Skip to main content
Springer logoLink to Springer
. 2023 Dec 31;65(1):9. doi: 10.1007/s10455-023-09937-6

Modular geodesics and wedge domains in non-compactly causal symmetric spaces

Vincenzo Morinelli 1, Karl-Hermann Neeb 2,, Gestur Ólafsson 3
PMCID: PMC10757968  PMID: 38169487

Abstract

We continue our investigation of the interplay between causal structures on symmetric spaces and geometric aspects of Algebraic Quantum Field Theory. We adopt the perspective that the geometric implementation of the modular group is given by the flow generated by an Euler element of the Lie algebra (an element defining a 3-grading). Since any Euler element of a semisimple Lie algebra specifies a canonical non-compactly causal symmetric space M=G/H, we turn in this paper to the geometry of this flow. Our main results concern the positivity region W of the flow (the corresponding wedge region): If G has trivial center, then W is connected, it coincides with the so-called observer domain, specified by a trajectory of the modular flow which at the same time is a causal geodesic. It can also be characterized in terms of a geometric KMS condition, and it has a natural structure of an equivariant fiber bundle over a Riemannian symmetric space that exhibits it as a real form of the crown domain of G/K. Among the tools that we need for these results are two observations of independent interest: a polar decomposition of the positivity domain and a convexity theorem for G-translates of open H-orbits in the minimal flag manifold specified by the 3-grading.

Keywords: Euler element, Wedge domain, Causal space, Modular geodesic

Introduction

A new Lie theoretical approach to localization on spacetimes involved in Algebraic Quantum Field Theory (AQFT) has been introduced in the recent years by the authors and collaborators in a series of works, see [38, 41, 4952, 54]. In the current paper, we continue the investigation of the structure of wedge regions in non-compactly causal symmetric spaces, started in [52]. First we briefly recall the motivation form AQFT, and then, we introduce tools and details to formulate our results.

Symmetric spaces are quotients M=G/H, where G is a Lie group, τ is an involutive automorphism of G and HGτ is an open subgroup (cf. [36]). A causal symmetric space carries a G-invariant field of pointed generating closed convex cones CmTm(M) in their tangent spaces. Typical examples are de Sitter space dSdSO1,d(R)e/SO1,d-1(R)e and anti-de Sitter space AdSdSO2,d-1(R)e/SO1,d-1(R)e as well as products dSd×Sk and AdSd×Hk with spheres and hyperbolic spaces, respectively. These are Lorentzian, but we do not require our causal structure to come from a Lorentzian metric, which creates much more flexibility and a richer variety of geometries. Causal symmetric spaces permit to study causality aspects of spacetimes in a highly symmetric environment. Here we shall always assume that M is non-compactly causal in the sense that the causal curves define a global order structure with compact order intervals (they are called globally hyperbolic), and in this context one can also prove the existence of a global “time function” with group theoretic methods (see [46]). We refer to the monograph [26] for more details and a complete exposition of the classification of irreducible causal symmetric spaces. A new perspective on the classification has been developed in [41].

Recent interest in causal symmetric spaces in relation to representation theory arose from their role as analogs of spacetime manifolds in the context of Algebraic Quantum Field Theory in the sense of Haag–Kastler. A model in AQFT is specified by a net of von Neumann algebras M(O) acting on a fixed Hilbert space indexed by open subsets O of the chosen spacetime M [20]. The hermitian elements of the algebra M(O) represent observables that can be measured in the “laboratory” O. These nets are supposed to satisfy fundamental quantum and relativistic assumptions:

  • (I)

    Isotony: O1O2 implies M(O1)M(O2).

  • (L)

    Locality: O1O2 implies M(O1)M(O2), where O is the “causal complement” of O, i.e., the maximal open subset that cannot be connected to O by causal curves.

  • (RS)

    Reeh–Schlieder property: There exists a unit vector ΩH that is cyclic for M(O) if O.

  • (Cov)

    Covariance: There is a Lie group G acting on M and a unitary representation U:GU(H) such that UgM(O)Ug-1=M(gO) for gG.

  • (BW)

    Bisognano–Wichmann property: Ω is separating for some “wedge region” WM and there exists an element hg with Δ-it/2π=U(expth) for tR, where Δ is the modular operator corresponding to (M(W),Ω) in the sense of the Tomita–Takesaki Theorem ([4, Thm. 2.5.14]).

  • (Vac)

    Invariance of the vacuum: U(g)Ω=Ω for every gG.

The (BW) property gives a geometrical meaning to the dynamics provided by the modular group (Δit)tR of the von Neumann algebra M(W) associated with wedge regions with respect to the vacuum state specified by Ω. On Minkowski/de Sitter spacetime, it provides an identification of the one-parameter group (ΛW(t))tR of boosts in the Poincaré/Lorentz group with the Tomita–Takesaki modular operator:

U(ΛW(t))=Δ-it/2π.

Here ΛW=gΛW1g-1 is a one-parameter group of boosts associated with W=g.W1, where W1={xM:|x0|<x1} is the standard right wedge and

ΛW1(t)=(cosh(t)x0+sinh(t)x1,cosh(t)x1+sinh(t)x0,x2,,xd)

describes the boosts associated with W1.

The homogeneous spacetimes occurring naturally in AQFT are causal symmetric spaces associated with their symmetry groups (Minkowski spacetime for the Poincaré group, de Sitter space for the Lorentz group and anti-de Sitter space for SO2,d(R)), and the localization in wedge regions is ruled by the acting group. The rich interplay between the geometric and algebraic objects in AQFT allowed a generalization of fundamental localization properties and the subsequent definition of fundamental models (second quantization fields), having as initial data a general Lie group with distinguished elements (Euler elements) in the Lie algebra. Given an AQFT on Minkowski spacetime M=R1,d (or de Sitter spacetime dSdR1,d), the Bisognano–Wichmann (BW) property allows an identification of geometric and algebraic objects in both free and interacting theories in all dimensions [3, 13, 44]. This plays a central role in many results in AQFT and is a building block of our discussion.

One can generalize the picture we get from these explicit AQFT models and construct nets of von Neumann algebras on causal symmetric spaces with representation theoretical methods. We start with a unitary representation U:GU(H) of a reductive Lie group G whose Lie algebra contains Euler elements. Then, one constructs so-called one-particle nets on causal symmetric spaces. These are isotonous, G-covariant maps that associate to non-empty open subsets of the causal symmetric space standard subspaces1 of the “one-particle space” H. For positive energy representations, we refer to [50] for left invariant nets on reductive Lie groups, to [55] for left invariant nets on non-reductive Lie groups, and to [51] for nets on compactly causal symmetric spaces. For general unitary representation, nets on non-compactly causal symmetric spaces have been constructed in [17] and on abstract wedge families in [38]; see also [40]. These constructions have the (BW) property as a fundamental input. Bosonic second quantization associates to a one-particle net an isotonous, G-covariant net of von Neumann algebras acting on the bosonic Fock space [5, 38].

These constructions naturally generalize the AQFT framework, re-construct the free second quantization AQFT models on the chiral conformal circle, on de Sitter and anti-de Sitter space, and provide several new models [17, 38, 50]. One can also recover free AQFT models on Minkowski spacetime as addressed in [3840]. If Z(G) is non-trivial, then a proper second quantization scheme to provide a (twisted-)local net of von Neumann algebras remains to be determined (cf. [11, 19]). We stress that our setting provides a general framework to study properties of AQFT that is not restricted to second quantization theories. It also provides results on the type of von Neumann algebras and on properties of wedge symmetries appearing in these models (see, e.g., [40]).

We know from [40] that, in the general context, the potential generators hg of the modular groups in (BW) are Euler elements, i.e., adh defines a 3-grading

g=g1(h)g0(h)g-1(h),wheregλ(h)=ker(adh-λ1).

This leads to the question how the existence and the choice of the Euler element affect the geometry of the associated symmetric space. The (BW) property establishes a one-to-one correspondence between “wedge regions” WM and the associated Euler elements. So these fundamental localization regions can be determined in terms of Euler elements. This allowed the following generalization of nets of von Neumann algebras on Minkowski/de Sitter spacetime:

  • Given a Lie group G with Lie algebra g, then the couples (h,τh), where hg is an Euler element and τh an involutive automorphism of G, inducing on g the involution τh=eπiadh, allow the definition of an ordered, G-covariant set of “abstract wedge regions” carrying also some locality information [38]. In particular, they encode the commutation relation property of the Tomita operators (modular operator and modular conjugation).

  • Causal symmetric spaces provide manifolds and a causal structure supporting nets of algebras. Here the wedge regions can be defined as open subsets in several ways. The equivalence of various characterizations has been shown in [51, 52]; see also the discussion below.

The whole picture complies with Minkowski, de Sitter and anti-de Sitter spacetimes and the associated free fields. A generalization of wedge regions of the Minkowski or de Sitter spacetime on general curved spacetimes has been proposed by many authors, see for instance [12] and references therein. In our framework, on non-compactly causal symmetric spaces, the rich geometric symmetries allow different characterizations of wedge regions, in particular in terms of positivity of the modular flow, or geometric KMS conditions and in terms of polar decompositions as described in [52]. Some of them directly accord with the literature, for instance for positivity of the modular flow, see [9, Defin. 3.1] and in particular [45] for the connection to thermodynamics on de Sitter space. To see how these definitions apply to wedges in de Sitter space, cf. [52, App. D.3] and [6]. For causal symmetric spaces all definitions of wedge regions discussed in [41, 52] specify the same regions, up to choosing connected components (cf. [52, Thm. 7.1]). In Theorem 7.1, we prove that the identification is actually complete for the adjoint groups since the wedge region defined in terms of positivity of the modular flow is connected. This contrasts the situation for compactly causal symmetric spaces, where wedge regions are in general not connected, as for anti-de Sitter space ([52, Lemma 11.2]).

To formulate our results, we recall some basic terminology concerning symmetric Lie algebras (see [52] for more details).

  • A symmetric Lie algebra is a pair (g,τ), where g is a finite-dimensional real Lie algebra and τ an involutive automorphism of g. We write h=gτ=ker(1-τ) and q=g-τ=ker(1+τ) for the τ-eigenspaces.

  • A causal symmetric Lie algebra is a triple (g,τ,C), where (g,τ) is a symmetric Lie algebra and Cq is a pointed generating closed convex cone invariant under the group Inng(h)=eadh acting in q. We call (g,τ,C) compactly causal (cc) if C is elliptic in the sense that, for xC (the interior of C in q), the operator adx is semisimple with purely imaginary spectrum. We call (g,τ,C) non-compactly causal (ncc) if C is hyperbolic in the sense that, for xC, the operator adx is diagonalizable.

As explained in detail in [41], Euler elements in reductive Lie algebras g lead naturally to ncc symmetric Lie algebras: For an Euler element hg, choose a Cartan involution θ of g with z(g)g-θ such that θ(h)=-h. Then τh:=eπiadh is an involutive automorphism of g commuting with θ, so that τ:=τhθ defines a symmetric Lie algebra (g,τ) and there exists a pointed generating Inng(h)-invariant hyperbolic cone C with hC. Under the assumption that h=gτ contains no non-zero ideal of g, there is a unique minimal cone Cqmin(h) with this property. It is generated by the orbit Inng(h)hq.

Let (g,τ,C) be an ncc symmetric Lie algebra and (G,τG,H) a corresponding symmetric Lie group, i.e., G is a connected Lie group, τG an involutive automorphism of G integrating τ, and HGG an open subgroup. If, in addition, Ad(H)C=C, then we call the quadruple (G,τG,H,C) a causal symmetric Lie group. On M=G/H, we then obtain the structure of a causal symmetric space, specified by the G-invariant field of open convex cones2

V+(gH):=g.CTgH(M).2 1.1

We further assume that

S:=Hexp(C)=exp(C)HG

is a closed subsemigroup for which the polar map H×CS,(h,x)hexpx is a homeomorphism. Then

g1Hg2Hifg2-1g1S 1.2

defines on M a partial order, called the causal order on M. According to Lawson’s Theorem [30] and Theorem C.1), this is always the case if z(g)q and exp|z(g) is injective. The second condition is always satisfied if G is simply connected.

For an Euler element hg, we consider the associated modular flow on M=G/H, defined by

αt(gH)=exp(th)gH. 1.3

We study orbits of this flow which are geodesics γ:RM with respect to the symmetric space structure and causal in the sense that γ(t)V+(γ(t)) for tR. We call them h-modular geodesics. All these are contained in the positivity domain

WM+(h):={mM:XhM(m)V+(m)} 1.4

of the vector field XhM generating the modular flow. We refer to [52] for a detailed analysis of the latter domain in the special situations where the modular flow on M has fixed points, which is equivalent to the adjoint orbit Oh=Inn(g)h intersecting h.

We show for ncc symmetric Lie algebras, which are direct sums of irreducible ones, that:

  • Causal modular geodesics exist if and only if the adjoint orbit Oh=Ad(G)hg intersects the interior of the cone Cq, and then the centralizer Gh={gG:Ad(g)h=h} of h acts transitively on the union of the corresponding curves (Proposition 3.2(c)).

  • Suppose that the cone is maximal, i.e., Cq=Cqmax (see (2.3) and [41, §3.5.2] for details). Let qk=qk for a Cartan decomposition g=kp with hqp and consider the domain
    Ωqk={xqk:ρ(adx)<π2},
    where ρ(adx) is the spectral radius of adx. Then, the connected component W:=WM+(h)eH of the base point eH in the positivity domain is
    W=Geh.ExpeH(Ωqk)
    (Theorem 3.6).
  • We associate to any modular geodesic a connected open subset W(γ)M; the corresponding observer domain. For de Sitter space dSd, we thus obtain the familiar wedge domain obtained by intersecting dSd with a Rindler wedge in Minkowski space R1,d (Example 5.3). In Theorem 5.7, we show that it coincides with W, provided that H=Khexp(hp) and Cq=Cqmax.

  • A key step in the proof of Theorem 5.7 is the following Convexity Theorem. Let
    P-:=exp(g-1(h))GhG
    be the “negative” parabolic subgroup of G specified by h and identity g1(h) with the open subset B:=exp(g1(h)).eP-G/P-. Then D:=H.0B is an open convex subset, and for any gG with g.DB, the subset g.DB is convex (Theorem 4.5).
  • In Sect. 6 we further show that, for C=Cqmax and g simple, that the real tube domain h+C intersects the set E(g) of Euler elements in a connected subset (Theorem 6.1). As a consequence, we derive that WM+(h) if and only if hOh (Corollary 6.3). In particular, only one conjugacy class of Euler elements possesses non-empty positivity regions. This is of particular relevance for locality properties of nets of local algebras. We plan to investigate this in subsequent work.

  • In Theorem 7.1 we show that the positivity domain WM+(h) is connected for G=Inn(g) and g simple, and this implies that
    W(γ)=W=WM+(h).
    From this in turn we derive that the stabilizer group GW={gG:g.W=W} coincides with Gh (Proposition 7.3), so that the wedge space W(M):={g.W:gG} of wedge regions in M can be identified, as a homogeneous G-space, with the adjoint orbit Oh=Ad(G)hG/Gh. In particular W(M) also is a symmetric space.
  • Finally, we show in Theorem 8.2 that W coincides with the KMS wedge domain
    WKMS={mM:αit(m)Ξfor0<t<π},
    where Ξ is the crown domain of the Riemannian symmetric space G/K.

We conclude this introduction with some more motivation from AQFT. The analysis of the properties of the modular flow on symmetric spaces is also motivated by the investigation of energy inequalities in quantum and relativistic theories. In General Relativity, there exist many solutions to the Einstein equation that, for various reasons, may not be physical. Energy conditions such as the pointwise non-negativity of the energy density, which ensures that the gravity force is attractive, can be required to discard non-physical models [14, 62]. In quantum and relativistic theories, the energy conditions need to be rewritten. For instance, it is well known that the energy density at individual spacetime points is unbounded from below, even if the energy density integrated over a Cauchy surface is non-negative (see [14, 15] and references therein).

Families of inequalities have been discussed in several models, employing different mathematical and physical approaches (see for instance [14, 16, 27, 29, 42, 61]). In recent years, operator algebraic techniques have been very fruitful for the study of the energy inequalities because of the central role played by the modular hamiltonian in some of these energy conditions. This object corresponds to the logarithm of the modular operator of a local algebra of a specific region, which in some cases can be identified with the generator of a one-parameter group of spacetime symmetries by the Bisognano–Wichmann property. In this regard, we mention the ANEC (Averaged Null Energy Condition) and the QNEC (Quantum Null Energy Condition) and their relation with the Araki relative entropy, an important quantum-information quantity, defined in terms of relative modular operators (see, for instance, [1, 9, 10, 3235, 43]). We stress that, in this analysis, the study of the modular flow on the manifold can be particularly relevant. Moreover, in order to find regions where energy inequalities hold, one may also need to deform the modular flow [8, 43]. In our abstract context, the Euler element specifies the flow that can be implemented by the modular operator, hence the modular Hamiltonian, when the Bisognano–Wichmann property holds. In particular, the identification of specific flows on symmetric spaces (modular flows), the characterization in terms of modular operators of covariant local subspaces attached to specific regions (wedges) motivate an analysis of modular flows on non-compactly causal symmetric spaces pursued in our project.

In this respect, the wedge regions are the first fundamental open subsets of spacetime to be studied in detail. Following General Relativity (see, for instance, [9, 12] and references therein), one can define them as an open connected, causally convex subregion W of a spacetime M, associated with a Killing flow Λ preserving W, which is timelike and time-oriented on W. On Minkowski spacetime the flow Λ, a one-parameter group of boosts, corresponds to the time-evolution of a uniformly accelerated observer moving within W. Then, W is a horizon for this observer: he cannot send a signal outside W and receive it back. Then the vacuum state becomes a thermal state for the algebra of observables inside the wedge region W by the Bisognano–Wichmann property [18, 21, 31]. In our general context, we recover the definition (and equivalent ones) of wedge regions. Then, by the Bisognano–Wichmann property, the thermal property of the vacuum state holds when nets of algebras or standard subspaces are considered [38, 41, 52]. In this paper, we focus on the related properties of the wedge regions in non-compactly causal symmetric spaces.

Notation

  • If M is a topological space and mM, then Mm denotes the connected component of M containing m. In particular, we write eG for the identity element in the Lie group G and Ge for its identity component.

  • Involutive automorphisms of G are typically denoted τG, and τ is the corresponding automorphism of the Lie algebra g=L(G). We write gτ=ker(1-τ) and g-τ=ker(1+τ).

  • For xg, we write Gx:={gG:Ad(g)x=x} for the stabilizer of x in the adjoint representation and Gex=(Gx)e for its identity component.

  • For hg and λR, we write gλ(h):=ker(adh-λ1) for the corresponding eigenspace in the adjoint representation.

  • If g is a Lie algebra, we write E(g) for the set of Euler elements hg, i.e., adh is non-zero and diagonalizable with Spec(adh){-1,0,1}. The corresponding involution is denoted τh=eπiadh.

  • For a Lie subalgebra sg, we write Inng(s)=eadsAut(g) for the subgroup generated by eads.

  • For a convex cone C in a vector space V, we write C:=intC-C(C) for the relative interior of C in its span.

  • We use the notation
    ρ(A):=sup{|λ|:λSpec(A)} 1.5
    for the spectral radius of a linear operator A.

Causal Euler elements and ncc symmetric spaces

In this section, we recall some basic results on Euler elements and their relation with non-compactly causal symmetric spaces. Most of these statements are discussed in detail in [41].

Recall from above that an Euler element in a Lie algebra g is an element h defining a 3-grading of g by g=g-1g0g+1 with gj=ker(adh-j1), j=-1,0,1. We write E(g) for the set of Euler elements in g. In this section, we recall some results on from [41] on Euler elements that are crucially used in the following.

Definition 2.1

Let g be a reductive Lie algebra.

  1. A Cartan involution of g is an involutive automorphism θ for which z(g)g-θ and gθ is maximal compactly embedded in the commutator algebra [g,g]. We then write, using the notation from the introduction,
    g=kpwithk=gθandp=g-θ
  2. If τ is another involution on g commuting with θ, h:=gτ and q:=g-τ, then we have
    h=hkhp,q=qkqpwithhk=hk,hp=hp,qk=qk,qp=qp.
  3. The Cartan dual of the symmetric Lie algebra (g,τ) is the symmetric Lie algebra (gc,τc) with
    gc=h+iqandτc(x+iy)=x-iyforxh,yq.
    Note that gc=(gC)τ¯ where τ¯ is the conjugate-linear extension of τ to gC; in particular gc is a real form of gC.

Definition 2.2

Let (g,τ) be a symmetric Lie algebra and hE(g)q. We say that h is causal if there exists an Inng(h)-invariant closed pointed generating convex cone C in q with hC. We write Ec(q)E(g)q for the set of causal Euler elements in q. Recall that the triple (g,τ,C) is ncc if C is hyperbolic.

Lemma 2.3

Let (g,τ,C) be a simple ncc symmetric Lie algebra and hq be a causal Euler element. Then, the following assertions hold:

  1. There exist closed convex pointed generating Inng(h)-invariant cones
    Cqmin(h)Cqmax(h)
    such that hCqmin(h) and either
    Cqmin(h)CCqmax(h)orCqmin(h)-CCqmax(h).
  2. If (G,τG,H) is a connected symmetric Lie group with symmetric Lie algebra (g,τ), then two mutually exclusive cases occur:
    • Ad(H)h=Ad(He)h and G/H is causal.
    • -hAd(H)h and G/H is not causal.

Proof

(a) follows from [41, Sect. 3.5.2] and (b) from [41, Prop. 4.18].

If h is an Euler element in the reductive Lie algebra g and θ a Cartan involution with θ(h)=-h, τ:=θτh and z(g)g-θ, then [41, Thm. 4.2] implies that there exists an Inng(h)-invariant pointed closed convex cone Cq with hC, so that (g,τ,C) is ncc. Further, all ideals of g contained in gτ=h are compact. We have a decomposition

g=gkgrgs, 2.1

where gs is the sum of all simple ideals not commuting with h (the strictly ncc part), gr is the sum of the center z(g) and all non-compact simple ideals commuting with h on which τ=θ (the non-compact Riemannian part), and gk is the sum of all simple compact ideals (they commute with h). All these ideals are invariant under θ and τ=τhθ, so that we obtain decompositions

gs=hsqs,gr=hrqrandgk=hk, 2.2

where hrhk is a compact ideal of h, gr=hrqr is a Cartan decomposition and qp=qp,sqr. In particular q=qsqr. Let ps:qqs be the projection onto qs with kernel qr. Then, [41, Prop. B.4] implies that every Inng(h)-invariant closed convex cone C satisfies

Cs:=ps(C)=CqsandCs=Cqs.

By Lemma 2.3(a), we obtain a pointed Inng(h)-invariant cone Cqsmin(h)qs, adapted to the decomposition into irreducible summands, whose dual cone Cqsmax(h) with respect to the Cartan–Killing form κ(x,y)=tr(adxady) satisfies Cqsmin(h)Cqsmax(h). Put

Cqmin(h):=Cqsmin(h)Cqmax(h):=Cqsmax(h)+qr. 2.3

Both cones are adapted to the decomposition of (g,τ) into irreducible summands. Further, each pointed generating Inng(h)-invariant cone C containing h satisfies

Cqmin(h)CCqmax(h). 2.4

Here the first inclusion is obvious, and the second one follows from the fact that h is also contained in the dual cone

C={yq:(xC)κ(x,y)0}.

This leads to Cqmin(h)C, and thus to CCqmin(h)=Cqmax(h) (cf. [41, §3.5] for more details).

Lemma 2.4

If x(Cqmax), then the centralizer zh(x)=hker(adx) is compactly embedded in g, i.e., consists of elliptic elements.

Proof

First we observe that the cone Cqmax is adapted to the decomposition g=(gk+gr)+gs and so is the centralizer of x=xr+xs in h=(gk+kr)+hs. Hence the assertion follows from the fact that gk+kr is compactly embedded and zhs(x)=zhs(xs) is compactly embedded because the cone Cqsmax is pointed ([47, Prop. V.5.11]).

Theorem 2.5

(Uniqueness of the causal involution) ([41, Thm. 4.5]) Let (g,τ,C) be a semisimple ncc symmetric Lie algebra for which all ideals of g contained in h are compact, gs the sum of all non-Riemannian ideals, qs:=gsq, Cs:=Cqs, and θ a Cartan involution commuting with τ. Then the following assertions hold:

  1. Csqp contains a unique Euler element h, and this Euler element satisfies τ=τhθ.

  2. Inng(h) acts transitively on CsE(g).

  3. For every Euler element hCs, the involution ττh is Cartan.

Proposition 2.6

Let (G,τG,H,C) be a connected semisimple ncc symmetric Lie group for which h=gτ contains no non-compact ideal of g (g=gr+gs) and let hCs (cf. Theorem 2.5) be a causal Euler element. Then the following assertions hold:

  1. H=HeHh, i.e., every connected component of H meets Hh.

  2. Ad(Hh)=Ad(H)h is a maximal compact subgroup of Ad(H).

  3. Ad(H)h=Ad(H)τh and τh:=eπiadh induces a Cartan involution on Ad(H).

  4. τ induces a Cartan involution on Ad(H)h for which Ad(Heh)τ=eadhk is connected.

Proof

The statements on the adjoint group Ad(G)=Inn(g) follow from [41, Cor. 4.6] because Ad(H)Inn(g)τ preserves C. Further, Ad(H)h=Ad(Hh) and Hh=Ad-1(Ad(H)h) imply with (a) (for Ad(G)) that H=HeHh.

Definition 2.7

If g is a simple hermitian Lie algebra, θ a Cartan involution of g and ap maximal abelian, then the restricted root system Σ(g,a) is either of type Cr or BCr. In the first case, we say that g is of tube type.

Recall that if (g,τ) is simple ncc, then either gc is simple hermitian or gchC, where h=gτ is simple hermitian ([41, Rem. 4.24]).

Proposition 2.8

([41, Lemma 5.1, Prop. 5.2]) Let (g,τ,C) be a simple ncc symmetric Lie algebra. Pick a causal Euler element hC and tqqk maximal abelian and set s:=dimtq. Then, the following assertions hold:

  1. The Lie algebra l generated by h and tq is reductive.

  2. The commutator algebra [l,l] is isomorphic to sl2(R)s

  3. z(l)=Rh0 for some hyperbolic element h0 satisfying τ(h0)=-h0=θ(h0) which is zero if and only if gc is of tube type.

  4. The Lie algebra l is τ-invariant and lτso1,1(R)s.

  5. For xtq, we have ρ(adx)=ρ(adx|s), where ρ denotes the spectral radius. With the basis
    zj=(0,,0,120-110,0,,0),j=1,,s,
    in so2(R)s we have for x=j=1sxjzj
    ρ(adx)=max{|xj|:j=1,,s}. 2.5

Note that (c) implies that l is semisimple, i.e., h[l,l], if and only if gc is of tube type.

Proposition 2.9

([41, Prop. 7.10]) Let (g,τ,C) be a semisimple modular non-compactly causal semisimple symmetric Lie algebra, where τ=τhθ, hqpC a causal Euler element,

G:=InngC(g)Inn(g),K:=Gθ=Inng(k),andGc:=InngC(gc).

Then H:=GGc satisfies

H=Khexp(hp)andHK=Kh.

In particular KhKτh=Kτ implies HGτ.

The positivity domain and modular geodesics

Let (G,τG,H,C) be a connected semisimple causal symmetric Lie group with ncc symmetric Lie algebra (g,τ,C). We fix a causal Euler element hC (Theorem 2.5) and write M=G/H for the associated symmetric space.

One of our goals in this paper is to describe the structure of the positivity domain

WM+(h):={mM:XhM(m)V+(m)}

of the vector field XhM generating the modular flow. Our first major result is the identification of the connected component W of the base point eH in the positivity domain WM+(h) as

W:=WM+(h)eH=GehExpeH(Ωqk) 3.1

(Theorem 3.6).

Some of the results in this section had been obtained in [52] for the special case of ncc symmetric Lie algebras for which h contains an Euler element, whereas here we are dealing with general non-compactly causal symmetric Lie algebras.

Modular geodesics

In this subsection, we introduce the concept of an h-modular geodesic in a non-compactly causal symmetric space M and discuss some of its immediate properties. We also show that, in compactly causal spaces, non-trivial causal modular geodesics do not exist.

Definition 3.1

(Geodesics and causality) Let M=G/H as above.

  • We call a geodesic γ:RM causal if γ(t)V+(γ(t)) for every tR (see (1.1)).

  • Let hg be an Euler element. The flow on M defined by
    αt(gH)=exp(th)gH=gexp(Ad(g-1)h)H 3.2
    is called the modular flow (associated to h). Its infinitesimal generator is denoted XhMV(M).
  • A geodesic γ:RM is called h-modular if γ(t)=αt(γ(0)) holds for all tR, i.e., γ is an integral curve of XhM.

Proposition 3.2

Suppose that (g,τ) is a direct sum of irreducible ncc symmetric Lie algebras (g=gs). The following assertions hold for any Euler element hE(g) and the corresponding modular flow αt(m)=exp(th).m on M=G/H:

  1. The orbit under the modular flow is a causal geodesic if and only if m is contained in
    MCh={gHG/H:Ad(g)-1hC}. 3.3
  2. All connected components of MCh are Riemannian symmetric space of non-compact type: For every mMCh, the exponential map
    Expm:Tm(MCh)(Mh)m
    is a diffeomorphism.
  3. h-modular causal geodesics exist if and only if Oh=Ad(G)h intersects C. In this case Gh acts transitively on MCh.

Proof

(a) Assume first that Ad(g)-1hq. Then (3.2) implies that the orbit of m=gH under the modular flow is a geodesic. The causality is by definition equivalent to Ad(g)-1hC.

Suppose, conversely, that tαt(gH) is a causal geodesic. Lemma B.1 implies that Ad(g)-1h=xh+xq, where [xh,xq]=0 and xqC. By Lemma 2.4, xh is elliptic and xq is hyperbolic because it is contained in C. Therefore, adxh+adxq is the unique Jordan decomposition of adx into elliptic and hyperbolic summand. As Ad(g)-1h is an Euler element, the elliptic summand vanishes, and thus, adxh=0, i.e., xhz(g)h={0} (recall that z(g)q). This shows that Ad(g)-1hq, so that gHMCh.

(b) Choosing m as a base point, we may assume that m=eH, so that (a) implies that hCq is a causal Euler element. Pick a Cartan involution θ commuting τ which satisfies θ(h)=-h (cf. [28]), i.e., hqp. Then τ=τhθ follows from Theorem 2.5(a). As (MCc)m=Geh.m by Lemma B.2, the assertion now follows from gh=hkqp.

(c) The first assertion follows immediately from (a). For the second assertion, suppose that m0=g0HMCh. As MCh is Gh-invariant, Gh.m0MCh. Let hc:=Ad(g0)-1h, so that E(g)C=Inng(h)hc by Theorem 2.5(b) (recall that C=Cs). If gHMCh, i.e.,

Ad(g)-1hCE(g)=2.5Inng(h)hc=Inng(h)Ad(g0)-1h,

then there exists an element g1Inng(h) with gg1g0-1Gh, so that gGhg0g1-1Ghg0H, and therefore gHGh.m0.

We record the following consequence of (3.2):

Lemma 3.3

For any causal Euler element hC, we have

WM+(h)={gHG/H:Ad(g)-1hTC},whereTC:=h+C.

Due to the hyperbolicity of Euler elements, modular causal geodesics do not exist for compactly causal symmetric spaces:

Proposition 3.4

If M=G/H is a compactly causal symmetric space, then non-trivial causal modular geodesics do not exist.

Proof

If there exists a modular causal geodesic and (g,τ,C) is the infinitesimal data of M, then there exists a gG such that the Euler element h satisfies Ad(g)-1h=xh+xq with xqC and [xh,xq]=0 (Lemma B.1). As C is elliptic, xq is elliptic. Further the pointedness of C implies that xhker(adxq) is elliptic. This implies that the Euler element Ad(g)-1h is elliptic, a contradiction.

The fiber bundle structure of the positivity domain

The main result of this section is Theorem 3.6 in which we exhibit a natural bundle structure on the wedge domain WM that is equivariant with respect to the connected group Geh, the base is the Riemannian symmetric space of this group, and the fiber is a bounded convex subset of qk.

Definition 3.5

Let hqpC be a causal Euler element, so that τ=τhθ. Then zh(h)=hτh=hk implies that

Ohq:=Inng(h)h=eadhph

is the non-compact Riemannian symmetric space associated with the symmetric Lie algebra (h,θ).

Theorem 3.6

(Positivity Domain Theorem) Suppose that (G,τG,C,H) is a connected semisimple non-compactly causal Lie group for which (g,τ) contains no τ-invariant Riemannian ideals (g=gs) and that h is a causal Euler element. Suppose that C:=Cqmax(h) is the maximal Inng(h)-invariant cone with hC. Then, the following assertions hold:

  1. The connected component W=WM+(h)eH of eH in the positivity domain WM+(h) is given by
    W=Geh.ExpeH(Ωqk),whereΩqk={xqk:ρ(adx)<π2}. 3.4
  2. The polar map Ψ:Geh×GehHΩqkW,[g,x]g.ExpeH(x) is a diffeomorphism

  3. W is contractible, hence in particular simply connected.

  4. GehH=Keh.

Proof

(a) Recall from [41, Thm. 6.7] that the connected component of h in the open subset OhTC of Oh is

Inng(h)eadΩqkh=Ad(He)eadΩqkhTC. 3.5

If xΩqk, then ρ(adx)<π/2, so that (3.5) implies that g=expx satisfies

Ad(g)-1h=e-adxhTC=h+C. 3.6

By Lemma 3.3

ExpeH(Ωqk)WM+(h),andthusGeh.ExpeH(Ωqk)W 3.7

by Gh-invariance of WM+(h).

Conversely, for gHW, the element Ad(g)-1hOhTC is contained in the connected component of h, so that (3.5) implies that it is contained in Ad(He)eadΩqkh. Therefore

gHeexp(Ωqk)Gh.

This is equivalent to gHeGhexp(Ωqk), which implies

gHGhexp(Ωqk).eH=GhExpeH(Ωqk)

and thus

WGh.ExpeH(Ωqk). 3.8

If gGh satisfies gExpeH(Ωqk)W, then g.W=W follows from (3.7) and the fact that g permutes the connected components of WM+(h). Therefore, (3.8), combined with (3.7), leads with GWh:={gGh:g.W=W} to

WGWh.ExpeH(Ωqk)GWh.W=W,

and this entails

W=GWh.ExpeH(Ωqk). 3.9

Next we observe that the exponential map ExpeH:qkM is regular in every xΩqk because ρ(adx)<π/2<π ([52, Lemma C.3(b)]). Thus [52, loc.cit.] further implies that the map

Φ:Gh×ΩqkM,(g,x)g.ExpeH(x)

is regular in (gx) because Spec(adx)(-π/2,π/2)i does not intersect (π2+Zπ)i for xΩqk. This implies that the differential of Φ is surjective in each point of Gh×Ωqk; hence, the image of every connected component is open. Now the connectedness of W implies that WGeh.ExpeH(Ωqk), and this completes the proof.

(b)–(d): The surjectivity of Ψ follows from Theorem 3.6. As gh=hkqp is a Cartan decomposition of gh, the polar map Keh×qpGeh,(k,x)kexpx is a diffeomorphism. In particular,

GehHGehGG=KehH

implies GehH=Keh and thus (b).

The space Geh×GehHΩqk is a fiber bundle over Geh/Keh whose fiber is the convex set Ωqk. Therefore, it is homotopy equivalent to the base Geh/Keh, which is also contractible because the exponential map ExpeH:qpGeh/Keh is a diffeomorphism.

It therefore suffices to show that Ψ is a diffeomorphism. The proof of (a) shows already that its differential is everywhere surjective, hence invertible by equality of the dimensions of both spaces. So it suffices to check injectivity, i.e., that Exp:=ExpeH:qM satisfies

g1.Exp(x1)=g2.Exp(x2)g2-1g1Keh,x2=Ad(g2-1g1)x1. 3.10

Step 1: Exp|Ωqk is injective. If Exp(x1)=Exp(x2), then applying the quadratic representation implies exp(2x1)=exp(2x2) in G. As x1 and x2 are both exp-regular, [25, Lemma 9.2.31] implies that

[x1,x2]=0andexp(2x1-2x2)=e.

We conclude that e2ad(x1-x2)=idg, and since the spectral radius of 2ad(x1-x2) is less than 2π, it follows that ad(x1-x2)=0, so that x1=x2.

Step 2: g.Exp(x1)=Exp(x2) with gGeh and x1,x2Ωqk implies gKeh. Applying the involution θM, we see that g.Exp(x1) is a fixed point, so that

g.Exp(x1)=θ(g).Exp(x1)

entails that θ(g)-1g fixes m1:=Exp(x1). We now write g=kexpz in terms of the polar decomposition of Geh and obtain

θ(g)-1g=exp(2z)Gm1.

Applying the quadratic representation, we get

exp(2z)exp(2x1)exp(2z)=exp(2x1), 3.11

which can be rewritten as

exp(e2adx12z)=exp(-2z).

Since adz has real spectrum, so has e2adx1z. Therefore the same arguments as in Step 1 above imply that

[z,e2adx1z]=0andexp(2e2adx1z+2z)=e,

and e2adx1z=-z. The vanishing h-component of this element is sinh(2adx1)z, and since ρ(2adx1)<π, it follows that [x1,z]=0. Now (3.11) leads to exp(4z)=e, and further to z=0, because the exponential function on qp is injective. This proves that g=kKeh.

Step 3: From (3.10), we derive

g2-1g1.Exp(x1)=Exp(x2),

so that Step 2 shows that k:=g2-1g1Keh. We thus obtain

Exp(x2)=k.Exp(x1)=Exp(Ad(k)x1),

and since Ad(k)x1Ωqk, we infer from Step 1 that Ad(k)x1=x2. This completes the proof.

The following corollary identifies the connected component of MCh containing eH as a submanifold (cf. Lemma B.2) of the wedge domain W.

Corollary 6.4

Assume that τhG exists and leaves H invariant, so that τhM exists and leaves the base point eHM invariant. Then τhM(W)=W and the fixed point set of τhM in W is the Riemannian symmetric space

WτhM=MeHh=Geh.eH=ExpeH(qp).

Proof

For gGh and xqk:

τhM(gExpeH(x))=gExpeH(τh(x))=gExpeH(τ(x))=gExpeH(-x).

So gExpeH(x) is a fixed point if and only if ExpeH(-x)=ExpeH(x), which is equivalent to exp(2x)H. Now τ(x)=-x implies exp(2x)=exp(-2x). As ρ(2adx)<π, [52, Lemma C.3] further shows that x-(-x)=2xz(g)={0}. Therefore, gExpeH(x) is a fixed point if and only if x=0.

From W=Geh.ExpeH(Ωqk) and the polar decomposition Geh=Kehexp(qp)=exp(qp)(HK)e (Theorem3.6(b)), we derive that the fixed point set is

WτhM=Geh.eH=ExpeH(qp)=MeHh.

The preceding corollary shows that the wedge domain WM=G/H contains the symmetric subspace MeHh=ExpeH(qp) as the fixed point set of an involution. Hence, the description of W from Theorem 3.6 as

W=Geh.ExpeH(Ωqk)

suggest to consider W as a real “crown domain” of the Riemannian symmetric space MeHhGh/Hh.

Remark 3.8

Theorem 3.6 has a trivial generalization to semisimple non-compactly causal Lie algebras of the form g=gkgrgs because then

Cqmax=qrCqsmaxandTCqmax=gk+gr+TCqsmax.

For h=hr+hs with hsCqs the relation Ad(g)-1hTCqmax is therefore equivalent to Ad(g)-1hsTCqsmax. If M=Inn(g)/Inn(g)τMr×Ms is the corresponding product decomposition, we obtain

WM+(h)=Mr×WMs+(hs)forC=Cqmax.

However, if gr{0}, then Cqmax is not pointed, and there are many pointed invariant cones C, which are not maximal, for which the domain WM+(h) may have a more complicated structure.

Example 3.9

We consider the reductive Lie algebra

g=gl2(R)=R1sl2(R).

Any Euler element in g is conjugate to some

h=λ00μwithλ-μ=1.

The Cartan involution θ(x)=-x on g then satisfies θ(h)=-h and τ:=θτh acts by

τabcd=-acb-d.

With the Euler element

h0:=120110,

we then have

h=Rh0=so1,1(R)andq=R1+Rh+Rzqswithz:=120-110.

The group G:=GL2(R)e acts by g.A:=gAg on the 3-dimensional space Sym2(R) of symmetric matrices and the stabilizer of I1,-1:=100-1 is the subgroup H:=SO1,1(R) with Lie algebra h. Therefore M:=G.I1,1G/H can be identified with the subspace Sym1,1(R) of indefinite symmetric matrices. Note that Re×1=Z(G)e acts by multiplication with λ2 and that R+××M1M,(λ,A)λA is a diffeomorphism, where

M1:={AM:det(A)=-1}SL2(R)/SO1,1(R)dS2

is a realization of 2-dimensional de Sitter space. Note that the determinant defines a quadratic form of signature (1, 2) on Sym2(R) which is invariant under the action of the subgroup

{gGL2(R):|det(g)|=1}SL2(R)

which acts as SO1,2(R).

For the Euler element hs:=12100-1, we have

[h0,hs]=z,[z,hs]=h0and[h0,z]=hs.

According to [53, Ex. 3.1(c)], all Ad(H)-invariant cones in q are Lorentzian of the form

Cm={x01+x1(hs+z)+x-1(hs-z):x1x-1-mx020,x±10}forsomem>0.

Actually C0=Cqmax contains R1 and is not pointed.

(a) We write

h=λ+μ21+hs

to see that hCm is equivalent to

m(λ+μ)21.

We also note that the “semisimple part” of Cm is

Cm,s=Cmqs=Cm(Rh+Rz)=cone(hs±z)

coincides with the projection of Cm to qs, so that Cm,s=Cmqs.

Write W(Cm,h) for the positivity domain of the Euler element h with respect to the causal structure specified by the cone Cm. Then Theorem 3.6 implies that

W(C0,hs)=Ges.ExpeH(Ωqk),whereΩqk=(-π2,π2)z.

For xΩqk we have

e-adx.hsCs+hCm+h

(see (3.6)) and Gh=Ghs, so that we have

W(C0,hs)=Geh.ExpeH(Ωqk)W(Cm,hs)W(C0,hs)

implies the equality

W(Cm,hs)=Geh.ExpeH(Ωqk)forallm>0.

We also note that

W(Cm,h)W(C0,h)=W(C0,hs)=Geh.ExpeH(Ωqk)

because Ad(g)-1h=hz+Ad(g)-1hsC0 if and only if Ad(g)-1hsCs.

To determine the domain W(Cm,h) in general, we write

h=hz+hs=λ+μ21+12100-1.

By Geh-invariance, we have to determine when ExpeH(tz), |t|<π2, is contained in W(Cm,h). For g=exp(tz), we have

Ad(g)-1h=hz+e-tadzhs=hz+cos(t)hs-sin(t)h0,

and

pq(Ad(g)-1h)=hz+cos(t)hs=λ+μ21+cos(t)2(hs+z)+cos(t)2(hs-z).

We then have

x0=λ+μ2andx±1=cos(t)2.

We conclude that, for |t|<π2, the inclusion hz+cos(t)hs(Cm) is equivalent to

x1x-1-mx02=14(cos2(t)-m(λ+μ)2)>0.

We thus obtain the condition

|t|<arccos(m|λ+μ|).

For m>0 and hhs, this is specifies a proper subinterval of (-π2,π2).

(b) To determine which cone Cm corresponds to the canonical order on the space Sym1,1(R), induced from the natural order of Sym2(R) (which is also Lorentzian), we evaluate the tangent map qSym2(R),xxI11+I11x to

1.I1,1=2I1,1,hs.I1,1=1,z.I1,1=0110.

We thus obtain for x=x01+x1(hs+z)+x-1(hs-z) that

x.I1,1=xI11+I11x=2x0+x1+x-1x1-x-1x1-x-1-2x0+x1+x-1.

By the Hurwitz criterion, this matrix is positive semidefinite if and only if

x1+x-1|2x0|

and

(x1+x-1)2-4x02-(x1-x-1)2=4(x1x-1-x02)0.

Is x1+x-10, then these two inequalities are equivalent to x1x-1-x020. As these two conditions imply that x±10, we see that the canonical order on M corresponds to the cone C1, i.e., to m=1.

(c) For the modular vector field Xh, we have

Xhabbd=2λa(λ+μ)b(λ+μ)b2μd.

The positivity domain of Xh depends on λ, and with this formula one can also determine the positivity domain quite directly for m=1, where C1 corresponds to the canonical order.

Example 3.10

(cf. [52, Exs. 2.11, 2.25]) Let G:=GLn(R)+ and K:=SOn(R). We consider the Riemannian symmetric space

Mr:=Symn(R)+GLn(R)/SOn(R)

and the corresponding irreducible subspace

Mr,s:={AM:det(A)=1}SLn(R)/SOn(R)

(here the index s refers to “semisimple”). On g=gln(R), we consider the Cartan involution given by θ(x)=-x and write n=p+q with p,q>0. Then

hsp:=1nq1p00-p1qsln(R)andhp:=hsp-qn1=0001qg 3.12

are Euler elements and τ:=τhpθ leads to a non-compactly causal symmetric Lie algebra (g,τ,C), where

h=sop,q(R)andq={ab-bd:a=a,d=d}

To identify G/H in the boundary of the crown domain in GC/KCGC.1Symn(C)×, where GC acts on Symn(C) by g.A:=gAg ([52, Thm. 5.4]), we observe that

exp(ithp).1=e2ithp=cos(2thp)+isin(2thp)=1p00(cos(2t)+isin(2t))1q,

so that we obtain for t=π2 the matrix

exp(πi2hp).1=Ip,q.

The G-orbit of this matrix is the open subset

M:=G.Ip,q={gIp,qg:gGLn(R)+}=Symp,q(R)

of symmetric matrices of signature (pq). We have

Xhpabbd=hpabbd+abbdhp=0bb2d.

These matrices are never positive definite. So we have to take hs instead to find non-trivial positivity domains.

For the case p=q=1 and n=2, this has been carried out in Example 3.9. We also write

h=λ1p00μ1qwithλ-μ=1.

Then

Xhabbd=2λa(λ+μ)b(λ+μ)b2μd,

so that

Xh(Ip,q)=2λ1p00-μ1q>>0if0<λ<1,

which is equivalent to λμ<0.

The connected components of MCh

The main result in this section is Proposition 3.11 on the subgroup HK of Kh. We then discuss several examples to clarify the situation.

Proposition 3.11

(Connected components of MCh) If G=Inn(g) and (g,τ) is irreducible ncc with causal Euler element h, then π0(MCh)Kh/HK contains at most two elements.

Proof

We recall from Proposition 3.2(c) that MCh=Gh.eH. With [36, Thm. IV.3.5] we see that the symmetric space Gh.eHGh/Hh is a vector bundle over Kh/HKh, hence in particular homotopy equivalent to Kh.eHKh/HKh. In view of Proposition 2.6(c), we have for G=Inn(g) that Hh=Hτh=HKGτ is a maximal compact subgroup of H. It follows in particular that Hh=HKhKh. We conclude that π0(MCh)π0(Kh/HK). From [41, §7], we know that π0(Gh)π0(Kh) has at most two elements.

Example 3.12

(The inclusion HKKh may be proper) We have Gh=Khexp(qp) and KG=KτhG because K=Gθ. Further HKKh by Proposition 2.6(a), so that the equality HK=Kh is equivalent to KhHK. This may fail for two reasons. One is failure in the adjoint group Inn(g) (Proposition 3.11), and the other reason is that Z(G) may be non-trivial.

Assume that g is semisimple and (g,τ,C) ncc. Let G be a corresponding connected Lie group on which τG exists (for τ=τhθ) and H:=GeG. For the connected group K:=Gθ, the intersection HK:=HK=exphk is connected but KhZ(G)HK is in general not connected because Z(G) need not be contained in HK.

This can be seen easily for g=sl2(R). For

h=12100-1,θ(x)=-x,wehaveτ=θτh,abc-a-acba. 3.13

For any connected Lie group G with Lie algebra g, the group K=Gθ is connected 1-dimensional and τ(k)=k-1 for kK. Moreover, Kh=Z(G) is a discrete subgroup which intersects H=exphR trivially. Even the inclusion KhGG fails if |Z(G)|3, i.e., if τ acts non-trivially on Z(G). Note that Z(G) is infinite if G is simply connected.

Example 3.13

(a) For g=sl2(R), we consider again the Euler element h from (3.13) and the Cartan involution θ(x)=-x. By Lemma B.1, the α-orbit of gH is a geodesic if and only if Ad(g)-1h commutes with τ(Ad(g)-1h)=-Ad(τ(g))-1h, i.e., if

Ad(gτ(g)-1)hzg(h)=Rh.

As OhRh={±h}, this leaves two possibilities:

  1. If Ad(gτ(g)-1)h=h, then Ad(τ(g))-1h=Ad(g)-1h implies Ad(g)-1hq.

  2. If Ad(gτ(g)-1)h=-h, then -Ad(τ(g))-1h=Ad(g)-1h implies Ad(g)-1hh. In this case gH is a fixed point of the modular flow.

(b) For g=sl2k(R) with the Cartan involution θ(x)=-x and the causal Euler element

h=121k00-1k,

we obtain h=sok,k(R) for τ=θτh. There exists a subalgebra ssl2(R)k, where the sl2-factors correspond to the coordinates xj and xj+k for 1jk. Accordingly, h=j=1khj, where the Euler elements hj in the sl2-factors are conjugate to Euler elements hj in h. Therefore, the “geodesic condition” is satisfied by all elements j=1kh~jOh, where h~j is either hj or hj.

The following example shows that modular geodesics also exist in symmetric spaces without causal structure. They can be “space-like” rather than “time-like”, resp., causal.

Example 3.14

The d-dimensional hyperbolic space

Hd:={x=(x0,x)R1,d:x02-x2=1,x0>0}SO1,d(R)/SOd(R)

carries a modular flow specified by any Euler element hqso1,d(R) (corresponding to a tangent vector of length 1). Every geodesic of Hd is an orbit of the flow generated by an Euler element of so1,d(R).

Remark 3.15

Let (g,τ,C) be a simple ncc symmetric Lie algebra. In general, we have for a causal Euler element hE(g)C a proper inclusion

Inng(h)h=OhCOhq.

By Lemma B.4, this implies that Mh is not connected and MChMh.

For instance, if g=hC and h is simple hermitian of tube type, then we obtain for any pointed generating invariant cone Chh a hyperbolic cone C:=iChq=ih. If hE(g)C is a causal Euler element, then -hAd(G)h follows from [38, Thm. 3.10] and the subsequent discussion, but -hC; see also [41, Thm. 5.4].

Example 3.16

(a) For de Sitter space M=dSd (cf. Example 4.6 and Appendix D), the subspace MeHh=ExpeH(Rh) is a single geodesic, hence in particular 1-dimensional. Note that dimqp=1 in this case. The modular flow on M has the fixed point set MαSd-2.

(b) For M=GC/G, g hermitian, we have MeGh=Expe(ik) with dual symmetric space the group K, considered as a symmetric space.

Open H-orbits in flag manifolds and a convexity theorem

In this section, we prove a convexity theorem that is vital to derive the equality W=W(γ) in the next section. Here, as above, W=WM+(h)eH.

Let P-:=exp(g-1(h))GhG be the “negative” parabolic subgroup of G specified by h and identity g1(h) with the open subset B:=exp(g1(h)).eP-G/P-. Then D:=H.0B is an open convex subset, and our convexity theorem (Theorem 4.5) asserts that, for any gG with g.DB, the subset g.DB is convex.

We consider a connected semisimple Lie group G with Lie algebra g and an Euler element hg. We put

n±:=g±1(h)andN±:=exp(n±),

and write

P±:={gG:Ad(g)g±1(h)=g±1(h)}=N±GhN±Gh

(see [2, Thm. 1.12] for the equality) for the corresponding maximal parabolic subgroups. We write

M±:=G/P

for the corresponding flag manifold. The abelian subgroup N+ has an open orbit B:=N+.eP-M+, which we call the open Bruhat cell. It carries a natural affine structure because the map

φ:n+B:=N+.eP-,xexp(x)P-

defines an open embedding. Below we shall always use these coordinates on B.

Choose a Cartan involution θ with θ(h)=-h and consider the involution τ:=θeπiadh. We write

H:=Khexp(hp)forhp=gτpandHK:=Kh.

Then

P±H=GhH=Kh,

so that

D+:=H.eP-H/HK 4.1

is an open H-orbit in BG/P-. It is a real bounded symmetric domain ([26, Thm. 5.1.8]) and coincides with the unit ball in the positive real Jordan triple

V:=nwith{x,y,z}:=xy.z=-12[[x,θ(y)],z] 4.2

(cf. [2, (4.6)])

The open H-orbits in G/P±

Lemma 4.1

([2, Cor. 1.10]) For yg-1(h) and xg1(h), we have exp(y).exp(x)P-B if and only if the Bergman operators

B+(x,y):=1+ad(x)ad(y)+14(adx)2(ady)2End(g1(h))

and

B-(y,x):=1+ad(y)ad(x)+14(ady)2(adx)2End(g-1(h))

are both invertible.

Remark 4.2

Note that

θB-(y,x)θ=1+ad(θ(y))ad(θ(x))+14(adθ(y))2(adθ(x))2=B+(θ(y),θ(x)).

Example 4.3

We consider the group G=SL2(R) with Lie algebra g=sl2(R) and the linear basis

h:=12100-1,e=0100,f=0010, 4.3

satisfying

[h,e]=e,[h,f]=-f,[e,f]=2h.

Then,

N+=1R01,N-=10R1,andGh={a00a-1:aR×},

so that

P-={a0ca-1:aR×,cR}={gG:ge2Re2}.

For K=SO2(R), we have Kh={±1}. Identifying G/P- with the projective space P(R2)=G.[e2], the Bruhat cell is

B={[x:1]:xR}R,

and G acts by

g.x=ax+bcx+dforax+b0.

In particular, we have

exp(yf).x=x1+xy. 4.4

We consider the Cartan involution θ(x)=-x, so that τ:=θeπiadh acts by

τabc-a=-acbaandτGabcd=dcba.

Then

H=SL2(R)G={abba:a2-b2=1}=SO1,1(R),

so that

D+=H.0={ba-1:a2-b2=1}=(-1,1). 4.5

Note that Ad(H)H/{±1} is connected.

The Jordan triple product satisfies

{e,e,e}=-12[[e,θ(e)],e]=-12[[e,-f],e]=12[2h,e]=e,

so that

{xe,ye,ze}=xyz·eforx,y,zR.

Further

(adf)e=-2h,(adf)2e=-2[f,h]=-2f,(ade)f=2h,(ade)2f=-2e

implies

B+(xe,yf)e=e+xyad(e)ad(f)e+x2y24(ade)2(adf)2e=e+xyad(e)(-h)+x2y24(ade)2(-2f)=e+xy2e+x2y244e=(1+2xy+x2y2)e=(1+xy)2e.

Moreover,

B+(θ(yf),θ(xe))e=B+(-ye,-xf)e=(1+(-y)(-x))2e.=B+(xe,yf)e.

As 1+xy is invertible for all x with |x|<1 if and only if |y|1, it follows that

exp(y).D+B|y|1.

Now back to the general case. In the following we write · for the spectral norm on the Jordan triple system g1(h)=n+. If x=j=1kxjcj with pairwise orthogonal tripotents cj, then

x=max{|x1|,,|xk|}. 4.6

If

Dg:={xg1(h):x<1}, 4.7

then we have

D+=exp(Dg)P-G/P- 4.8

([26, Thm. 5.1.8]).

Proposition 4.4

The following assertions hold:

  1. g.D+B is equivalent to gP+exp(y) for yn- with y1.

  2. g.D+B is relatively compact if and only if gP+exp(y) for yn- with y<1.

Proof

The condition g.eP-B is equivalent to gN+P-=N+GhN-=P+N-. Let yn- with gP+exp(y). Then the invariance of B under P+ implies that g.D+B is equivalent to exp(y).D+B.

(a) Suppose first that exp(y).D+B. By the Spectral Theorem for positive Jordan triples ([59, Thm. VI.2.3]3), there exist pairwise orthogonal tripotents c1,,ck and β1,,βkR with

y=j=1kβjθ(cj)

([59, Thm. VI.2.3]). For x=jαjcj and z=jγjcj, we then have

{x,θ(y),z}=j=1kαjβjγj·cj

([59, Prop. V.3.1]). As xDg is equivalent to

x=max{|αj|:j=1,,k}<1,

the calculations in Example 4.3 show that exp(y).D+B implies y1.4

To prove the converse, suppose first that y<1. Then

exp(-y)=θ(exp(-θ(y))θ(HP-)=HP+

implies exp(y)P+H, so that

exp(y).D+P+H.D+=P+.D+B.

Now we assume that y=1. We observe that

exp(y).x=exp(-th)exp(th)exp(y).x=exp(-th)exp(e-ty).(etx),

so that, for r>0, exp(y).xB is equivalent to exp(r-1y).(rx)B. For xDg, we pick r>1 with rxDg. Then r-1y<1 implies exp(r-1.y).exp(rx)B, and thus exp(y).exp(x)B. This shows that exp(y).D+B.

(b) If y<1, then the argument under (a) shows that exp(y).D+P+.D+ is relatively compact.

Now we assume that y=1. We show that this implies that exp(y).D+ is unbounded. As above, we use the Spectral Theorem to write

y=j=1kβjθ(cj)

and observe that there exists an {1,,k} with |b|=1. For x=jαjcjDg, we then obtain with (4.4)

exp(y).x=j=1kαj1+αjβjcj.

For x=αc we get in particular

exp(y).x=α1+αβc.

For α-sgn(β) these element leave every compact subset of B. Therefore, exp(y).D+ is unbounded.

Theorem 4.5

(Convexity theorem for conformal balls) If gG is such that g.D+B, then gD+ is convex. If g.D+ is relatively compact in B, then there exists an element pP+ with g.D+=p.D+, so that g.D+ is an affine image of D+.

Proof

If g.D+B is relatively compact, then Proposition 4.4(b) and its proof imply the existence of pP+ with g.D+=p.D+. In particular g.D+ is an affine image of D+ and therefore convex.

If g.D+B is not relatively compact, then we put rn:=1-1n. Now

expy.D+=nNexpyexp(rnDg)P-

is an increasing union. Therefore it suffices to show that the subsets expyexp(rnDg)P- are convex. For rn=et we have

e-t·(expyexp(rnDg)P-)=exp(-th).(exp(y)exp(rnDg)P-)=exp(ety).exp(e-trnDg)P-=exp(rny).D+,

and these sets are convex by the preceding argument.

Example 4.6

We consider G=SO1,d(R)e as the identity component of the conformal group of the Euclidean space Rd-1, H=Ge1=SO1,d-1(R), and the Euler element hso1,1(R) with h.e0=e1 and h.e1=e0. As Z(SO1,d-1(R)){±1} and G preserves the positive light cone, the center of G is trivial.

The symmetric space M=G/HG.e1dSd is d-dimensional de Sitter space, P=N-Gh is the stabilizer of the positive light ray R+(e0-e1), and G/PSd-1 is the sphere of positive light rays. On the sphere Sd-1, the subgroup H has two open orbits which are positive half-spheres separated by the sphere Sd-2 of positive light rays in the subspace e1.

In the sphere the Bruhat cells are the point complements and if g.D+BRd-1, then the convexity of g.D+ is well-known from conformal geometry because conformal images of balls are balls or half spaces.

The subset realization of the ordered space M=G/H

As before G is assumed to be a connected semisimple Lie group. To simplify the notation, we write M for M+=G/P-. Recall the following fact about the compression semigroup of the H-orbit D+=H.eP-M+, which is the Riemannian symmetric space H/HK.

Lemma 4.7

The compression semigroup of the open H-orbit D+=H.eP-G/P- is

comp(D+)={gG:g.D+D+}=Hexp(-Cqmax(h)), 4.9

Proof

This result was announced in [57, 58], and a detailed proof was given in [24, Thm. VI.11] for the case where GGC, GC is simply connected and H=Gτ. In this case Ad(Gτ) preserves Cqmax(h), so that GτKhexp(hp). Conversely, Kh leaves D+ invariant, so that we obtain Gτ=H=Khexp(hp) in this particular case.

To see that the lemma also holds in the general case, note that the center of G acts trivially on G/P- and that Z(G)KhH. Therefore, the general assertion follows if the equality (4.9) holds at least for one connected Lie group G with Lie algebra g. Hence, it follows from the special case discussed above.

We now use this to realize G/H as an ordered symmetric space as a set of subsets of M and describe the ordering in that realization.

Proposition 4.8

(The subset realization of ncc symmetric spaces) Let G be a connected semisimple Lie group, hg an Euler element, θ a Cartan involution with θ(h)=-h and τ:=θτh, so that (g,τ) is a ncc symmetric Lie algebras with g=gs. Let D+G/P- be the open orbit of the base point under H:=Khexp(hp). We endow the homogeneous space

MD+:={g.D+:gG},

consisting of subsets of M, with the inclusion order. Then the stabilizer subgroup GD+ of the base point is H. The map gHg.D+ induces an isomorphism

(MD+,)(G/H,Cqmax(h))

of ncc symmetric spaces, where Cqmax(h) is the unique maximal Ad(H)-invariant cone in q containing h in its interior.

In this identification, the set {xG/H:xeH} is mapped to {s-1D+:scomp(D+)} and {xG/H:xeH} is mapped to {sD+:scomp(D+)}. In particular, gHeH is equivalent to D+g.D+ and eHgH to g.D+D+.

Proof

This follows from Lemma 4.7.

Remark 4.9

(The Riemannian case) Let (g,θ) be a Riemannian symmetric Lie algebra, i.e., g=gr. Then H=K, M=G/K and h=0. Thus G=P- and G/P- is a single point. Hence comp(D+)=G and MD+ is a single point. Therefore, Riemannian summands cannot be permitted in Proposition 4.8.

Example 4.10

Let G=SL2(R) and h=12100-1. Then, the canonical action of G on P1(R)=P(R2)S1=R{} is given by

abcd.x=ax+bcx+d

and the stabilizer of 0 is

P-={a0ca-1:a0,cR}=exp(g-1(h))Gh,Gh={a00a-1}R×.

The 1-parameter group

at=coshtsinhtsinhtcosht

fixes ±1 and the orbit D+=H.0 of H={±at:tR} is the open unit interval D+=(-1,1). The maximal cone in q is generated by Ad(H)R+h.

Since elements of P1(R) represent one-dimensional linear subspaces of R2 and SL2(R) acts transitively on triples of such subspaces, it follows easily that it acts transitively on the set of non-dense open intervals IS1, the ordered space G/H can be identified with the ordered set of open non-dense intervals in S1.

Example 4.11

A special case of the above construction is the “complex case” where H is a connected semisimple Lie group of hermitian type contained in a complex Lie group G with Lie algebra hC=hih. Then, G/H is a ncc symmetric space. Let θH be a Cartan involution on H. Then θH extends to a Cartan involution θ on G. Denote the corresponding maximal compact subgroup of G by K. Then HK is a maximal compact subgroup of H and the Riemannian symmetric space H/HK can be realized as complex symmetric bounded domain D+G/P-. Let z0z(hk) be the element determining the complex structure on H/HK. Then h=-iz0 is an Euler element in q=ih. Now (4.1) is the Harish–Chandra realization of H/HK as D+ in G/P- (see [60, p. 58] or [22, Ch. VII] for details).

Suppose that the complex conjugation τ of g with respect to h integrates to an involution τG on G. This is the case if G is simply connected or if G=Inng. We then assume that H=GeG. If G is simply connected, then H=GG is connected and [24, Thm. VI.11] implies that H=GD+, where GD+ is the stabilizer of the base point D+.

But in general, if G is not simply connected, then GD+ and GG may differ.

As an example, consider H=PSL2(R)G=PSL2(C)Inn(g) and note that τG(g)=τgτ in this case. Then

GG=PSL2(C)GPGL2(R)Aut(sl2(R)),

which is not connected because it also contains the image of i00-i. The domain D+=H.iC (the Riemann sphere) is the upper half plane and the stabilizer subgroup of D+ is

GD+=PSL2(R)PGL2(R)=GG.

The reflections in GL2(R) exchange the two open H-orbits.

Remark 4.12

The flag manifolds M=G/P-K/Kh appearing in this section are compact symmetric spaces on which the maximal compactly embedded subgroup KG acts by automorphisms. These spaces are called symmetric R-spaces.

Defining a symmetric R-space as a compact symmetric space M which is a real flag manifold, Loos shows in [37, Satz 1] that this implies the existence of an Euler element hE(g) such that MG/P-, so that MK/KP-=K/Kh as a Riemannian symmetric space (see [41, §7.2] for more details).

If G is hermitian of tube type, then MK/Kh can be identified with the Šhilov boundary of the corresponding bounded symmetric domain DGG/K, and this leads to a G-invariant causal structure on M. As dimZ(K)=1, with respect to the K-action, we have a natural 1-parameter family of K-invariant Lorentzian structures on M. They correspond to Kh-invariant Lorentzian forms on TeKh(M)qk=z(k)[hk,qk] which are positive definite on z(k) and negative definite on its orthogonal space [hk,qk].

Observer domains associated with modular geodesics

In this section, we associate to any modular causal geodesic γ in an ncc semisimple symmetric space M=G/H an observer domain W(γ). It is an open connected subset of M invariant under the centralizer Gh of the corresponding causal Euler element h. We then show that, for hC, the domain W(γ) coincides with the connected component WWM+(h) of the base point eH of the corresponding positivity domain. In Sect. 6, we show that WM+(h) is connected for G=Inn(g), which implies that W=WM+(h) in this case.

Definition 5.1

Let (G,τG,H,C) be a non-compactly causal symmetric Lie group and M=G/H be the corresponding ncc symmetric space. We assume that g=gs, i.e., that (g,τ) is a direct sum of irreducible ncc symmetric Lie algebras.

(a) We write for the order on M defined by the closed Olshanski semigroup S=Hexp(C)=exp(C)H which always exists because z(g)={0} ([30, Thm. 3.1] or Theorem C.1in Appendix C) via

g1Hg2Hifg1-1g2SH/H=ExpeH(C)

and write order intervals as

[x,y]={zM:xzy}=xy,

where

x={zM:xz}andy={zM:zy}.

(b) A subset XM is called order convex if

[a,b]Xfora,bX.

As the intersection of order convex subsets is order convex, we can defined the order convex hull

oconv(D):={DM:DD,Dorderconvex}.

Clearly oconv(D) is the smallest order convex subset of X containing D.

(c) For a modular geodesic γ:RM, we call

W(γ):=γ(R)γ(R)=t<s[γ(t),γ(s)]

the observer domain associated to γ. Note that this domain depends on the cone Cq specifying the order on M.

Lemma 5.2

The subset W(γ) has the following properties:

  1. W(γ)M is open and connected.

  2. W(γ)=oconv(MeHh) for MeHh=ExpeH(qp).

  3. Suppose that HK=Kh and C=Cqmax and identify M=G/H with MD+ (Proposition 4.8). Then
    W(γ)={g.D+:0g.D+,g.D+bounded} 5.1
    and this domain is Gh-invariant.

Proof

(a) To see that W(γ) is open, we first observe that γ(s)(γ(t)) for t<s. For real numbers tjR with t1<t2<t3<t4, this implies that

[γ(t2),γ(t3)][γ(t1),γ(t4)].

This shows that W(γ) is open.

To see that W(γ) is connected, we recall that the order on M is globally hyperbolic, in particular all order intervals [xy] are compact. As all elements z[x,y] lie on causal curves from x to y ([23, Thm. 4.29]), the order intervals are pathwise connected. As an increasing union of the order intervals [γ(-n),γ(n)], the wedge domain W(γ) is connected.

(b) Order intervals are convex and directed unions of convex sets of convex. Therefore,

W(γ)=t<s[γ(t),γ(s)]

is convex, whence W(γ)=oconv(γ(R)).

From the fact that h is central in hk+qp, it easily follows that, in the symmetric space MeHh=ExpeH(qp) the geodesic line γ(R) is cofinal in both directions because we have in q:

s<t(sh+C)(th-C)q.

For xqp, we thus find s,tR with xsh+C and xst-C. Then

Exp(sh)<Exp(x)<Exp(th)

in Meh. This implies that

W(γ)ExpeH(qp)=MeHh=Geh.eHγ(R).

This completes the proof.

(c) The modular group acts on BN+.eP-G/P- by exp(th).x=etx. Therefore γ(t)=etD+ enlarges D+ for t>0 and shrinks D+ for t<0 (Theorem 4.5). As γ is strictly increasing, this implies that

γ(R)={g.D+:(tR)g.D+etD+}={g.D+:g.D+boundedinB}.

Further

γ(R)={g.D+:(tR)g.D+etD+}={g.D+:0g.D+},

so that (5.1) follows. As any gGh=P+P- acts by linear maps on the Bruhat cell Bg1(h), (5.1) implies that Gh leaves the set W(γ) of all bounded domains g.D+ containing 0 invariant.

Example 5.3

(de Sitter space) We consider de Sitter space

M=dSd={(x0,x)R1,d:β(x,x)=-1},whereβ(x,y)=x0y0-xy

is the canonical Lorentzian form on R1,d (cf. Sect. D). Here

G=SO1,d(R)=SO1,d(R)e,H=Ge1=SO1,d-1(R)

and

CTe1(M)=e1givenbyC={(x0,x):x1=0,x00,x02x2}.

We claim that, for the modular geodesic

γ(t)=cosh(t)e1+sinh(t)e0=ethe1,

we have

W(γ)=WdSd(h)={xdSd:x1>|x0|}=WRdSd, 5.2

where WR={(x0,x):x1>|x0|} (cf. Appendix D in [52]). As the right wedge WRR1,d is causally complete, we clearly have W(γ)WRdSd=WdSd(h). For the converse inclusion, let xWR. We have to find a tR with xγ(t), i.e.,

x0<γ(t)0=sinh(t)

and

0<β(γ(t)-x,γ(t)-x)=(sinh(t)-x0)2-(cosh(t)-x1)2-x22--xd2.

Since β(γ(t),γ(t))=-1, we obtain for the right hand side

β(γ(t)-x,γ(t)-x)=β(γ(t),γ(t))-2β(γ(t),x)+β(x,x)=-1-2β(γ(t),x)+β(x,x).

Further

-2β(γ(t),x)=2x1cosh(t)-2x0sinh(t)et(x1-x0)fort>>0,

and if x1>|x0|, this expression is arbitrarily large for t. This shows that WRγ(R), and we likewise see that WRγ(R).

Proposition 5.4

If HK=Kh and C=Cqmax, then

  1. W(γ)W=WM+(h)eH.

  2. h+Dgh+C.

Proof

If gHW(γ)G/H, then the corresponding subset g.D+B is convex by Theorem 4.5, and it contains 0 by (5.1). Therefore the curve

η:RM,η(t):=exp(th)gH

is increasing because tetg.D+ is an increasing family of subsets of B. The invariance of the order thus implies that

g-1.η(0)=pq(Ad(g)-1h)Cqmax.

We also know that g.D+P+.D+ (Theorem 4.5 and Lemma 5.2(c)), so that there exist g1Gh and yg1(h) with g.H=g1exp(y).H. Thus

Ad(g)-1hAd(H)e-adyhh+Cqmax, 5.3

and therefore

e-adyh=h-[y,h]=h+yh+Cqmax.

Recall the definition of Dg in (4.7). The condition

eP-g.D+=g1exp(y).D+=g1.exp(y+Dg)P-

is equivalent to -yDg=-Dg, showing that

W(γ)=Ghexp(Dg).D+MD+ 5.4

(cf. Lemma 5.2(c)). We therefore derive from (5.3) that h+Dgh+Cqmax, and since hCqmax, and D+ is starlike with respect to 0, we obtain

h+Dgh+Cqmax,. 5.5

We thus obtain Ad(g)-1.hh+Cqmax,, i.e., gHWM+(h). This shows that W(γ)WM+(h), and the connectedness of W(γ) (Lemma 5.2(a)) yields W(γ)W.

Remark 5.5

From (5.4) it follows that, as a subset of M,

W(γ)=Ghexp(Dg).H=Gehexp(Dg).H. 5.6

For the quotient map q:GG/H, this means that

q-1(W(γ))=Ghexp(Dg)HG.

This is a Gh×H-invariant domain in G specified by its intersection with the abelian subgroup N+=exp(g1(h)); see [41, Rem. 6.2].

Combined with Theorem 7.1, that asserts the connectedness of WM+(h), the following result implies that WM+(h)W(γ).

Proposition 5.6

If HK=Kh and C=Cqmax, then WW(γ).

Proof

As both sides are Geh-invariant (Lemma 5.2), the Positivity Domain Theorem (Theorem 3.6) implies that we have to verify the inclusion

ExpeH(Ωqk)W(γ).

Invariance of both sides under (HK)e and Ad((HK)e)tq=qk further reduce the problem to the inclusion

ExpeH(Ωtq)W(γ). 5.7

To this end, we use the Lie subalgebra lg generated by h and tq (Proposition 2.8). Then [l,l]sl2(R)s and tqso2(R)s. This reduces the verification of the inclusion (5.7) to the case where g=sl2(R)s, h=so1,1(R)s and tqso2(R)s.

As this is a product situation, it suffices to consider the case where

g=sl2(R)h=so1,1(R),tq=so2(R)andh=12100-1.

By (5.6), we have to show that

exp(tx)Gehexp(Dg)Hfor|t|<π/2andx=120-110. 5.8

We identify sl2(R) with 3-dimensional Minkowski space R1,2, via

e0:=120-110,e1:=120110,e2:=h=12100-1.

In the centerfree group G:=Inn(g)SO1,2(R)e, we have

K:=Ge0SO2(R)andKh={e}=HK,

so that H:=Ge1=exp(hp)=SO1,1(R)e is connected. Therefore, G/HG.e1=dS2 (de Sitter space) and exp(tx)H corresponds to

exp(tx).e1=cos(t)e1+sin(t)e2.

Now |t|<π/2 implies cos(t)>0, hence that

exp(tx).e1WdS2(γ)forγ(t)=cos(t)e1+sin(t)e0 5.9

(Example 5.3). We write elements of D+ as y=se, |s|<1 (see Example 4.3). Then exp(y).e1 corresponds to

eady.12(e+f)=12eadse(e+f)=12(e+eadsef)=12(e+f+s[e,f]+s22[e,[e,f]])=12(e+f)+sh+s22[e,h]=12(e+f)+sh-s22e,

so that

exp(y).e1=e1+se2-s22(e1-e0)=(s22,1-s22,s).

This element lies in the wedge domain WdS2(h) if and only if 1-s2/2>s2/2 (Example 5.3), which is equivalent to |s|<1. Then its Geh-orbit contains the element (0,1-s2,s). For |t|<π/2, the element exp(tx).e1 is of this form, showing that exp(tx)Ghexp(y)H. This completes the proof.

Combining the preceding two propositions, we get the main result of this section. It shows that the observer domain W(γ) coincides with a connected component of the positivity domain WM+(h). This result provides two complementary perspectives on this domain.

Theorem 5.7

(Observer Domain Theorem) Let (g,τ,C) be a non-compactly causal semisimple symmetric Lie algebra with causal Euler element hCqp with τ=τhθ and let G be a connected Lie group with Lie algebra g and H:=Khexp(hp). If C=Cqmax, then W=W(γ).

We can even extend this result to coverings:

Corollary 7.2

If HH=Khexp(hp) is an open subgroup and C=Cqmax, then W:=WM+(h)eH=W(γ~) holds in M=G/H for γ~(t):=ExpeH(th).

Proof

Let q:M=G/HG/HMD+ be the canonical equivariant covering from [41, Lemma 7.11].

First we show that WM is order convex. So let xyz in M with x,zW and let η:[0,2]M be a causal curve with

η(0)=x,η(1)=y,η(2)=z.

Then q(η(t))[q(x),q(z)]W for t[0,2] holds because W=W(γ) is order convex in M.

As W is contractible by Theorem3.6(b), it is in particular simply connected. Therefore, q-1(W) is a disjoint union of open subsets (Wj)jJ mapped by q diffeomorphically onto W. By definition, W is one such connected component, so that

qW:=q|W:WW

is a diffeomorphism. Therefore η is the unique continuous lift of qη in M, hence contained in W. This implies that yW, so that W is order convex.

As qW:WW is an isomorphism of causal manifolds, it also is an order isomorphism. Finally W(γ)=oconv(γ(R))=W implies that W(γ)=oconv(γ~(R))=W.

Remark 5.9

It is not clear to which extent W(γ) depends on the specific cone C. In particular it would be interesting to see if the minimal and maximal cones lead to the same domain W(γ). We have already seen that the positivity domain WM+(h) depends non-trivially on the cone C ([41, Ex. 6.8]) so one may expect that this is also the case for W(γ).

Lemma 5.10

The involution τM on M defined by τM(gH)=τG(g)H satisfies

τM(WM+(h))=WM+(h)andτM(W(γ))=W(γ). 5.10

Proof

(a) The condition gHWM+(h) is equivalent to Ad(g)-1hTC by (5.10), and this implies that

Ad(τ(g))-1(-h)=τ(Ad(g)-1h)τ(TC)=-TC,

so that Ad(τ(g))-1hTC, i.e., τM(gH)WM+(h). As τM is an involution, it follows that τM(WM+(h))=WM+(h).

(b) As τ(C)=-C, the involution τM reverses the causal structure on M. Moreover, τM(γ(t))=γ(-t), so that

τM(W(γ))=t<s[τM(γ(s)),τM(γ(t))]=t<s[γ(-s),γ(-t)]=W(γ).

We have seen above that, for the modular geodesic γ(t)=ExpeH(th) in M, we have W(γ)=W. The modular geodesic γ is a specific orbit of the modular flow inside W. Now we show that all other α-orbits in W lead to the same “observer domain”.

Proposition 5.11

Let mW and consider the curve

β:RW,β(t)=αt(m)=exp(th).m.

Then

W=W(β)=s<t[β(s),β(t)]. 5.11

Proof

Using the subset realization of M=G/H as MD+={g.D+:gG} from Proposition 4.8, we have

W(γ)={gD+:0g.D+,g.D+boundedinexp(g1(h)).P-}

(Lemma 5.2(c)) and W=W(γ) by Theorem 5.7. So we can write

β(t)=et.DforsomeDW(γ).

As β(R)W(γ), the order convex hull W(β) of β(R) is contained in W(γ)=W. To verify the converse inclusion, let DW. Then 0D, and since D is bounded, there exists a tR with β(t)D. Likewise the boundedness of D implies the existence of some sR with Dβ(s). Hence, D[β(t),β(s)]W(β). This shows that WW(β), and hence equality in (5.11).

Remark 5.12

A similar result also holds in Minkowski space. If

xWR={yR1,d:y1>|y0|}

and

β(t)=ethx=(cosh(t)x0+sinh(t)x1,cosh(t)x1+sinh(t)x0,x2,,xd),

then any other element yWR satisfies y[β(t),β(s)] for suitable t<s, i.e., y-β(t)V+ and β(s)-yV+. In fact, β(t)0et(x0+x1) for t and β(t0)e-t(x0-x1)- for t-. Moreover, for s

(cosh(s)x0+sinh(s)x1-y0)2-(cosh(s)x1+sinh(s)x0-y1)2(esx0+x12-y0)2-(esx0+x12-y1)2es(x0+x1)(y1-y0)

and, for t-,

(cosh(t)x0+sinh(t)x1-y0)2-(cosh(t)x1+sinh(t)x0-y1)2(e-tx0-x12-y0)2-(e-tx1-x02-y1)2e-t(x0-x1)(-y0)-e-t(x1-x0)(-y1)=e-t(x1-x0)(y0+y1).

This shows that W(β)=WR for all integral curves of the modular flow in WR.

Remark 5.13

On the de Sitter space M=dSdR1.d, the involution τh can be implemented naturally by

τh,M(x)=(-x0,-x1,x2,,xd).

This involution does not fix the base point e1, it reverses the causal structure and it commutes with modular flow. Accordingly, we have the relation

τh,M(W+(h))=W+(-h).

As we shall see in the next section, such a relation can only be realized because -hAd(G)h, i.e., the direction of the boost can be reversed by an element of G. If -hAd(G) (h is not symmetric), then we shall see in Corollary 6.3 below that W+(-h)=, so that there is no involution on M mapping W+(h) to W+(-h).

However, as τh=τθ (as involutions on g), and there are natural implementations τM and θM on M=G/H, both fixing the base points, the involution τMθM implements the involution τh on M and fixes the base point, but it also fixes the wedge region

τMθM(W+(h))=W+(h)

because it preserves h and the causal structure. This is not desirable because we would prefer that τh maps W+(h) to some “opposite” wedge region (cf. [38]). Possible ways to resolve this problem and ideas how to implement locality conditions on non-compactly causal symmetric spaces are briefly discussed in [41, §4.3].

Existence of positivity domains for Euler elements

In this section, we show that, for the maximal cone C=Cqmax and a simple Lie algebra g, the real tube domain TC=h+C intersects the set E(g) of Euler elements in a connected subset (Theorem 6.1). This implies that, for an Euler element hg, the positivity domain WM+(h) is non-empty if and only if h and h are conjugate (Corollary 6.3).

Theorem 6.1

Suppose that (g,τ,C) is an irreducible simple ncc symmetric Lie algebra with C=Cqmax, TC:=h+C, G=Inn(g), H=Khexp(hp) and M=G/H. Then E(g)TC is connected and a subset of Oh. More precisely,

E(g)TC=OhTC=Ad(He)(h+Dg), 6.1

where Dg={ug1(h):u<1} is the open unit ball for which exp(Dg)P-=H.P-G/P-.

Proof

We recall from Proposition 4.8 the open subsets D±:=H.ePG/P which are the open orbits of the base point under H=Khexp(hp). Then

comp(D±)=Hexp(C)

follows from Proposition 4.8, applied to the causal Euler element h and its negative. These semigroups have the Lie wedges

L(comp(D±))=hC.

Let xE(g)TC for TC=h+C=L(comp(D-)). We then have st:=exp(tx)comp(D-) for t>0. We conclude that st(D-¯)D- and that there exists a complete metric on D- for which each st is a strict contraction (cf. [48, Thm. II.4]),5 so that the Banach Fixed Point Theorem implies the existence of a unique attracting fixed point m-D- for the vector field XxG/P+V(G/P+) defined by x. We now have

m-D-=H.eP+=He.eP+.

Hence there exists g1He with g1.m-=eP+, and thus

y:=Ad(g1)xp+=g1(h)g0(h). 6.2

Then yTCp+ is an Euler element, and a similar argument shows that the vector field XyG/P- has a unique repelling fixed point m+D+. So m+=exp(-z)P- for some zg1(h), and exp(z).m+=eP-. Hence the base point eP-G/P- is a repelling fixed point of the Euler element y:=eadzyg0(h), and eP+ is an attracting fixed point in G/P+. The attracting and repelling properties of the fixed points imply that

g1(h)g1(y)andg-1(h)g-1(y),

so that we also have

g0(h)=[g1(h),g-1(h)]g0(y).

As h and y are Euler elements, this entails that gλ(h)=gλ(y) for λ=-1,0,1. This shows that adh=ady and hence that y=h because z(g)={0}.

We conclude that

x=Ad(g1)-1y=Ad(g1)-1e-adzhwithg1He,zDg.

Conversely, we have seen in Proposition 5.4 that

eadDgh=h+DgTC. 6.3

We finally obtain (6.1).

Remark 6.2

Note that the preceding proof is based on the natural embedding

OhG/GhG/P-×G/P+

which maps the Euler element Ad(g)H to (m+,m-), where m+ is the unique repelling fixed point of the flow defined by h in G/P- and m-G/P+ is the unique attracting fixed point.

Corollary 8.3

(The set of positivity domains in M) If h1E(g) is an Euler element for which the positivity domain

WM+(h1)={mM=G/H:X1M(m)Cm}

is non-empty, then there exists a gG with h1=Ad(g)h and

WM+(h1)=g.WM+(h).

Proof

As X1M(g1H)Cg1H is equivalent to Ad(g1)-1h1h+C by (see Lemma 3.3), Theorem 6.1 implies that h1=Ad(g)hOh for some gG. The relation WM+(h1)=g.WM+(h) now follows directly from the definitions.

The preceding corollary shows that any wedge domain of the type WM+(h1)M, h1E(g), is a G-translate of the wedge domain WM+(h), where hCqp is a causal Euler element. So the action of G on the “wedge space” W(M) of M is transitive.

Corollary 8.4

If the causal Euler element h is not symmetric, then WM+(-h)=.

Remark 6.5

(Extensions to the non-simple case) If (g,τ) is a direct sum of irreducible ncc symmetric Lie algebra (gj,τj) and h=jhj accordingly, then

Cqmax(h)=jCqjmax(hj)

(cf. (2.3)). Projecting to the ideals gj, we obtain with Theorem 6.1 for C=Cqmax(h) and Cj=Cqjmax(hj) the relation

E(g)TCjE(gj)TCjjOhj=Oh. 6.4

Further,

OhTC=jOhjTCj

and Dg=jDgj imply (6.1) for this case.

Note that the situation corresponds to g=gs (see (2.2)). In the general situation, where we assume only that all ideals of g contained in h are compact, we have

g=gkgrgs,

where gkh is compact, gr is a direct sum of Riemannian symmetric Lie algebras and gs is a direct sum of irreducible ncc symmetric Lie algebras. All Euler elements are contained in gr+gs. If g is only reductive, we assume z(g)g-θ, so that z(g)gr. Then h=hr+hs and

Cqmax(h)=qrCqsmax(hs).

We conclude that

E(g)TCE(g)TCqmax(h)=((E(gr){0})×E(gs))TCqsmax(hs)(E(gr){0})×Ohs.

This shows that, for any Euler element kg with WM+(k) we must have ksOhs, but there is no restriction on the Riemannian component krE(gr).

Connectedness of the positivity domain

In this section, we show that if GInn(g) is the adjoint group, then the positivity domain WM+(h) is connected. This contrasts the situation for compactly causal symmetric spaces, where wedge regions are in general not connected. A typical example is anti-de Sitter spacetime (cf. [52, Lemma 11.2]).

Theorem 7.1

(Connectedness of positivity domains) Suppose that (g,τ,C) is an irreducible simple ncc symmetric Lie algebra with C=Cqmax and the causal Euler element hCqp. Let M=G/H for G=Inn(g) and H=Khexp(hp). Then the positivity domain WM+(h) is connected.

Proof

From Theorem 6.1 we derive that

G+(h):={gG:Ad(g)-1hTC}=Ghexp(Dg)He,

and this leads with Lemma 3.3 to

WM+(h)=G+(h).eH=Ghexp(Dg).eH.

Since Gh has at most two connected components, this set is either connected or has two connected components ([41, Thm. 7.8]). As Gh=Khexp(qp), we have Gh=KhGeh, and Ad(Kh) preserves the open unit ball in g1(h). We thus derive from Kh=HK:

WM+(h)=Ghexp(Dg).eH=GehKhexp(Dg).eH=Gehexp(Dg)Kh.eH=Gehexp(Dg).eH,

which is connected.

Corollary 3.7

W(γ)=WM+(h).

Proposition 7.3

(The stabilizer group of the observer domain) If g=gs, then Gh coincides with the stabilizer group

GW(γ):={gG:gW(γ)=W(γ)}

of the observer domain W(γ)M=G/H.

Proof

We work with the subset realization of M=G/H as MD+={g.D+:gG} from Proposition 4.8. Then

W(γ)={gD+:0g.D+,g.D+boundedinexp(g1(h)).P-}

(Lemma 5.2(c)). Since exp(Rh) acts on exp(g1(h)) by dilations, it follows that

gD+W(γ)gD+=tRetD+={eP-}. 7.1

Therefore, gW(γ)=W(γ) for the action of g on G/HP(G/P-) implies that g preserves the intersection {eP-} of all subsets contained in W(γ). This shows that g fixes eP-, so that gP-.

Next we recall that the involution τM on M defined by τM(gH)=τ(g)H leaves W(γ) invariant (Lemma 5.10), and this leads to

GW(γ)=τ(GW(γ))P-τ(P-)=P-P+=Gh.

The preceding proposition shows that the set W=W(M) of wedge domains in M=G/H coincides with

W=G.W(γ)G/GhOh. 7.2

In particular, it is a symmetric space. Recall that, by Corollary 7.2, the observer domain coincides with the positivity domain WM+(h).

KMS wedge regions

With the structural results obtained so far, we have good control over the positivity domains WM+(h) in ncc symmetric spaces M=G/H. So one may wonder if they also have an interpretation in terms of a KMS like condition. In [52], this has been shown for modular flows with fixed points, using such a fixed point as a base point. In this section we extend the characterization of the wedge domain W in terms of a geometric KMS condition to general ncc spaces.

To simplify references, we list our assumptions and the relevant notation below:

  • g is simple,

  • G=Inn(g)GC=Inn(gC)e (by (GP) and (Eff), [52, Lemma 2.12])

  • σ:GCGC denotes the complex conjugation with respect to G.

  • H=GcG, where Gc=(GCτ¯)e and KCGCθ is an open subgroup. Note that HGτ.

  • Ξ=G.ExpeK(iΩp)GC/KC.

  • HCGCτ is open with GHC=H (see §5), so that M=G/HGC/HC.

  • τhG(HC)=HC for the holomorphic involution of GC integrating the complex linear extension of τ.

  • σ(HC)=HC for the conjugation of GC with respect to G.

  • κh=e-πi2adh integrates to the automorphism κhG(g)=exp(-πi2h)gexp(πi2h) of GC.

Note that

τhG:=(κhG)2 8.1

is a holomorphic involutive automorphism of GC inducing τh on the Lie algebra g.

Let

Ξ:=G.ExpeK(iΩp)GC/KC

be the crown of G/K. The involution τh on G preserves K, hence induces an involution on G/K, and we extend it to an antiholomorphic involution τ¯h on GC/KC. The canonical map G×KiΩpΞ is a diffeomorphism ([52, Prop. 4.7]) and

τ¯h(g.Exp(ix))=τh(g).Exp(-iτh(x))

implies that

Ξτ¯h=Gτh.Exp(iΩp-τh)=exp(qp).Exp(iΩhp)Geh×KehiΩhp. 8.2

(see the proof of [52, Thm. 6.1] for details). This describes the fixed point as a “real crown domain” of the Riemannian symmetric space (G/K)τh=Exp(qp).

For an open subgroup HCGCτ (where τ denotes the holomorphic involution) with GHC=H, we obtain an embedding M=G/HGC/HC. Then the stabilizer of

mK:=ExpeH(πi2h)=exp(πi2h)HCGC/HC

coincides with K, so that G.mKG/K ([52, Thm. 5.4]). Accordingly,

KC:=(κhG)-1(HC)

is an open subgroup of GCθ that coincides with the stabilizer GCK. In this sense GC/HCGC/KC, but with different base points mH:=eHC and mK. Recall that τ=eπiadhθ=eπi2adhθe-πi2adh implies θ=(κhG)-1τκhG. The invariance of HC under τhG implies that

HC=(κhG)-1(KC),

so that KC and HC are exchanged by the order-4 automorphism κhG and invariant under τhG.

As τhG commutes with κhG, it also leaves KC invariant. Moreover, σκhGσ=(κhG)-1 entails

σ(KC)=κhG(σ(HC))=κhG(HC)=KC.

Therefore, the antiholomorphic extension τ¯hG also preserves KC and induces on GC/KCGC/HC an antiholomorphic involution τ¯h fixing the base point mK with stabilizer KC. Then

mH:=τ¯h(mH)=τ¯h(exp(-πi2h).mK)=exp(πi2h).mK=exp(πih).mH,

may be different from mH.

Remark 8.1

The condition mH=mH is equivalent to exp(πih)HC. Note that eπiadh=τhAut(gC) is an involution that commutes with τ, so that the choice of HC determines whether exp(πih) is contained in HC or not.

For g=sl2(R), G=Inn(g), and h=12100-1, we obtain on SL2(R)SL2(C) the involution

τabcd=dcba.

For gSL2(C), the condition τAd(g)τ=Ad(g) is equivalent to τ(g)g-1ker(Ad)={±1}. As

eπiadh=Ad(expπih)=Adi00-iandτi00-ii00-i-1=-1,

it follows that

τh=eπiadhGCτ\(GCτ)e.

In particular, KC and HC have two connected components in GCPSL2(C).

In GPSL2(R), a similar argument shows that θ=Ad0-110Gτ\Geτ. So Gτ also has two connected components, but only its identity component Geτ acts causally on q. Therefore, H=Geτ, but for HC we have two choices, GCτ, or its identity component.

Comparing with the arguments in [52, Lemma 6.3], where απi=τh on M, we have to be more careful in the present context. Here τ¯h restricts to a map

M=G.mHM:=G.mH=exp(πih).M,

and these two copies of G/H may not be identical. However, the antiholomorphic map

σM:=απiτ¯h

maps M to itself, fixes the base point mH and commutes with the G-action. Hence it fixes M pointwise and describes a “complex conjugation” with respect to M. In particular, the two maps

τ¯h:MMandαπi:MM

coincide on M.

We define the KMS wedge domain

WKMS:={mM:αit(m)Ξfor0<t<π}. 8.3

Theorem 8.2

WKMS=WM+(h)eH.

Proof

”: For zC and pM, we first observe that

τ¯h(αz(p))=αz¯(τ¯h(p))=αz¯απi(p)=απi+z¯(p).

For z=πi2, we thus obtain απi/2(p)MCh. We conclude that

απi2(WKMS)Ξτ¯h=exp(qp).Exp(iΩhp). 8.4

Hence,

WKMSκh(Ξτ¯h)=Geh.ExpeH(κh(iΩhp)),

where

κh(iΩhp)={xqk:adx<π2}=:Ωqk.

This suggest to define a “polar wedge domain” as

WMpol(h):=Geh.Exp(Ωqk)M.

We actually know from Theorem 3.6 that this is the connected component W=WM+(h)eHWM+(h) containing the base point. We thus obtain

WKMSW=WM+(h)eH. 8.5

”: To see that WM+(h)eHWKMS, we first recall from the first part of the proof that

WM+(h)eH=κh(Ξτ¯h)=Geh.ExpeH(κh(iΩhp))=Geh.α-πi/2(ExpeK(iΩhp)).

To see that this domain is contained in the Geh-invariant domain WKMSM, we thus have to show that, for xΩhp, we have

αit.ExpeK(ix)Ξfor|t|<π/2.

Let tqqk is a maximal abelian subspace (they are all conjugate under (HK)e). Then, ah:=iκh(tq)hp is also maximal abelian and Ωhp=eadhk.Ωah. So it suffices to show that, for xΩah and |t|<π/2, we have αit.ExpeK(ix)Ξ. By Proposition 2.8, tq is contained in a τ-invariant subalgebra ssl2(R)s, where Rh+s is generated by h and tq and h=h0+h1++hs, where hj, j=1,,s, is an Euler element in a simple ideal sjsl2(R) of s. Then ah=iκh(tq)a is spanned by s Euler elements x1,,xs and

Ωah={j=1stjxj:(j)|tj|<π/2}.

Let S:=exps and ΞS:=S.Exp(i(Ωps))Ξ. Then the discussion in Remark D.1 implies that, for |t|<π/2 and x=jtjxjΩah, we have αit(ExpeK(ix))ΞSΞ.

The preceding proof implies in particular the following interesting observation:

Corollary 5.8

For every mΞτ¯h, we have αit(m)Ξ for |t|<π/2, so that the orbit map αm extends to a holomorphic map S±π/2Ξ.

Corollary 6.3

απi2:WKMSΞτ¯h is a diffeomorphism that induces an equivalence of fiber bundles

WKMSGeh×KehΩqkGeh×KehiΩhpΞτ¯h.

Proof

Theorem 8.2 implies in particular that απi2:WKMSΞτ¯h is bijective. Since WKMS=WM+(h)eH is an open subset of M and Ξτ¯h an open subset of MCh, it actually is a diffeomorphism. The second assertion follows from the fact that it commutes with the action of the subgroup Geh.

Acknowledgements

We thank the late Joseph A. Wolf for discussions concerning the Convexity Theorem 4.5.

Irreducible ncc symmetric Lie algebras

The following table lists all irreducible non-compactly causal symmetric Lie algebras (g,τ) according to the following types:

  • Complex type: g=hC and τ is complex conjugation with respect to h. In this case gch2, so that rkR(gc)=2rkR(h).

  • Cayley type (CT): τ=τh1 for an Euler element h1h. Then rkR(gc)=rkR(g)=rkR(h).

  • Split type (ST): ττh1 for all h1hE(g) and rkRh=rkRgc:

  • Non-split type (NST): ττh1 for all h1hE(g) and rkRh=rkRgc2:

In Table 1, we write r=rkR(gc) and s=rkR(h). Further ap is maximal abelian of dimension r. For root systems Σ(g,a) of type An-1, there are n-1 Euler elements h1,,hn-1, but for the other root systems there are less; see [38, Thm. 3.10] for the concrete list. For 1j<n we write j:=min(j,n-j).

Table 1.

Irreducible ncc symmetric Lie algebras with corresponding causal Euler elements ha

graphic file with name 10455_2023_9937_Tab1_HTML.jpg

Geodesics in symmetric spaces

This appendix contains some elementary observations concerning geodesics in symmetric spaces.

Lemma B.1

Let M=G/H be a symmetric space with symmetric Lie algebra (g,τ), xg and yq. Then

exp(tx)H=exp(ty)HforalltR B.1

holds if and only if pq(x)=y and [x,y]=0.

In particular, γ(t):=exp(tx)H is a geodesic in M if and only if [x,τ(x)]=0.

Proof

The relation (B.1) is equivalent to

exp(-ty)exp(tx)HGGforalltR.

Applying τG, we obtain

exp(ty)exp(tτ(x))=exp(-ty)exp(tx),

which leads to exp(2ty)=exp(tx)exp(-tτ(x)). Evaluating the derivative of this curve in the right trivialization of T(G), we get

2y=x+etadx(-τ(x))=x-etadx(τ(x))foralltR.

For t=0 we get pq(x)=y, and taking derivatives in 0 shows that [x,τ(x)]=0.

If, conversely, this condition is satisfied, then x=xh+xq with xhh and xqq, where

0=[x,τ(x)]=2[xh,xq].

Therefore,

exp(tx)H=exp(txq)exp(txh)H=exp(txq)H=ExpeH(txq)

is a geodesic in M.

The following lemma provides important information on the subset Mx.

Lemma B.2

Let xg and write

Mx:={gHM:Ad(g)-1xq}.

Then Mx is a submanifold of M which is invariant under the action of Gx, and the orbits or Gex are the connected components of Mx.

Proof

Let m0=g0HMx and xc:=Ad(g0)-1x. For yq we have

Expm0(g0.y)=g0.ExpeH(y)=g0(expy)H=exp(Ad(g0)y).m0

and

Ad(g0exp(y))-1x=e-adyxc=cosh(ady)xc-sinh(ady)xc=cosh(ady)xcq-sinh(ady)ady[y,xc]h.

Let Uq be a 0-neighborhood for which ExpeH|U is a diffeomorphism onto an open subset of M and the spectral radius of ady is smaller than π for yU. Then sinh(ady)ady:hh is invertible. With the above formula, we thus conclude for yU that Expm0(g0.y)Mx is equivalent to [y,xc]=0, which is equivalent to Ad(g0)ygx. This shows that Mx is a submanifold of M.

As Expm0(g0.y)=exp(Ad(g0)y).m0exp(gx).m0, it further follows that the orbit of m0 under the connected group Gex contains a neighborhood of m0. This shows that the orbits of Gex in Mx are connected open subsets, hence coinciding with its connected components.

Remark B.3

For xq the centralizer gx is τ-invariant, so that gx=hxqx and the dimension of the Gex-orbit through eH is dimqx. We have

MeHxGex/(HGex),

and Lemma B.2 shows that the geodesic ExpeH(Rx) is central in the symmetric space MeHx in the sense that its tangent space Rx is central in the Lie algebra gx (cf. [36]).

Lemma B.4

For yq, the equality My=Gy.eH is equivalent to

Oyq=Ad(H)y. B.2

Proof

As yq, the base point eH is contained in My, and thus Gy.eHMy. So the equality My=Gy.eH means that MyGy.eH, i.e., Ad(g)-1yq implies gHGy.eH, resp., gGyH. This in turn is equivalent to Ad(g)-1yAd(H)y.

Lawson’s Theorem

Let (G,τG,H,C) be a causal symmetric Lie group, i.e., τG is an involutive automorphism of G, HGG an open subgroup and Cq a hyperbolic pointed generating Ad(H)-invariant closed convex cone. We write g=hq for the corresponding decomposition of g=L(G).

According to [30, Lemma 2.3], exp|C is injective if and only if ΓZ:=z(g)expG-1(e) satisfies

ΓZq=ΓZ(C-C)={0}.

If z(g)q, this condition is satisfied if and only if exp|z(g) is injective. This condition is always satisfied if g is semisimple because z(g)={0} in this case.

Suppose that ΓZ={0}. By [30, Lemma 2.4], exp|C is a homeomorphism onto a closed subset of G if and only if, for no non-zero xCz(g), the subgroup exp(Rx)¯ is compact. By [30, Thm. 3.1], this in turn is equivalent to the polar map

Φ:C×GGexp(C)GG

being a homeomorphism onto a closed subset of G. [30, Thm. 3.1] further shows that exp(C)GG is a subsemigroup of G. If G is 1-connected, then the subgroup GG is connected and Z(G) is simply connected, so that all requirements from above are satisfied ([30, Cor. 3.2]).

Theorem C.1

(Lawson’s Theorem) Let (G,τG,H,C) be a non-compactly causal reductive symmetric Lie group. Suppose that z(g)q and that ΓZ={0}. Then S:=exp(C)H is a closed subsemigroup of G with Lie wedge L(S)=h+C.

Proof

Our assumption implies that exp|z(g):z(g)Z(G)e is bijective, hence a diffeomorphism onto the closed subgroup Z(G)e. It follows in particular that exp(Rx)R is non-compact for each non-zero xz(g). Therefore the polar map Φ is a homeomorphism onto a closed subset and the remaining assertions follow from [30, Thm. 3.1].

Remark C.2

(a) If G is reductive, then G=(G,G)eZ(G)e and if xz(g) satisfies expz(G,G)e, then expzZ((G,G)e)-τG, which is a discrete group. We shall see below that this group may be infinite, even if ΓZ={0}.

(b) If M=GC/G is of complex type and G is hermitian, then Z(GC) is finite.

(c) If M is of non-complex type and irreducible, then gc is simple hermitian with z(kc)=Rih, where hE(g) is a causal Euler element. If Z(G) is infinite, then g is also hermitian, hence of tube type because it contains an Euler element [38]. Then all Euler elements in g are conjugate and this implies that (g,τ) is of Cayley type. So z(k)qk and thus Z(G)-τ is infinite if G is simply connected. This shows that it is possible that (G,G)eZ(G)e is infinite.

A concrete example is the group

G:=(SL~2(R)×R)/D,

where DZ(SL~2(R))×RZ×R is the graph of a non-zero homomorphism γ:Z(SL~2(R))R. Then Z(G)R and Z(G)(G,G)Z(SL~2(R))Z.

Remark C.3

Suppose that h0hE(g) is such that -τh0(C)=C, then Cz(g) is contained in C-C={0}. Therefore the condition on Cz(g) in Lawson’s Theorem (cf. Appendix C) is satisfied.

de Sitter space

In this appendix we collect some concrete observations concerning de Sitter space dSd, which is an important example of a non-compactly causal symmetric space. Some facts on 2-dimensional de Sitter space are used in particular in some of our proofs to verify the corresponding assertions for g=sl2(R).

In (d+1)-dimensional Minkowski space R1,d, we write the Lorentzian form as

β(x,y)=x0y0-xyforx=(x0,x),y=(y0,y).

We consider d-dimensional de Sitter space

M:=dSd:={x=(x0,x)R1,d:x02-x2=-1},

G=SO1,d(R)e and the Euler element hso1,d(R), defined by

h.(x0,,xd)=(x1,x0,0,,0).

It generates the Lorentz boost in the x0-x1-plane. The fixed point set of the modular flow in M=dSd is

Mα=Mspan{e2,,ed}={(0,0,x2,,xd):x22++xd2=1}Sd-2.

This submanifold is connected for d>2 and consists of two points for d=2. The corresponding wedge domain is the connected subset

WM+(h)=MWR={xdSd:x1>|x0|}.

By [52, Prop. D.3], the timelike geodesics of M of velocity 1 take the form

γ(t)=Expx(tv)=cosh(t)x+sinh(t)v,β(v,v)=1,β(x,v)=0,β(x,x)=1

whereas the trajectories of the modular flow are

αt(x)=ethx=(cosh(t)x0+sinh(t)x1,cosh(t)x1+sinh(t)x0,x2,,xd).

Comparing both expressions leads for h-modular geodesics to the conditions

x2==xd=0andv=h.x=(x1,x0,0,,0).

Therefore exactly two orbits of the modular flow are timelike geodesics. If we also ask for the geodesic to be positive with respect to the causal structure, then x1>0 determines the geodesic uniquely.

We infer from [52, Prop. D.3] that

Expe2(te1)=cos(t)e2+sin(t)e1

is a closed space-like geodesics. For 0<t<π, its values are contained in WM+(h), and this geodesic arc connects the two fixed points e2 to -e2 of the modular flow.

Remark D.1

In addition to h, we also consider the Euler elements defined by

hd(x0,,xd)=(xd,0,,0,x0).

The involution corresponding to h acts on R1+d by

τh(x0,x1,,xd)=(-x0,-x1,x2,,xd),

and its antilinear extension acts on C1+d by

τ¯h(z0,z1,,zd)=(-z0¯,-z1¯,z2¯,,zd¯).

Note that, in g, we have τh(h)=h and τh(hd)=-hd, so that hqp and hdhp.

In C1+d, we consider the domain

Ξ:={z=x+iyC1+d:y0>0,y02>y12++yd2}.

On Ξ the antiholomorphic involution τ¯h has the fixed point set

Ξτ¯h=Ξ(iR2Rd-1)={(ix0,ix1,x2,,xd):x0>|x1|,-x02+x12-x22--xd2=-1}={(ix0,ix1,x2,,xd):x0>|x1|,x02-x12>0+x22++xd2=1}.

It follows in particular that

x02-x12(0,1].

The analytic extension of the modular flow (αt)tR acts on Ξ by

αit(z0,,zd)=(cost·z0+isint·z1,isint·z0+cost·z1,z2,,zd).

Starting with a τ¯h-fixed element z=(ix0,ix1,x2,,xd) in Ξ, this leads to

αit(ix0,ix1,x2,,xd)=(cost·ix0-sint·x1,-sint·x0+cost·ix1,x2,,xd)

with imaginary part

(x0cost,x1cost,0,,0),

so that we obtain for |t|<π/2 that

|x0cost|=x0cost>|x1|cost,

which implies that

αit(z)ΞforzΞτ¯hand|t|<π/2. D.1

Example D.2

For the special case d=2, we have sl2(R)R1,2 and the Euler element

h0=12100-1sl2(R)R1,2

corresponds to the base point e2 (see [52]), so that

sl2(R)OhdS2R1,2.

Accordingly,

C=cone(e0,f0),Cc=cone(e0,-f0),andx0:=12(e0-f0)=1201-10Cc.

For gt:=exp(tx0) we then have

Ad(gπ/2)h0=-h1,Ad(gπ/2)h1=h0andAd(gπ)h0=-h0.

We also note that, for 0<t<π, the Lie algebra element Ad(gt)h0 corresponds to Expe2(te1)WM+(h0). Note that

gπKτhG=KG.

Author Contributions

VM, KHN, GO wrote the manuscript on equal parts.

Funding

Open Access funding enabled and organized by Projekt DEAL.

Declaration

Conflicts of interest

The authors declare no competing interests.

Footnotes

1

A closed real subspace H of a complex Hilbert space H is called standard if H+iH¯=H and HiH={0}

2

Note that the cones V+(m) are open, whereas C is closed.

3

This theorem is stated for complex hermitian Jordan triple systems, but V=g1(h) is a real form of the complex JTS VC=gC,1(h) on which we have an antilinear isomorphism σ with V=VCσ. Therefore, the uniqueness in the spectral decomposition shows that, for xV, the corresponding spectral tripotents are contained in V.

4

As D+ is invariant under the group (HK)e which acts linearly, and this group acts transitively on the set of all maximal flat subtriples of V ([59, Lemma VI.3.1]), it suffices to shows that an element with a spectral resolution x=j=1rxjcj is contained in Dg if and only if |xj|<1 for every j. This follows easily from (4.5).

5

This reference deals with bounded symmetric domains in complex spaces, but D can be embedded into such a domain DC by embedding ggCgCc. If Cgcgc is an invariant cone with C=giCgc, then (g,τ,C)(gC,τgC,iCgc) is a causal embedding and D+=H.eP-Gc.ePC-=D+C is a real form of a complex bounded symmetric domain D+C; see [56, Lem. 1.4] or [26, Lem 5.1.11] for more details.

The research of V. Morinelli VM was partially supported by a Humboldt Research Fellowship for Experienced Researchers; the University of Rome through the MIUR Excellence Department Project 2023–2027, the “Tor Vergata” CUP E83C23000330006 and “Tor Vergata” “Beyond Borders” CUP E84I19002200005, Fondi di Ricerca Scientifica d’Ateneo 2021, OAQM, CUP E83C22001800005, the European Research Council Advanced Grant 669240 QUEST and INdAM-GNAMPA. The research of K.-H. Neeb was partially supported by DFG-grants NE 413/10-1 and NE 413/10-2. The research of G. Ólafsson was partially supported by Simons grant 586106.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

References

  • 1.Araki H. Relative entropy of states of von Neumann algebras. Publ. RIMS, Kyoto Univ. 1976;11:809–833. [Google Scholar]
  • 2.Bertram W, Neeb K-H. Projective completions of Jordan pairs, Part I. The generalized projective geometry of a Lie algebra. J. Algebra. 2004;277(2):474–519. [Google Scholar]
  • 3.Bisognano J, Wichmann EH. On the duality condition for quantum fields. J. Math. Phys. 1976;17:303–321. [Google Scholar]
  • 4.Bratteli, O., Robinson, D. W.: Operator Algebras and Quantum Statistical Mechanics I, 2nd ed., Texts and Monographs in Physics, Springer (1987)
  • 5.Bratteli, O., Robinson, D. W.: Operator Algebras and Quantum Statistical Mechanics II, 2nd ed., Texts and Monographs in Physics, Springer (1996)
  • 6.Bros J, Moschella U. Two-point functions and quantum fields in de Sitter universe. Rev. Math. Phys. 1996;8(3):327–391. [Google Scholar]
  • 7.Brunetti R, Guido D, Longo R. Modular localization and Wigner particles. Rev. Math. Phys. 2002;14:759–785. [Google Scholar]
  • 8.Ceyhan F, Faulkner T. Recovering the QNEC from the ANEC. Commun. Math. Phys. 2020;377(2):999–1045. [Google Scholar]
  • 9.Ciolli, F., Longo, R., Ranallo, A., Ruzzi, G.: Relative entropy and curved spacetimes, J. Geom. and Phys. 172 (2022), Paper No. 104416, 16 pp
  • 10.Ciolli F, Longo R, Ruzzi G. The information in a wave. Commun. Math. Phys. 2020;379(3):979–1000. [Google Scholar]
  • 11.Correa da Silva R, Lechner G. Modular structure and inclusions of twisted Araki-Woods algebras. Commun. Math. Phys. 2023;402(3):2339–2386. [Google Scholar]
  • 12.Dappiaggi C, Lechner G, Morfa-Morales E. Deformations of quantum field theories on spacetimes with Killing vector fields. Commun. Math. Phys. 2011;305:99–130. [Google Scholar]
  • 13.Dybalski W, Morinelli V. Bisognano-Wichmann property for asymptotically complete massless QFT. Commun. Math. Phys. 2020;380:1267–1294. [Google Scholar]
  • 14.Fewster, C.: Lectures on quantum energy inequalities. In: Lobo, Francisco S. N. (ed.), Wormholes, Warp Drives and Energy Conditions. Springer, Fundamental Theories of Physics, Cham, 189, 215–254, (2017). arXiv:1208.5399
  • 15.Fewster CJ, Hollands S. Quantum energy inequalities in two-dimensional conformal field theory. Rev. Math. Phys. 2005;17(5):577–612. [Google Scholar]
  • 16.Faulkner, T., Leigh, R. G., Parrikar, O., Wang, H.: Modular Hamiltonians for deformed half-spaces and the averaged null energy condition, J. High Energy Phys. 2016:9 (2016), Paper 38, 35 p. arXiv:1605.08072
  • 17.Frahm, J., Neeb, K.-H., Ólafsson, G.: Nets of standard subspaces on non-compactly causal symmetric spaces, to appear in “Toshiyuki Kobayashi Festschrift,” Progress in Mathematics, Springer-Nature; arxiv:2303.10065
  • 18.Guido D, Longo R. A converse Hawking-Unruh effect and dS2-CFT correspondence. Ann. Henri Poincaré. 2003;4:1169–1218. [Google Scholar]
  • 19.Guido D, Longo R. An algebraic spin and statistics theorem. Commun. Math. Phys. 1995;172(3):517–533. [Google Scholar]
  • 20.Haag, R.: Local Quantum Physics. Fields, Particles, Algebras. Second edition, Texts and Monographs in Physics, Springer, Berlin (1996)
  • 21.Haag R, Hugenholtz NM, Winnink M. On the equilibrium states in quantum statistical mechanics. Commun. Math. Phys. 1967;5:215. [Google Scholar]
  • 22.Helgason S. Differential Geometry, Lie Groups, and Symmetric Spaces. London: Academic Press; 1978. [Google Scholar]
  • 23.Hilgert, J., Neeb, K.-H.: Lie Semigroups and Their Applications. Lecture Notes in Math, vol. 1552. Springer Verlag, Berlin, Heidelberg, New York (1993)
  • 24.Hilgert J, Neeb K-H. Compression semigroups of open orbits on real flag manifolds. Monatshefte für Math. 1995;119:187–214. [Google Scholar]
  • 25.Hilgert J, Neeb K-H. Structure and Geometry of Lie Groups. Berlin: Springer; 2012. [Google Scholar]
  • 26.Hilgert J, Ólafsson G. Causal Symmetric Spaces, Geometry and Harmonic Analysis, Perspectives in Mathematics. Cambridge: Academic Press; 1997. [Google Scholar]
  • 27.Kontou EA, Sanders K. Energy conditions in general relativity and quantum field theory. Class. Quant. Grav. 2020;37:193001. [Google Scholar]
  • 28.Krötz B, Neeb K-H. On hyperbolic cones and mixed symmetric spaces. J. Lie Theory. 1996;6(1):69–146. [Google Scholar]
  • 29.Koeller J, Leichenauer S, Levine A, Shahbazi-Moghaddam A. Local modular Hamiltonians from the quantum null energy condition. Phys. Rev. D. 2018;97(6):065011. [Google Scholar]
  • 30.Lawson JD. Polar and Ol’shanskiĭ decompositions. J. Reine Ang. Math. 1994;448:191–219. [Google Scholar]
  • 31.Longo R. An analogue of the Kac-Wakimoto Formula and black hole conditional entropy. Commun. Math. Phys. 1997;186:451–479. [Google Scholar]
  • 32.Longo R. Entropy of coherent excitations. Lett. Math. Phys. 2019;109(12):2587–2600. [Google Scholar]
  • 33.Longo R. Entropy distribution of localised states. Commun. Math. Phys. 2020;373(2):473–505. [Google Scholar]
  • 34.Longo R, Morsella G. The Massless Modular Hamiltonian. Commun. Math. Phys. 2023;400:1181–1201. [Google Scholar]
  • 35.Longo R, Feng X. Relative entropy in CFT. Adv. Math. 2018;337:139–170. [Google Scholar]
  • 36.Loos O. Symmetric Spaces I: General Theory. New York: W.A. Benjamin Inc; 1969. [Google Scholar]
  • 37.Loos O. Charakterisierung symmetrischer R-Räume durch ihre Einheitsgitter. Math. Zeitschrift. 1985;189:211–226. [Google Scholar]
  • 38.Morinelli, V., Neeb, K.-H.: Covariant homogeneous nets of standard subspaces. Commun. Math. Phys. 386, 305–358 (2021). arXiv:2010.07128
  • 39.Morinelli V, Neeb K-H. A family of non-modular covariant AQFTs. Anal. Math. Phys. 2022;12:124. [Google Scholar]
  • 40.Morinelli, V., Neeb, K.-H.: From local nets to Euler elements, in preparation
  • 41.Morinelli V, Neeb K-H, Ólafsson G. From Euler elements and 3-gradings to non-compactly causal symmetric spaces. J. Lie Theory. 2023;33(1):377–432. [Google Scholar]
  • 42.Much, A., Passegger, A. G., Verch, R.: An approximate local modular quantum energy inequality in general quantum field theory, arXiv:2210.01145
  • 43.Morinelli V, Tanimoto Y, Wegener B. Modular operator for null plane algebras in free fields. Commun. Math. Phys. 2022;395:331–363. doi: 10.1007/s00220-022-04432-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Mund J. The Bisognano-Wichmann theorem for massive theories. Ann. Henri Poincaré. 2001;2:907–926. [Google Scholar]
  • 45.Narnhofer H, Peter I, Thirring W. How hot is the de Sitter space?, “Memorial Issue for H. Umezawa,” Internat. J. Mod. Phys. B. 1996;13–14(10):1507–1520. [Google Scholar]
  • 46.Neeb K-H. Conal orders on homogeneous spaces. Inventiones Math. 1991;134:467–496. [Google Scholar]
  • 47.Neeb, K.-H.: Holomorphy and Convexity in Lie Theory. Expositions in Math. 28, de Gruyter Verlag, Berlin, 2000
  • 48.Neeb K-H. Compressions of infinite-dimensional bounded symmetric domains. Semigroup Forum. 2001;63(1):71–105. [Google Scholar]
  • 49.Neeb, K.-H., Ólafsson, G.: Antiunitary representations and modular theory. In: Grabowska, K., Grabowski, J., Fialowski, A., Neeb, K.-H. (eds) 50th Sophus Lie Seminar, Banach Center Publications 113, 291–362 (2017) arXiv:1704.01336
  • 50.Neeb K-H, Ólafsson G. Nets of standard subspaces on Lie groups. Adv. Math. 2021;384:107715. [Google Scholar]
  • 51.Neeb K-H, Ólafsson G. Wedge domains in compactly causal symmetric spaces. Int. Math. Res. Not. 2023;2023(12):10209–10312. [Google Scholar]
  • 52.Neeb, K.-H., Ólafsson, G.: Wedge domains in non-compactly causal symmetric spaces. Geometriae Dedicata 217(2), 30 (2023) arXiv:2205.07685 [DOI] [PMC free article] [PubMed]
  • 53.Neeb K-H, Ólafsson G, Ørsted B. Standard subspaces of Hilbert spaces of holomorphic functions on tube domains. Commun. Math. Phys. 2021;386:1437–1487. [Google Scholar]
  • 54.Oeh D. Nets of standard subspaces induced by unitary representations of admissible Lie groups. J. Lie Theory. 2022;32:29–74. [Google Scholar]
  • 55.Oeh D. Classification of 3-graded causal subalgebras of real simple Lie algebras. Transf. Groups. 2022;27(4):1393–1430. [Google Scholar]
  • 56.Ólafsson G. Symmetric spaces of hermitian type. Diff. Geom. Its Appl. 1991;1:195–233. [Google Scholar]
  • 57.Olshanski GI. Invariant cones in Lie algebras, Lie semigroups and the holomorphic discrete series Function. Anal. Appl. 1981;15:275–285. [Google Scholar]
  • 58.Olshanski GI. Convex cones in symmetric Lie algebras, Lie semigroups and invariant causal (order) structures on pseudo-Riemannian symmetric spaces. Soviet Math. Dokl. 1982;26:97–101. [Google Scholar]
  • 59.Roos, G., Jordan triple systems. In: Faraut, J. et al (eds.) Analysis and Geometry on Complex Homogeneous Domains, Progress in Math. 185, Birkhäuser, Boston (2000)
  • 60.Satake, I.: Algebraic Structures of Symmetric Domains. Publ. Math. Soc. Japan 14, Princeton Univ. Press (1980)
  • 61.Verch R. The averaged null energy condition for general quantum field theories in two dimensions. J. Math. Phys. 2000;41(1):206–217. [Google Scholar]
  • 62.Wald RM. General Relativity. Chicago, IL: University of Chicago Press; 1984. [Google Scholar]

Articles from Annals of Global Analysis and Geometry are provided here courtesy of Springer

RESOURCES