝔫λmN.

Proof.

We will use induction on dim(𝔫). If dim𝔫=1, then 𝔫 is abelian and the result holds trivially.

Next, assume that m>1 and assume that the result has been established for 𝔫 with dim𝔫<m. Assume that 𝔫𝒩 has dimension m.

Denote by 𝔫1 the center of 𝔫, which is non-trivial. If 𝔫1=𝔫, then 𝔫 is abelian and the result is trivially true. Thus, we may as well assume that 0𝔫1𝔫. In particular, this implies that both 𝔫1 and 𝔫/𝔫1 have dimensions at most m-1. Put l:=dim𝔫1.

The ideal 𝔫1 is stable under the action of 𝔥 and it is readily verified that 𝔫1 and 𝔫/𝔫2 with the natural 𝔥-representations belong to 𝒩. Furthermore, since all weight spaces are 1-dimensional, we see that

Λ(𝔫)=Λ(𝔫1)Λ(𝔫/𝔫1).

We will first prove that ψ is a diffeomorphism under the assumption that the 𝔥-weights in 𝔫 are numbered in such a way that

Λ(𝔫1)={λ1,,λl}andΛ(𝔫/𝔫1)={λl+1,,λm}.

Since N is simply-connected, the map exp:𝔫N is a diffeomorphism; hence, N1:=exp(𝔫1) is the connected subgroup of N with Lie algebra 𝔫1. In particular, N1 is simply connected as well. Since 𝔫1 is an ideal, N/N1 has a unique structure of Lie group for which the natural map NN/N1 is a Lie group homomorphism. We now observe that NN/N1 is a principal fiber bundle with fiber N1. By standard homotopy theory, we have a natural exact sequence

π1(N)π1(N/N1)π0(N1).

Since N is simply-connected, and N1 connected, we conclude that N/N1 is the simply connected group with Lie algebra 𝔫/𝔫1.

By the induction hypothesis, the maps

ψ𝔫1:𝔫λ1××𝔫λlN1,
ψ𝔫/𝔫1:(𝔫/𝔫1)λl+1××(𝔫/𝔫1)λmN/N1

are diffeomorphisms. For every j{l+1,,m} the canonical projection 𝔫𝔫/𝔫1 induces the isomorphisms of weight spaces 𝔫λj(𝔫/𝔫1)λj. Let ψ¯:𝔫λl+1××𝔫λmN/N1 be defined by

ψ¯(Xl+1,,Xm)=expXl+1expXmN1.

Then the following diagram commutes:

From this we infer that ψ¯ is a diffeomorphism. We now obtain that the map ψ~:𝔫λl+1××𝔫λm×N1N,

(Xl+1,,Xm,n1)(expXl+1expXm)n1,

is a diffeomorphism onto N. Since

ψ(X1,,Xl,Xl+1,,Xm)=ψ~(Xl+1,,Xm,ψ𝔫1(X1,,Xl)),

it follows that ψ is a diffeomorphism as well. Clearly, the above proof works for every enumeration of the weights in Λ(𝔫/𝔫1). Since the weight spaces (𝔫1)λ for λΛ(𝔫1) are all central in 𝔫, we conclude that the result holds for any enumeration of the weights in Λ(𝔫).

Corollary A.2

Let G be a connected complex semisimple Lie group and let nB be the nilpotent radical of a Borel subalgebra b of g. Let h be a Cartan subalgebra contained in b. Let n1,,nk be linearly independent subalgebras of nB, each of which is a direct sum of h-root spaces, and assume that their direct sum n:=n1nk is again a subalgebra. Put N:=expn and Nj:=exp(nj) for 1jk. Then the multiplication map

μ:N1××NkN

is a diffeomorphism.

Proof.

This is an immediate consequence of Lemma A.1. ∎

Proof of Lemma 2.11.

We assume that G is a real reductive Lie group of the Harish-Chandra class. Define

𝔤1:=[𝔤,𝔤],

the semisimple part of the Lie algebra of G. Let G1 be the analytic subgroup of G with Lie algebra 𝔤1. Since the nilpotent radical NP of P is completely contained in G1, we may assume from the start that G=G1, i.e., G is connected semisimple with finite center.

Since Ad is a finite covering homomorphism from G onto Aut(𝔤), mapping N diffeomorphically onto Ad(N), whereas Aut(𝔤) is a connected real form of Int(𝔤), we may assume that G is a connected real form of a connected complex semisimple Lie group G with trivial center. Let τ be the conjugation on G, such that

G=(Gτ).

Let 𝔤 denote the Lie algebra of G, then 𝔤=𝔤i𝔤. Note that the complexification 𝔫P of 𝔫P equals 𝔫Pi𝔫P and that

NP=(NP)τ.

Take a Cartan subalgebra of 𝔤, containing 𝔞=𝔞i𝔞. It is of the form

𝔥=𝔱𝔞,

where 𝔱 is a maximal abelian subspace of 𝔪:=Z𝔨(𝔞). Since 𝔱 centralizes 𝔞, all 𝔞-root spaces are invariant under ad(𝔱). This implies that the subalgebras 𝔫j:=𝔫ji𝔫j (j{1,,k}) of 𝔫P are direct sums of 𝔥-root spaces. Furthermore, their direct sum equals 𝔫=𝔫i𝔫, hence is a subalgebra. Finally, there exists a Borel subalgebra containing 𝔥+𝔫. By Corollary A.2, the multiplication map

μ:N1××NkN

is a diffeomorphism. It readily follows that μ restricts to a bijection from (N1)τ××(Nk)τ onto (N)τ. Since

(N)τ=Nand(Nj)τ=Njfor all 1jk,

it follows that μ is a bijective embedding from N1××Nk onto N, hence a diffeomorphism. ∎

B The case of the group

Every semisimple Lie group G can be viewed as a semisimple symmetric space for the group G×G. In this section we investigate what our convexity theorem means for this particular example. An independent proof for this case is presented in [3, Section 3.2.2].

More generally, let G be a real reductive group of the Harish-Chandra class, θ a Cartan involution, K:=Gθ the associated maximal compact subgroup, and 𝔤=𝔨𝔭 the associated Cartan decomposition as in Section 1. Let 𝔞 be a maximal abelian subspace of 𝔭, A=exp𝔞 and Σ(𝔤,𝔞) the associated root system.

Let

G:=G×G.

Then θ:=θ×θ is a Cartan decomposition of G with associated maximal compact subgroup K:=K×K. The involution

σ:GG,(x,y)(y,x),

commutes with θ. Its fixed point group H equals the diagonal in G×G and is essentially connected in G, see [3, Example 2.3.7].

The associated space 𝔭𝔮 equals {(X,-X):X𝔭} and has

𝔞q:={(X,-X):X𝔞}

as a maximal abelian subspace. Its root system is given by

Σ(𝔤,𝔞q)=Σ(𝔤,𝔞)×{0}{0}×Σ(𝔤,𝔞).

Finally, 𝔞q is contained in the maximal abelian subspace 𝔞:=𝔞×𝔞 of 𝔭. We put A:=exp(𝔞)=A×A. Note that the projection map prq:𝔞𝔞q is given by

(B.1)prq(U,V)=(12(U-V),12(V-U)).

Let P and Q be minimal parabolic subgroups of G containing A, i.e., P,Q𝒫(A). Then P×Q is a minimal parabolic subgroup of G containing A. Moreover, any minimal parabolic subgroup of G containing A is of this form. The positive system of 𝔞-roots associated with P×Q is given by

Σ(P×Q):=Σ(P)×{0}{0}×Σ(Q),

where Σ(P) and Σ(Q) are positive systems for Σ(𝔤,𝔞) corresponding to the minimal parabolic subgroups P and Q.

In the present setting, our main result, Theorem 10.1, tells us that for aAq, we have

prqP×Q(aH)=conv(WKHloga)+Γ(P×Q).

In order to determine the cone Γ(P×Q), we need to determine the set Σ(P×Q,σθ) of roots γΣ(P×Q) for which σθγΣ(P×Q). Let γ=(α,0) be such a root. Then αΣ(P) and σθγ=(0,-α) must be an element of {0}×Σ(Q) so that αΣ(P)Σ(Q¯). Likewise, if (0,β) belongs to this set, then βΣ(Q)Σ(P¯). We thus see that

Σ(P×Q,σθ)=(Σ(P)Σ(Q¯))×{0}{0}×(Σ(P¯)Σ(Q)).

Notice that there are no roots γΣ(P×Q) for which σθγ=γ. Thus, Σ(P×Q)-=Σ(P×Q,σθ) and we conclude that

Γ(P×Q)=Γ𝔞q(Σ(P×Q,σθ))=γΣ(P×Q,σθ)0prqHγ.

If γ is of the form (α,0), then Hγ=(Hα,0) and if γ=(0,α), then Hγ=(0,Hα). In view of (B.1), we now obtain

Γ(P×Q)=αΣ(P)Σ(Q¯)0(12Hα,-12Hα)+αΣ(P¯)Σ(Q)0(-12Hα,12Hα)
=αΣ(P)Σ(Q¯)0(Hα,-Hα)+αΣ(P¯)Σ(Q)0(H-α,-H-α)
=αΣ(P)Σ(Q¯)0(Hα,-Hα)+αΣ(P)Σ(Q¯)0(Hα,-Hα)
=αΣ(P)Σ(Q¯)0(Hα,-Hα).

We will identify 𝔞q with 𝔞 via the map (X,-X)X. Thus,

Γ𝔞q(Σ(P×Q))=Γ𝔞(Σ(P)Σ(Q¯)).

For Q=P, the resulting cone is the zero one, and we retrieve the non-linear convexity theorem of Kostant [20] for the group G. At the other extreme, for Q=P¯, the resulting cone is maximal, and we retrieve the convexity theorem of [4] for the pair (G,H).

Taking a=e we obtain, with the same identification 𝔞q𝔞,

(B.2)prqP×Q(H)=Γ𝔞(Σ(P)Σ(Q¯)).

On the other hand,

prqP×Q(H)=prqP×Q(diag(G×G))
=prq({(P(g),Q(g)):gG})
=prq({(P(kanp),Q(kanp)):kK,aA,npNP})
=prq({(loga,Q(anpa-1)+loga):aA,npNP})
=prq({(loga,Q(np)+loga):aA,npNP})
={(-12Q(np),12Q(np)):npNP}.

Using the same identification 𝔞q𝔞 as above, we conclude that

prqP×Q(H)=-12Q(NP)
=-12Q((NPN¯Q)(NPNQ))
=-12Q(NPN¯Q).

Thus, by equation (B.2), we obtain that

-12Q(NPN¯Q)=Γ𝔞(Σ(P)Σ(Q¯)),

which is equivalent to

Q(NPN¯Q)=Γ𝔞(Σ(P¯)Σ(Q)).

Thus, we retrieve the identity of Lemma 4.9, which, of course, was used in the proof of our main theorem.

Acknowledgements

The authors would like to thank Job Kuit for the proof of Proposition 2.14 and his useful comments. One of the authors, Dana Bălibanu, thanks Ioan Mǎrcuţ for his interest in this work and all his good suggestions on how to improve it.

References

[1] Andersen N. B., Flensted-Jensen M. and Schlichtkrull H., Cuspidal discrete series for semisimple symmetric spaces, J. Funct. Anal. 263 (2012), no. 8, 2384–2408. 10.1016/j.jfa.2012.07.009Suche in Google Scholar

[2] Atiyah M. F., Convexity and commuting Hamiltonians, Bull. Lond. Math. Soc. 14 (1982), no. 1, 1–15. 10.1112/blms/14.1.1Suche in Google Scholar

[3] Bălibanu D., Convexity theorems for symmetric spaces and representations of n-Lie algebras. Two studies in Lie theory, Ph.D. thesis, Universiteit Utrecht, 2014. Suche in Google Scholar

[4] van den Ban E. P., A convexity theorem for semisimple symmetric spaces, Pacific J. Math. 124 (1986), no. 1, 21–55. 10.2140/pjm.1986.124.21Suche in Google Scholar

[5] Duistermaat J. J., Convexity and tightness for restrictions of Hamiltonian functions to fixed point sets of an antisymplectic involution, Trans. Amer. Math. Soc. 275 (1983), no. 1, 417–429. 10.1090/S0002-9947-1983-0678361-2Suche in Google Scholar

[6] Duistermaat J. J., On the similarity between the Iwasawa projection and the diagonal part, Mém. Soc. Math. Fr. (N.S.) 15 (1984), 129–138. 10.24033/msmf.301Suche in Google Scholar

[7] Duistermaat J. J., Kolk J. A. C. and Varadarajan V. S., Functions, flows and oscillatory integrals on flag manifolds and conjugacy classes in real semisimple Lie groups, Compos. Math. 49 (1983), no. 3, 309–398. Suche in Google Scholar

[8] Flensted-Jensen M., Spherical functions of a real semisimple Lie group. A method of reduction to the complex case, J. Funct. Anal. 30 (1978), no. 1, 106–146. 10.1016/0022-1236(78)90058-7Suche in Google Scholar

[9] Foth P. and Otto M., A symplectic approach to van den Ban’s convexity theorem, Doc. Math. 11 (2006), 407–424. . 10.4171/dm/216Suche in Google Scholar

[10] Gindikin S. G. and Karpelevič F. I., Plancherel measure for symmetric Riemannian spaces of non-positive curvature, Dokl. Akad. Nauk 145 (1962), 252–255. Suche in Google Scholar

[11] Guillemin V. and Sternberg S., Convexity properties of the moment mapping, Invent. Math. 67 (1982), no. 3, 491–513. 10.1007/BF01398933Suche in Google Scholar

[12] Guillemin V. and Sternberg S., Convexity properties of the moment mapping. II, Invent. Math. 77 (1984), no. 3, 533–546. 10.1007/BF01388837Suche in Google Scholar

[13] Harish-Chandra , Spherical functions on a semisimple Lie group. I, Amer. J. Math. 80 (1958), 241–310. 10.2307/2372786Suche in Google Scholar

[14] Heckman G. J., Projections of orbits and asymptotic behavior of multiplicities for compact L,ie groups Ph.D. thesis, Rijksuniversiteit Leiden, 1980. Suche in Google Scholar

[15] Helgason S., Differential Geometry, Lie Groups, and Symmetric Spaces, Pure Appl. Math. 80, Academic Press, New York, 1978. Suche in Google Scholar

[16] Helgason S., Groups and Geometric Analysis, Pure Appl. Math. 113, Academic Press, Orlando, 1984. Suche in Google Scholar

[17] Hilgert J. and Neeb K.-H., Poisson Lie groups and non-linear convexity theorems, Math. Nachr. 191 (1998), 153–187. 10.1002/mana.19981910108Suche in Google Scholar

[18] Kirwan F., Convexity properties of the moment mapping. III, Invent. Math. 77 (1984), no. 3, 547–552. 10.1007/BF01388838Suche in Google Scholar

[19] Knapp A. W., Lie Groups Beyond an Introduction, 2nd ed., Progr. Math. 140, Birkhäuser, Boston, 2002. Suche in Google Scholar

[20] Kostant B., On convexity, the Weyl group and the Iwasawa decomposition, Ann. Sci. Éc. Norm. Supér (4) 6 (1973), 413–455. 10.24033/asens.1254Suche in Google Scholar

[21] Lazard M., Groupes semi-simples: structure de B et de G/B, Sém. Claude Chevalley 1 (1956–58), no. 13, 1–13. Suche in Google Scholar

[22] Lu J.-H. and Ratiu T., On the nonlinear convexity theorem of Kostant, J. Amer. Math. Soc. 4 (1991), no. 2, 349–363. 10.1090/S0894-0347-1991-1086967-2Suche in Google Scholar

[23] Mostow G. D., Some new decomposition theorems for semi-simple groups, Mem. Amer. Math. Soc. 1955 (1955), no. 14, 31–54. Suche in Google Scholar

[24] Rossmann W., The structure of semisimple symmetric spaces, Canad. J. Math. 31 (1979), no. 1, 157–180. 10.4153/CJM-1979-017-6Suche in Google Scholar

Received: 2015-4-28
Published Online: 2016-1-10
Published in Print: 2016-11-1

© 2016 by De Gruyter

Heruntergeladen am 30.10.2025 von https://www.degruyterbrill.com/document/doi/10.1515/forum-2015-0079/html
Button zum nach oben scrollen