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 , 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, 49–52, 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 , where G is a Lie group, is an involutive automorphism of G and is an open subgroup (cf. [36]). A causal symmetric space carries a G-invariant field of pointed generating closed convex cones in their tangent spaces. Typical examples are de Sitter space and anti-de Sitter space as well as products and 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 acting on a fixed Hilbert space indexed by open subsets of the chosen spacetime M [20]. The hermitian elements of the algebra represent observables that can be measured in the “laboratory” . These nets are supposed to satisfy fundamental quantum and relativistic assumptions:
-
(I)
Isotony: implies
-
(L)
Locality: implies , where is the “causal complement” of , i.e., the maximal open subset that cannot be connected to by causal curves.
-
(RS)
Reeh–Schlieder property: There exists a unit vector that is cyclic for if .
-
(Cov)
Covariance: There is a Lie group G acting on M and a unitary representation such that for .
-
(BW)
Bisognano–Wichmann property: is separating for some “wedge region” and there exists an element with for , where is the modular operator corresponding to in the sense of the Tomita–Takesaki Theorem ([4, Thm. 2.5.14]).
-
(Vac)
Invariance of the vacuum: for every .
The (BW) property gives a geometrical meaning to the dynamics provided by the modular group of the von Neumann algebra associated with wedge regions with respect to the vacuum state specified by . On Minkowski/de Sitter spacetime, it provides an identification of the one-parameter group of boosts in the Poincaré/Lorentz group with the Tomita–Takesaki modular operator:
Here is a one-parameter group of boosts associated with , where is the standard right wedge and
describes the boosts associated with .
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 ), 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 (or de Sitter spacetime ), 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 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” . 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 [38–40]. 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 of the modular groups in (BW) are Euler elements, i.e., defines a 3-grading
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” 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 , then the couples , where is an Euler element and an involutive automorphism of G, inducing on the involution , 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 , where is a finite-dimensional real Lie algebra and an involutive automorphism of . We write and for the -eigenspaces.
A causal symmetric Lie algebra is a triple , where is a symmetric Lie algebra and is a pointed generating closed convex cone invariant under the group acting in . We call compactly causal (cc) if C is elliptic in the sense that, for (the interior of C in ), the operator is semisimple with purely imaginary spectrum. We call non-compactly causal (ncc) if C is hyperbolic in the sense that, for , the operator is diagonalizable.
As explained in detail in [41], Euler elements in reductive Lie algebras lead naturally to ncc symmetric Lie algebras: For an Euler element , choose a Cartan involution of with such that . Then is an involutive automorphism of commuting with , so that defines a symmetric Lie algebra and there exists a pointed generating -invariant hyperbolic cone C with . Under the assumption that contains no non-zero ideal of , there is a unique minimal cone with this property. It is generated by the orbit .
Let be an ncc symmetric Lie algebra and a corresponding symmetric Lie group, i.e., G is a connected Lie group, an involutive automorphism of G integrating , and an open subgroup. If, in addition, , then we call the quadruple a causal symmetric Lie group. On , we then obtain the structure of a causal symmetric space, specified by the G-invariant field of open convex cones2
1.1 |
We further assume that
is a closed subsemigroup for which the polar map is a homeomorphism. Then
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 and is injective. The second condition is always satisfied if G is simply connected.
For an Euler element , we consider the associated modular flow on , defined by
1.3 |
We study orbits of this flow which are geodesics with respect to the symmetric space structure and causal in the sense that for . We call them h-modular geodesics. All these are contained in the positivity domain
1.4 |
of the vector field 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 intersecting .
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 intersects the interior of the cone , and then the centralizer of h acts transitively on the union of the corresponding curves (Proposition 3.2(c)).
We associate to any modular geodesic a connected open subset ; the corresponding observer domain. For de Sitter space , we thus obtain the familiar wedge domain obtained by intersecting with a Rindler wedge in Minkowski space (Example 5.3). In Theorem 5.7, we show that it coincides with W, provided that and .
In Sect. 6 we further show that, for and simple, that the real tube domain intersects the set of Euler elements in a connected subset (Theorem 6.1). As a consequence, we derive that if and only if (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 is connected for and simple, and this implies that
From this in turn we derive that the stabilizer group coincides with (Proposition 7.3), so that the wedge space of wedge regions in M can be identified, as a homogeneous G-space, with the adjoint orbit . In particular also is a symmetric space. - Finally, we show in Theorem 8.2 that W coincides with the KMS wedge domain
where 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, 32–35, 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 preserving W, which is timelike and time-oriented on W. On Minkowski spacetime the flow , 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 , then denotes the connected component of M containing m. In particular, we write for the identity element in the Lie group G and for its identity component.
Involutive automorphisms of G are typically denoted , and is the corresponding automorphism of the Lie algebra . We write and .
For , we write for the stabilizer of x in the adjoint representation and for its identity component.
For and , we write for the corresponding eigenspace in the adjoint representation.
If is a Lie algebra, we write for the set of Euler elements , i.e., is non-zero and diagonalizable with . The corresponding involution is denoted .
For a Lie subalgebra , we write for the subgroup generated by .
For a convex cone C in a vector space V, we write for the relative interior of C in its span.
- We use the notation
for the spectral radius of a linear operator A.1.5
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 is an element h defining a 3-grading of by with , . We write for the set of Euler elements in . In this section, we recall some results on from [41] on Euler elements that are crucially used in the following.
Definition 2.1
Let be a reductive Lie algebra.
- A Cartan involution of is an involutive automorphism for which and is maximal compactly embedded in the commutator algebra . We then write, using the notation from the introduction,
- If is another involution on commuting with , and , then we have
- The Cartan dual of the symmetric Lie algebra is the symmetric Lie algebra with
Note that where is the conjugate-linear extension of to ; in particular is a real form of .
Definition 2.2
Let be a symmetric Lie algebra and . We say that h is causal if there exists an -invariant closed pointed generating convex cone C in with . We write for the set of causal Euler elements in . Recall that the triple is ncc if C is hyperbolic.
Lemma 2.3
Let be a simple ncc symmetric Lie algebra and be a causal Euler element. Then, the following assertions hold:
- There exist closed convex pointed generating -invariant cones
such that and either - If is a connected symmetric Lie group with symmetric Lie algebra , then two mutually exclusive cases occur:
- and G/H is causal.
- 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 and a Cartan involution with , and , then [41, Thm. 4.2] implies that there exists an -invariant pointed closed convex cone with , so that is ncc. Further, all ideals of contained in are compact. We have a decomposition
2.1 |
where is the sum of all simple ideals not commuting with h (the strictly ncc part), is the sum of the center and all non-compact simple ideals commuting with h on which (the non-compact Riemannian part), and is the sum of all simple compact ideals (they commute with h). All these ideals are invariant under and , so that we obtain decompositions
2.2 |
where is a compact ideal of , is a Cartan decomposition and . In particular . Let be the projection onto with kernel . Then, [41, Prop. B.4] implies that every -invariant closed convex cone C satisfies
By Lemma 2.3(a), we obtain a pointed -invariant cone , adapted to the decomposition into irreducible summands, whose dual cone with respect to the Cartan–Killing form satisfies . Put
2.3 |
Both cones are adapted to the decomposition of into irreducible summands. Further, each pointed generating -invariant cone C containing h satisfies
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
This leads to , and thus to (cf. [41, §3.5] for more details).
Lemma 2.4
If , then the centralizer is compactly embedded in , i.e., consists of elliptic elements.
Proof
First we observe that the cone is adapted to the decomposition and so is the centralizer of in . Hence the assertion follows from the fact that is compactly embedded and is compactly embedded because the cone is pointed ([47, Prop. V.5.11]).
Theorem 2.5
(Uniqueness of the causal involution) ([41, Thm. 4.5]) Let be a semisimple ncc symmetric Lie algebra for which all ideals of contained in are compact, the sum of all non-Riemannian ideals, , , and a Cartan involution commuting with . Then the following assertions hold:
contains a unique Euler element h, and this Euler element satisfies .
acts transitively on .
For every Euler element , the involution is Cartan.
Proposition 2.6
Let be a connected semisimple ncc symmetric Lie group for which contains no non-compact ideal of () and let (cf. Theorem 2.5) be a causal Euler element. Then the following assertions hold:
, i.e., every connected component of H meets .
is a maximal compact subgroup of .
and induces a Cartan involution on .
induces a Cartan involution on for which is connected.
Proof
The statements on the adjoint group follow from [41, Cor. 4.6] because preserves C. Further, and imply with (a) (for ) that .
Definition 2.7
If is a simple hermitian Lie algebra, a Cartan involution of and maximal abelian, then the restricted root system is either of type or . In the first case, we say that is of tube type.
Recall that if is simple ncc, then either is simple hermitian or , where is simple hermitian ([41, Rem. 4.24]).
Proposition 2.8
([41, Lemma 5.1, Prop. 5.2]) Let be a simple ncc symmetric Lie algebra. Pick a causal Euler element and maximal abelian and set . Then, the following assertions hold:
The Lie algebra generated by h and is reductive.
The commutator algebra is isomorphic to
for some hyperbolic element satisfying which is zero if and only if is of tube type.
The Lie algebra is -invariant and .
- For , we have where denotes the spectral radius. With the basis
in we have for2.5
Note that (c) implies that is semisimple, i.e., , if and only if is of tube type.
Proposition 2.9
([41, Prop. 7.10]) Let be a semisimple modular non-compactly causal semisimple symmetric Lie algebra, where , a causal Euler element,
Then satisfies
In particular implies .
The positivity domain and modular geodesics
Let be a connected semisimple causal symmetric Lie group with ncc symmetric Lie algebra . We fix a causal Euler element (Theorem 2.5) and write for the associated symmetric space.
One of our goals in this paper is to describe the structure of the positivity domain
of the vector field 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 as
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 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 as above.
We call a geodesic causal if for every (see (1.1)).
- Let be an Euler element. The flow on M defined by
is called the modular flow (associated to h). Its infinitesimal generator is denoted .3.2 A geodesic is called h-modular if holds for all , i.e., is an integral curve of .
Proposition 3.2
Suppose that is a direct sum of irreducible ncc symmetric Lie algebras (). The following assertions hold for any Euler element and the corresponding modular flow on :
- The orbit under the modular flow is a causal geodesic if and only if m is contained in
3.3 - All connected components of are Riemannian symmetric space of non-compact type: For every , the exponential map
is a diffeomorphism. h-modular causal geodesics exist if and only if intersects . In this case acts transitively on .
Proof
(a) Assume first that . Then (3.2) implies that the orbit of under the modular flow is a geodesic. The causality is by definition equivalent to .
Suppose, conversely, that is a causal geodesic. Lemma B.1 implies that , where and . By Lemma 2.4, is elliptic and is hyperbolic because it is contained in . Therefore, is the unique Jordan decomposition of into elliptic and hyperbolic summand. As is an Euler element, the elliptic summand vanishes, and thus, , i.e., (recall that ). This shows that , so that .
(b) Choosing m as a base point, we may assume that , so that (a) implies that is a causal Euler element. Pick a Cartan involution commuting which satisfies (cf. [28]), i.e., . Then follows from Theorem 2.5(a). As by Lemma B.2, the assertion now follows from .
(c) The first assertion follows immediately from (a). For the second assertion, suppose that . As is -invariant, . Let , so that by Theorem 2.5(b) (recall that ). If , i.e.,
then there exists an element with , so that , and therefore .
We record the following consequence of (3.2):
Lemma 3.3
For any causal Euler element , we have
Due to the hyperbolicity of Euler elements, modular causal geodesics do not exist for compactly causal symmetric spaces:
Proposition 3.4
If is a compactly causal symmetric space, then non-trivial causal modular geodesics do not exist.
Proof
If there exists a modular causal geodesic and is the infinitesimal data of M, then there exists a such that the Euler element h satisfies with and (Lemma B.1). As C is elliptic, is elliptic. Further the pointedness of C implies that is elliptic. This implies that the Euler element 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 that is equivariant with respect to the connected group , the base is the Riemannian symmetric space of this group, and the fiber is a bounded convex subset of .
Definition 3.5
Let be a causal Euler element, so that . Then implies that
is the non-compact Riemannian symmetric space associated with the symmetric Lie algebra .
Theorem 3.6
(Positivity Domain Theorem) Suppose that is a connected semisimple non-compactly causal Lie group for which contains no -invariant Riemannian ideals () and that h is a causal Euler element. Suppose that is the maximal -invariant cone with . Then, the following assertions hold:
- The connected component of eH in the positivity domain is given by
3.4 The polar map is a diffeomorphism
W is contractible, hence in particular simply connected.
.
Proof
(a) Recall from [41, Thm. 6.7] that the connected component of h in the open subset of is
3.5 |
If , then , so that (3.5) implies that satisfies
3.6 |
By Lemma 3.3
3.7 |
by -invariance of .
Conversely, for , the element is contained in the connected component of h, so that (3.5) implies that it is contained in . Therefore
This is equivalent to which implies
and thus
3.8 |
If satisfies then follows from (3.7) and the fact that g permutes the connected components of . Therefore, (3.8), combined with (3.7), leads with to
and this entails
3.9 |
Next we observe that the exponential map is regular in every because ([52, Lemma C.3(b)]). Thus [52, loc.cit.] further implies that the map
is regular in (g, x) because does not intersect for . This implies that the differential of is surjective in each point of ; hence, the image of every connected component is open. Now the connectedness of W implies that and this completes the proof.
(b)–(d): The surjectivity of follows from Theorem 3.6. As is a Cartan decomposition of , the polar map is a diffeomorphism. In particular,
implies and thus (b).
The space is a fiber bundle over whose fiber is the convex set . Therefore, it is homotopy equivalent to the base , which is also contractible because the exponential map is a diffeomorphism.
It therefore suffices to show that 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 satisfies
3.10 |
Step 1: is injective. If , then applying the quadratic representation implies in G. As and are both -regular, [25, Lemma 9.2.31] implies that
We conclude that , and since the spectral radius of is less than , it follows that , so that .
Step 2: with and implies . Applying the involution , we see that is a fixed point, so that
entails that fixes . We now write in terms of the polar decomposition of and obtain
Applying the quadratic representation, we get
3.11 |
which can be rewritten as
Since has real spectrum, so has . Therefore the same arguments as in Step 1 above imply that
and . The vanishing -component of this element is , and since , it follows that . Now (3.11) leads to , and further to , because the exponential function on is injective. This proves that .
Step 3: From (3.10), we derive
so that Step 2 shows that . We thus obtain
and since , we infer from Step 1 that . This completes the proof.
The following corollary identifies the connected component of containing eH as a submanifold (cf. Lemma B.2) of the wedge domain W.
Corollary 6.4
Assume that exists and leaves H invariant, so that exists and leaves the base point invariant. Then and the fixed point set of in W is the Riemannian symmetric space
Proof
For and :
So is a fixed point if and only if which is equivalent to . Now implies . As , [52, Lemma C.3] further shows that . Therefore, is a fixed point if and only if .
From and the polar decomposition (Theorem3.6(b)), we derive that the fixed point set is
The preceding corollary shows that the wedge domain contains the symmetric subspace as the fixed point set of an involution. Hence, the description of W from Theorem 3.6 as
suggest to consider W as a real “crown domain” of the Riemannian symmetric space .
Remark 3.8
Theorem 3.6 has a trivial generalization to semisimple non-compactly causal Lie algebras of the form because then
For with the relation is therefore equivalent to . If is the corresponding product decomposition, we obtain
However, if , then is not pointed, and there are many pointed invariant cones C, which are not maximal, for which the domain may have a more complicated structure.
Example 3.9
We consider the reductive Lie algebra
Any Euler element in is conjugate to some
The Cartan involution on then satisfies and acts by
With the Euler element
we then have
The group acts by on the 3-dimensional space of symmetric matrices and the stabilizer of is the subgroup with Lie algebra . Therefore can be identified with the subspace of indefinite symmetric matrices. Note that acts by multiplication with and that is a diffeomorphism, where
is a realization of 2-dimensional de Sitter space. Note that the determinant defines a quadratic form of signature (1, 2) on which is invariant under the action of the subgroup
which acts as .
For the Euler element , we have
According to [53, Ex. 3.1(c)], all -invariant cones in are Lorentzian of the form
Actually contains and is not pointed.
(a) We write
to see that is equivalent to
We also note that the “semisimple part” of is
coincides with the projection of to , so that .
Write for the positivity domain of the Euler element h with respect to the causal structure specified by the cone . Then Theorem 3.6 implies that
For we have
(see (3.6)) and , so that we have
implies the equality
We also note that
because if and only if .
To determine the domain in general, we write
By -invariance, we have to determine when , , is contained in . For , we have
and
We then have
We conclude that, for , the inclusion is equivalent to
We thus obtain the condition
For and , this is specifies a proper subinterval of .
(b) To determine which cone corresponds to the canonical order on the space , induced from the natural order of (which is also Lorentzian), we evaluate the tangent map to
We thus obtain for that
By the Hurwitz criterion, this matrix is positive semidefinite if and only if
and
Is , then these two inequalities are equivalent to . As these two conditions imply that , we see that the canonical order on M corresponds to the cone , i.e., to .
(c) For the modular vector field , we have
The positivity domain of depends on , and with this formula one can also determine the positivity domain quite directly for , where corresponds to the canonical order.
Example 3.10
(cf. [52, Exs. 2.11, 2.25]) Let and . We consider the Riemannian symmetric space
and the corresponding irreducible subspace
(here the index s refers to “semisimple”). On , we consider the Cartan involution given by and write with . Then
3.12 |
are Euler elements and leads to a non-compactly causal symmetric Lie algebra , where
To identify G/H in the boundary of the crown domain in , where acts on by ([52, Thm. 5.4]), we observe that
so that we obtain for the matrix
The G-orbit of this matrix is the open subset
of symmetric matrices of signature (p, q). We have
These matrices are never positive definite. So we have to take instead to find non-trivial positivity domains.
For the case and , this has been carried out in Example 3.9. We also write
Then
so that
which is equivalent to .
The connected components of
The main result in this section is Proposition 3.11 on the subgroup of . We then discuss several examples to clarify the situation.
Proposition 3.11
(Connected components of ) If and is irreducible ncc with causal Euler element h, then contains at most two elements.
Proof
We recall from Proposition 3.2(c) that . With [36, Thm. IV.3.5] we see that the symmetric space is a vector bundle over , hence in particular homotopy equivalent to . In view of Proposition 2.6(c), we have for that is a maximal compact subgroup of H. It follows in particular that . We conclude that . From [41, §7], we know that has at most two elements.
Example 3.12
(The inclusion may be proper) We have and because . Further by Proposition 2.6(a), so that the equality is equivalent to . This may fail for two reasons. One is failure in the adjoint group (Proposition 3.11), and the other reason is that Z(G) may be non-trivial.
Assume that is semisimple and ncc. Let G be a corresponding connected Lie group on which exists (for ) and . For the connected group , the intersection is connected but is in general not connected because Z(G) need not be contained in .
This can be seen easily for . For
3.13 |
For any connected Lie group G with Lie algebra , the group is connected 1-dimensional and for . Moreover, is a discrete subgroup which intersects trivially. Even the inclusion fails if , i.e., if acts non-trivially on Z(G). Note that Z(G) is infinite if G is simply connected.
Example 3.13
(a) For , we consider again the Euler element h from (3.13) and the Cartan involution . By Lemma B.1, the -orbit of gH is a geodesic if and only if commutes with , i.e., if
As , this leaves two possibilities:
If , then implies .
If , then implies . In this case gH is a fixed point of the modular flow.
(b) For with the Cartan involution and the causal Euler element
we obtain for . There exists a subalgebra , where the -factors correspond to the coordinates and for . Accordingly, , where the Euler elements in the -factors are conjugate to Euler elements in . Therefore, the “geodesic condition” is satisfied by all elements , where is either or .
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
carries a modular flow specified by any Euler element (corresponding to a tangent vector of length 1). Every geodesic of is an orbit of the flow generated by an Euler element of .
Remark 3.15
Let be a simple ncc symmetric Lie algebra. In general, we have for a causal Euler element a proper inclusion
By Lemma B.4, this implies that is not connected and .
For instance, if and is simple hermitian of tube type, then we obtain for any pointed generating invariant cone a hyperbolic cone . If is a causal Euler element, then follows from [38, Thm. 3.10] and the subsequent discussion, but ; see also [41, Thm. 5.4].
Example 3.16
(a) For de Sitter space (cf. Example 4.6 and Appendix D), the subspace is a single geodesic, hence in particular 1-dimensional. Note that in this case. The modular flow on M has the fixed point set .
(b) For , hermitian, we have 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 in the next section. Here, as above, .
Let be the “negative” parabolic subgroup of G specified by h and identity with the open subset . Then is an open convex subset, and our convexity theorem (Theorem 4.5) asserts that, for any with , the subset is convex.
We consider a connected semisimple Lie group G with Lie algebra and an Euler element . We put
and write
(see [2, Thm. 1.12] for the equality) for the corresponding maximal parabolic subgroups. We write
for the corresponding flag manifold. The abelian subgroup has an open orbit , which we call the open Bruhat cell. It carries a natural affine structure because the map
defines an open embedding. Below we shall always use these coordinates on .
Choose a Cartan involution with and consider the involution . We write
Then
so that
4.1 |
is an open H-orbit in . It is a real bounded symmetric domain ([26, Thm. 5.1.8]) and coincides with the unit ball in the positive real Jordan triple
4.2 |
(cf. [2, (4.6)])
The open H-orbits in
Lemma 4.1
([2, Cor. 1.10]) For and , we have if and only if the Bergman operators
and
are both invertible.
Remark 4.2
Note that
Example 4.3
We consider the group with Lie algebra and the linear basis
4.3 |
satisfying
Then,
so that
For , we have . Identifying with the projective space , the Bruhat cell is
and G acts by
In particular, we have
4.4 |
We consider the Cartan involution , so that acts by
Then
so that
4.5 |
Note that is connected.
The Jordan triple product satisfies
so that
Further
implies
Moreover,
As is invertible for all x with if and only if , it follows that
Now back to the general case. In the following we write for the spectral norm on the Jordan triple system . If with pairwise orthogonal tripotents , then
4.6 |
If
4.7 |
then we have
4.8 |
([26, Thm. 5.1.8]).
Proposition 4.4
The following assertions hold:
is equivalent to for with .
is relatively compact if and only if for with .
Proof
The condition is equivalent to Let with . Then the invariance of under implies that is equivalent to .
(a) Suppose first that . By the Spectral Theorem for positive Jordan triples ([59, Thm. VI.2.3]3), there exist pairwise orthogonal tripotents and with
([59, Thm. VI.2.3]). For and , we then have
([59, Prop. V.3.1]). As is equivalent to
the calculations in Example 4.3 show that implies .4
To prove the converse, suppose first that . Then
implies so that
Now we assume that . We observe that
so that, for , is equivalent to . For , we pick with . Then implies and thus . This shows that .
(b) If , then the argument under (a) shows that is relatively compact.
Now we assume that . We show that this implies that is unbounded. As above, we use the Spectral Theorem to write
and observe that there exists an with . For , we then obtain with (4.4)
For we get in particular
For these element leave every compact subset of . Therefore, is unbounded.
Theorem 4.5
(Convexity theorem for conformal balls) If is such that , then is convex. If is relatively compact in , then there exists an element with , so that is an affine image of .
Proof
If is relatively compact, then Proposition 4.4(b) and its proof imply the existence of with . In particular is an affine image of and therefore convex.
If is not relatively compact, then we put . Now
is an increasing union. Therefore it suffices to show that the subsets are convex. For we have
and these sets are convex by the preceding argument.
Example 4.6
We consider as the identity component of the conformal group of the Euclidean space , , and the Euler element with and . As and G preserves the positive light cone, the center of G is trivial.
The symmetric space is d-dimensional de Sitter space, is the stabilizer of the positive light ray , and is the sphere of positive light rays. On the sphere , the subgroup H has two open orbits which are positive half-spheres separated by the sphere of positive light rays in the subspace .
In the sphere the Bruhat cells are the point complements and if , then the convexity of is well-known from conformal geometry because conformal images of balls are balls or half spaces.
The subset realization of the ordered space
As before G is assumed to be a connected semisimple Lie group. To simplify the notation, we write for . Recall the following fact about the compression semigroup of the H-orbit , which is the Riemannian symmetric space .
Lemma 4.7
The compression semigroup of the open H-orbit is
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 , is simply connected and . In this case preserves , so that . Conversely, leaves invariant, so that we obtain in this particular case.
To see that the lemma also holds in the general case, note that the center of G acts trivially on and that . Therefore, the general assertion follows if the equality (4.9) holds at least for one connected Lie group G with Lie algebra . 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 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, an Euler element, a Cartan involution with and , so that is a ncc symmetric Lie algebras with . Let be the open orbit of the base point under . We endow the homogeneous space
consisting of subsets of , with the inclusion order. Then the stabilizer subgroup of the base point is H. The map induces an isomorphism
of ncc symmetric spaces, where is the unique maximal -invariant cone in containing h in its interior.
In this identification, the set is mapped to and is mapped to . In particular, is equivalent to and to .
Proof
This follows from Lemma 4.7.
Remark 4.9
(The Riemannian case) Let be a Riemannian symmetric Lie algebra, i.e., . Then , and . Thus and is a single point. Hence and is a single point. Therefore, Riemannian summands cannot be permitted in Proposition 4.8.
Example 4.10
Let and . Then, the canonical action of G on is given by
and the stabilizer of 0 is
The 1-parameter group
fixes and the orbit of is the open unit interval . The maximal cone in is generated by .
Since elements of represent one-dimensional linear subspaces of and acts transitively on triples of such subspaces, it follows easily that it acts transitively on the set of non-dense open intervals , the ordered space G/H can be identified with the ordered set of open non-dense intervals in .
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 . Then, G/H is a ncc symmetric space. Let be a Cartan involution on H. Then extends to a Cartan involution on G. Denote the corresponding maximal compact subgroup of G by K. Then is a maximal compact subgroup of H and the Riemannian symmetric space can be realized as complex symmetric bounded domain . Let be the element determining the complex structure on . Then is an Euler element in . Now (4.1) is the Harish–Chandra realization of as in (see [60, p. 58] or [22, Ch. VII] for details).
Suppose that the complex conjugation of with respect to integrates to an involution on G. This is the case if G is simply connected or if . We then assume that . If G is simply connected, then is connected and [24, Thm. VI.11] implies that , where is the stabilizer of the base point .
But in general, if G is not simply connected, then and may differ.
As an example, consider and note that in this case. Then
which is not connected because it also contains the image of . The domain (the Riemann sphere) is the upper half plane and the stabilizer subgroup of is
The reflections in exchange the two open H-orbits.
Remark 4.12
The flag manifolds appearing in this section are compact symmetric spaces on which the maximal compactly embedded subgroup acts by automorphisms. These spaces are called symmetric R-spaces.
Defining a symmetric R-space as a compact symmetric space which is a real flag manifold, Loos shows in [37, Satz 1] that this implies the existence of an Euler element such that , so that as a Riemannian symmetric space (see [41, §7.2] for more details).
If G is hermitian of tube type, then can be identified with the Šhilov boundary of the corresponding bounded symmetric domain , and this leads to a G-invariant causal structure on . As , with respect to the K-action, we have a natural 1-parameter family of K-invariant Lorentzian structures on . They correspond to -invariant Lorentzian forms on which are positive definite on and negative definite on its orthogonal space .
Observer domains associated with modular geodesics
In this section, we associate to any modular causal geodesic in an ncc semisimple symmetric space an observer domain . It is an open connected subset of M invariant under the centralizer of the corresponding causal Euler element h. We then show that, for , the domain coincides with the connected component of the base point eH of the corresponding positivity domain. In Sect. 6, we show that is connected for , which implies that in this case.
Definition 5.1
Let be a non-compactly causal symmetric Lie group and be the corresponding ncc symmetric space. We assume that , i.e., that is a direct sum of irreducible ncc symmetric Lie algebras.
(a) We write for the order on M defined by the closed Olshanski semigroup which always exists because ([30, Thm. 3.1] or Theorem C.1in Appendix C) via
and write order intervals as
where
(b) A subset is called order convex if
As the intersection of order convex subsets is order convex, we can defined the order convex hull
Clearly is the smallest order convex subset of X containing D.
(c) For a modular geodesic , we call
the observer domain associated to . Note that this domain depends on the cone specifying the order on M.
Lemma 5.2
The subset has the following properties:
is open and connected.
for .
- Suppose that and and identify with (Proposition 4.8). Then
and this domain is -invariant.5.1
Proof
(a) To see that is open, we first observe that for . For real numbers with , this implies that
This shows that is open.
To see that is connected, we recall that the order on M is globally hyperbolic, in particular all order intervals [x, y] are compact. As all elements 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 , the wedge domain is connected.
(b) Order intervals are convex and directed unions of convex sets of convex. Therefore,
is convex, whence
From the fact that h is central in , it easily follows that, in the symmetric space the geodesic line is cofinal in both directions because we have in :
For , we thus find with and . Then
in . This implies that
This completes the proof.
(c) The modular group acts on by . Therefore enlarges for and shrinks for (Theorem 4.5). As is strictly increasing, this implies that
Further
so that (5.1) follows. As any acts by linear maps on the Bruhat cell , (5.1) implies that leaves the set of all bounded domains containing 0 invariant.
Example 5.3
(de Sitter space) We consider de Sitter space
is the canonical Lorentzian form on (cf. Sect. D). Here
and
We claim that, for the modular geodesic
we have
5.2 |
where (cf. Appendix D in [52]). As the right wedge is causally complete, we clearly have . For the converse inclusion, let . We have to find a with , i.e.,
and
Since , we obtain for the right hand side
Further
and if , this expression is arbitrarily large for . This shows that , and we likewise see that .
Proposition 5.4
If and , then
.
.
Proof
If , then the corresponding subset is convex by Theorem 4.5, and it contains 0 by (5.1). Therefore the curve
is increasing because is an increasing family of subsets of . The invariance of the order thus implies that
We also know that (Theorem 4.5 and Lemma 5.2(c)), so that there exist and with . Thus
5.3 |
and therefore
Recall the definition of in (4.7). The condition
is equivalent to , showing that
5.4 |
(cf. Lemma 5.2(c)). We therefore derive from (5.3) that and since and is starlike with respect to 0, we obtain
5.5 |
We thus obtain , i.e., . This shows that , and the connectedness of (Lemma 5.2(a)) yields .
Remark 5.5
From (5.4) it follows that, as a subset of M,
5.6 |
For the quotient map , this means that
This is a -invariant domain in G specified by its intersection with the abelian subgroup ; see [41, Rem. 6.2].
Combined with Theorem 7.1, that asserts the connectedness of , the following result implies that .
Proposition 5.6
If and , then .
Proof
As both sides are -invariant (Lemma 5.2), the Positivity Domain Theorem (Theorem 3.6) implies that we have to verify the inclusion
Invariance of both sides under and further reduce the problem to the inclusion
5.7 |
To this end, we use the Lie subalgebra generated by h and (Proposition 2.8). Then and . This reduces the verification of the inclusion (5.7) to the case where , and .
As this is a product situation, it suffices to consider the case where
By (5.6), we have to show that
5.8 |
We identify with 3-dimensional Minkowski space , via
In the centerfree group , we have
so that is connected. Therefore, (de Sitter space) and corresponds to
Now implies , hence that
5.9 |
(Example 5.3). We write elements of as , (see Example 4.3). Then corresponds to
so that
This element lies in the wedge domain if and only if (Example 5.3), which is equivalent to . Then its -orbit contains the element . For , the element is of this form, showing that . This completes the proof.
Combining the preceding two propositions, we get the main result of this section. It shows that the observer domain coincides with a connected component of the positivity domain . This result provides two complementary perspectives on this domain.
Theorem 5.7
(Observer Domain Theorem) Let be a non-compactly causal semisimple symmetric Lie algebra with causal Euler element with and let G be a connected Lie group with Lie algebra and . If , then .
We can even extend this result to coverings:
Corollary 7.2
If is an open subgroup and , then holds in for .
Proof
Let be the canonical equivariant covering from [41, Lemma 7.11].
First we show that is order convex. So let in with and let be a causal curve with
Then for holds because is order convex in M.
As W is contractible by Theorem3.6(b), it is in particular simply connected. Therefore, is a disjoint union of open subsets mapped by q diffeomorphically onto W. By definition, is one such connected component, so that
is a diffeomorphism. Therefore is the unique continuous lift of in , hence contained in . This implies that , so that is order convex.
As is an isomorphism of causal manifolds, it also is an order isomorphism. Finally implies that .
Remark 5.9
It is not clear to which extent 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 . We have already seen that the positivity domain depends non-trivially on the cone C ([41, Ex. 6.8]) so one may expect that this is also the case for .
Lemma 5.10
The involution on M defined by satisfies
5.10 |
Proof
(a) The condition is equivalent to by (5.10), and this implies that
so that , i.e., . As is an involution, it follows that .
(b) As , the involution reverses the causal structure on M. Moreover, , so that
We have seen above that, for the modular geodesic in M, we have . The modular geodesic is a specific orbit of the modular flow inside W. Now we show that all other -orbits in W lead to the same “observer domain”.
Proposition 5.11
Let and consider the curve
Then
5.11 |
Proof
Using the subset realization of as from Proposition 4.8, we have
(Lemma 5.2(c)) and by Theorem 5.7. So we can write
As , the order convex hull of is contained in . To verify the converse inclusion, let . Then , and since is bounded, there exists a with . Likewise the boundedness of implies the existence of some with . Hence, . This shows that , and hence equality in (5.11).
Remark 5.12
A similar result also holds in Minkowski space. If
and
then any other element satisfies for suitable , i.e., and . In fact, for and for . Moreover, for
and, for ,
This shows that for all integral curves of the modular flow in .
Remark 5.13
On the de Sitter space , the involution can be implemented naturally by
This involution does not fix the base point , it reverses the causal structure and it commutes with modular flow. Accordingly, we have the relation
As we shall see in the next section, such a relation can only be realized because , i.e., the direction of the boost can be reversed by an element of G. If (h is not symmetric), then we shall see in Corollary 6.3 below that , so that there is no involution on M mapping to .
However, as (as involutions on ), and there are natural implementations and on , both fixing the base points, the involution implements the involution on M and fixes the base point, but it also fixes the wedge region
because it preserves h and the causal structure. This is not desirable because we would prefer that maps 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 and a simple Lie algebra , the real tube domain intersects the set of Euler elements in a connected subset (Theorem 6.1). This implies that, for an Euler element , the positivity domain is non-empty if and only if and h are conjugate (Corollary 6.3).
Theorem 6.1
Suppose that is an irreducible simple ncc symmetric Lie algebra with , , , and . Then is connected and a subset of . More precisely,
6.1 |
where is the open unit ball for which
Proof
We recall from Proposition 4.8 the open subsets which are the open orbits of the base point under . Then
follows from Proposition 4.8, applied to the causal Euler element h and its negative. These semigroups have the Lie wedges
Let for . We then have for . We conclude that and that there exists a complete metric on for which each 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 for the vector field defined by x. We now have
Hence there exists with , and thus
6.2 |
Then is an Euler element, and a similar argument shows that the vector field has a unique repelling fixed point . So for some , and . Hence the base point is a repelling fixed point of the Euler element , and is an attracting fixed point in . The attracting and repelling properties of the fixed points imply that
so that we also have
As h and are Euler elements, this entails that for . This shows that and hence that because .
We conclude that
Conversely, we have seen in Proposition 5.4 that
6.3 |
We finally obtain (6.1).
Remark 6.2
Note that the preceding proof is based on the natural embedding
which maps the Euler element to , where is the unique repelling fixed point of the flow defined by h in and is the unique attracting fixed point.
Corollary 8.3
(The set of positivity domains in M) If is an Euler element for which the positivity domain
is non-empty, then there exists a with and
Proof
As is equivalent to by (see Lemma 3.3), Theorem 6.1 implies that for some . The relation now follows directly from the definitions.
The preceding corollary shows that any wedge domain of the type , , is a G-translate of the wedge domain , where is a causal Euler element. So the action of G on the “wedge space” of M is transitive.
Corollary 8.4
If the causal Euler element h is not symmetric, then .
Remark 6.5
(Extensions to the non-simple case) If is a direct sum of irreducible ncc symmetric Lie algebra and accordingly, then
(cf. (2.3)). Projecting to the ideals , we obtain with Theorem 6.1 for and the relation
6.4 |
Further,
and imply (6.1) for this case.
Note that the situation corresponds to (see (2.2)). In the general situation, where we assume only that all ideals of contained in are compact, we have
where is compact, is a direct sum of Riemannian symmetric Lie algebras and is a direct sum of irreducible ncc symmetric Lie algebras. All Euler elements are contained in . If is only reductive, we assume , so that . Then and
We conclude that
This shows that, for any Euler element with we must have , but there is no restriction on the Riemannian component .
Connectedness of the positivity domain
In this section, we show that if is the adjoint group, then the positivity domain 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 is an irreducible simple ncc symmetric Lie algebra with and the causal Euler element . Let for and . Then the positivity domain is connected.
Proof
From Theorem 6.1 we derive that
and this leads with Lemma 3.3 to
Since has at most two connected components, this set is either connected or has two connected components ([41, Thm. 7.8]). As , we have , and preserves the open unit ball in . We thus derive from :
which is connected.
Corollary 3.7
.
Proposition 7.3
(The stabilizer group of the observer domain) If , then coincides with the stabilizer group
of the observer domain .
Proof
We work with the subset realization of as from Proposition 4.8. Then
(Lemma 5.2(c)). Since acts on by dilations, it follows that
7.1 |
Therefore, for the action of g on implies that g preserves the intersection of all subsets contained in . This shows that g fixes , so that .
Next we recall that the involution on M defined by leaves invariant (Lemma 5.10), and this leads to
The preceding proposition shows that the set of wedge domains in coincides with
7.2 |
In particular, it is a symmetric space. Recall that, by Corollary 7.2, the observer domain coincides with the positivity domain .
KMS wedge regions
With the structural results obtained so far, we have good control over the positivity domains in ncc symmetric spaces . 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:
is simple,
(by (GP) and (Eff), [52, Lemma 2.12])
denotes the complex conjugation with respect to G.
, where and is an open subgroup. Note that .
.
is open with (see §5), so that .
for the holomorphic involution of integrating the complex linear extension of .
for the conjugation of with respect to G.
integrates to the automorphism of .
Note that
8.1 |
is a holomorphic involutive automorphism of inducing on the Lie algebra .
Let
be the crown of G/K. The involution on G preserves K, hence induces an involution on G/K, and we extend it to an antiholomorphic involution on . The canonical map is a diffeomorphism ([52, Prop. 4.7]) and
implies that
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 .
For an open subgroup (where denotes the holomorphic involution) with , we obtain an embedding . Then the stabilizer of
coincides with K, so that ([52, Thm. 5.4]). Accordingly,
is an open subgroup of that coincides with the stabilizer . In this sense , but with different base points and . Recall that implies . The invariance of under implies that
so that and are exchanged by the order-4 automorphism and invariant under .
As commutes with , it also leaves invariant. Moreover, entails
Therefore, the antiholomorphic extension also preserves and induces on an antiholomorphic involution fixing the base point with stabilizer . Then
may be different from .
Remark 8.1
The condition is equivalent to . Note that is an involution that commutes with , so that the choice of determines whether is contained in or not.
For , , and , we obtain on the involution
For , the condition is equivalent to . As
it follows that
In particular, and have two connected components in .
In , a similar argument shows that . So also has two connected components, but only its identity component acts causally on . Therefore, , but for we have two choices, , or its identity component.
Comparing with the arguments in [52, Lemma 6.3], where on M, we have to be more careful in the present context. Here restricts to a map
and these two copies of G/H may not be identical. However, the antiholomorphic map
maps M to itself, fixes the base point 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
coincide on M.
We define the KMS wedge domain
8.3 |
Theorem 8.2
.
Proof
“”: For and , we first observe that
For , we thus obtain . We conclude that
8.4 |
Hence,
where
This suggest to define a “polar wedge domain” as
We actually know from Theorem 3.6 that this is the connected component containing the base point. We thus obtain
8.5 |
“”: To see that , we first recall from the first part of the proof that
To see that this domain is contained in the -invariant domain , we thus have to show that, for , we have
Let is a maximal abelian subspace (they are all conjugate under ). Then, is also maximal abelian and . So it suffices to show that, for and , we have By Proposition 2.8, is contained in a -invariant subalgebra , where is generated by h and and , where , , is an Euler element in a simple ideal of . Then is spanned by s Euler elements and
Let and . Then the discussion in Remark D.1 implies that, for and , we have
The preceding proof implies in particular the following interesting observation:
Corollary 5.8
For every , we have for , so that the orbit map extends to a holomorphic map .
Corollary 6.3
is a diffeomorphism that induces an equivalence of fiber bundles
Proof
Theorem 8.2 implies in particular that is bijective. Since is an open subset of M and an open subset of , it actually is a diffeomorphism. The second assertion follows from the fact that it commutes with the action of the subgroup .
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 according to the following types:
Complex type: and is complex conjugation with respect to . In this case , so that .
Cayley type (CT): for an Euler element . Then .
Split type (ST): for all and :
Non-split type (NST): for all and :
In Table 1, we write and . Further is maximal abelian of dimension r. For root systems of type , there are Euler elements , but for the other root systems there are less; see [38, Thm. 3.10] for the concrete list. For we write .
Table 1.
Geodesics in symmetric spaces
This appendix contains some elementary observations concerning geodesics in symmetric spaces.
Lemma B.1
Let be a symmetric space with symmetric Lie algebra , and . Then
B.1 |
holds if and only if and .
In particular, is a geodesic in M if and only if .
Proof
The relation (B.1) is equivalent to
Applying , we obtain
which leads to Evaluating the derivative of this curve in the right trivialization of T(G), we get
For we get , and taking derivatives in 0 shows that
If, conversely, this condition is satisfied, then with and , where
Therefore,
is a geodesic in M.
The following lemma provides important information on the subset .
Lemma B.2
Let and write
Then is a submanifold of M which is invariant under the action of , and the orbits or are the connected components of .
Proof
Let and . For we have
and
Let be a 0-neighborhood for which is a diffeomorphism onto an open subset of M and the spectral radius of is smaller than for . Then is invertible. With the above formula, we thus conclude for that is equivalent to , which is equivalent to . This shows that is a submanifold of M.
As , it further follows that the orbit of under the connected group contains a neighborhood of . This shows that the orbits of in are connected open subsets, hence coinciding with its connected components.
Remark B.3
For the centralizer is -invariant, so that and the dimension of the -orbit through eH is . We have
and Lemma B.2 shows that the geodesic is central in the symmetric space in the sense that its tangent space is central in the Lie algebra (cf. [36]).
Lemma B.4
For , the equality is equivalent to
B.2 |
Proof
As , the base point eH is contained in , and thus . So the equality means that , i.e., implies , resp., . This in turn is equivalent to .
Lawson’s Theorem
Let be a causal symmetric Lie group, i.e., is an involutive automorphism of G, an open subgroup and a hyperbolic pointed generating -invariant closed convex cone. We write for the corresponding decomposition of .
According to [30, Lemma 2.3], is injective if and only if satisfies
If , this condition is satisfied if and only if is injective. This condition is always satisfied if is semisimple because in this case.
Suppose that . By [30, Lemma 2.4], is a homeomorphism onto a closed subset of G if and only if, for no non-zero , the subgroup is compact. By [30, Thm. 3.1], this in turn is equivalent to the polar map
being a homeomorphism onto a closed subset of G. [30, Thm. 3.1] further shows that is a subsemigroup of G. If G is 1-connected, then the subgroup 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 be a non-compactly causal reductive symmetric Lie group. Suppose that and that . Then is a closed subsemigroup of G with Lie wedge .
Proof
Our assumption implies that is bijective, hence a diffeomorphism onto the closed subgroup . It follows in particular that is non-compact for each non-zero . Therefore the polar map 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 and if satisfies , then , which is a discrete group. We shall see below that this group may be infinite, even if .
(b) If is of complex type and G is hermitian, then is finite.
(c) If M is of non-complex type and irreducible, then is simple hermitian with , where is a causal Euler element. If Z(G) is infinite, then is also hermitian, hence of tube type because it contains an Euler element [38]. Then all Euler elements in are conjugate and this implies that is of Cayley type. So and thus is infinite if G is simply connected. This shows that it is possible that is infinite.
A concrete example is the group
where is the graph of a non-zero homomorphism . Then and .
Remark C.3
Suppose that is such that , then is contained in . Therefore the condition on in Lawson’s Theorem (cf. Appendix C) is satisfied.
de Sitter space
In this appendix we collect some concrete observations concerning de Sitter space , 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 .
In -dimensional Minkowski space , we write the Lorentzian form as
We consider d-dimensional de Sitter space
and the Euler element , defined by
It generates the Lorentz boost in the --plane. The fixed point set of the modular flow in is
This submanifold is connected for and consists of two points for . The corresponding wedge domain is the connected subset
By [52, Prop. D.3], the timelike geodesics of M of velocity 1 take the form
whereas the trajectories of the modular flow are
Comparing both expressions leads for h-modular geodesics to the conditions
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 determines the geodesic uniquely.
We infer from [52, Prop. D.3] that
is a closed space-like geodesics. For , its values are contained in , and this geodesic arc connects the two fixed points to of the modular flow.
Remark D.1
In addition to h, we also consider the Euler elements defined by
The involution corresponding to h acts on by
and its antilinear extension acts on by
Note that, in , we have and , so that and .
In , we consider the domain
On the antiholomorphic involution has the fixed point set
It follows in particular that
The analytic extension of the modular flow acts on by
Starting with a -fixed element in , this leads to
with imaginary part
so that we obtain for that
which implies that
D.1 |
Example D.2
For the special case , we have and the Euler element
corresponds to the base point (see [52]), so that
Accordingly,
For we then have
We also note that, for , the Lie algebra element corresponds to . Note that
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
A closed real subspace of a complex Hilbert space is called standard if and
Note that the cones are open, whereas C is closed.
This theorem is stated for complex hermitian Jordan triple systems, but is a real form of the complex JTS on which we have an antilinear isomorphism with . Therefore, the uniqueness in the spectral decomposition shows that, for , the corresponding spectral tripotents are contained in V.
As is invariant under the group 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 is contained in if and only if for every j. This follows easily from (4.5).
This reference deals with bounded symmetric domains in complex spaces, but can be embedded into such a domain by embedding . If is an invariant cone with , then is a causal embedding and is a real form of a complex bounded symmetric domain ; 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 -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]