Log in

Wedge domains in non-compactly causal symmetric spaces

  • Original Paper
  • Published:
Geometriae Dedicata Aims and scope Submit manuscript

Abstract

This article is part of an ongoing project aiming at the connections between causal structures on homogeneous spaces, Algebraic Quantum Field Theory, modular theory of operator algebras and unitary representations of Lie groups. In this article we concentrate on non-compactly causal symmetric spaces G/H. This class contains de Sitter space but also other spaces with invariant partial ordering. The central ingredient is an Euler element h in the Lie algebra of \({{\mathfrak {g}}}\). We define three different kinds of wedge domains depending on h and the causal structure on G/H. Our main result is that the connected component containing the base point eH of these seemingly different domains all agree. Furthermore we discuss the connectedness of those wedge domains. We show that each of these spaces has a natural extension to a non-compactly causal symmetric space of the form \(G_{{\mathbb {C}}}/G^c\) where \(G^c\) is a certain real form of the complexification \(G_{{\mathbb {C}}}\) of G. As \(G_{{\mathbb {C}}}/G^c\) is non-compactly causal, it also contains three types of wedge domains. Our results says that the intersection of these domains with G/H agree with the wedge domains in G/H.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price includes VAT (Germany)

Instant access to the full article PDF.

Similar content being viewed by others

Data availibility

Data sharing is not applicable to this article as no dataset were used or generated in the article.

Notes

  1. We write \(C^\circ \) for the relative interior of the cone C in its span.

  2. As in the proof of Lemma 4.8 below, this follows from the \(\mathop {\mathrm{Inn}}\nolimits _{{\mathfrak {g}}}({{\mathfrak {h}}}_1)\)-invariance of the characteristic function of \(C_1\) by taking \(D = \varphi ^{-1}([1,\infty ))\).

  3. A subspace is called hyperbolic if it consists of hyperbolic elements.

  4. We want to keep some flexibility in choosing \(K_{{\mathbb {C}}}\) and hence the complexification \(M^r_{{\mathbb {C}}}\) because the crown domain, which is simply connected, can be realized in many “complexifications”.

References

  1. Akhiezer, D.N., Gindikin, S.G.: On Stein extensions of real symmetric spaces. Math. Ann. 286, 1–12 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  2. Araki, H., Woods, E.J.: Representations of the canonical commutation relations describing a nonrelativistic infinite free Bose gas. J. Math. Phys. 4, 637–662 (1963)

    Article  MathSciNet  Google Scholar 

  3. Baumgärtel, H., Jurke, M., Lledo, F.: Twisted duality of the CAR-algebra. J. Math. Phys. 43(8), 4158–4179 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  4. Borchers, H.-J.: On the net of von Neumann algebras assiciated with a wedge and wedge-causal manifold (2009). Preprint, available at http://www.theorie.physik.uni-goettingen.de/forschung/qft/publications/2009

  5. Borchers, H.-J., Buchholz, D.: Global properties of vacuum states in de Sitter space. Annales Poincare Phys. Theor. A 70, 23–40 (1999)

    MathSciNet  MATH  Google Scholar 

  6. Bratteli, O., Robinson, D.W.: Operator Algebras and Quantum Statistical Mechanics I, 2nd edn., Texts and Monographs in Physics. Springer, Berlin (1987)

  7. Bros, J., Moschella, U.: Two-point functions and quantum fields in de Sitter universe. Rev. Math. Phys. 8, 327–391 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  8. Buchholz, D., Mund, J., Summers, S.J.: Transplantation of local nets and geometric modular action on Robertson-Walker Space-Times. Fields Inst. Commun. 30, 65–81 (2001)

    MathSciNet  MATH  Google Scholar 

  9. Buchholz, D., Summers, S.J.: Stable quantum systems in anti-de Sitter space: causality, independence and spectral properties. J. Math. Phys. 45, 4810–4831 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  10. Ciolli, F., Longo, R., Ranallo, A., Ruzzi, G.: Relative entropy and curved spacetimes. J. Geom. Phys. 172, Paper No. 104416, 16 pp (2022)

  11. Dappiaggi, C., Lechner, G., Morfa-Morales, E.: Deformations of quantum field theories on spacetimes with Killing vector fields. Commun. Math. Phys. 305(1), 99–130 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  12. Dörr, N., Neeb, K.-H.: On wedges in Lie triple systems and ordered symmetric spaces. Geometriae Dedicata 46, 1–34 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  13. Faraut, J., Koranyi, A.: Analysis on Symmetric Cones. Oxford Math Monographs, Oxford University Press, Oxford (1994)

    MATH  Google Scholar 

  14. Gindikin, S., Krötz, B.: Complex crowns of Riemannian symmetric spaces and non-compactly causal symmetric spaces. Trans. Am. Math. Soc. 354(8), 3299–3327 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  15. Haag, R.: Local Quantum Physics. Fields, Particles, Algebras, 2nd edn, Texts and Monographs in Physics. Springer, Berlin (1996)

  16. Hilgert, J., Neeb, K.-H.: Lie Semigroups and Their Applications, Lecture Notes in Math, vol. 1552. Springer, Berlin (1993)

    Book  MATH  Google Scholar 

  17. Hilgert, J., Neeb, K.-H.: Structure and Geometry of Lie Groups. Springer, Berlin (2012)

    Book  MATH  Google Scholar 

  18. Hilgert, J., Ólafsson, G.: Causal Symmetric Spaces, Geometry and Harmonic Analysis, Perspectives in Mathematics, vol. 18. Academic Press, New York (1997)

    MATH  Google Scholar 

  19. Krötz, B.: Corner view on the crown domain. Jpn. J. Math. 2, 303–311 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  20. Krötz, B., Neeb, K.-H.: On hyperbolic cones and mixed symmetric spaces. J. Lie Theory 6(1), 69–146 (1996)

    MathSciNet  MATH  Google Scholar 

  21. Krötz, B., Stanton, R.J.: Holomorphic extensions of representations. II. Geometry and harmonic analysis. Geom. Funct. Anal. 15(1), 190–245 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  22. Lauridsen-Ribeiro, P.: Structural and Dynamical Aspects of the AdS/CFT Correspondence: a Rigorous Approach. PhD thesis, Sao Paulo (2007) ar**v:0712.0401v3 [math-ph] (2008)

  23. Lechner, G.: Algebraic Constructive Quantum Field Theory: Integrable Models and Deformation Techniques. In: Brunetti, R., et al. (eds.) Advances in Algebraic Quantum Field Theory, Math. Phys. Stud, pp. 397–449. Springer, Berlin (2015). ar**v:1503.03822

  24. Loos, O.: Symmetric Spaces I: General Theory. W. A. Benjamin Inc, New York (1969)

    MATH  Google Scholar 

  25. Morinelli, V., Neeb, K.-H.: Covariant homogeneous nets of standard subspaces. Commun. Math. Phys. 386, 305–358 (2021). ar**v:2010.07128

  26. Morinelli, V., Neeb, K.-H., Ólafsson, G.: From Euler elements and \(3\)-gradings to non-compactly causal symmetric spaces. J. Lie Theory 33(1) (2023) (to appear). ar**v:2207.14034

  27. Morinelli, V., Neeb, K.-H., Ólafsson, G.: Modular geodesics and wedge domains in general non-compactly causal symmetric spaces (in preparation)

  28. Neeb, K.-H.: On the complex geometry of invariant domains in complexified symmetric spaces. Annales de l’Inst. Fourier 49(1), 177–225 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  29. Neeb, K.-H.: Holomorphy and Convexity in Lie Theory, Expositions in Mathematics, vol. 28. de Gruyter Verlag, Berlin (2000)

    Book  Google Scholar 

  30. Neeb, K.-H.: On differentiable vectors for representations of infinite dimensional Lie groups. J. Funct. Anal. 259, 2814–2855 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  31. Neeb, K.-H., Ólafsson, G.: Antiunitary representations and modular theory. In: Grabowska, K., et al. (eds.) 50th Sophus Lie Seminar, Banach Center Publications, vol. 113, pp. 291–362. ar**v:1704.01336

  32. Neeb, K.-H., Ólafsson, G.: Wedge domains in compactly causal symmetric spaces. Int. Math. Res. Notices (to appear). ar**v:2107.13288

  33. Neeb, K.-H., Ólafsson, G., Ørsted, B.: Standard subspaces of Hilbert spaces of holomorphic functions on tube domains. Commun. Math. Phys. 386, 1437–1487 (2021). ar**v:2007.14797

  34. Neumann, A., Ólafsson, G.: Minimal and maximal semigroups related to causal symmetric spaces. Semigroup Forum 61, 57–85 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  35. Oeh, D.: Classification of 3-graded causal subalgebras of real simple Lie algebras appeared. Trans. Groups 27, 1393–1430 (2022)

  36. Ólafsson, G.: Symmetric spaces of Hermitian type. Differ. Geom. Appl. 1, 195–233 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  37. Ólafsson, G., Ørsted, B.: The holomorphic discrete series of an affine symmetric space and representations with reproducing kernels. Trans. Am. Math. Soc. 326, 385–405 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  38. Schroer, B.: Wigner representation theory of the Poincaré group, localization, statistics and the S-matrix. Nuclear Phys. B 499–3, 519–546 (1997)

    Article  MATH  Google Scholar 

  39. Simon, B.: The \(P(\Phi )_2\) Euclidean (Quantum) Field Theory. Princeton University Press, Princeton (1974)

    Google Scholar 

  40. Thomas, L.J., Wichmann, E.H.: On the causal structure of Minkowski spacetime. J. Math. Phys. 38, 5044–5086 (1997)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Gestur Ólafsson.

Additional information

Publisher's Note

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

The research of K.-H. Neeb was partially supported by DFG-Grant NE 413/10-1. The research of G. Ólafsson was partially supported by Simons Grant 586106.

Appendices

Appendix

1.1 A Irreducible modular ncc symmetric Lie algebras

See Table 1.

Table 1 Irreducible ncc symmetric Lie algebras \(({{\mathfrak {g}}},\tau )\) with \({\mathcal {E}}({{\mathfrak {g}}}) \cap {{\mathfrak {h}}}\not =\emptyset \)

B Some calculations in \(\mathop {{\mathfrak {sl}}}\nolimits _2({{\mathbb {R}}})\)

Arguments are often reduced to relatively simple \(\mathop {{\mathfrak {sl}}}\nolimits _2({{\mathbb {R}}})\) calculations. We therefore collect the basic notations and calculations here in one place for reference. For \({{\mathfrak {g}}}= \mathop {{\mathfrak {sl}}}\nolimits _2({{\mathbb {R}}})\), we fix the Cartan involution \(\theta (x) = - x^\top \), so that

$$\begin{aligned} {{\mathfrak {k}}}= \mathop {{\mathfrak {so}}}\nolimits _2({{\mathbb {R}}}) \quad \text{ and } \quad {{\mathfrak {p}}}= \{ x \in \mathop {{\mathfrak {sl}}}\nolimits _2({{\mathbb {R}}}) : x^\top = x \}.\end{aligned}$$

The basis elements

$$\begin{aligned} h^0 := \frac{1}{2} \begin{pmatrix} 1 &{} 0 \\ 0 &{} -1 \end{pmatrix},\quad e^0 = \begin{pmatrix} 0 &{} 1 \\ 0 &{} 0 \end{pmatrix}\quad \text{ and } \quad f^0 = \begin{pmatrix} 0 &{} 0 \\ 1 &{} 0 \end{pmatrix} \end{aligned}$$
(B.1)

and

$$\begin{aligned} h^1 = \frac{1}{2} \begin{pmatrix} 0 &{} 1 \\ 1 &{} 0 \end{pmatrix} = \frac{1}{2}(e^0 + f^0),\quad e^1 = \frac{1}{2}\begin{pmatrix} -1 &{} 1 \\ -1 &{} 1 \end{pmatrix}\quad \text{ and } \quad f^1= \frac{1}{2} \begin{pmatrix} -1 &{} -1 \\ 1 &{} 1 \end{pmatrix} \end{aligned}$$
(B.2)

satisfy

$$\begin{aligned}{}[h^j, e^j] = e^j, \quad [h^j, f^j] = -f^j, \quad {[}e^j, f^j] = 2 h^j \quad \text{ and } \quad \theta (e^j) = -f^j. \end{aligned}$$

For the involution

$$\begin{aligned} \tau \begin{pmatrix} a &{} b \\ c &{} d \end{pmatrix} = \begin{pmatrix} a &{} -b \\ -c &{} d \end{pmatrix} \end{aligned}$$

we have

$$\begin{aligned} {{\mathfrak {h}}}= {{\mathfrak {g}}}^\tau = {{\mathbb {R}}}h^0, \quad {{\mathfrak {q}}}= {{\mathfrak {g}}}^{-\tau } = {{\mathbb {R}}}h^1 + {{\mathbb {R}}}(e^0 - f^0) \quad \text {and} \quad C = [0,\infty ) e^0 + [0,\infty ) f^0 \end{aligned}$$

is a hyperbolic \(\mathop {\mathrm{Inn}}\nolimits ({{\mathfrak {h}}})\)-invariant cone in \({{\mathfrak {q}}}\), containing \(h^1\) as a causal Euler element. Further

$$\begin{aligned} C_+ = [0,\infty ) e^0 =\left\{ \begin{pmatrix} 0 &{} x\\ 0 &{} 0\end{pmatrix} : x\ge 0\right\} \quad \text{ and } \quad C_- = - [0,\infty ) f^0 =\left\{ \begin{pmatrix} 0 &{} 0\\ -y &{} 0\end{pmatrix}: y\ge 0\right\} \end{aligned}$$

lead to

$$\begin{aligned} C_+ + C_- = [0,\infty ) e^0 - [0,\infty ) f^0 =\left\{ \begin{pmatrix} 0 &{} x \\ -y &{} 0\end{pmatrix}: x,y\ge 0\right\} ,\end{aligned}$$

so that

$$\begin{aligned} (C_+ + C_-)^\pi \cap {{\mathbb {R}}}(e^0 - f^0) = \Big \{ t (e^0 - f^0) : 0< t < \frac{\pi }{2}\Big \}. \end{aligned}$$
(B.3)

The subspace \({{\mathfrak {t}}}_{{\mathfrak {q}}}:= {{\mathbb {R}}}(e^0 - f^0) = \mathop {{\mathfrak {so}}}\nolimits _2({{\mathbb {R}}})\) of \({{\mathfrak {q}}}\) is maximal elliptic. For

$$\begin{aligned} x_0 := \frac{\pi }{4} (e^0 - f^0) = \frac{\pi }{4} \begin{pmatrix} 0 &{} 1 \\ -1 &{} 0 \end{pmatrix} \in {{\mathfrak {t}}}_{{\mathfrak {q}}}\end{aligned}$$

and

$$\begin{aligned} g_0 := \exp (x_0) = \begin{pmatrix} \cos \big (\frac{\pi }{4}\big ) &{} \sin \big (\frac{\pi }{4}\big )\\ -\sin \big (\frac{\pi }{4}\big ) &{} \cos \big (\frac{\pi }{4}\big ) \end{pmatrix} = \frac{1}{\sqrt{2}} \begin{pmatrix} 1 &{} 1 \\ -1 &{} 1 \end{pmatrix},\end{aligned}$$

we then have

$$\begin{aligned} \mathop {\mathrm{Ad}}\nolimits (g_0) h^1 = h^0 \quad \text{ and } \quad \mathop {\mathrm{Ad}}\nolimits (g_0) h^0 = - h^1, \end{aligned}$$
(B.4)

More generally, we have for \(t \in {{\mathbb {R}}}\)

$$\begin{aligned} e^{t\mathop {\mathrm{ad}}\nolimits (e^0 - f^0)} h^1 = \cos (2t) h^1 + \sin (2t) h^0 \end{aligned}$$
(B.5)

because \( [e^0 - f^0, h^1] = 2 h^0, \quad [e^0 - f^0, h^0] = -2 h^1. \)

Now fix the Euler element

$$\begin{aligned} h = h^0 = \frac{1}{2} \begin{pmatrix} 1 &{} 0 \\ 0 &{} -1 \end{pmatrix} \quad \text{ and } \quad \tau _h = e^{\pi i \mathop {\mathrm{ad}}\nolimits h},\end{aligned}$$

and similarly drop the \({}^0\) in other places. Then we have

$$\begin{aligned} {{\mathfrak {g}}}_1(h) = {{\mathbb {R}}}e, \quad e = \begin{pmatrix} 0 &{} 1 \\ 0 &{} 0 \end{pmatrix} \quad \text{ and } \qquad {{\mathfrak {g}}}_{-1}(h) = {{\mathbb {R}}}f, \quad f = \begin{pmatrix} 0 &{} 0 \\ 1 &{} 0 \end{pmatrix}.\end{aligned}$$

Recall that

$$\begin{aligned}{}[h,e] = e, \qquad [h,f] = -f \quad \text{ and } \quad [e,f] = 2h. \end{aligned}$$
(B.6)

We have

$$\begin{aligned} (C_{{\mathfrak {g}}}^\circ )^{-\tau _h} = \underbrace{{{\mathbb {R}}}_+ e}_{C_+^\circ } - \underbrace{{{\mathbb {R}}}_+ f}_{C_-^\circ } \quad \text{ with } \quad C_+^\circ = {{\mathbb {R}}}_+ e, \quad C_-^\circ = {{\mathbb {R}}}_+ f.\end{aligned}$$

For \(y = \lambda e + \mu f \in {{\mathfrak {g}}}^{-\tau _h}\), we have

$$\begin{aligned}{}[y,h] = -\lambda e + \mu f \quad \text{ and } \quad [y,[y,h]] = [\lambda e + \mu f, -\lambda e + \mu f] = \lambda \mu 2[e,f] = 4 \lambda \mu h.\end{aligned}$$

We define the function \(S : {{\mathbb {C}}}\rightarrow {{\mathbb {C}}}\) by \(S(z) = \sum _{k = 0}^\infty \frac{(-1)^k}{(2k+1)!} z^k\), so that \(\sin (z) = z S(z^2).\) Then

$$\begin{aligned} \sin (\mathop {\mathrm{ad}}\nolimits y)h&= S((\mathop {\mathrm{ad}}\nolimits y)^2) [y,h] = S(4 \lambda \mu ) \cdot (-\lambda e+ \mu f). \end{aligned}$$

This element is contained in \(-C_{{\mathfrak {g}}}^\circ \) if \(\lambda , \mu > 0\) and \(\sin (2\sqrt{\lambda \mu }) > 0\). This is satisfied for \(0< 4\mu \lambda < \pi ^2\). Then y is hyperbolic with \(\mathrm{Spec}(\mathop {\mathrm{ad}}\nolimits y) = \{ \pm 2 \sqrt{\lambda \mu }\} \subseteq (-\pi , \pi )\).

Example B.1

(cf. Example 6.2) Let G be a connected Lie group with Lie algebra \({{\mathfrak {g}}}\) and \(h \in {{\mathfrak {g}}}\) an Euler element. In general \(G^h\) may be much larger than \(G^{\tau _h}\). This can be seen for \(G={\widetilde{\mathop {\mathrm{SL}}\nolimits }}_2({{\mathbb {R}}})\) (the simply connected covering group of \(\mathop {\mathrm{SL}}\nolimits _2({{\mathbb {R}}})\)) and the Euler element \(h = \mathop {\mathrm{diag}}\nolimits (1/2,-1/2)\). The group \(G^h\) contains \(Z(G) \cong {{\mathbb {Z}}}\) and \(G^h_e = \exp ({{\mathbb {R}}}h)\), which shows that \(\pi _0(G^h) \cong {{\mathbb {Z}}}\). The involution \(\tau _h\) acts on \(Z(G)\subseteq \exp (\mathop {{\mathfrak {so}}}\nolimits _2({{\mathbb {R}}}))\) by inversion (cf. [25, Ex. 2.10(d)]).

For \(G = \mathop {\mathrm{SL}}\nolimits _2({{\mathbb {R}}})\) and the same Euler element h, the centralizer \(G^h\) is the non-connected subgroup of diagonal matrices, which coincides with the fixed point group \(G^{\tau _h}\) of the corresponding involution

$$\begin{aligned} \tau _h\begin{pmatrix} a &{} b \\ c &{} d \end{pmatrix} = \begin{pmatrix} a &{} -b \\ -c &{} d \end{pmatrix}.\end{aligned}$$

On the other hand, if we consider the centerfree group \(G=\mathop {\mathrm{SO}}\nolimits _{1,2}({{\mathbb {R}}})_e \cong \mathop {\mathrm{PSL}}\nolimits _2({{\mathbb {R}}})\), then \(G^h=G^h_e\), but the subgroup \(G^{\tau _h}\) has 2 connected components: \(G^{\tau _h}= G^h\cup \theta G^h\), where \(\theta \) is a Cartan involution with \(\theta (h)=-h\).

C Polar maps

In this section we discuss polar maps associated to an involution on a symmetric space, resp., to a pair of commuting involutions on a Lie group. Key properties are collected in Lemma C.4. These results are used in particular in Sect. 4.2 to obtain a polar decomposition of the crown domain of the Riemannian symmetric space \(M^r = G/K\) and in the characterization of this domain as a subset of the tube domain of the cone C (Theorem 4.10).

1.1 C.1 Some spectral theory

Let V be a finite dimensional real vector space and \(A \in \mathop {\mathrm{End}}\nolimits (V)\).

Lemma C.1

\(\ker \big (\frac{\sinh (A)}{A}\big ) = \bigoplus _{n \not =0} \ker (A^2 + n^2\pi ^2 \mathbf{1 })\).

Proof

We may w.l.o.g. assume that V is complex. Then \(B := \sinh (A)/A\) is invertible on all generalized eigenspace corresponding to eigenvalues \(\lambda \not = \pi n i\), \(n\in {{\mathbb {Z}}}\setminus \{0\}\). We may therefore assume that V has only one eigenvalue \(\lambda = n \pi i\), \(n \not =0\). Then A is invertible, so that

$$\begin{aligned} \ker (B) = \ker (\sinh (A)) = \ker (e^A- e^{-A}) = \ker (e^{2A} - \mathbf{1 }).\end{aligned}$$

Writing \(A = A_s + A_n\) for the Jordan decomposition of A, it follows that

$$\begin{aligned} e^{2A} = e^{2 A_s} e^{2 A_n} = e^{2 n \pi i} e^{2 A_n} = e^{2A_n}.\end{aligned}$$

As \(\ker (e^{2A_n}-\mathbf{1 }) = \ker (A_n)\) follows from \(2A_n = \log (e^{2A_n})\) as a polynomial in \(e^{2A_n}-1\), we see that \(\ker (B) = \ker (A_n)\) is the \(\lambda \)-eigenspace of A. \(\square \)

With similar arguments, or by replacing A by \(A - \frac{\pi i}{2}\mathbf{1 }\), we get:

Lemma C.2

\(\ker (\cosh (A)) = \ker (e^{-2A} + \mathbf{1 }) = \bigoplus _{n \in {{\mathbb {N}}}}\ker \Big (A^2 + \big (n + \frac{1}{2}\big )^2 \pi ^2 \mathbf{1 }\Big )\).

1.2 C.2 Fine points on polar maps

In this subsection, we consider two commuting involutions \(\sigma \) and \(\tau \) on a connected, not necessarily reductive, Lie group G and an open \(\sigma \)-invariant subgroup \(H \subseteq G^\tau \). We shall study the polar map

$$\begin{aligned} \Phi : G^\sigma \times _{H^\sigma } {{\mathfrak {q}}}^{-\sigma } \rightarrow M, \quad [g,x] \mapsto g.\mathop {\mathrm{Exp}}\nolimits _{eH}(x) = g \exp (x)H\in G/H \end{aligned}$$
(C.1)

and its applications.

As H is invariant under \(\tau \) and \(\sigma \), both define commuting involutions on M and their fixed point manifolds intersect transversally in eH. The map

$$\begin{aligned} \Psi : G^\sigma \times _{H^\sigma } {{\mathfrak {q}}}^{-\sigma }\rightarrow N(G^\sigma /H^\sigma ), \quad \ [g,x] \mapsto g.x\end{aligned}$$

is a diffeomorphism onto the normal bundle \(N(G^\sigma /H^\sigma )\) of the subspace \(G^\sigma .eH \cong G^\sigma /H^\sigma \) and \(\Phi = \mathop {\mathrm{Exp}}\nolimits \circ \Psi \), where \(\mathop {\mathrm{Exp}}\nolimits : T(M) \rightarrow M\) is the exponential map.

First, we determine the regular points of \(\Phi \). As \(\Phi \) is \(G^\sigma \)-equivariant, it suffices to determine for which points [ex] the tangent map \(T_{[e,x]}(\Phi )\) is injective, hence bijective for dimensional reasons. In the following calculation, we shall use the formula

$$\begin{aligned} T_x(\mathop {\mathrm{Exp}}\nolimits _{eH})y = \exp (x).\frac{\sinh (\mathop {\mathrm{ad}}\nolimits x)}{\mathop {\mathrm{ad}}\nolimits x} y \quad \text{ for } \quad x,y \in {{\mathfrak {q}}}\end{aligned}$$
(C.2)

for the differential of \(\mathop {\mathrm{Exp}}\nolimits \) [12, Lemma 4.6], where \({{\mathfrak {q}}}\rightarrow T_{\mathop {\mathrm{Exp}}\nolimits _{eH}(x)}, v \mapsto \exp x.v\), is the linear isomorphism induced by the action of \(\exp x \in G\) on M. For \(a \in {{\mathfrak {g}}}^\sigma , x,b \in {{\mathfrak {q}}}^{-\sigma }\), we obtain

$$\begin{aligned} T_{[e,x]}(\Phi )(a,b)&= a.\mathop {\mathrm{Exp}}\nolimits (x) + T_{x}(\mathop {\mathrm{Exp}}\nolimits _{eH})(b) \nonumber \\&= \exp (x).\Big (p_{{{\mathfrak {q}}}}(e^{-\mathop {\mathrm{ad}}\nolimits x} a) + \frac{\sinh (\mathop {\mathrm{ad}}\nolimits x)}{\mathop {\mathrm{ad}}\nolimits x} b\Big ) \end{aligned}$$
(C.3)

Note that \(e^{-\mathop {\mathrm{ad}}\nolimits x}a = \cosh (\mathop {\mathrm{ad}}\nolimits x) a - \sinh (\mathop {\mathrm{ad}}\nolimits x) a\). If \(a\in {{\mathfrak {h}}}^{\sigma }\) then \(p_{{\mathfrak {q}}}(e^{-\mathop {\mathrm{ad}}\nolimits x}a) = - \sinh (\mathop {\mathrm{ad}}\nolimits x)a\), and if \(a \in {{\mathfrak {q}}}^\sigma \), then \(p_{{\mathfrak {q}}}(e^{-\mathop {\mathrm{ad}}\nolimits x}a) = \cosh (\mathop {\mathrm{ad}}\nolimits x)a\). Writing \(a = a_{{\mathfrak {h}}}+ a_{{\mathfrak {q}}}\) with \(a_{{\mathfrak {h}}}\in {{\mathfrak {h}}}^\sigma \) and \(a_{{\mathfrak {q}}}\in {{\mathfrak {q}}}^\sigma \), we thus obtain

$$\begin{aligned} T_{[e,x]}(\Phi )(a,b) = \exp (x).\Big (\underbrace{\cosh (\mathop {\mathrm{ad}}\nolimits x)a_{{\mathfrak {q}}}}_{\in {{\mathfrak {q}}}^\sigma } + \underbrace{\frac{\sinh (\mathop {\mathrm{ad}}\nolimits x)}{\mathop {\mathrm{ad}}\nolimits x} b - \sinh (\mathop {\mathrm{ad}}\nolimits x)a_{{\mathfrak {h}}}}_{\in {{\mathfrak {q}}}^{-\sigma }} \Big ). \end{aligned}$$
(C.4)

The following lemma provides a characterization of the regular points.

Lemma C.3

For \(x \in {{\mathfrak {q}}}\), the following assertions hold:

  1. (a)

    \(\mathop {\mathrm{Exp}}\nolimits _{eH}\) is regular in x if and only if the map \(\frac{\sinh (\mathop {\mathrm{ad}}\nolimits x)}{\mathop {\mathrm{ad}}\nolimits x} : {{\mathfrak {q}}}\rightarrow {{\mathfrak {q}}}\) is invertible, which is equivalent to

    $$\begin{aligned} \mathrm{Spec}(\mathop {\mathrm{ad}}\nolimits x\vert _{{{\mathfrak {q}}}_L}) \cap {{\mathbb {Z}}}\pi i \subseteq \{0\}, \quad \text{ where } \quad {{\mathfrak {q}}}_L := {{\mathfrak {q}}}+ [{{\mathfrak {q}}},{{\mathfrak {q}}}]. \end{aligned}$$
    (C.5)
  2. (b)

    If \(\mathop {\mathrm{Exp}}\nolimits _{eH}\vert _{{{\mathfrak {q}}}^{-\sigma }}\) is regular in \(x \in {{\mathfrak {q}}}^{-\sigma }\), then the polar map \(\Phi \) in (C.2) is regular in [gx] if and only if, in addition, \(\cosh (\mathop {\mathrm{ad}}\nolimits x) : {{\mathfrak {q}}}^\sigma \rightarrow {{\mathfrak {q}}}^\sigma \) is invertible, which is equivalent to

    $$\begin{aligned} \mathrm{Spec}(\mathop {\mathrm{ad}}\nolimits x\vert _{{{\mathfrak {q}}}_L}) \cap \Big (\frac{\pi }{2} + {{\mathbb {Z}}}\pi \Big ) i = \emptyset . \end{aligned}$$
    (C.6)

Proof

(a) follows from the spectral theoretic description of the kernel of \(\frac{\sinh (\mathop {\mathrm{ad}}\nolimits x)}{\mathop {\mathrm{ad}}\nolimits x}\big |_{{{\mathfrak {q}}}}\) as the intersection of \({{\mathfrak {q}}}\) with the sum of the eigenspaces of \(\mathop {\mathrm{ad}}\nolimits x\) in \({{\mathfrak {g}}}_{{\mathbb {C}}}\) for the eigenvalues \(\lambda \in \pi i {{\mathbb {Z}}}\setminus \{0\}\) (Lemma C.1).

(b) Suppose that the restriction of \(\mathop {\mathrm{Exp}}\nolimits _{eH}\) to \({{\mathfrak {q}}}^{-\sigma }\) is regular, i.e., that \(\frac{\sinh (\mathop {\mathrm{ad}}\nolimits x)}{\mathop {\mathrm{ad}}\nolimits x} : {{\mathfrak {q}}}^{-\sigma } \rightarrow {{\mathfrak {q}}}^{-\sigma }\) is invertible. Then (C.4) shows that \(\Phi \) is regular in [ex] if and only if \(\cosh (\mathop {\mathrm{ad}}\nolimits x) : {{\mathfrak {q}}}^\sigma \rightarrow {{\mathfrak {q}}}^\sigma \) is invertible, and this is equivalent to the condition on \(\mathrm{Spec}(\mathop {\mathrm{ad}}\nolimits x\vert _{{{\mathfrak {q}}}_L})\) stated in (b). \(\square \)

The following lemma contains a wealth of information on singular points of the polar map \(\Phi \).

Lemma C.4

Let \(\Omega \subseteq {{\mathfrak {q}}}^{-\sigma }\) be an open \(H^\sigma \)-invariant subset consisting of \(\mathop {\mathrm{Exp}}\nolimits \)-regular elliptic elements, and consider the polar map

$$\begin{aligned} \Phi : G^\sigma _e \times _{H^\sigma } \Omega \rightarrow M, \quad [g,y] \mapsto g.\mathop {\mathrm{Exp}}\nolimits _{eH}(y).\end{aligned}$$

Let \(x \in \Omega \), write \(m := \mathop {\mathrm{Exp}}\nolimits _{eH}(x) \in M = G/H\), \({\mathcal {O}}_m := G^\sigma _e.m\) for its orbit, and put

$$\begin{aligned} \sigma _x := e^{-2 \mathop {\mathrm{ad}}\nolimits x}, \qquad \zeta _x := e^{-\mathop {\mathrm{ad}}\nolimits x} \in \mathop {\mathrm{Aut}}\nolimits ({{\mathfrak {g}}}).\end{aligned}$$

Then the following assertions hold:

  1. (a)

    \({{\mathfrak {g}}}^{-\sigma _x} := \{ y \in {{\mathfrak {g}}}: \sigma _x(y) = - y \} = \ker (\cosh (\mathop {\mathrm{ad}}\nolimits x))\).

  2. (b)

    \({{\mathfrak {q}}}^{\sigma , -\sigma _x}\) complements the subspace \(\exp (-x).\mathop {\mathrm{im}}\nolimits (T_{[e,x]}(\Phi )) = \cosh (\mathop {\mathrm{ad}}\nolimits x){{\mathfrak {q}}}^\sigma \oplus {{\mathfrak {q}}}^{-\sigma }\) in \({{\mathfrak {q}}}\).

  3. (c)

    The eigenspaces \({{\mathfrak {g}}}^{\pm \sigma _x}\) are \(\tau \)-invariant, and on the Lie subalgebra \({{\mathfrak {g}}}^{\sigma _x^2} = {{\mathfrak {g}}}^{\sigma _x} \oplus {{\mathfrak {g}}}^{-\sigma _x}\), the involution \(\tau \) commutes with \(\sigma _x\).

  4. (d)

    The eigenspaces \({{\mathfrak {g}}}^{\pm \sigma _x}\) are \(\zeta _x\)-invariant, on \({{\mathfrak {g}}}^{\sigma _x}\) the automorphisms \(\zeta _x\) and \(\tau \) commute, and on on \({{\mathfrak {g}}}^{-\sigma _x}\) the complex structure \(\zeta _x\) and \(\tau \) anticommute. In particular, we have

    $$\begin{aligned} \zeta _x({{\mathfrak {h}}}^{\sigma _x}) = {{\mathfrak {h}}}^{\sigma _x}, \quad \zeta _x({{\mathfrak {q}}}^{\sigma _x}) = {{\mathfrak {q}}}^{\sigma _x}, \quad \zeta _x({{\mathfrak {h}}}^{-\sigma _x}) = {{\mathfrak {q}}}^{-\sigma _x}, \quad \zeta _x({{\mathfrak {q}}}^{-\sigma _x}) = {{\mathfrak {h}}}^{-\sigma _x}.\end{aligned}$$
  5. (e)

    The eigenspaces \({{\mathfrak {g}}}^{\pm \sigma _x}\) are \(\sigma \)-invariant, and on the Lie subalgebra \({{\mathfrak {g}}}^{\sigma _x^2}\), the involution \(\sigma \) commutes with \(\sigma _x\). On \({{\mathfrak {g}}}^{\sigma _x}\), the automorphisms \(\zeta _x\) and \(\sigma \) commute, and on on \({{\mathfrak {g}}}^{-\sigma _x}\) the complex structure \(\zeta _x\) anticommute with \(\sigma \). In particular, we have

    $$\begin{aligned} \zeta _x({{\mathfrak {g}}}^{\sigma ,\sigma _x}) = {{\mathfrak {g}}}^{\sigma ,\sigma _x}, \quad \zeta _x({{\mathfrak {g}}}^{-\sigma ,\sigma _x}) = {{\mathfrak {g}}}^{-\sigma , \sigma _x}, \quad \zeta _x({{\mathfrak {g}}}^{\pm \sigma , -\sigma _x}) = {{\mathfrak {g}}}^{{\mp }\sigma , -\sigma _x}.\end{aligned}$$
  6. (f)

    The stabilizer Lie algebra of m in \({{\mathfrak {g}}}\) is

    $$\begin{aligned} {{\mathfrak {g}}}_m = {{\mathfrak {g}}}^{\tau \sigma _x}= \{ y \in {{\mathfrak {g}}}: \sigma _x(y) = \tau (y)\} \quad \text{ and } \quad {{\mathfrak {g}}}_m^{\sigma _x^2} = {{\mathfrak {h}}}^{\sigma _x} \oplus {{\mathfrak {q}}}^{-\sigma _x}.\end{aligned}$$

    The stabilizer Lie algebra in \({{\mathfrak {g}}}^\sigma \) is

    $$\begin{aligned} {{\mathfrak {g}}}_m^\sigma ={{\mathfrak {h}}}^{\sigma ,\sigma _x} \oplus {{\mathfrak {q}}}^{\sigma , -\sigma _x}. \end{aligned}$$
    (C.7)

    The stabilizer group \(G_m\) acts on

    $$\begin{aligned} T_m(M) = \exp x.{{\mathfrak {q}}}\, by \, g.(\exp x.y) = \exp x.\big (\mathop {\mathrm{Ad}}\nolimits (\zeta _x^G(g))y\big ). \end{aligned}$$
  7. (g)

    The tangent space of the orbit \({\mathcal {O}}_m\) is

    $$\begin{aligned} T_m({\mathcal {O}}_m) = \exp (x).\big (\cosh (\mathop {\mathrm{ad}}\nolimits x){{\mathfrak {q}}}^\sigma + [x,{{\mathfrak {h}}}^\sigma ]\big ).\end{aligned}$$
  8. (h)

    Let \({{\mathfrak {q}}}_x := {{\mathfrak {q}}}^{\sigma ,-\sigma _x} + {{\mathfrak {q}}}^{-\sigma ,\sigma _x}\) and \({{\mathfrak {h}}}_x := [{{\mathfrak {q}}}_x,{{\mathfrak {q}}}_x]\). If the group \(\mathop {\mathrm{Inn}}\nolimits _{{\mathfrak {g}}}({{\mathfrak {h}}}_x)\) acts as a relatively compact group on \({{\mathfrak {q}}}_x\) and \(\mathop {\mathrm{Inn}}\nolimits _{{\mathfrak {g}}}({{\mathfrak {h}}}_x){{\mathfrak {q}}}^{-\sigma ,\sigma _x} = {{\mathfrak {q}}}_x\), then m is an interior point of \(\mathop {\mathrm{im}}\nolimits (\Phi )\).

Proof

(a) follows directly from \(2\cosh (\mathop {\mathrm{ad}}\nolimits x) = e^{\mathop {\mathrm{ad}}\nolimits x} + e^{-\mathop {\mathrm{ad}}\nolimits x}\).

(b) With (C.3) we see that the image of \(T_{[e,x]}(\Phi )\) is the subspace

$$\begin{aligned} \exp (x).\Big (\underbrace{\cosh (\mathop {\mathrm{ad}}\nolimits x){{\mathfrak {q}}}^{\sigma }}_{\subseteq {{\mathfrak {q}}}^{\sigma }} + \underbrace{\frac{\sinh (\mathop {\mathrm{ad}}\nolimits x)}{\mathop {\mathrm{ad}}\nolimits x} {{\mathfrak {q}}}^{-\sigma }}_{\subseteq {{\mathfrak {q}}}^{-\sigma }}\Big ).\end{aligned}$$

As \(\mathop {\mathrm{ad}}\nolimits x\) is semisimple, a complement of \(\exp (-x).\mathop {\mathrm{im}}\nolimits (T_{[e,x]}(\Phi ))\) in \({{\mathfrak {q}}}\) is

$$\begin{aligned} \ker \big (\cosh (\mathop {\mathrm{ad}}\nolimits x)\vert _{{{\mathfrak {q}}}^{\sigma }}\big ) = {{\mathfrak {q}}}^{\sigma , -\sigma _x}. \end{aligned}$$
(C.8)

(c) In view of \(\tau \sigma _x^2 \tau = \sigma _x^{-2}\), the fixed point space \({{\mathfrak {g}}}^{\sigma _x^2}\) is \(\tau \)-invariant. On this subspace \(\sigma _x= \sigma _x^{-1} = \tau \sigma _x \tau \), so that \(\tau \) preserves the two eigenspaces \({{\mathfrak {g}}}^{\pm \sigma _x}\) of \(\sigma _x\) on \({{\mathfrak {g}}}^{\sigma _x^2}\).

(d) As \(\zeta _x\) commutes with \(\sigma _x\), the eigenspaces \({{\mathfrak {g}}}^{\pm \sigma _x}\) are \(\zeta _x\)-invariant. Further, (c) implies that, on \({{\mathfrak {g}}}^{\sigma _x}\), we have \(\tau \zeta _x \tau = \zeta _x^{-1} = \zeta _x\).

(e) is shown with similar arguments as (c) and (d).

(f) In \(M = G/H\), the point \(m = \mathop {\mathrm{Exp}}\nolimits _{eH}(\exp x) = \exp x H\) is obtained by acting with \(\exp x\) on the base point eH. Therefore its stabilizer group is \(G_m = \exp x H \exp (-x)\) with the Lie algebra

$$\begin{aligned} {{\mathfrak {g}}}_m = e^{\mathop {\mathrm{ad}}\nolimits x} {{\mathfrak {h}}}= \mathop {\mathrm{Fix}}\nolimits (e^{\mathop {\mathrm{ad}}\nolimits x} \tau e^{-\mathop {\mathrm{ad}}\nolimits x}) = \mathop {\mathrm{Fix}}\nolimits (\tau e^{-2\mathop {\mathrm{ad}}\nolimits x}) = \mathop {\mathrm{Fix}}\nolimits (\tau \sigma _x).\end{aligned}$$

Now the \(\tau \)-invariance of \({{\mathfrak {g}}}^{\sigma _x^2}\) implies that

$$\begin{aligned} {{\mathfrak {g}}}^{\sigma _x^2}_m = {{\mathfrak {g}}}_m \cap {{\mathfrak {g}}}^{\sigma _x^2} = {{\mathfrak {g}}}^{\sigma _x,\tau } \oplus {{\mathfrak {g}}}^{-\sigma _x,-\tau } = {{\mathfrak {h}}}^{\sigma _x} \oplus {{\mathfrak {q}}}^{-\sigma _x}.\end{aligned}$$

To verify (C.7), let \(y = y_{{\mathfrak {h}}}+ y_{{\mathfrak {q}}}\in {{\mathfrak {g}}}^\sigma \) with \(y_{{\mathfrak {h}}}\in {{\mathfrak {h}}}^\sigma \) and \(y_{{\mathfrak {q}}}\in {{\mathfrak {q}}}^\sigma \). Then the corresponding vector field \(Y_M\) on M satisfies

$$\begin{aligned} Y_M(m)&= y.m = \exp (x).p_{{{\mathfrak {q}}}}(e^{-\mathop {\mathrm{ad}}\nolimits x}y) = \exp (x).\Big (\frac{1}{2}(e^{-\mathop {\mathrm{ad}}\nolimits x}y - \tau (e^{-\mathop {\mathrm{ad}}\nolimits x}y)\Big ) \nonumber \\&= \exp (x).\Big (\frac{1}{2}(e^{-\mathop {\mathrm{ad}}\nolimits x}y - e^{\mathop {\mathrm{ad}}\nolimits x}\tau (y))\Big ) = \exp (x).\big (\cosh (\mathop {\mathrm{ad}}\nolimits x)y_{{\mathfrak {q}}}- \sinh (\mathop {\mathrm{ad}}\nolimits x)y_{{\mathfrak {h}}}\big ). \end{aligned}$$
(C.9)

Therefore \(Y_M(m) = 0\) is equivalent to

$$\begin{aligned} 0 = \underbrace{\cosh (\mathop {\mathrm{ad}}\nolimits x)y_{{\mathfrak {q}}}}_{\in {{\mathfrak {q}}}^{\sigma }} - \underbrace{\sinh (\mathop {\mathrm{ad}}\nolimits x) y_{{\mathfrak {h}}}}_{\in {{\mathfrak {q}}}^{-\sigma }}.\end{aligned}$$

Thus both summands have to vanish, which is equivalent to

$$\begin{aligned} e^{-2\mathop {\mathrm{ad}}\nolimits x} y_{{\mathfrak {q}}}= - y_{{\mathfrak {q}}}\quad \text{ and } \quad e^{-2\mathop {\mathrm{ad}}\nolimits x} y_{{\mathfrak {h}}}= y_{{\mathfrak {h}}}.\end{aligned}$$

This implies (C.7).

To complete the proof of (f), we note that, for \(v \in {{\mathfrak {q}}}\) and \(g \in G_m\), we have

$$\begin{aligned} g.(\exp (x).y) = \exp (x).\big (\mathop {\mathrm{Ad}}\nolimits (\zeta ^G_x(g))y\big ), \end{aligned}$$
(C.10)

where \(\zeta ^G_x(G_m) = H\) acts on \(T_{eH}(M) \cong {{\mathfrak {q}}}\) by the adjoint representation.

(g) From (C.9) it follows that

$$\begin{aligned} \exp (-x).T_m({\mathcal {O}}_m)&= \cosh (\mathop {\mathrm{ad}}\nolimits x) {{\mathfrak {q}}}^\sigma + \sinh (\mathop {\mathrm{ad}}\nolimits x) {{\mathfrak {h}}}^\sigma = \cosh (\mathop {\mathrm{ad}}\nolimits x) {{\mathfrak {q}}}^\sigma + \frac{\sinh (\mathop {\mathrm{ad}}\nolimits x)}{\mathop {\mathrm{ad}}\nolimits x} [x,{{\mathfrak {h}}}^\sigma ]\\&{\buildrel ! \over =} \cosh (\mathop {\mathrm{ad}}\nolimits x) {{\mathfrak {q}}}^\sigma + [x,{{\mathfrak {h}}}^\sigma ]. \end{aligned}$$

Here the last equality follows from \((\mathop {\mathrm{ad}}\nolimits x)^2 {{\mathfrak {h}}}^\sigma \subseteq {{\mathfrak {h}}}^\sigma \) and the \(\mathop {\mathrm{Exp}}\nolimits \)-regularity of x, which, by Lemma C.3, is equivalent to the invertibility of \(\frac{\sinh (\mathop {\mathrm{ad}}\nolimits x}{\mathop {\mathrm{ad}}\nolimits x}\) on \({{\mathfrak {q}}}\).

(h) By (C.8), a natural complement of

$$\begin{aligned} \exp (-x).T_m({\mathcal {O}}_m) = \cosh (\mathop {\mathrm{ad}}\nolimits x) {{\mathfrak {q}}}^\sigma + [x,{{\mathfrak {h}}}^\sigma ] \end{aligned}$$

in \({{\mathfrak {q}}}\) is the subspace

$$\begin{aligned} {{\mathfrak {q}}}_x := {{\mathfrak {q}}}^{\sigma ,-\sigma _x} \oplus ({{\mathfrak {q}}}^{-\sigma } \cap \ker (\mathop {\mathrm{ad}}\nolimits x)) = {{\mathfrak {q}}}^{\sigma ,-\sigma _x} \oplus {{\mathfrak {q}}}^{-\sigma , \sigma _x},\end{aligned}$$

where the last equality follows from

$$\begin{aligned} {{\mathfrak {q}}}^{\sigma _x} = \bigoplus _{n \in {{\mathbb {Z}}}} \ker ( (\mathop {\mathrm{ad}}\nolimits x)^2 + \pi ^2 n^2 \mathop {\mathrm{id}}\nolimits _{{\mathfrak {q}}}) \quad \text{ and } \quad \mathrm{Spec}(\mathop {\mathrm{ad}}\nolimits x) \cap {{\mathbb {Z}}}\pi i \subseteq \{0\}.\end{aligned}$$

As x is \(\mathop {\mathrm{Exp}}\nolimits _{eH}\)-regular and

$$\begin{aligned} \mathop {\mathrm{Exp}}\nolimits _{eH}(x + y) = \exp (x).\mathop {\mathrm{Exp}}\nolimits _m(y) \quad \text{ for } \quad y \in {{\mathfrak {q}}}^{-\sigma } \cap \ker (\mathop {\mathrm{ad}}\nolimits x) = {{\mathfrak {q}}}^{-\sigma ,\sigma _x},\end{aligned}$$

the subset

$$\begin{aligned} \Omega := \{ y \in {{\mathfrak {q}}}_x : \mathop {\mathrm{Exp}}\nolimits _m(\exp (x).y) \in \mathop {\mathrm{im}}\nolimits (\Phi ) \} \end{aligned}$$

contains a 0-neighborhood \(U_0\) in \({{\mathfrak {q}}}^{-\sigma ,\sigma _x}\). Now our assumption \(\mathop {\mathrm{Inn}}\nolimits _{{\mathfrak {g}}}({{\mathfrak {h}}}_x) {{\mathfrak {q}}}^{-\sigma ,\sigma _x} ={{\mathfrak {q}}}_x\) and the relative compactness of the group \(\mathop {\mathrm{Inn}}\nolimits _{{\mathfrak {g}}}({{\mathfrak {h}}}_x)\) on \({{\mathfrak {q}}}_x\) imply that \(U_1 := \mathop {\mathrm{Inn}}\nolimits _{{\mathfrak {g}}}({{\mathfrak {h}}}_x) U_0\) is a 0-neighborhood in \({{\mathfrak {q}}}_x\).

Finally, we obtain from (C.7)

$$\begin{aligned} \zeta _x^{-1}({{\mathfrak {h}}}_x) \subseteq \zeta _x^{-1}({{\mathfrak {h}}}^{\sigma ,\sigma _x} + {{\mathfrak {h}}}^{-\sigma ,-\sigma _x}) \ {\buildrel {(d),(e)} \over \subseteq }\ {{\mathfrak {h}}}^{\sigma , \sigma _x} + {{\mathfrak {q}}}^{\sigma , -\sigma _x} \ {\buildrel (C.7) \over =}\ {{\mathfrak {g}}}_m^\sigma ,\end{aligned}$$

so that

$$\begin{aligned} \mathop {\mathrm{Exp}}\nolimits _m(\exp (x).U_1) = \mathop {\mathrm{Exp}}\nolimits _m(\exp (x).\mathop {\mathrm{Inn}}\nolimits _{{\mathfrak {g}}}({{\mathfrak {h}}}_x) U_0) = (G^\sigma _e)_m.\mathop {\mathrm{Exp}}\nolimits _m(\exp (x).U_0) \subseteq \mathop {\mathrm{im}}\nolimits (\Phi ).\end{aligned}$$

This means that \(\Omega \) is a 0-neighborhood in \({{\mathfrak {q}}}_x\), and hence that \(\mathop {\mathrm{im}}\nolimits (\Phi )\) is a neighborhood of m because the map

$$\begin{aligned} G^\sigma _e \times {{\mathfrak {q}}}_x \rightarrow M, \quad (g,y) \mapsto g.\mathop {\mathrm{Exp}}\nolimits _m(\exp (x).y) \end{aligned}$$

has surjective differential in (e, 0). \(\square \)

Remark C.5

If \(\rho _i(\mathop {\mathrm{ad}}\nolimits x) < \pi /2\), then the polar map \(\Phi \) in the preceding lemma is regular in [ex] (Lemma C.3). In this case \(\ker (\cosh (\mathop {\mathrm{ad}}\nolimits x)) = {{\mathfrak {g}}}^{-\sigma _x}\) is trivial, so that

$$\begin{aligned} {{\mathfrak {g}}}_m^\sigma = {{\mathfrak {h}}}^{\sigma ,\sigma _x} = {{\mathfrak {h}}}^\sigma \cap \ker (\mathop {\mathrm{ad}}\nolimits x) = {{\mathfrak {z}}}_{{{\mathfrak {h}}}^\sigma }(x)\end{aligned}$$

follows from Lemma C.4(f).

The next lemma is useful to verify condition (h) in the preceding lemma.

Lemma C.6

Suppose that \(({{\mathfrak {g}}},\tau )\) is a reductive symmetric Lie algebra such that \({{\mathfrak {q}}}\) consist of elliptic elements, \(\sigma \tau = \tau \sigma \), and that \({{\mathfrak {a}}}\subseteq {{\mathfrak {q}}}^{-\sigma }\) is a maximal abelian subspace in \({{\mathfrak {q}}}\). Then

$$\begin{aligned} {{\mathfrak {q}}}^\sigma =[{{\mathfrak {h}}}^{-\sigma }, {{\mathfrak {q}}}^{-\sigma }], \qquad \mathop {\mathrm{Inn}}\nolimits _{{\mathfrak {g}}}([{{\mathfrak {q}}},{{\mathfrak {q}}}]){{\mathfrak {q}}}^{-\sigma } = {{\mathfrak {q}}},\end{aligned}$$

and the group \(\mathop {\mathrm{Inn}}\nolimits _{{\mathfrak {g}}}([{{\mathfrak {q}}},{{\mathfrak {q}}}])\) is compact.

Proof

Let \({{\mathfrak {a}}}\subset {{\mathfrak {q}}}\) be maximal abelian in \({{\mathfrak {q}}}\). Then \({{\mathfrak {a}}}\) contains \({{\mathfrak {z}}}({{\mathfrak {q}}})\). We then note that \({{\mathfrak {z}}}_{{\mathfrak {g}}}({{\mathfrak {a}}}) = {{\mathfrak {a}}}\oplus {{\mathfrak {z}}}_{{\mathfrak {h}}}({{\mathfrak {a}}})\) is \(\tau \)- and \(\sigma \)-invariant. Further \({{\mathfrak {g}}}= {{\mathfrak {z}}}_{{\mathfrak {g}}}({{\mathfrak {a}}}) \oplus [{{\mathfrak {a}}},{{\mathfrak {g}}}]\) implies that \({{\mathfrak {q}}}= {{\mathfrak {a}}}\oplus [{{\mathfrak {a}}},{{\mathfrak {h}}}].\) As

$$\begin{aligned}{}[{{\mathfrak {a}}},{{\mathfrak {h}}}] = [{{\mathfrak {a}}}, {{\mathfrak {h}}}^\sigma \oplus {{\mathfrak {h}}}^{-\sigma }] \subseteq {{\mathfrak {q}}}^{-\sigma } \oplus {{\mathfrak {q}}}^{\sigma },\end{aligned}$$

this leads to

$$\begin{aligned} {{\mathfrak {q}}}^{\sigma } = [{{\mathfrak {a}}}, {{\mathfrak {h}}}^{-\sigma }] \subseteq [{{\mathfrak {q}}}^{-\sigma }, {{\mathfrak {h}}}^{-\sigma }] \subseteq {{\mathfrak {q}}}^\sigma .\end{aligned}$$

The second assertion follows from \(\mathop {\mathrm{Inn}}\nolimits _{{\mathfrak {g}}}([{{\mathfrak {q}}},{{\mathfrak {q}}}]){{\mathfrak {a}}}= {{\mathfrak {q}}}\), and the third from the fact that the reductive Lie algebra \({{\mathfrak {q}}}+ [{{\mathfrak {q}}},{{\mathfrak {q}}}]\) (it is an ideal of \({{\mathfrak {g}}}\)) is compact. \(\square \)

1.3 C.3 Fibers of the polar map

For the polar map we have to analyze the relation

$$\begin{aligned} \mathop {\mathrm{Exp}}\nolimits (x) = g.\mathop {\mathrm{Exp}}\nolimits (y).\end{aligned}$$

Applying the quadratic representation yields

$$\begin{aligned} \exp (2x) = g \exp (2y) g^\sharp = \exp (2 \mathop {\mathrm{Ad}}\nolimits (g)y) gg^\sharp .\end{aligned}$$

For \(x,y \in {{\mathfrak {q}}}^{-\sigma }\) and \(g \in G^\sigma \), we also have \(gg^\sharp \in G^\sigma \), so that

$$\begin{aligned} \exp (4x) = \exp (2x) \sigma (\exp (2x))^{-1} = \exp (4 \mathop {\mathrm{Ad}}\nolimits (g)y).\end{aligned}$$

If x and y are sufficiently small (imaginary spectral radius \(< \frac{\pi }{4}\)), we thus obtain \(\mathop {\mathrm{Ad}}\nolimits (g)y = x\), and thus \(gg^\sharp = e\), i.e., \(\tau (g) = g\).

See [16, Cor. 7.35] for similar arguments.

1.4 C.4 Fibers of \(\mathop {\mathrm{Exp}}\nolimits \)

Suppose that \(x, y \in {{\mathfrak {q}}}\) have the same exponential image \(\mathop {\mathrm{Exp}}\nolimits (x) = \mathop {\mathrm{Exp}}\nolimits (y)\) in M. We further assume that \(\mathrm{Spec}(\mathop {\mathrm{ad}}\nolimits x) \cap i \pi {{\mathbb {Z}}}\subseteq \{0\}\), so that \(\mathop {\mathrm{Exp}}\nolimits \) is regular in x. Then we obtain in G the identity

$$\begin{aligned} \exp (2x) = Q(\mathop {\mathrm{Exp}}\nolimits x) = Q(\mathop {\mathrm{Exp}}\nolimits y) = \exp (2y),\end{aligned}$$

and since \(\mathrm{Spec}(\mathop {\mathrm{ad}}\nolimits (2x)) \cap 2 \pi i {{\mathbb {Z}}}\subseteq \{0\}\), \(\exp \) is regular in x. Therefore [17, Lemma 9.2.31] implies that

$$\begin{aligned}{}[x,y]= 0 \quad \text{ and } \quad \exp (2x - 2y) = e.\end{aligned}$$

We conclude that \(\exp (x-y) = \exp (y-x)\), which leads to

$$\begin{aligned} \mathop {\mathrm{Exp}}\nolimits (y-x) = \tau _M(\mathop {\mathrm{Exp}}\nolimits (x-y)) = \exp (y-x)H = \exp (x-y)H = \mathop {\mathrm{Exp}}\nolimits (x-y),\end{aligned}$$

so that \(\mathop {\mathrm{Exp}}\nolimits (y-x) \in M^\tau \), and \(\mathop {\mathrm{Exp}}\nolimits ({{\mathbb {R}}}(x-y)) \subseteq M\) is a closed geodesic.

We also conclude that \(x-y\) is elliptic with \(\mathrm{Spec}(\mathop {\mathrm{ad}}\nolimits (x-y))\subseteq \pi i {{\mathbb {Z}}}\).

Lemma C.7

If \(\rho _i(\mathop {\mathrm{ad}}\nolimits x), \rho _i(\mathop {\mathrm{ad}}\nolimits y) < \pi ,\) then \(\exp (x) = \exp (y)\) implies \(x-y \in {{\mathfrak {z}}}({{\mathfrak {g}}})\).

If, in addition, G is simply connected or \({{\mathfrak {g}}}\) is semisimple, then \(x = y\).

Proof

The preceding discussion implies that \([x,y] = 0\). Now \(\rho _i(\mathop {\mathrm{ad}}\nolimits (x-y)) < 2\pi \) leads to \(\mathop {\mathrm{ad}}\nolimits (x-y) = 0\), i.e., to \(x-y \in {{\mathfrak {z}}}({{\mathfrak {g}}})\). \(\square \)

D Quadrics as symmetric spaces

In this appendix we discuss an important example of a non-compactly causal symmetric spaces: d-dimensional de Sitter space \(\mathop {\mathrm{dS}}\nolimits ^d\), realized as a hyperboloid in Minkowski space.

1.1 D.1 Quadrics as symmetric spaces

Definition D.1

[24] (a) Let M be a smooth manifold and

$$\begin{aligned} \mu : M \times M \rightarrow M, \quad (x,y) \mapsto x \cdot y =: s_x(y) \end{aligned}$$

be a smooth map with the following properties: each \(s_x\) is an involution for which x is an isolated fixed point and

$$\begin{aligned} s_x(y \cdot z) = s_x(y)\cdot s_x(z) \quad \text{ for } \text{ all } \quad x,y \in M. \end{aligned}$$
(D.1)

Then we call \((M,\mu )\) a symmetric space.

(b) A morphism of symmetric spaces M and N is a smooth map \(\varphi : M\rightarrow N\) such that \(\varphi (x\cdot y)= \varphi (x)\cdot \varphi (y)\) for \(x, y \in M\).

(c) A geodesic of a symmetric space is a morphism \(\gamma : {{\mathbb {R}}}\rightarrow M\) of symmetric spaces, i.e.,

$$\begin{aligned} \gamma (2x-y) = \gamma (x) \cdot \gamma (y) \quad \text{ for } \quad x,y \in {{\mathbb {R}}}.\end{aligned}$$

Any geodesic is uniquely determined by \(\gamma '(0) \in T_{\gamma (0)}(M)\), and, conversely, every \(v \in T_p(M)\) generates a unique geodesic \(\gamma _v\) with \(\gamma _v(0) = p\) and \(\gamma _v'(0) = v\). Accordingly, geodesics are encoded in the exponential functions

$$\begin{aligned} \mathop {\mathrm{Exp}}\nolimits _p : T_p(M) \rightarrow M, \quad \mathop {\mathrm{Exp}}\nolimits _p(v) := \gamma _v(1).\end{aligned}$$

We then have

$$\begin{aligned} \gamma _v(t) = \mathop {\mathrm{Exp}}\nolimits _p(tv) \quad \text{ for } \quad t \in {{\mathbb {R}}}.\end{aligned}$$

Example D.2

Let \((V,\beta )\) be a finite dimensional real vector space, endowed with a non-degenerate symmetric bilinear form \(\beta \). Then every anisotropic element \(x \in V\) defines an involution

$$\begin{aligned} s_x(y) := -y + 2 \frac{\beta (x,y)}{\beta (x,x)} x \end{aligned}$$

fixing x and satisfying \(\mathop {\mathrm{Fix}}\nolimits (-s_x) = x^\bot = \{ y \in V: \beta (x,y) = 0\}\).

For \(c \in {{\mathbb {R}}}^\times \), let

$$\begin{aligned} Q_c := Q_c(V,\beta ) := \{ v \in V : \beta (v,v) = c\} \end{aligned}$$

denote the corresponding quadric in \((V,\beta )\). Then \((Q_c, \mu )\) with \(\mu (x,y) = s_x(y)\) is a symmetric space (Definition D.1) and \(\mathop {\mathrm{dim}}\nolimits Q_c = \mathop {\mathrm{dim}}\nolimits V -1\). Note that dilation by \(r \in {{\mathbb {R}}}^\times \) is an isomorphism of symmetric spaces from \(Q_c\) to \(Q_{r^2c}\).

For \(p \in Q_c\), the tangent space is \(T_p(Q_c) = p^\bot = \{ v \in V : \beta (p,v) = 0\}\). To describe the exponential function of the symmetric space \(Q_c\), we use the entire functions \(C, S : {{\mathbb {C}}}\rightarrow {{\mathbb {C}}}\) defined by

$$\begin{aligned} C(z) := \sum _{k = 0}^\infty \frac{(-1)^k}{(2k)!} z^{k} \quad \text{ and } \quad S(z) := \sum _{k = 0}^\infty \frac{(-1)^k}{(2k+1)!} z^{k} \end{aligned}$$
(D.2)

which satisfy

$$\begin{aligned} \cos z = C(z^2) \quad \text{ and } \quad \sin z = z S(z^2) \quad \text{ for } \quad z \in {{\mathbb {C}}}\end{aligned}$$
(D.3)

and

$$\begin{aligned} \cosh z = C(-z^2) \quad \text{ and } \quad \sinh z = z S(-z^2) \quad \text{ for } \quad z \in {{\mathbb {C}}}. \end{aligned}$$
(D.4)

Note that

$$\begin{aligned} 1 = C(z)^2 + z S(z)^2 \quad \text{ for } \quad z \in {{\mathbb {C}}}\end{aligned}$$
(D.5)

follows from \(1 = \cos ^2 z + \sin ^2 z\) and the surjectivity of the square map on \({{\mathbb {C}}}\).

Proposition D.3

The exponential function of the symmetric space \(Q_c\) is given by

$$\begin{aligned} \mathop {\mathrm{Exp}}\nolimits _p(v) = C\Big (\frac{\beta (v,v)}{\beta (p,p)}\Big ) p + S\Big (\frac{\beta (v,v)}{\beta (p,p)}\Big ) v \quad \text{ for } \quad p \in Q_c, v \in T_p(Q_c) = p^\bot . \end{aligned}$$
(D.6)

Proof

To verify this claim, abbreviate \(\varepsilon := \frac{\beta (v,v)}{\beta (p,p)}\). By (D.5), on the right hand side of (D.6), the quadratic form \(\beta (\cdot ,\cdot )\) has the value

$$\begin{aligned} C(\varepsilon )^2 \beta (p,p) + S(\varepsilon )^2 \beta (v,v) = \big (C(\varepsilon )^2 + S(\varepsilon )^2 \varepsilon \big ) \beta (p,p) = \beta (p,p) = c.\end{aligned}$$

Therefore the right hand side of (D.6) is contained in \(Q_c\).

It remains to show that, for \(\varepsilon \in \{-1,0,1\}\), the curve

$$\begin{aligned} \gamma _v(t) := C(t^2\varepsilon ) p + S(t^2\varepsilon ) t v, \end{aligned}$$

which satisfies \(\gamma _v'(0) = v\), actually is a geodesic, i.e.,

$$\begin{aligned} \gamma _v(2t-s) = \gamma _v(t) \cdot \gamma _v(s)\quad \text{ for } \quad t,s \in {{\mathbb {R}}}.\end{aligned}$$

We consider three cases:

\(\varepsilon = 0\): Then \(\gamma _v(t) = p + tv\) and

$$\begin{aligned} \gamma _v(t) \cdot \gamma _v(s) = - \gamma _v(s) + \frac{2}{c} \beta (\gamma _v(t),\gamma _v(s)) \gamma _v(t) = - (p + sv) + 2(p + tv) = p + (2t-s)v.\end{aligned}$$

\(\varepsilon = 1\): Then

$$\begin{aligned} \gamma _v(t) = \cos (t) p + \sin (t) v\end{aligned}$$

by (D.3), and

$$\begin{aligned} \gamma _v(t) \cdot \gamma _v(s)&= - \gamma _v(s) + \frac{2}{c} \beta (\gamma _v(t),\gamma _v(s)) \gamma _v(t) \\&= - (\cos (s) p + \sin (s) v) + 2(\cos (t)\cos (s) + \sin (t) \sin (s)) (\cos (t)p + \sin (t)v) \\&= (-\cos (s) + 2\cos (t)^2\cos (s) + 2\sin (t)\sin (s)\cos (t)) p \\&\quad + (-\sin (s) + 2\sin (t)\cos (t)\cos (s) +2 \sin (t)^2\sin (s)) v \\&= (\cos (s) - 2\sin (t)^2\cos (s) + 2\sin (t)\sin (s)\cos (t)) p \\&\quad + (-\sin (s) + 2\sin (t)\cos (t)\cos (s) +2 \sin (t)^2\sin (s)) v \\&= \cos (2t-s) p + \sin (2t-s) v = \gamma _v(2t-s). \end{aligned}$$

\(\varepsilon = -1\): Then

$$\begin{aligned} \gamma _v(t) = \cosh (t) p + \sinh (t) v\end{aligned}$$

by (D.4), and

$$\begin{aligned} \gamma _v(t) \cdot \gamma _v(s)&= - \gamma _v(s) + \frac{2}{c} \beta (\gamma _v(t),\gamma _v(s)) \gamma _v(t) \\&= - (\cosh (s) p + \sinh (s) v) + 2(\cosh (t)\cosh (s) \\&\quad + \sinh (t) \sinh (s)) (\cosh (t)p + \sinh (t)v) \\&= (-\cosh (s) + 2\cosh (t)^2\cosh (s)\\&\quad + 2\sinh (t)\sinh (s)\cosh (t)) p \\&\quad + (-\sinh (s) + 2\sinh (t)\cosh (t)\cosh (s) +2 \sinh (t)^2\sinh (s)) v \\&= (\cosh (s) +2\sinh (t)^2\cosh (s) + 2\sinh (t)\sinh (s)\cosh (t)) p \\&\quad + (-\sinh (s) + 2\sinh (t)\cosh (t)\cosh (s) +2 \sinh (t)^2\sinh (s)) v \\&= \cosh (2t-s) p + \sinh (2t-s) v = \gamma _v(2t-s). \end{aligned}$$

\(\square \)

1.2 D.2 Minkowski space

On \(V := {{\mathbb {R}}}^{1,d}\), we consider the Lorentzian form

$$\begin{aligned}{}[x,y] := x_0 y_0 - {\mathbf{x}}{\mathbf{y}},\end{aligned}$$

the future light cone

$$\begin{aligned} V_+ := \{ x = (x_0, {\mathbf{x}}) \in V : x_0> 0, x_0^2 - {\mathbf{x}}^2 > 0\}, \end{aligned}$$

and the tube domain

$$\begin{aligned} {\mathcal {T}}_V = V + i V_+ \subseteq V_{{\mathbb {C}}}.\end{aligned}$$

The standard boost vector field \(X_h(v) = hv\) is defined by \(h \in \mathop {{\mathfrak {so}}}\nolimits _{1,d}({{\mathbb {R}}})\), given by

$$\begin{aligned} hx = (x_1, x_0, 0,\ldots , 0). \end{aligned}$$
(D.7)

It generates the flow

$$\begin{aligned} \alpha _t(x) = e^{th} x = (\cosh t \cdot x_0 + \sinh t \cdot x_1, \cosh t \cdot x_1 + \sinh t \cdot x_0, x_2, \ldots , x_d)\end{aligned}$$

and defines the involution

$$\begin{aligned}\tau _h(x) := e^{\pi i h} x= (-x_0, -x_1, x_2, \ldots , x_d), \end{aligned}$$

that we extend to an antilinear involution \(\overline{\tau }_h\) on \(V_{{\mathbb {C}}}\). It also defines a Wick rotation

$$\begin{aligned} \kappa _h(x) = e^{-\frac{\pi i}{2} h} x = (-i x_1, -i x_0, x_2, \ldots , x_d) \end{aligned}$$

satisfying \(\kappa _h^2 = \tau _h\).

Lemma D.4

The following subsets of V are equal:

  1. (a)

    The standard right wedge \(W_R := \{ x \in V : x_1 > |x_0| \}\).

  2. (b)

    The positivity domain \(W_V^+(h) := \{ v \in V : X_h(v) \in V_+ \}\) of \(X_h(v) = hv\).

  3. (c)

    The KMS domain of \(\alpha \): \(\{ x \in V : (\forall t \in (0,\pi ))\ \alpha _{it}(x) \in {\mathcal {T}}_V\}\).

  4. (d)

    The Wick rotation \(\kappa _h({\mathcal {T}}_V^{\overline{\tau }_h})\) of the fixed point set \({\mathcal {T}}_V^{\overline{\tau }_h} = {\mathcal {T}}_V \cap (V^{\tau _h} + i V^{-\tau _h})\) of the antiholomorphic involution \(\overline{\tau }_h\) on \({\mathcal {T}}_V\).

Proof

The equality of \(W_R\) and \(W_V^+(h)\) follows immediately from (D.7). For \(0< t < \pi \), we have

$$\begin{aligned} \mathop {\mathrm{Im}}\nolimits (\alpha _{it}(x))= & {} \mathop {\mathrm{Im}}\nolimits (\cos t \cdot x_0 + \sin t \cdot i x_1, \cos t \cdot x_1 + \sin t \cdot i x_0, x_2, \ldots , x_d)\\= & {} \sin t\cdot (x_1, x_0, 0,\ldots ,0).\end{aligned}$$

This implies the equality of the sets in (a) and (c). Finally, we observe that

$$\begin{aligned} {\mathcal {T}}_V^{\overline{\tau }_h} = (V + i V_+)^{\overline{\tau }_h} = V^{\tau _h} + i V_+^{-\tau _h} =\{ (ix_0, i x_1, x_2, \ldots , x_d) \in V : x_0 > |x_1|\}\end{aligned}$$

implies that \(\kappa _h({\mathcal {T}}_V^{\overline{\tau }_h}) = W_R\). \(\square \)

Remark D.5

For the closed light cone \(C := \overline{V_+}\) and the h-eigenspaces

$$\begin{aligned} V_{\pm 1}(h) = {{\mathbb {R}}}({\mathbf{e}}_1 \pm {\mathbf{e}}_0), \quad \text{ we } \text{ put } \quad C_\pm := \pm C \cap V_{\pm 1}(h) = [0,\infty ) ({\mathbf{e}}_1 \pm {\mathbf{e}}_0),\end{aligned}$$

so that we obtain the following description of the standard right wedge

$$\begin{aligned} W_R = W_V^+(h) = V_0(h) + C_+^\circ + C_-^\circ = V_0(h) + {{\mathbb {R}}}_+ ({\mathbf{e}}_1 + {\mathbf{e}}_0) + {{\mathbb {R}}}_+ ({\mathbf{e}}_1 - {\mathbf{e}}_0).\end{aligned}$$

1.3 D.3 De Sitter space \(\mathop {\mathrm{dS}}\nolimits ^d, d \ge 2\)

We write \(G := \mathop {\mathrm{SO}}\nolimits _{1,d}({{\mathbb {R}}})^{\uparrow }\) for the connected Lorentz group acting on Minkowski space \((V,[\cdot ,\cdot ])\) and consider the de Sitter space

$$\begin{aligned} M := \mathop {\mathrm{dS}}\nolimits ^d := \{ x \in V : [x,x] = -1\} = G.{\mathbf{e}}_1.\end{aligned}$$

The purpose of this appendix is derive our main results for the example \(\mathop {\mathrm{dS}}\nolimits ^d\) by direct calculations without the elaborate structure theory developed for the general case. By Proposition D.3, the exponential function of the symmetric space \(\mathop {\mathrm{dS}}\nolimits ^d\) is given by

$$\begin{aligned} \mathop {\mathrm{Exp}}\nolimits _p(v) = C(-[v,v]) p + S(-[v,v]) v,\end{aligned}$$

so that spacelike vectors v generate closed geodesics.

De Sitter space inherits the structure of a non-compactly causal symmetric space from the embedding into Minkowski space \((V,V_+)\). In particular, the positive cone in a point \(x \in \mathop {\mathrm{dS}}\nolimits ^d\) is given by

$$\begin{aligned} V_+(x) := V_+ \cap T_x(\mathop {\mathrm{dS}}\nolimits ^d) = V_+ \cap x^\bot .\end{aligned}$$

We keep the notation h, \(\alpha _t\) and \(X_h\) from our discussion of Minkowski space in Subsection D.2. As \(\mathop {\mathrm{dS}}\nolimits ^d\) is \(\alpha \)-invariant, the vector field \(X_h\) is tangential to \(\mathop {\mathrm{dS}}\nolimits ^d\), and Lemma D.4 immediately implies that its positivity domain is given by

$$\begin{aligned} W_{\mathop {\mathrm{dS}}\nolimits ^d}^+(h) := \{ x \in \mathop {\mathrm{dS}}\nolimits ^d : X_h(x) \in V_+(x) \} = W_R \cap \mathop {\mathrm{dS}}\nolimits ^d = \{ x \in \mathop {\mathrm{dS}}\nolimits ^d : x_1 > |x_0|\}.\end{aligned}$$

The centralizer of h in G is the subgroup

$$\begin{aligned} G^h = \exp ({{\mathbb {R}}}h) \mathop {\mathrm{SO}}\nolimits _{d-1}({{\mathbb {R}}}) \cong \mathop {\mathrm{SO}}\nolimits _{1,1}({{\mathbb {R}}})^{\uparrow }\times \mathop {\mathrm{SO}}\nolimits _{d-1}({{\mathbb {R}}}) \end{aligned}$$
(D.8)

(cf. [31, Lemma 4.12]).

Lemma D.6

For

$$\begin{aligned} V_+^\pi ({\mathbf{e}}_1) := \{ x \in T_{{\mathbf{e}}_1}(\mathop {\mathrm{dS}}\nolimits ^d) \cong {\mathbf{e}}_1^\bot : x_0 > 0, 0< [x,x] < \pi ^2 \}, \end{aligned}$$

we have

$$\begin{aligned} (V + i V_+) \cap \mathop {\mathrm{dS}}\nolimits ^d_{{\mathbb {C}}}= \{ z \in V_{{\mathbb {C}}}: [z,z] = -1, \mathop {\mathrm{Im}}\nolimits z \in V_+\} = G.\mathop {\mathrm{Exp}}\nolimits _{{\mathbf{e}}_1}(iV_+^\pi ({\mathbf{e}}_1)). \end{aligned}$$
(D.9)

Proof

First we observe that both sides of (D.9) are G-invariant.

\(\supseteq \)”: We have to show that any \(x \in V_+^\pi ({\mathbf{e}}_1)\) satisfies \(\mathop {\mathrm{Exp}}\nolimits _{{\mathbf{e}}_1}(ix) \in V + i V_+\). Since the orbit of x under \(G_{{\mathbf{e}}_1}\) contains an element in \({{\mathbb {R}}}{\mathbf{e}}_0\), we may assume that \(x = x_0 {\mathbf{e}}_0\) with \(0< x_0 < \pi \). Then

$$\begin{aligned} \mathop {\mathrm{Exp}}\nolimits _{{\mathbf{e}}_1}(ix) = \mathop {\mathrm{Exp}}\nolimits _{{\mathbf{e}}_1}(ix_0 {\mathbf{e}}_0) = C(x_0^2) {\mathbf{e}}_1 + S(x_0^2) x_0 i {\mathbf{e}}_0 = \cos (x_0) {\mathbf{e}}_1 + \sin (x_0) i {\mathbf{e}}_0 \in V + i V_+ \end{aligned}$$

follows from \(\sin (x_0) > 0\).

\(\subseteq \)”: If \(z = x + i y\in \mathop {\mathrm{dS}}\nolimits ^d_{{\mathbb {C}}}\cap (V + i V_+)\), then \([z,z] = -1\) and \(y \in V_+\). Acting with G, we may thus assume that \(y = y_0 {\mathbf{e}}_0\) with \(y_0 > 0\). Then \(x \in y^\bot = {\mathbf{e}}_0^\bot \) follows from \(\mathop {\mathrm{Im}}\nolimits [z,z] = 0\), so that we may further assume that \(x = x_1 {\mathbf{e}}_1\) for some \(x_1 \ge 0\). Now \(z = i y_0 {\mathbf{e}}_0 + x_1 {\mathbf{e}}_1 \in \mathop {\mathrm{dS}}\nolimits ^d_{{\mathbb {C}}}\) implies that

$$\begin{aligned} -1 = [z,z] = - y_0^2 - x_1^2.\end{aligned}$$

Hence there exists a \(t \in (0,\pi )\) with \(y_0 = \sin t\) and \(x_1 = \cos t\). This leads to

$$\begin{aligned} \mathop {\mathrm{Exp}}\nolimits _{{\mathbf{e}}_1}(i t {\mathbf{e}}_0) = \cos (t) {\mathbf{e}}_1 + \sin (t) i {\mathbf{e}}_0 = x_1 {\mathbf{e}}_1 + y_0 i {\mathbf{e}}_0 = z. \end{aligned}$$

\(\square \)

Definition D.7

The complex manifold

$$\begin{aligned} {\mathcal {T}}_M = {\mathcal {T}}_{\mathop {\mathrm{dS}}\nolimits ^d} := G.\mathop {\mathrm{Exp}}\nolimits _{{\mathbf{e}}_1}(i V_+^\pi ({\mathbf{e}}_1)) = \mathop {\mathrm{dS}}\nolimits ^d_{{\mathbb {C}}}\cap (V + i V_+) \end{aligned}$$
(D.10)

is called the tube domain of \(\mathop {\mathrm{dS}}\nolimits ^d\). If coincides with \(G.\mathop {\mathrm{Exp}}\nolimits _m(i V_+^\pi (m))\) for every \(m \in \mathop {\mathrm{dS}}\nolimits ^d\).

Starting from the relation \({\mathcal {T}}_{\mathop {\mathrm{dS}}\nolimits ^d} = G.\mathop {\mathrm{Exp}}\nolimits _{{\mathbf{e}}_2}(i V_+^\pi ({\mathbf{e}}_2))\), the invariance of \({\mathbf{e}}_2\) under \(\tau _h = e^{\pi i h}\) permits us to obtain a nice description of the fixed point set of \(\overline{\tau }_h\) on \({\mathcal {T}}_{\mathop {\mathrm{dS}}\nolimits ^d}\).

Lemma D.8

\({\mathcal {T}}_{\mathop {\mathrm{dS}}\nolimits ^d}^{\overline{\tau }_h} = G^h.\mathop {\mathrm{Exp}}\nolimits _{{\mathbf{e}}_2}(i V_+^\pi ({\mathbf{e}}_2)^{-\tau _h}).\)

Proof

We clearly have

$$\begin{aligned} {\mathcal {T}}_{\mathop {\mathrm{dS}}\nolimits ^d}^{\overline{\tau }_h} \supseteq G^{\tau _h}.\mathop {\mathrm{Exp}}\nolimits _{{\mathbf{e}}_2}(i V_+^\pi ({\mathbf{e}}_2)^{-\tau _h}).\end{aligned}$$

We also note that

$$\begin{aligned} V_+^\pi ({\mathbf{e}}_2)^{-\tau _h}= & {} V_+^\pi ({\mathbf{e}}_2) \cap ({{\mathbb {R}}}{\mathbf{e}}_0 + {{\mathbb {R}}}{\mathbf{e}}_1)\\= & {} \{ x_0 {\mathbf{e}}_0 + x_1 {\mathbf{e}}_1 : x_0 > 0, 0< x_0^2 - x_1^2 < \pi ^2\}.\end{aligned}$$

Next we observe that

$$\begin{aligned} {\mathcal {T}}_{\mathop {\mathrm{dS}}\nolimits ^d}^{\overline{\tau }_h}= & {} (V + i V_+)^{\overline{\tau }_h} \cap \mathop {\mathrm{dS}}\nolimits ^d_{{\mathbb {C}}}= (V^{\tau _h} + i V_+^{-\tau _h}) \cap \mathop {\mathrm{dS}}\nolimits ^d_{{\mathbb {C}}}\\= & {} \{ (i x_0, i x_1, x_2, \ldots , x_d) \in \mathop {\mathrm{dS}}\nolimits ^d_{{\mathbb {C}}}: x_0 > |x_1|\}.\end{aligned}$$

Hence the elements \(z \in {\mathcal {T}}_{\mathop {\mathrm{dS}}\nolimits ^d}^{\overline{\tau }_h}\) are of the form

$$\begin{aligned} z = (ix_0, ix_1, x_2, \ldots , x_d), \quad x_0 > |x_1|, \ x_0^2 - x_1^2 + x_2^2 + \cdots + x_d^2 = 1.\end{aligned}$$

Therefore the orbit of z under \(G^h = \exp ({{\mathbb {R}}}h) \mathop {\mathrm{SO}}\nolimits _{d-1}({{\mathbb {R}}})\) (see (D.8)) contains an element of the form \(y = (iy_0, 0, y_2,0,\ldots ,0)\) with \(y_0 > 0\) and \(y_0^2 + y_2^2 = 1\). Hence there exists a \(t \in (0,\pi )\) with \(y_0 = \sin (t)\) and \(y_2 = \cos (t)\). Then

$$\begin{aligned} \mathop {\mathrm{Exp}}\nolimits _{{\mathbf{e}}_2} (it{\mathbf{e}}_0) = \cos (t) {\mathbf{e}}_2 + \sin (t) i {\mathbf{e}}_0 = y_2 {\mathbf{e}}_2 + i y_0 {\mathbf{e}}_0 = y.\end{aligned}$$

This implies that

$$\begin{aligned} {\mathcal {T}}_{\mathop {\mathrm{dS}}\nolimits ^d}^{\overline{\tau }_h} \subseteq G^h.\mathop {\mathrm{Exp}}\nolimits _{{\mathbf{e}}_2}(i (0,\pi ) {\mathbf{e}}_0) = G^h.\mathop {\mathrm{Exp}}\nolimits _{{\mathbf{e}}_2}(i V_+^\pi ({\mathbf{e}}_2)^{-\tau _h}). \end{aligned}$$

\(\square \)

Proposition D.9

The following subsets of de Sitter space \(M =\mathop {\mathrm{dS}}\nolimits ^d\) are equal:

  1. (a)

    \(W_R \cap \mathop {\mathrm{dS}}\nolimits ^d = \{ x \in \mathop {\mathrm{dS}}\nolimits ^d : x_1 > |x_0|\}\).

  2. (b)

    \(W_M^+(h) := \{ x \in \mathop {\mathrm{dS}}\nolimits ^d : X_h(x) \in V_+(x) \}\) (the positivity domain of \(X_h\)).

  3. (c)

    \(W_M^{\mathrm{KMS}}(h) = \{ x \in \mathop {\mathrm{dS}}\nolimits ^d : (\forall t \in (0,\pi )) \ \alpha _{it}(x) \in {\mathcal {T}}_M\}\) (the KMS domain of \(\alpha \)).

  4. (d)

    \(\kappa _h({\mathcal {T}}_M^{\overline{\tau }_h})\).

  5. (e)

    \(W_M(h) = G^h.\mathop {\mathrm{Exp}}\nolimits _{{\mathbf{e}}_2}((C_+^\circ + C_-^\circ )^\pi )\).

Proof

The equality of the sets under (a) and (b) follows from Lemma D.4.

As \(\alpha _{it}(m) \in M_{{\mathbb {C}}}\) for every \(t \in {{\mathbb {R}}}\) and \(m \in M\), Lemma D.6 shows that the set under (c) coincides with \(W_V^+(h) \cap M = W_M ^+(h)\).

From the proof of Lemma D.8, we recall that

$$\begin{aligned} {\mathcal {T}}_{\mathop {\mathrm{dS}}\nolimits ^d}^{\overline{\tau }_h} = \{ (i x_0, i x_1, x_2, \ldots , x_d) \in \mathop {\mathrm{dS}}\nolimits ^d_{{\mathbb {C}}}: x_0 > |x_1|\}.\end{aligned}$$

Now \(\kappa _h(x) = (-i x_1, -i x_0, x_2, \ldots , x_d)\) implies

$$\begin{aligned} \kappa _h({\mathcal {T}}_{\mathop {\mathrm{dS}}\nolimits ^d}^{\overline{\tau }_h}) = \{ (x_0, x_1, x_2, \ldots , x_d) \in \mathop {\mathrm{dS}}\nolimits ^d : x_1 > |x_0|\} = W_R \cap \mathop {\mathrm{dS}}\nolimits ^d = W_M(h).\end{aligned}$$

Combining Lemma D.8 with (d), we finally obtain

$$\begin{aligned} W_M^+(h)&= \kappa _h({\mathcal {T}}_{\mathop {\mathrm{dS}}\nolimits ^d}^{\overline{\tau }_h}) = G^h.\mathop {\mathrm{Exp}}\nolimits _{{\mathbf{e}}_2}( \kappa _h(iV_+^\pi ({\mathbf{e}}_2)^{-\tau _h})) = G^h.\mathop {\mathrm{Exp}}\nolimits _{{\mathbf{e}}_2}((C_+^\circ + C_-^\circ )^\pi ) \end{aligned}$$

for \(C_\pm ^\circ = {{\mathbb {R}}}_+ ({\mathbf{e}}_1 \pm {\mathbf{e}}_0).\) \(\square \)

Remark D.10

For the \(\alpha \)-fixed base point \({\mathbf{e}}_2 \in \mathop {\mathrm{dS}}\nolimits ^d\) the cone

$$\begin{aligned} W_R({\mathbf{e}}_2) := W_R \cap T_{{\mathbf{e}}_2}(\mathop {\mathrm{dS}}\nolimits ^d) = C_+^\circ + C_-^\circ + T_{{\mathbf{e}}_2}(\mathop {\mathrm{dS}}\nolimits ^d)^{\tau _h}, \quad \text{ where } \quad C_\pm ^\circ = {{\mathbb {R}}}_+ ({\mathbf{e}}_1 \pm {\mathbf{e}}_0)\end{aligned}$$

is an infinitesimal version of the wedge domain in \(M = \mathop {\mathrm{dS}}\nolimits ^d\).

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Neeb, KH., Ólafsson, G. Wedge domains in non-compactly causal symmetric spaces. Geom Dedicata 217, 30 (2023). https://doi.org/10.1007/s10711-022-00755-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10711-022-00755-x

Keywords

Mathematics Subject Classification

Navigation