Skip to main content
Log in

Homogeneous Einstein metrics and butterflies

  • Published:
Annals of Global Analysis and Geometry Aims and scope Submit manuscript

Abstract

In 2012, M. M. Graev associated to a compact homogeneous space G/H a nerve \({\text {X}}_{G/H}\), whose non-contractibility implies the existence of a G-invariant Einstein metric on G/H. The nerve \({\text {X}}_{G/H}\) is a compact, semi-algebraic set, defined purely Lie theoretically by intermediate subgroups. In this paper we present a detailed description of the work of Graev and the curvature estimates given by Böhm in 2004.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

Data availability

Data sharing is not applicable to this article as no datasets were generated or analyzed during this work.

References

  1. Alekseevsky, D.V., Kimel’fel’d, B.N.: Structure of homogeneous Riemann spaces with zero Ricci curvature. Funct. Anal. Appl. 9, 97–102 (1975)

    MATH  Google Scholar 

  2. Alekseevsky, D.V., Dotti, I., Ferraris, C.: Homogeneous Ricci positive \(5\)-manifolds. Pacific J. Math. 175, 1–12 (1996)

    MathSciNet  MATH  Google Scholar 

  3. Arvanitoyeorgos, A.: New invariant Einstein metrics on generalized flag manifolds. Trans. Amer. Math. Soc. 337, 981–995 (1993)

    MathSciNet  MATH  Google Scholar 

  4. Aubin, T.: Équations du type Monge-Ampère sur les variétés kähleriennes compactes. Bull. Sci. Math. (2) 102(1), 63–95 (1978)

    MathSciNet  MATH  Google Scholar 

  5. Belgun, F., Cortés, V., Haupt, A.S., Lindemann, D.: Left-invariant Einstein metrics on \(S^3\times S^3\). J. Geom. Phys. 128, 128–139 (2018)

    MathSciNet  MATH  Google Scholar 

  6. Ballmann, W., Thorbergsson, G., Ziller, W.: Existence of closed geodesics on positively curved manifolds. J. Differential Geom. 18, 221–252 (1983)

    MathSciNet  MATH  Google Scholar 

  7. Benedetti, R., Risler, J.J.: Real Algebraic and Semi-algebraic Sets. Actualités Mathématiques, Hermann, Paris (1990)

    MATH  Google Scholar 

  8. Berger, M.: Quelques formules de variation pour une structure Riemannienne. Ann. Sci. Ec. Norm. Super. 3, \(4^e\) serie, 285–294 (1970)

  9. Besse, A.L.: Einstein Manifolds. Springer, Berlin (1987)

    MATH  Google Scholar 

  10. Böhm, C.: Inhomogeneous Einstein metrics on low-dimensional spheres and other low-dimensional spaces. Invent. Math. 134, 145–176 (1998)

    MathSciNet  MATH  Google Scholar 

  11. Böhm, C.: Homogeneous Einstein metrics and simplicial complexes. J. Differential Geom. 67, 79–165 (2004)

    MathSciNet  MATH  Google Scholar 

  12. Böhm, C.: Non-existence of homogeneous Einstein metrics. Comment. Math. Helv. 80, 123–146 (2005)

    MathSciNet  MATH  Google Scholar 

  13. Böhm, C.: On the long time behavior of homogeneous Ricci flows. Comment. Math. Helv. 90, 543–571 (2015)

    MathSciNet  MATH  Google Scholar 

  14. Böhm, C., Kerr, M.M.: Low-dimensional homogeneous Einstein manifolds. Trans. Amer. Math. Soc. 358, 1455–1468 (2006)

    MathSciNet  MATH  Google Scholar 

  15. Böhm, C., Lafuente, R.: Homogeneous Einstein metrics on Euclidean spaces are Einstein solvmanifolds. Geom. Topol. 26, 899–936 (2022)

    MathSciNet  MATH  Google Scholar 

  16. Böhm, C., Lafuente, R.: Non-compact Einstein manifolds with symmetry. JAMS 36, 591–651 (2023)

    Google Scholar 

  17. Böhm, C., Wang, M.Y., Ziller, W.: A variational approach for compact homogeneous Einstein manifolds. GAFA 14, 681–733 (2004)

    MathSciNet  MATH  Google Scholar 

  18. Boyer, C.P., Galicki, K.: On Sasakian–Einstein geometry. Internat. J. Math. 11, 873–909 (2000)

    MathSciNet  MATH  Google Scholar 

  19. Boyer, C.P., Galicki, K., Kollár, J.: Einstein metrics on spheres. Ann. Math. 557–580 (2005)

  20. Bredon, G.E.: Introduction to Compact Transformation Groups, Pure and Applied Mathematics, vol. 46. Academic Press, New York (1972)

    MATH  Google Scholar 

  21. Bröcker, T., tom Dieck, T.: Graduate Texts in Mathematics. Representations of Compact Lie Groups, vol. 98. Springer, Berlin (1995)

    MATH  Google Scholar 

  22. Cheeger, J., Colding, T.H.: Lower bounds on Ricci curvature and the almost rigidity of warped products. Ann. Math. 144, 189–237 (1996)

    MathSciNet  MATH  Google Scholar 

  23. Cheeger, J., Colding, T.H.: On the structure of spaces with Ricci curvature bounded below I. J. Differential Geom. 46, 406–480 (1997)

    MathSciNet  MATH  Google Scholar 

  24. Chen, X., Donaldson, S., Sun, S.: Kähler–Einstein metrics on Fano manifolds. I: approximation of metrics with cone singularities. J. Amer. Math. Soc. 28, 183–197 (2015)

    MathSciNet  MATH  Google Scholar 

  25. Chen, X., Donaldson, S., Sun, S.: Kähler–Einstein metrics on Fano manifolds. II: Limits with cone angle less than \(2\pi \). J. Amer. Math. Soc. 28, 199–234 (2015)

    MathSciNet  MATH  Google Scholar 

  26. Chen, X., Donaldson, S., Sun, S.: Kähler-Einstein metrics on Fano manifolds. III: Limits as cone angle approaches \(2\pi \) and completion of the main proof. J. Amer. Math. Soc. 28, 235–278 (2015)

    MathSciNet  MATH  Google Scholar 

  27. Chrysikos, I., Sakane, Y.: Non-naturally reductive Einstein metrics on exceptional Lie groups. J. Geom. Phys. 116, 152–186 (2017)

    MathSciNet  MATH  Google Scholar 

  28. D’Atri, J., Ziller, W.: Naturally reductive metrics and Einstein metrics on compact Lie groups, Memoirs of the AMS, 215 (1979)

  29. Delfs, H., Knebusch, M.: Semialgebraic topology over a real closed field II: Basic theory of semialgebraic spaces. Math. Z. 178, 175–213 (1981)

    MathSciNet  MATH  Google Scholar 

  30. Delfs, H., Knebusch, M.: Separation, Retractions and Homotopy extension in semi-algebraic spaces. Pacific J. Math. 114, 47–71 (1984)

    MathSciNet  MATH  Google Scholar 

  31. Delfs, H., Knebusch, M.: Locally Semialgebraic Spaces. Lecture Notes in Math, vol. 1173. Springer, Berlin (1985)

    MATH  Google Scholar 

  32. Dickinson, W., Kerr, M.M.: The geometry of compact homogeneous spaces with two isotropy summands. Ann. Global Anal. Geom. 34, 329–350 (2008)

    MathSciNet  MATH  Google Scholar 

  33. Dynkin, E.B.: Maximal subgroups of the classical groups. Transl. Amer. Math. Soc. Series 2 6, 245–378 (1957)

    MATH  Google Scholar 

  34. Eschenburg, J.-H.: Lecture notes on symmetric spaces, preprint (1997)

  35. Foscolo, L., Haskins, M.: New \(G_2\)-holonomy cones and exotic nearly Kähler structures on \(S^6\) and \(S^3\times S^3\). Ann. Math. 2, 59–130 (2017)

    MATH  Google Scholar 

  36. Fulton, W., Harris, J.: Representation Theory: A First Course Graduate Texts in Mathematics, vol. 129. Springer, Berlin (1991)

    MATH  Google Scholar 

  37. Ghoussoub, N.: Location, multiplicity and Morse indices of min-max critical points. J. Reine Angew. Math. 417, 27–76 (1991)

    MathSciNet  MATH  Google Scholar 

  38. Graev, M.M.: On the number of invariant Einstein metrics on a compact homogeneous space, Newton polytopes and contractions of Lie algebras. Int. J. Geom. Methods Mod. Phys. 3(5–6), 1047–1075 (2006)

    MathSciNet  MATH  Google Scholar 

  39. Graev, M.M.: The existence of invariant Einstein metrics on a compact homogeneous space. Trans. Moscow Math. Soc. 73, 1–28 (2012)

    MathSciNet  MATH  Google Scholar 

  40. Graev, M.M.: On the compactness of the set of invariant Einstein metrics. Ann. Global Anal. Geom. 44(4), 471–500 (2013)

    MathSciNet  MATH  Google Scholar 

  41. Hatcher, A.: Algebraic Topology. Cambridge University Press, Cambridge (2002)

    MATH  Google Scholar 

  42. He, C.: Cohomogeneity one manifolds with a small family of invariant metrics. Geom. Dedicata 157, 41–90 (2012)

    MathSciNet  MATH  Google Scholar 

  43. Heber, J.: Noncompact homogeneous Einstein spaces. Invent. Math. 133, 279–352 (1998)

    MathSciNet  MATH  Google Scholar 

  44. Helgason, S.: Differential Geometry, Lie Groups, and Symmetric Spaces, 2nd edn. Academic Press, New York (1978)

    MATH  Google Scholar 

  45. Hilbert, D.: Die Grundlagen der Physik, Nachr. Akad. Wiss. Gött., pp. 395–407 (1915)

  46. Hironaka, H.: Triangulations of algebraic sets. Proc. Sympos. Pure Math. 29, 165–185 (1975)

    MathSciNet  MATH  Google Scholar 

  47. Jensen, G.: Homogeneous Einstein spaces of dimension 4. J. Differential Geom. 3, 309–349 (1969)

    MathSciNet  MATH  Google Scholar 

  48. Joyce, D.D.: Compact \(8\)-manifolds with holonomy \({\rm Spin}(7)\). Invent. Math. 123, 507–552 (1996)

    MathSciNet  MATH  Google Scholar 

  49. Joyce, D.D.: Compact Riemannian \(7\)-manifolds with holonomy \(G_2\). I, II. J. Differ. Geom. 43, 291–328, 329–375 (1996)

  50. Joyce, D.: Compact Manifolds with Special Holonomy. Oxford Mathematical Monographs. Oxford University Press, Oxford (2000)

    Google Scholar 

  51. Kerr, M.M.: Some new homogeneous Einstein metrics on symmetric spaces. Trans. Amer. Math. Soc. 348, 153–171 (1996)

    MathSciNet  MATH  Google Scholar 

  52. Klaus, S.: Einfach-zusammenhängende Kompakte Homogene Räume bis zur Dimension Neun. Diplomarbeit, Johannes Gutenberg Universität (1988)

  53. Kreck, M., Stolz, S.: A diffeomorphism classification of 7-dimensional homogeneous Einstein manifolds with \(SU(3) \times SU(2) \times U(1)\)-symmetry. Ann. Math. (2) 127(2), 373–388 (1988)

    MathSciNet  MATH  Google Scholar 

  54. Lauret, J.: Ricci soliton homogeneous nilmanifolds. Math. Ann. 319, 715–733 (2001)

    MathSciNet  MATH  Google Scholar 

  55. Lauret, J.: Einstein solvmanifolds are standard. Ann. Math. (2) 172(3), 1859–1877 (2010)

    MathSciNet  MATH  Google Scholar 

  56. Lauret, J.: On the stability of homogeneous Einstein manifolds, preprint arXiv:2105.06336 (2021) Asien J. Math. (to appear)

  57. Lauret, E.A., Lauret, J.: The stability of standard homogeneous Einstein manifolds. Math. Z. 303(1), 36 (2023)

    MathSciNet  MATH  Google Scholar 

  58. Lauret, J., Will, C.: On the stability of homogeneous Einstein manifolds, II. J. Lond. Math. Soc. 106, 3638–3669 (2022)

    MathSciNet  Google Scholar 

  59. Lauret, J., Will, C.: Homogeneous Einstein metrics and local maxima of the Hilbert action. J. Geom. Phys. 178, 8 (2022)

    MathSciNet  MATH  Google Scholar 

  60. Maunder, C.R.F.: Algebraic Topology. Cambridge University Press, Cambridge (1980)

    MATH  Google Scholar 

  61. Milnor, J.: Construction of Universal Bundles, II. Ann. Math. 63, 430–436 (1956)

    MathSciNet  MATH  Google Scholar 

  62. Nienhaus, J., Wink, M.: Einstein metrics on the Ten-sphere (2023). arXiv:2303.04832

  63. Nikonorov, Yu.G., Rodionov, E.D.: Compact \(6\)-dimensional homogeneous Einstein manifolds. Dokl. Math. RAS 336, 599–601 (1999)

    MathSciNet  MATH  Google Scholar 

  64. Nikonorov, Yu.G.: Compact homogeneous Einstein \(7\)-manifolds. Geom. Dedicata 109, 7–30 (2004)

    MathSciNet  MATH  Google Scholar 

  65. Nikonorov, Yu.G., Classification of generalized Wallach spaces. Geom. Dedicata 181, 193–212 (2016), Correction Geom. Dedicata 214, 849–851 (2021)

  66. Nikonorov, Yu.G.: On left-invariant Einstein Riemannian metrics that are not geodesic orbit. Transform. Groups 24(2), 511–530 (2019)

    MathSciNet  MATH  Google Scholar 

  67. Park, J.-S., Sakane, Y.: Invariant Einstein metrics on certain homogeneous spaces. Tokyo J. Math. 20, 51–61 (1997)

    MathSciNet  MATH  Google Scholar 

  68. Quillen, D.: Homotopy Properties of the Poset of Nontrivial p-Subgroups of a Group. Adv. Math. 28, 101–128 (1978)

    MathSciNet  MATH  Google Scholar 

  69. Rauße, C.: Simplicial complexes of compact homogeneous spaces. Ph.D.-thesis Münster, (2016)

  70. Schwahn, P., Semmelmann, U., Weingart, G.: Stability of the Non-Symmetric Space \(E_7/{\rm PSO}(8)\) (2022). arXiv:2203.10138

  71. Spiro, A.: A remark on locally homogeneous Riemannian spaces. Results Math. 24, 318–325 (1993)

    MathSciNet  MATH  Google Scholar 

  72. Tian, G.: K-stability and Kähler-Einstein metrics. Commun. Pure Appl. Math. 68, 1085–1156 (2015). Corrigendum 68, 2082–2083 (2015)

  73. Wang, M.Y.: Some examples of homogeneous Einstein manifolds in dimension seven. Duke Math. J. 49, 23–28 (1982)

    MathSciNet  MATH  Google Scholar 

  74. Wang, M.Y.: Einstein metrics from symmetry and bundle construction in Surveys in differential geometry VI: essays on Einstein manifolds. C. LeBrun, M. Y. Wang (eds.), pp. 287–325. International Press, Boston (1999)

  75. Wang, M.Y.: Einstein metrics from symmetry and bundle constructions: a sequel. Adv. Lect. Math. 22, 253–309 (2012)

    MathSciNet  MATH  Google Scholar 

  76. Wang, M.Y., Ziller, W.: On normal homogeneous Einstein manifolds. Ann. scient. Éc. Norm. Sup. \(4^e\) série t. 18, 563–633 (1985)

  77. Wang, M.Y., Ziller, W.: Existence and non-existence of homogeneous Einstein metrics. Invent. Math. 84, 177–194 (1986)

    MathSciNet  MATH  Google Scholar 

  78. Wang, M.Y., Ziller, W.: Einstein metrics on principal torus bundles. J. Differential Geom. 31, 215–248 (1990)

    MathSciNet  MATH  Google Scholar 

  79. Wang, M.Y., Ziller, W.: On isotropy irreducible Riemannian manifolds. Acta Math. 199, 223–261 (1991)

    MathSciNet  MATH  Google Scholar 

  80. Wolf, J.A.: The geometry and structure of isotropy irreducible homogeneous spaces. Acta. Math. 120, 59–148 (1968); correction. Acta. Math. 152, 141–142 (1984)

  81. Yau, S.T.: On the Ricci curvature of a compact Kähler manifold and the complex Monge-Ampère equation I. Commun. Pure Appl. Math. 31, 339–411 (1978)

    MATH  Google Scholar 

  82. Ziller, W.: Homogeneous Einstein metrics on spheres and projective spaces. Math. Ann. 259, 351–358 (1982)

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

Funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy EXC 2044—390685587, Mathematics Münster: Dynamics–Geometry–Structure, and the Collaborative Research Centre CRC 1442, Geometry: Deformations and Rigidity.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Megan M. Kerr.

Ethics declarations

Conflict of interest

The authors have no financial or proprietary interests in any material discussed in this article.

Additional information

This paper is dedicated to the memory of M. M. Graev.

Publisher's Note

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

Appendix

Appendix

In this section we will make a brief digression, first into semi-algebraic geometry. We also provide helpful background on the isotropy representation of a compact homogeneous space G/H.

1.1 Semi-algebraic sets

We refer the reader to Benedetti and Risler for more details [7]. A set \(X\subset {{{\mathbb {R}}}}^m\) is called semi-algebraic if X is defined by finitely many polynomial equations and inequalities. The inequalities are allowed to be both strict and non-strict.

Given two semi-algebraic sets \(X\subset {{{\mathbb {R}}}}^m\) and \(Y\subset {{{\mathbb {R}}}}^n\), we say a map \(f:X\rightarrow Y\) is semi-algebraic if the graph of f is a semi-algebraic subset of \({{{\mathbb {R}}}}^{m+n}\): see [7, Definition 2.3.2]. A semi-algebraic map is not necessarily continuous, see [7, Example 2.7.3], though some authors include a requirement of continuity in their definition of a semi-algebraic map: see [46, 29, Theorem 7.6], [31, I, Proposition 3.13]. The paradigm for semi-algebraic maps are polynomials.

A fundamental result in semi-algebraic geometry is the following:

Theorem 7.1

(Tarski-Seidenberg) The image of a semi-algebraic subset of \({{{\mathbb {R}}}}^n\times {{{\mathbb {R}}}}^m\) under the projection \(\pi :{{{\mathbb {R}}}}^n \times {{{\mathbb {R}}}}^m \rightarrow {{{\mathbb {R}}}}^n\) such that \((x,y)\mapsto x\) is a semi-algebraic set.

The theorem implies that intersections, unions, and complements of semi-algebraic sets are semi-algebraic. Moreover, it is possible to use first-order formulae (in the language of ordered fields with parameters in \({{{\mathbb {R}}}}\)) to obtain new semi-algebraic sets from known ones. For instance, given \(X_1\subset {{{\mathbb {R}}}}^n,X_2 \subset {{{\mathbb {R}}}}^m\) semi-algebraic sets and a semi-algebraic map \(F:{{{\mathbb {R}}}}^n\times {{{\mathbb {R}}}}^m \rightarrow {{{\mathbb {R}}}}^k\), the set

$$\begin{aligned} \{ x_1 \in X_1\mid \exists x_2 \in X_2 \text { s.t. } F(x_1,x_2) \le 0\} \end{aligned}$$

is semi-algebraic. As another example, given endomorphisms AB of \({{{\mathbb {R}}}}^n\), the condition \(\ker (A) \subsetneq \ker (B)\) is a semi-algebraic condition because it is equivalent to

$$\begin{aligned} \{ \forall x \in {{{\mathbb {R}}}}^n \mid A(x)=0 \Rightarrow B(x)=0\} \wedge \{ \exists y \in {{{\mathbb {R}}}}^n \mid B(y)=0 \wedge A(y) \ne 0\}. \end{aligned}$$

Connected components of semi-algebraic sets are semi-algebraic, and the closure, interior, or boundary of a semi-algebraic set is semi-algebraic.

The key property of compact semi-algebraic sets used in Sects. 3.1 and 3.2 is that such sets are absolute neighborhood retracts. The following result states a useful generalization of that.

Lemma 7.2

[39] Let \(X,Y \subset {{{\mathbb {R}}}}^N\) be compact, semi-algebraic sets with \(X \subset Y\) and \(X \ne \emptyset \). Suppose that for some \(\delta _0>0\), for all \(\delta \in (0,\delta _0)\) there exists a continuous map

$$\begin{aligned} H_\delta :[0,1]\times Y \rightarrow Y \end{aligned}$$

with the following properties: \(H_\delta (0,y)=y\) for all \(y \in Y\), \(H_\delta (t,x)=x\) for all \(x \in X\) and all \(t \in [0,1]\), and

$$\begin{aligned} d(H_\delta (1,y),X)< \delta \text { for all } y \in Y. \end{aligned}$$

Then X is a strong deformation retract of Y.

Proof

It follows from Theorem 1 of [30], see also [31, III, Theorem 1.1], that there exists an open semi-algebraic neighborhood U of X in Y and a semi-algebraic, continuous map \(G:[0,1]\times {\overline{U}}\rightarrow {\overline{U}}\), such that the restriction \(G\vert _{[0,1]\times U}\) yields a strong deformation retraction from U to X.

By compactness of X, there exists \(\delta _0>0\) such that for all \(0<\delta \le \delta _0\), we have

$$\begin{aligned} T_\delta (X):=\{y \in Y \mid d(y,X)<\delta \} \subset U. \end{aligned}$$

Hence we can define

$$\begin{aligned} H:[0,1]\times Y \rightarrow Y \text { s.t. } (t,y) \mapsto \left\{ \begin{array}{ll} H_{\delta }(2t,y) &{} \text { for } 0\le t \le \tfrac{1}{2} \\ G(2t-1,H_\delta (1,y)) &{}\text { for } \tfrac{1}{2}\le t \le 1. \end{array} \right. \end{aligned}$$

The function H is continuous since for \(t=\tfrac{1}{2}\) we have \(G(2t-1,H_\delta (1,y))=H_{\delta }(2t,y)\). Moreover \(H(t,x)=x\) for all \(t \in [0,1]\) and all \(x \in X\) and \(H(1,Y)=X\). This shows the claim. \(\square \)

1.2 Lie-theory I

In this section we provide details concerning the isotropy representation of a compact homogeneous space G/H and the normalizer \(N_G(H)\) of H in G. We assume, as we have throughout the paper, that G and H are connected.

Definition 7.3

For any compact homogeneous space G/H, we denote by \({{{\mathfrak {m}}}}_0\) the subspace of \({{{\mathfrak {m}}}}\) on which \({\text {Ad}}(H)\) acts trivially.

Notice that if \(\text {rk} \,G=\text {rk} \,H\), then \({{{\mathfrak {m}}}}_0 = \{0\}\), while if \(H=\{e\}\), \({{{\mathfrak {m}}}}_0={{{\mathfrak {g}}}}\). Recall also that

$$\begin{aligned} \exp ({\text {ad}}(X))={\text {Ad}}(\exp _G(X)) \end{aligned}$$
(7.1)

for all \(X \in {{{\mathfrak {g}}}}\), where \(\exp _G:{{{\mathfrak {g}}}}\rightarrow G\) denotes the exponential of G and \(\exp (D)=\sum _{k=0}^\infty \tfrac{D^k}{k!}\) for any endomorphism \(D:{{{\mathfrak {g}}}}\rightarrow {{{\mathfrak {g}}}}\) [44, II].

Lemma 7.4

A subspace \(\tilde{{{\mathfrak {m}}}}\) of \({{{\mathfrak {m}}}}\) is \({\text {Ad}}(H)\)-invariant if and only if \([{{{\mathfrak {h}}}},\tilde{{{{\mathfrak {m}}}}}]\subset \tilde{{{{\mathfrak {m}}}}}\). Moreover, \([{{{\mathfrak {h}}}},{{{\mathfrak {m}}}}_0]=0\) and for any \({\text {Ad}}(H)\)-invariant subspace \({\tilde{{{{\mathfrak {m}}}}}}\) of \({{{\mathfrak {m}}}}\ominus {{{\mathfrak {m}}}}_0\), we have \(\{0\} \ne [{{{\mathfrak {h}}}},{\tilde{{{{\mathfrak {m}}}}}}]\).

Proof

If \(\tilde{{{{\mathfrak {m}}}}}\) is \({{\text {Ad}}}(H)\)-invariant, then differentiation yields \([{{{\mathfrak {h}}}},\tilde{{{{\mathfrak {m}}}}}]\subset \tilde{{{{\mathfrak {m}}}}}\). Conversely, suppose we know \([{{{\mathfrak {h}}}},\tilde{{{{\mathfrak {m}}}}}]\subset \tilde{{{{\mathfrak {m}}}}}\). Since H is connected, for any \(Z \in {{{\mathfrak {h}}}}\), \({{\text {Ad}}}(\exp (Z))=\exp ({{\text {ad}}}(Z))\). Hence, \(\tilde{{{{\mathfrak {m}}}}}\) is \({{\text {Ad}}}(H)\)-invariant. This shows the first claim. The second and the third claim follow immediately. \(\square \)

Let \({{{\mathfrak {n}}}}({{{\mathfrak {h}}}})\) denote the Lie algebra of the normalizer \(N_G(H)\) of H in G.

Lemma 7.5

Let G/H be a compact homogeneous space. Then \({{{\mathfrak {n}}}}({{{\mathfrak {h}}}})= {{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_0\). Moreover, \({{{\mathfrak {m}}}}_0\) is a compact subalgebra.

Proof

Since \([{{{\mathfrak {h}}}},{{{\mathfrak {m}}}}_0]=0 \in {{{\mathfrak {h}}}}\), see above, we conclude \({{{\mathfrak {m}}}}_0 \subset {{{\mathfrak {n}}}}({{{\mathfrak {h}}}})\). Conversely, suppose that \(\tilde{{{\mathfrak {m}}}}\subset {{{\mathfrak {n}}}}({{{\mathfrak {h}}}}) \cap {{{\mathfrak {m}}}}\). Then on the one hand, \([{{{\mathfrak {h}}}}, \tilde{{{\mathfrak {m}}}}]\subset {{{\mathfrak {h}}}}\) by definition of \({{{\mathfrak {n}}}}({{{\mathfrak {h}}}})\), while on the other hand, \([{{{\mathfrak {h}}}}, \tilde{{{\mathfrak {m}}}}] \subset {{{\mathfrak {m}}}}\) because \([{{{\mathfrak {h}}}},{{{\mathfrak {m}}}}]\subset {{{\mathfrak {m}}}}\). This shows \([{{{\mathfrak {h}}}}, \tilde{{{\mathfrak {m}}}}]=0\), and thus \({{{\mathfrak {m}}}}\cap {{{\mathfrak {n}}}}({{{\mathfrak {h}}}})\subset {{{\mathfrak {m}}}}_0\), proving our equality.

Since \([{{{\mathfrak {m}}}}_0,{{{\mathfrak {m}}}}_0] \subset {{{\mathfrak {n}}}}({{{\mathfrak {h}}}})={{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_0\) and \([{{{\mathfrak {h}}}},{{{\mathfrak {m}}}}_0]=0\), then because Q is self-adjoint (4.3), we know that \({{{\mathfrak {m}}}}_0\) is a subalgebra of \({{{\mathfrak {n}}}}({{{\mathfrak {h}}}})\) and thus is a subalgebra of \({{{\mathfrak {g}}}}\). In order to show that the corresponding connected subgroup of G is compact we now use the special property of the biinvariant metric Q chosen in Sect. 2.1. We decompose the compact Lie algebra \({{{\mathfrak {n}}}}({{{\mathfrak {h}}}})={{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_0\) into the Q-orthogonal sum of its semisimple part \(({{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_0)_s= {{{\mathfrak {h}}}}_s \oplus ({{{\mathfrak {m}}}}_0)_s\) and its center \({{{\mathfrak {z}}}}({{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_0)\). Each of the Lie algebras \({{{\mathfrak {h}}}}_s\), \(({{{\mathfrak {m}}}}_0)_s\) and \({{{\mathfrak {z}}}}({{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_0)\) is compact, and so \({{{\mathfrak {h}}}}\cap {{{\mathfrak {z}}}}({{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_0)\) is also compact. It remains to show that the Q-orthogonal complement of \({{{\mathfrak {h}}}}\cap {{{\mathfrak {z}}}}({{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_0)\) in \({{{\mathfrak {z}}}}({{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_0)\) is compact.

By the definition of Q, the compact, abelian subalgebras \({{{\mathfrak {h}}}}\cap {{{\mathfrak {z}}}}({{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_0)\) and \({{{\mathfrak {z}}}}({{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_0)\) can be viewed as subalgebras of \({{{\mathfrak {s}} {\mathfrak {o}}}}(6N)\). Notice we can choose a Q-orthonormal (standard) basis \(({\hat{e}}_1,\ldots ,{\hat{e}}_{3N})\) for the standard maximal (diagonal) torus \({{{\mathfrak {t}}}}_{6N}\) of \({{{\mathfrak {s}} {\mathfrak {o}}}}(6N)\) so that every compact subtorus \({{{\mathfrak {t}}}}\) of \({{{\mathfrak {t}}}}_{6N}\) has a basis \((e_1,\ldots ,e_{\dim {{{{\mathfrak {t}}}}}})\), for which each \(e_i\) is a rational linear combination of the above basis elements \(({\hat{e}}_i)\) of \({{{\mathfrak {s}} {\mathfrak {o}}}}(6N)\). Then, the Q-orthogonal complement in \({{{\mathfrak {t}}}}_{6N}\) of the a subtorus \({{{\mathfrak {t}}}}\) must also have such a basis, and hence it corresponds to a compact subtorus in \({{{\mathfrak {s}} {\mathfrak {o}}}}(6N)\). As a consequence, the Q-orthogonal complement of \({{{\mathfrak {h}}}}\cap {{{\mathfrak {z}}}}({{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_0)\) in \({{{\mathfrak {z}}}}({{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_0)\) is the intersection of two compact subalgebras, thus a compact subalgebra. \(\square \)

Remark 7.6

We decompose \({{{\mathfrak {m}}}}\) into its \({\text {Ad}}(H)\)-invariant isotypical summands (see [21, II, Proposition 6.9]):

$$\begin{aligned} {{{\mathfrak {m}}}}=\bigoplus _{i=1}^{\ell _{\text {iso} }}\,{{{\mathfrak {p}}}}_i . \end{aligned}$$

Each \({{{\mathfrak {p}}}}_i\) is a direct sum of \({\text {Ad}}(H)\)-irreducible summands which are equivalent (as \({\text {Ad}}(H)\)-representations), while, for \(i \ne j\), irreducible summands in \({{{\mathfrak {p}}}}_i\) and \({{{\mathfrak {p}}}}_j\) are inequivalent. Note by Lemma 7.4, if \({{{\mathfrak {m}}}}_0\) is non-trivial, then \({{{\mathfrak {m}}}}_0\) is one of the isotypical summands of \({{{\mathfrak {m}}}}\).

Thus by Schur’s Lemma, for every \(P_g \in \mathcal {M}^G\) we have

$$\begin{aligned} Q(P_g\,\cdot ,\,\cdot \,)= Q((P_g)_1\,\cdot ,\,\cdot \,)\vert _{{{{\mathfrak {p}}}}_1} \perp \cdots \perp Q((P_g)_{\ell _{\text {iso} }}\,\cdot ,\,\cdot \,) \vert _{{{{\mathfrak {p}}}}_{\ell _{\text {iso} }}} \end{aligned}$$
(7.2)

where, for each \(1 \le i \le \ell _{\text {iso} }\), \((P_g)_i:{{{\mathfrak {p}}}}_i \rightarrow {{{\mathfrak {p}}}}_i\) is an \({\text {Ad}}(H)\)-equivariant, self-adjoint, positive definite endomorphism. Consequently, every \(P_g \in \mathcal {M}^G\) respects the decomposition \({{{\mathfrak {m}}}}=\oplus _{i=1}^{\ell _{\text {iso} }}\,{{{\mathfrak {p}}}}_i\). Notice this is also true for every \(v \in \text {Sym} _{{{{\mathfrak {m}}}}}^{H}\).

For \({{{\mathfrak {k}}}}\) a subalgebra in \(\mathcal {K}\), let \({{{\mathfrak {m}}}}_{{{{\mathfrak {k}}}}}:={{{\mathfrak {m}}}}\cap {{{\mathfrak {k}}}}\). For convenience, \({{{\mathfrak {m}}}}_{{{{{\mathfrak {h}}}}}}:= {{{\mathfrak {m}}}}\).

Lemma 7.7

Let \({{{\mathfrak {k}}}}\) be a subalgebra, \({{{\mathfrak {h}}}}\le {{{\mathfrak {k}}}}< {{{\mathfrak {g}}}}\), let K be the connected subgroup of G corresponding to \({{{\mathfrak {k}}}}\), and let \(A \in \text {Sym} _{{{{\mathfrak {g}}}}}^{H}\) with \({{{\mathfrak {k}}}}\subset \ker (A)\). Then, the condition \([A,{\text {ad}}({{{\mathfrak {k}}}})]=0\) is equivalent to the \({\text {Ad}}(K)\)-equivariance of A. Furthermore, \([A,{\text {ad}}({{{\mathfrak {k}}}})]=0\) if and only if \([A\vert _{{{{\mathfrak {k}}}}^\perp },{\text {ad}}({{{\mathfrak {m}}}}_{{{{\mathfrak {k}}}}})\vert _{{{{\mathfrak {k}}}}^\perp }]=0\) if and only if \([A\vert _{{{{{\mathfrak {m}}}}}},{\text {ad}}({{{\mathfrak {m}}}}_{{{{\mathfrak {k}}}}})\vert _{{{{{\mathfrak {m}}}}}}]=0\).

Proof

Let K be the connected subgroup of G with Lie algebra \({{{\mathfrak {k}}}}\). We know by differentiation that \({\text {Ad}}(K)\)-equivariance of A implies \([A,{\text {ad}}({{{\mathfrak {k}}}})]=0\). To see the converse, let \(A \in \text {Sym} _{{{{\mathfrak {g}}}}}^{H}\) and suppose that \([A,{\text {ad}}({{{\mathfrak {k}}}})]=0\). Using (7.1) shows that A is \({\text {Ad}}(K)\)-equivariant.

Next, let \(A \in \text {Sym} _{{{{\mathfrak {g}}}}}^{H}\) with \({{{\mathfrak {k}}}}\subset \ker (A)\). Since \({\text {Ad}}(K)\), \({\text {ad}}({{{\mathfrak {k}}}})\) respect the decomposition \({{{\mathfrak {g}}}}= {{{\mathfrak {k}}}}\oplus {{{\mathfrak {k}}}}^\perp \), it follows that \(A\vert _{{{{\mathfrak {k}}}}^\perp }\) is \({\text {Ad}}(K)\vert _{{{{\mathfrak {k}}}}^\perp }\)-equivariant if and only if \([A\vert _{{{{\mathfrak {k}}}}^\perp },{\text {ad}}({{{\mathfrak {k}}}})\vert _{{{{\mathfrak {k}}}}^\perp }]=0\).

Since \(A \in \text {Sym} _{{{{\mathfrak {g}}}}}^{H}\) with \({{{\mathfrak {k}}}}\subset \ker (A)\) is \({\text {Ad}}(K)\)-equivariant if and only if \(A\vert _{{{{\mathfrak {k}}}}^\perp }\) is \({\text {Ad}}(K)\vert _{{{{\mathfrak {k}}}}^\perp }\)-equivariant we deduce that \([A,{\text {ad}}({{{\mathfrak {k}}}})]=0\) is equivalent to \([A\vert _{{{{\mathfrak {k}}}}^\perp },{\text {ad}}({{{\mathfrak {k}}}})\vert _{{{{\mathfrak {k}}}}^\perp }]=0\).

Since for \(A \in \text {Sym} _{{{{\mathfrak {g}}}}}^{H}\) with \({{{\mathfrak {k}}}}\subset \ker (A)\) A preserves \({{{\mathfrak {m}}}}\) we deduce from \([A,{\text {ad}}({{{\mathfrak {k}}}})]=0\) that \([A\vert _{{{{{\mathfrak {m}}}}}},{\text {ad}}({{{\mathfrak {k}}}})\vert _{{{{{\mathfrak {m}}}}}}]=0\), which implies \([A\vert _{{{{{\mathfrak {m}}}}}},{\text {ad}}({{{\mathfrak {m}}}}_{{{{\mathfrak {k}}}}})\vert _{{{{{\mathfrak {m}}}}}}]=0\) since \({{{\mathfrak {m}}}}_{{{{\mathfrak {k}}}}}\subset {{{\mathfrak {k}}}}\).

Suppose now \([A\vert _{{{{{\mathfrak {m}}}}}},{\text {ad}}({{{\mathfrak {m}}}}_{{{{\mathfrak {k}}}}})\vert _{{{{{\mathfrak {m}}}}}}]=0\). All elements of \(\text {Sym} _{{{{\mathfrak {g}}}}}^{H}\) are \({\text {Ad}}(H)\)-equivariant, by definition, thus \([A\vert _{{{{{\mathfrak {m}}}}}},{\text {ad}}({{{\mathfrak {h}}}})\vert _{{{{{\mathfrak {m}}}}}}]=0\). Using \({{{\mathfrak {k}}}}={{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_{{{{\mathfrak {k}}}}}\) we conclude \([A\vert _{{{{{\mathfrak {m}}}}}},{\text {ad}}({{{\mathfrak {k}}}})\vert _{{{{{\mathfrak {m}}}}}}]=0\). Since \({\text {ad}}({{{\mathfrak {k}}}})\) preserves the decomposition \({{{\mathfrak {g}}}}={{{\mathfrak {k}}}}\oplus {{{\mathfrak {k}}}}^\perp \) it follows that \({\text {ad}}({{{\mathfrak {k}}}})\vert _{{{{{\mathfrak {m}}}}}}\) preserves the decomposition \({{{\mathfrak {m}}}}={{{\mathfrak {m}}}}_{{{{\mathfrak {k}}}}} \oplus {{{\mathfrak {k}}}}^\perp \). We deduce \([A\vert _{{{{\mathfrak {k}}}}^\perp },{\text {ad}}({{{\mathfrak {k}}}})\vert _{{{{\mathfrak {k}}}}^\perp }]=0\). Using that for \(A \in \text {Sym} _{{{{\mathfrak {g}}}}}^{H}\) with \({{{\mathfrak {k}}}}\subset \ker (A)\), this is equivalent to \([A,{\text {ad}}({{{\mathfrak {k}}}})]=0\), proving the claim. \(\square \)

Any such A induces a submersion metric on G/H with respect to \(K/H \rightarrow G/H \rightarrow G/K\). Notice however, that K need not be a compact subgroup of G.

Corollary 7.8

Let \(A \in \text {Sym} _{{{{\mathfrak {g}}}}}^{H}\) (\(v \in {{\text {S}}}\)) and let \({{{\mathfrak {m}}}}_{I_1},\dots , {{{\mathfrak {m}}}}_{I_\ell } \subset {{{\mathfrak {m}}}}\) denote the eigenspaces of \(A\vert _{{{{\mathfrak {m}}}}}\) corresponding to distinct eigenvalues in increasing order.

  1. (i)

    Then \([{{{\mathfrak {m}}}}_{I_1},{{{\mathfrak {m}}}}_{I_1}]\perp {{{\mathfrak {m}}}}_{I_p}\) (i.e. \([I_1 I_1 I_p]=0\)) for all \(1<p \le \ell \) if and only if \({{{\mathfrak {k}}}}= {{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_{I_1}\) is a subalgebra.

  2. (ii)

    Furthermore, when \({{{\mathfrak {k}}}}= {{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_{I_1}\) is a subalgebra, \(A \in {\text {D}}({{{\mathfrak {k}}}})\) if and only if \({{{\mathfrak {k}}}}\le \ker (A)\) and for all \(1 \le p \ne q \le \ell \), we have \([{{{\mathfrak {m}}}}_{I_1},{{{\mathfrak {m}}}}_{I_p}]\perp {{{\mathfrak {m}}}}_{I_q}\) (i.e. \([I_1 I_p I_q]=0\)).

Proof

(i) If for all \(1 < p \) we have \([{{{\mathfrak {m}}}}_{I_1},{{{\mathfrak {m}}}}_{I_1}]\perp {{{\mathfrak {m}}}}_{I_p}\), then since \([{{{\mathfrak {h}}}},{{{\mathfrak {m}}}}_{I_1}] \subset {{{\mathfrak {m}}}}_{I_1}\) we have that \({{{\mathfrak {k}}}}= {{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_{I_1}\) is a subalgebra of \({{{\mathfrak {g}}}}\). Conversely, if \({{{\mathfrak {k}}}}= {{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_{I_1}\in \mathcal {K}\) is a subalgebra, then we know \([{{{\mathfrak {m}}}}_{I_1},{{{\mathfrak {m}}}}_{I_1}] \subset {{{\mathfrak {k}}}}\perp {{{\mathfrak {m}}}}_{I_p}\) for all \(1<p\).

(ii) Suppose \({{{\mathfrak {k}}}}= {{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_{I_1}\in \mathcal {K}\) is a subalgebra, and \(A \in {\text {D}}({{{\mathfrak {k}}}})\) or equivalently, \(v \in {\text {D}}({{{\mathfrak {k}}}})^{{\text {S}}}\). Since \(A \in {\text {D}}({{{\mathfrak {k}}}})\) we know \([A,{\text {ad}}({{{\mathfrak {k}}}})]=0\). Thus, by Lemma 7.7 we deduce \([A\vert _{{{{\mathfrak {k}}}}^\perp },{\text {ad}}({{{\mathfrak {m}}}}_{{{{\mathfrak {k}}}}})\vert _{{{{\mathfrak {k}}}}^\perp }]=0\). This shows that \({\text {ad}}({{{\mathfrak {m}}}}_{{{{\mathfrak {k}}}}})\) respects the eigenspaces of \(A\vert _{{{{{\mathfrak {m}}}}}}\). That is, \([{{{\mathfrak {m}}}}_{I_1}, {{{\mathfrak {m}}}}_{I_p}] \subset {{{\mathfrak {m}}}}_{I_p} \perp {{{\mathfrak {m}}}}_{I_q}\) for \(p \ne q\). Conversely, suppose \({{{\mathfrak {k}}}}= {{{\mathfrak {h}}}}\oplus {{{\mathfrak {m}}}}_{I_1} \le \ker (A)\) and for all \(p \ne q\), we have \([{{{\mathfrak {m}}}}_{I_1},{{{\mathfrak {m}}}}_{I_p}]\perp {{{\mathfrak {m}}}}_{I_q}\). Thus \({\text {ad}}({{{\mathfrak {k}}}})\vert _{{{{\mathfrak {k}}}}^\perp }\) preserves the eigenspaces of \(A\vert _{{{{\mathfrak {k}}}}^\perp }\). By Lemma 7.7, to see that \(A \in {\text {D}}({{{\mathfrak {k}}}})\) we need that \([A\vert _{{{{\mathfrak {k}}}}^\perp },{\text {ad}}({{{\mathfrak {k}}}})\vert _{{{{\mathfrak {k}}}}^\perp }]=0\) which follows from \([A|_{{{{\mathfrak {m}}}}_{I_p}},{\text {ad}}({{{\mathfrak {m}}}}_{I_1})|_{{{{\mathfrak {m}}}}_{I_p}}] =0\) for every \(p>1\). Let \(X_1 \in {{{\mathfrak {m}}}}_{I_i}\) and \(X_p \in {{{\mathfrak {m}}}}_{I_p}\), \(p>1\). Then

$$\begin{aligned} {[}A|_{{{{\mathfrak {m}}}}_{I_p}},{\text {ad}}(X_1)|_{{{{\mathfrak {m}}}}_{I_p}}](X_p) = A[X_1,X_p] - [X_1,A X_p]=0, \end{aligned}$$

since by hypothesis, \([{{{\mathfrak {m}}}}_{I_1},{{{\mathfrak {m}}}}_{I_p}] \subset {{{\mathfrak {m}}}}_{I_p}\). Thus our equality holds. \(\square \)

Here we provide a brief overview of representation theory of compact Lie groups K. Let V be a vector space over \({{{\mathbb {C}}}}\). We call a representation \(\rho _{{{{{\mathbb {C}}}}}}:K \rightarrow {\text {U}}(V) \subset \textrm{Gl}(V,{{{\mathbb {C}}}})\) an irreducible complex representation if there exists no proper, non-trivial \({{{\mathbb {C}}}}\)-vector subspace \({\tilde{V}}\) of V which is \(\rho _{{{{{\mathbb {C}}}}}}(K)\)-invariant. The realization of an irreducible complex representation \(\rho _{{{{{\mathbb {C}}}}}}:K \rightarrow {\text {U}}(V,{{{\mathbb {C}}}})\) is given by \(\rho _{{{{{\mathbb {R}}}}}}:K\rightarrow {\text {SO}}(W) \subset \text {Gl} (W,{{{\mathbb {R}}}})\,;\,\, k \mapsto \rho _{{{{{\mathbb {C}}}}}}(k)\), where W is the \({{{\mathbb {R}}}}\)-vector space induced by V; that is, \(W=V\) as sets, but \(\dim _{{{{{\mathbb {R}}}}}} W =2 \dim _{{{{{\mathbb {C}}}}}} V\). Such representations are called real representations. Notice that the realization of an irreducible complex representation may not be irreducible.

An irreducible real representation \(\rho _{{{{{\mathbb {R}}}}}}:K \rightarrow {\text {SO}}(W)\) is called of real type, complex type or quaternionic type, respectively, if the group of intertwining operators

$$\begin{aligned} G(W,K):=\{ A \in \text {End} (W,{{{\mathbb {R}}}}) \mid A \cdot \rho _{{{{{\mathbb {R}}}}}}(k) = \rho _{{{{{\mathbb {R}}}}}}(k) \cdot A \text { for all } k\in K\} \end{aligned}$$

is isomorphic to \({{{\mathbb {R}}}}\), \({{{\mathbb {C}}}}\) or \({{{\mathbb {H}}}}\), respectively: see [21, II, 6.2, II, Theorem 6.7].

Lemma 7.9

Let \(\rho _{{{{{\mathbb {C}}}}}}:{\text {SU}}(n)\rightarrow {\text {U}}(V,{{{\mathbb {C}}}})\) be an irreducible complex representation, \(n \ge 2\). Suppose that \(\rho _{{{{{\mathbb {C}}}}}}\) can be extended to a representation of \({\text {U}}(n)\) such that the center of \({\text {U}}(n)\) acts non-trivially on V by multiples of the identity. Then the realization \(\rho _{{{{{\mathbb {R}}}}}}:{\text {U}}(n)\rightarrow {\text {SO}}(W)\), \(W=V\), is an irreducible real representation.

Proof

For a contradiction, suppose that \({\tilde{W}}\) is a non-trivial proper real subspace of W which is \(\rho _{{{{{\mathbb {R}}}}}}({\text {U}}(n))\)-invariant. By assumption there exists some \(\alpha \in {{{\mathbb {R}}}}\) such that for every \(e^{i\varphi } \cdot I_n\) in the center of \({\text {U}}(n)\), we have \(\rho _{{{{{\mathbb {C}}}}}}(e^{i\varphi } \cdot I_n)=e^{i\varphi \alpha }\cdot I_V\). Thus whenever \({\tilde{w}} \in {\tilde{W}}\), then \(i\cdot {\tilde{w}} \in {\tilde{W}}\) as well. Consequently \({\tilde{W}}\) is an \({\text {SU}}(n)\)-invariant complex subspace of \(W=V\), contradicting our hypothesis. Consequently W is an irreducible real representation. \(\square \)

1.3 Lie-theory II

In this section we describe the Casimir operator of the isotropy representation of a compact homogeneous space G/H and its relation to the Killing form and the structure constants.

Let \((Z_m)_{1\le m \le \dim H}\) denote a Q-orthonormal basis of \({{{\mathfrak {h}}}}\). Let \({{{\mathfrak {m}}}}_i\) be an \({\text {Ad}}(H)\)-irreducible summand of \({{{\mathfrak {m}}}}\). The Casimir operator on \({{{\mathfrak {m}}}}_i\) is given by

$$\begin{aligned} C_{{{{{\mathfrak {m}}}}}_i,Q\vert _{{{{{\mathfrak {h}}}}}}}:=-\sum _{m}{{\text {ad}}}(Z_m)\circ {{\text {ad}}}(Z_m):{{{\mathfrak {m}}}}_i \rightarrow {{{\mathfrak {m}}}}_i \end{aligned}$$

and it satisfies

$$\begin{aligned} C_{{{{{\mathfrak {m}}}}}_i,Q\vert _{{{{{\mathfrak {h}}}}}}}=c_i\cdot {{\text {Id}}}_{{{{{\mathfrak {m}}}}}_i}. \end{aligned}$$

A short computation using (4.3) shows \(c_i\ge 0\). Clearly, \(c_i=0\) if and only if \({{{\mathfrak {m}}}}_i \subset {{{\mathfrak {m}}}}_0\): see Lemma 7.5.

Recall, the definitions of \(d_i\), \(b_i\) and [ijk] can be found in Sect. 4.1 and that these numbers depend on the decomposition \(f=\oplus _{i=1}^\ell {{{\mathfrak {m}}}}_i\) of \({{{\mathfrak {m}}}}\) choosen.

Lemma 7.10

[77] Let G/H be a compact homogeneous space and let f be a decomposition of \({{{\mathfrak {m}}}}\). Then for all \(1 \le i \le \ell \),

$$\begin{aligned} d_ib_i-\tfrac{1}{2}\sum _{j,k=1}^\ell [ijk] =2d_ic_i+\tfrac{1}{2}\sum _{j,k=1}^\ell [ijk] \ge 0. \end{aligned}$$

Moreover, the expression is zero if and only if \({{{\mathfrak {m}}}}_i \subset {{{\mathfrak {z}}}}({{{\mathfrak {g}}}})\).

Proof

The equality is proved in [77, La.1.5].

Suppose that \(i=1\) and the left-hand side is zero. Then \(c_1=0\), thus \([{{{\mathfrak {h}}}},{{{\mathfrak {m}}}}_1]=0\) and \({{{\mathfrak {m}}}}_1 \subset {{{\mathfrak {m}}}}_0\) by Lemma 7.4. Since \([11j]=0\) for all \(1\le j\le \ell \), the module \({{{\mathfrak {m}}}}_1\) is an abelian subalgebra of \({{{\mathfrak {m}}}}_0\). Moreover, since for all \(2 \le j,k\le \ell \) we have \([1jk]=0\), the subspace \({{{\mathfrak {g}}}}':={{{\mathfrak {h}}}}\oplus (\bigoplus _{j=2}^\ell {{{\mathfrak {m}}}}_j)\) is a subalgebra of \({{{\mathfrak {g}}}}\) commuting with \({{{\mathfrak {m}}}}_1\): see Lemma 5.21. It follows that \({{{\mathfrak {m}}}}_1 \subset {{{\mathfrak {z}}}}({{{\mathfrak {g}}}})\).

The converse direction is clear. \(\square \)

Lemma 7.11

Let \({{{\mathfrak {k}}}}\in \mathcal {K}\) and let f be a decomposition of \({{{\mathfrak {m}}}}\) such that \({{{\mathfrak {k}}}}\) is f-adapted, and let \(I_1=I^{{{{\mathfrak {k}}}}}_1\). Then

$$\begin{aligned} a_{{{{\mathfrak {k}}}}}:=\sum _{j\in I_1}d_jc_j+\tfrac{1}{4}[I_1I_1I_1] \ge 0. \end{aligned}$$

Furthermore, \({{{\mathfrak {k}}}}\) is toral if and only if \(a_{{{{\mathfrak {k}}}}} =0\).

Proof

Suppose that \({{{\mathfrak {k}}}}={{{\mathfrak {h}}}}\oplus {{{\mathfrak {a}}}}\) is a toral subalgebra of \({{{\mathfrak {g}}}}\), that is, \({{{\mathfrak {a}}}}\subset {{{\mathfrak {m}}}}_0\). Then clearly \(a_{{{{\mathfrak {k}}}}}=0\). Conversely, when \(a_{{{{\mathfrak {k}}}}}=0\), then \(\sum _{j\in I_1}d_jc_j=0\), so we know \({{{\mathfrak {m}}}}_{I_1}:={{{\mathfrak {m}}}}_{{{{\mathfrak {k}}}}}\subset {{{\mathfrak {m}}}}_0\). Because \({{{\mathfrak {k}}}}\) is a subalgebra we conclude \([{{{\mathfrak {m}}}}_{I_1},{{{\mathfrak {m}}}}_{I_1}]\subset {{{\mathfrak {m}}}}_{I_1}\). Since \([I_1I_1I_1]=0\), the claim follows. \(\square \)

Corollary 7.12

For every a compact homogeneous space G/H, there exists a constant \(n_{G/H}>0\) such that the following holds: given any non-toral subalgebra \({{{\mathfrak {k}}}}\), with f a decomposition of \({{{\mathfrak {m}}}}\) such that \({{{\mathfrak {k}}}}\) is f-adapted, and letting \(I_1=I^{{{{\mathfrak {k}}}}}_1\), then

$$\begin{aligned} a_{{{{\mathfrak {k}}}}}=\sum _{j\in I_1}d_j c_j+\tfrac{1}{4}[I_1I_1I_1] \ge n_{G/H}. \end{aligned}$$

Proof

For a contradiction, suppose that there exists a sequence of non-toral subalgebras \(({{{\mathfrak {k}}}}_\alpha )_{\alpha \in {{{\mathbb {N}}}}}\) and a sequence \((f_\alpha )_{\alpha \in {{{\mathbb {N}}}}}\) of decompositions of \({{{\mathfrak {m}}}}\), such that for each \(\alpha \in {{{\mathbb {N}}}}\), \({{{\mathfrak {k}}}}_\alpha \) is \(f_\alpha \)-adapted, with \(a_{{{{\mathfrak {k}}}}_\alpha } \rightarrow 0\) as \(\alpha \rightarrow \infty \). Because the Casimir operator \(C\vert _{{{{\mathfrak {m}}}}}:{{{\mathfrak {m}}}}\rightarrow {{{\mathfrak {m}}}}\) is non-negative with kernel \({{{\mathfrak {m}}}}_0\), for \(\alpha \) large we must have \({{{\mathfrak {m}}}}_{{{{{\mathfrak {k}}}}}_\alpha }\subset {{{\mathfrak {m}}}}_0\). But since \({{{\mathfrak {m}}}}_0\) and \({{{\mathfrak {k}}}}_\alpha \) are subalgebras of \({{{\mathfrak {g}}}}\), so is \({{{\mathfrak {k}}}}_\alpha ^0:={{{\mathfrak {m}}}}_{{{{{\mathfrak {k}}}}}_\alpha }={{{\mathfrak {m}}}}\cap {{{\mathfrak {k}}}}_{\alpha }\). Since each \({{{\mathfrak {k}}}}_\alpha \) is non-toral by assumption, \({{{\mathfrak {k}}}}_\alpha ^0\) is a non-abelian subalgebra of \({{{\mathfrak {m}}}}_0\) and consequently, \({{{\mathfrak {k}}}}_\alpha \) has a nonvanishing semisimple part \(({{{\mathfrak {k}}}}_\alpha ^0)^{ss}\). Passing to a subsequence, we may assume that these semisimple subalgebras converge to a limit subalgebra \({{{\mathfrak {k}}}}_\infty ^0\) of \({{{\mathfrak {m}}}}_0 \subset {{{\mathfrak {g}}}}\), and also that \(f_\alpha \rightarrow f_\infty \) as \(\alpha \rightarrow \infty \). Up to conjugation, \({{{\mathfrak {g}}}}\) has only finitely many semisimple subalgebras, see [17, Corollary 4.5]; thus \({{{\mathfrak {k}}}}_\infty ^0\) must be semisimple. Hence \([I_1I_1I_1]_{f_\infty }>0\), a contradiction. \(\square \)

Corollary 7.13

Let G/H be a compact homogeneous space with finite fundamental group and let f be a decomposition of \({{{\mathfrak {m}}}}\) into \({\text {Ad}}(H)\)-irreducible summands. For each \(1\le i \le \ell \),

$$\begin{aligned} d_ib_i-\tfrac{1}{2}\sum _{j,k=1}^\ell [ijk]>0, \end{aligned}$$

and consequently

$$\begin{aligned} b_{G/H}:=\sum _{i=1}^{\ell }d_ib_i> 0. \end{aligned}$$
(7.3)

Proof

By Lemma 7.10 we know if \(2d_ic_i+\frac{1}{2}\sum _{j,k=1}^\ell [ijk]=0\), then \({{{\mathfrak {m}}}}_i\) is an abelian subalgebra in \({{{\mathfrak {g}}}}\) and \({{{\mathfrak {m}}}}_i\) commutes with the subalgebra \({{{\mathfrak {h}}}}\oplus (\bigoplus _{j=1,j\ne i}^\ell {{{\mathfrak {m}}}}_j)\). This would imply that G/H has infinite fundamental group, see Lemma 5.21, contradicting our hypothesis. This proves the claim. \(\square \)

1.4 Lie theory III: A Łojasiewicz inequality for structure constants

We recall the Łojasiewicz inequality in semi-algebraic geometry (cf. [7, Proposition 2.3.11]): Let \(K \subset {{{\mathbb {R}}}}^n\) be a compact semi-algebraic set and let \(f,g:K \rightarrow {{{\mathbb {R}}}}\) be continuous and semi-algebraic. Suppose that for all \(x \in K\) we have: \(f(x)=0 \Rightarrow g(x)=0\). Thus \(f^{-1}(0) \subset g^{-1}(0)\). Then there exists \(N\in {{{\mathbb {N}}}}\) and \(C=C(f,g,K)\ge 0\) such that for all \(x \in K\) we have

$$\begin{aligned} \vert g(x)\vert ^N \le C\cdot \vert f(x) \vert . \end{aligned}$$

The Łojasiewicz inequality stated in Proposition 7.14 provides the key estimate for the proof of Theorem 4.10. Note that standard estimation methods cannot be applied to the above situation.

Let \(I^{\textsf {c}}=\{1,\ldots ,\dim {{{\mathfrak {g}}}}\}\backslash I\).

Proposition 7.14

Let G be a compact, connected Lie group endowed with a biinvariant metric Q. Let \(I\subset \{1,\ldots ,\dim \,{{{\mathfrak {g}}}}\}\) with \(1\in I\). Let \(Z_{\vert I\vert }\) denote the set of all Q-orthonormal bases \(b:=(e_1,\ldots ,e_{\dim {{{{\mathfrak {g}}}}}})\) of \(({{{\mathfrak {g}}}},Q)\) such that

$$\begin{aligned} {{{\mathfrak {t}}}}:=\oplus _{i\in I} \langle e_i \rangle \end{aligned}$$

is an abelian subalgebra. Then there exists a constant \(C=C(G,Q)>0\) and an open neighborhood \(U_{\vert I\vert }\) of \(Z_{\vert I\vert }\) in the space of all Q-orthonormal bases, such that for all \(\tilde{b}:=({\tilde{e}}_1,\ldots ,{\tilde{e}}_{\dim {{{{\mathfrak {g}}}}}}) \in U_{\vert I\vert }\) we have

$$\begin{aligned} \sum _{j,k \in I}Q([{\tilde{e}}_1,{\tilde{e}}_j],{\tilde{e}}_k)^2 \le C \cdot \!\!\!\!\sum _{j\in I,k \in I^{\textsf {c}}} Q([{\tilde{e}}_1,{\tilde{e}}_j],\tilde{e}_k)^2. \end{aligned}$$
(7.4)

Proof

First note that if \(I^{\textsf {c}}=\emptyset \), then \({{{\mathfrak {g}}}}\) is abelian and the above claim is trivially true. Hence we may assume that \(I^{\textsf {c}}\ne \emptyset \) and that \({{{\mathfrak {g}}}}\) is not abelian. We prove the result by induction on \(\vert I\vert \). It is clear that for \(\vert I\vert =1,2\) the above claim is true, since the left hand side of (7.4) is zero by \({\text {Ad}}\)-invariance of Q. Thus we may assume that \(\vert I\vert \ge 3\).

We first show that for an arbitrary choice of Q-orthonormal basis \(b \in Z_{\vert I\vert }\) and an arbitrary choice of sequence \((b_\alpha )_{\alpha \in {{{\mathbb {N}}}}}:=(e_1(\alpha ),\ldots ,e_{\dim {{{{\mathfrak {g}}}}}}(\alpha ))_{\alpha \in {{{\mathbb {N}}}}}\) of Q-orthonormal bases for \({{{\mathfrak {g}}}}\) with \(\lim _{\alpha \rightarrow \infty }b_\alpha =b\), such a constant \(C>0\) exists (a priori, C may depend on b and the sequence \((b_\alpha )\)).

For a contradiction, suppose that no such constant \(C>0\) exists. In what follows we will pass to a subsequence whenever convenient, without explicitly mentioning it. For \(1 \le j,k \le \dim {{{\mathfrak {g}}}}\), we set

$$\begin{aligned} {[}1jk]_\alpha :=Q([e_1(\alpha ),e_j(\alpha )], e_k(\alpha ))^2 . \end{aligned}$$

We set \(I':= I\backslash \{1\}\) and get

$$\begin{aligned} \sum _{j,k \in I'} [1jk]_\alpha> & {} g(\alpha ) \cdot \sum _{j\in I',k \in I^{\textsf {c}}} [1jk]_\alpha \end{aligned}$$
(7.5)

where \(g:{{{\mathbb {R}}}}\rightarrow {{{\mathbb {R}}}}\) with \(\lim _{\alpha \rightarrow \infty }g(\alpha )=+\infty \).

Suppose that for each \(\alpha \in {{{\mathbb {N}}}}\), \(e_1(\alpha )\in {{{\mathfrak {t}}}}(\alpha )\), where \({{{\mathfrak {t}}}}(\alpha )\) denotes a maximal abelian subalgebra of \({{{\mathfrak {g}}}}\). Since maximal abelian subalgebras of \({{{\mathfrak {g}}}}\) are conjugate, there exists a maximal torus \({{{\mathfrak {t}}}}(\infty )\) and a sequence \((g_\alpha )_{\alpha \in {{{\mathbb {N}}}}}\) of group elements in G with \(\textrm{Ad}(g_\alpha )({{{\mathfrak {t}}}}(\alpha ))={{{\mathfrak {t}}}}(\infty )\). We obtain a sequence \((b_\alpha )\) of Q-orthonormal bases of \({{{\mathfrak {g}}}}\) converging to b, with \(e_1(\alpha )\in {{{\mathfrak {t}}}}(\infty )\) for all \(\alpha \in {{{\mathbb {N}}}}\), which we again denote by \(b_\alpha =(e_1(\alpha ),\ldots , e_{\dim {{{{\mathfrak {g}}}}}}(\alpha ))\). Since all terms in (7.5) are \(\textrm{Ad}(G)\)-invariant, Eq. (7.5) holds for each \(b_\alpha \) in this sequence as well.

Step 1: Let \(V_1(\alpha ):=\langle e_2(\alpha ),\ldots , e_{\vert I\vert }(\alpha )\rangle _{{{{{\mathbb {R}}}}}}\), and let \(V_1(\alpha )^\perp \) denote the Q-orthogonal complement of \(V_1(\alpha )\) in \({{{\mathfrak {g}}}}\). Let \(j\in I'\). Since \(e_1(\alpha )\in {{{\mathfrak {t}}}}(\infty )\) we have

$$\begin{aligned}{}[e_1(\alpha ),e_j(\alpha )]= & {} \text {pr} _{V_1(\alpha )} ([e_1(\alpha ),e_j(\alpha )]) + \text {pr} _{V_1(\alpha )^\perp }([e_1(\alpha ),e_j(\alpha )])\\= & {} \sum _{k\in I'\backslash \{j\}}s_{jk}(\alpha ) \cdot e_k(\alpha )+ \textrm{pr}_{V_1(\alpha )^\perp }([e_1(\alpha ),e_j(\alpha )]) \,\,\,\in \,\,\,{{{\mathfrak {t}}}}(\infty )^\perp . \end{aligned}$$

As mentioned above, we prove our proposition by induction on \(\vert I\vert \). Suppose that, after renumbering, \(I=\{1,\ldots ,\vert I\vert \}\), and that (after passing to a subsequence)

$$\begin{aligned} \vert s_{23}(\alpha )\vert \ge \vert s_{jk}(\alpha )\vert \end{aligned}$$

for all \(j,k\in I'\) and all \(\alpha \in {{{\mathbb {N}}}}\). By (7.5) we know that \(\vert s_{23}(\alpha )\vert >0\) for all sufficiently large \(\alpha \). Thus, for all \(j,k\in I'\)

$$\begin{aligned} \lim _{\alpha \rightarrow \infty }\tfrac{\vert s_{jk}(\alpha ) \vert }{\vert s_{23}(\alpha )\vert }\in [0,1]. \end{aligned}$$
(7.6)

Note that \(\lim _{\alpha \rightarrow \infty } s_{23}(\alpha )=0\), since \({{{\mathfrak {t}}}}=\oplus _{i=1}^{\vert I \vert }\langle e_i \rangle \) is abelian.

By (7.5) we claim that for each \(j\in I'\) there exists some \(k\in I'\) such that the limit in (7.6) is positive. For \(\vert I\vert =3\) this is trivially true. To see the claim, suppose that there exists \(j_0 \in I'\backslash \{2,3\}\) such that this limit is zero for all \(k\in I'\). It follows that there exists a function \({\tilde{g}}:{{{\mathbb {R}}}}\rightarrow {{{\mathbb {R}}}}\) with \({\displaystyle \lim _{\alpha \rightarrow \infty } {\tilde{g}}(\alpha )=+\infty }\) and

$$\begin{aligned} \tfrac{1}{ \vert I\vert } \cdot [123]_\alpha = \tfrac{1}{ \vert I\vert }\cdot \vert s_{23}(\alpha ) \vert ^2 \ge \vert s_{j_0k}(\alpha )\vert ^2 \cdot {\tilde{g}}(\alpha ) =[1j_0k]_\alpha \cdot {\tilde{g}}(\alpha ) \end{aligned}$$
(7.7)

for all \(k \in I'\) and all \(\alpha \in {{{\mathbb {N}}}}\). Let \(I_{j_0}:= I \backslash \{j_0\}\) and \(I'_{j_0}:= I'\backslash \{j_0\}\). Then

$$\begin{aligned} \sum _{j,k \in I'} [1jk]_\alpha =\sum _{j,k \in I'_{j_0}} [1jk]_\alpha +2 \sum _{k \in I'_{j_0}}[1j_0k]_\alpha , \end{aligned}$$

and

$$\begin{aligned} \sum _{j\in I', k\in I^{\textsf {c}}} [1jk]_\alpha= & {} \sum _{j \in I_{j_0}', k \in I^{\textsf {c}}} [1jk]_\alpha + \sum _{k\in I^{\textsf {c}}} [1j_0k]_\alpha \\= & {} \sum _{j \in I_{j_0}', k \in (I_{j_0})^\textsf {c}} [1jk]_\alpha + \sum _{k\in I^{\textsf {c}}} [1j_0k]_\alpha - \sum _{j \in I'_{j_0}} [1jj_0]_\alpha . \end{aligned}$$

We deduce from (7.5)

$$\begin{aligned} \sum _{j,k \in I'_{j_0}} [1jk]_\alpha + (2 +g(\alpha )) \cdot \sum _{k \in I'_{j_0}} [1j_0k]_\alpha> & {} g(\alpha ) \cdot \sum _{j \in I'_{j_0}, k \in (I_{j_0})^\textsf {c}} [1jk]_\alpha . \end{aligned}$$

By (7.7) this yields

$$\begin{aligned} {\tilde{g}}(\alpha ) \cdot \sum _{j,k \in I'_{j_0}} [1jk]_\alpha + (2 + g(\alpha )) \cdot [123]_\alpha> & {} {\tilde{g}}(\alpha ) \cdot g(\alpha ) \cdot \sum _{j \in I'_{j_0}, k \in (I_{j_0})^\textsf {c}} [1jk]_\alpha . \end{aligned}$$

Since \(\sum _{j,k \in I'_{j_0}} [1jk]_\alpha \ge [123]_\alpha \), using \(j_0 \ne 2,3\), this yields

$$\begin{aligned} \sum _{j,k \in I'_{j_0}} [1jk]_\alpha> & {} \tfrac{{\tilde{g}}(\alpha ) \cdot g(\alpha ) }{{\tilde{g}}(\alpha )+2+g(\alpha )} \cdot \sum _{j \in I'_{j_0}, k \in (I_{j_0})^\textsf {c}} [1jk]_\alpha . \end{aligned}$$

This shows that for the torus \({\displaystyle {{{\mathfrak {t}}}}_{j_0}:=\oplus _{i\in I_{j_0}} \langle e_i \rangle }\) of dimension \(\vert I\vert -1\) we obtain an inequality like (7.5), because as \(\alpha \rightarrow \infty \),

$$\begin{aligned} \tfrac{{\tilde{g}}(\alpha ) \cdot g(\alpha ) }{\tilde{g}(\alpha )+2+g(\alpha )} \rightarrow +\infty . \end{aligned}$$

By the induction hypothesis, we obtain a contradiction. This proves for each \(j\in I'\) there exists some \(k\in I'\) such that the limit in (7.6) is positive.

Step 2: For every \(j\in I'\) there exists some \(k(j)\in I'\backslash \{j\}\) with

$$\begin{aligned} \vert s_{jk(j)}(\alpha )\vert \ge \vert s_{jk}(\alpha )\vert \end{aligned}$$

for all \(k\in I'\) and \(\alpha \in {{{\mathbb {N}}}}\). By Step 1 we have

$$\begin{aligned} \lim _{\alpha \rightarrow \infty } \tfrac{\vert s_{jk(j)}(\alpha )\vert }{\vert s_{23} (\alpha )\vert } >0\,, \end{aligned}$$
(7.8)

which yields \(\vert s_{jk(j)}(\alpha )\vert >0\) for all sufficiently large \(\alpha \in {{{\mathbb {N}}}}\). Thus we obtain, for all \(j \in I'\),

$$\begin{aligned}{}[e_1(\alpha ),e_j(\alpha )]= s_{jk(j)}(\alpha )\cdot (E_j^1(\alpha )+X_j^1(\alpha ))\in {{{\mathfrak {t}}}}(\infty )^\perp , \end{aligned}$$
(7.9)

with

$$\begin{aligned} E_j^1(\alpha ):=\tfrac{ \textrm{pr}_{V_1(\alpha )}([e_1(\alpha ),e_j(\alpha )])}{ s_{jk(j)}(\alpha )} \quad \text { and } \quad X_j^1(\alpha ):=\tfrac{ \textrm{pr}_{V_1(\alpha )^\perp }([e_1(\alpha ),e_j(\alpha )])}{ s_{jk(j)}(\alpha )}. \end{aligned}$$

The left hand side of (7.5) is bounded from above by \(C' \cdot (s_{23}(\alpha ))^2\) by the very definition of \(s_{23}(\alpha )\), where \(C'\) is independent of \(\alpha \). Meanwhile, \(\Vert X_j^1(\alpha )\Vert ^2 \cdot s_{jk(j)}(\alpha )^2 \) contributes to the right hand side. Since by (7.8), the limit behavior of \(s_{jk(j)}(\alpha )\) is like that of \(s_{23}(\alpha )\), in order for the inequality in (7.5) to hold while \(g(\alpha ) \rightarrow \infty \), we must have

$$\begin{aligned} \lim _{\alpha \rightarrow \infty } X_j^1(\alpha )= 0. \end{aligned}$$
(7.10)

Next, let \({{{\mathfrak {c}}}}(e_1)\le {{{\mathfrak {g}}}}\) denote the centralizer of \(e_1\). Since \(e_1 \in {{{\mathfrak {t}}}}(\infty )\) and \([e_1,{{{\mathfrak {t}}}}']=0\), where \({\displaystyle {{{\mathfrak {t}}}}':=\oplus _{i\in I'} \langle e_j\rangle \subset {{{\mathfrak {t}}}}}\) is abelian, we obtain

$$\begin{aligned} {{{\mathfrak {t}}}}(\infty ) \subset {{{\mathfrak {t}}}}(\infty ) \oplus {{{\mathfrak {t}}}}' \subset {{{\mathfrak {c}}}}(e_1). \end{aligned}$$
(7.11)

Using that \(1 \le \Vert E_j^1(\alpha )\Vert \le C''\) by the definition of \(s_{jk(j)}(\alpha )\) this implies

$$\begin{aligned} \underbrace{E_j(\infty )}_{\ne 0}:= \lim _{\alpha \rightarrow \infty } E_j^1(\alpha )=\lim _{\alpha \rightarrow \infty } (E_j^1(\alpha )+X_j^1(\alpha )) \in {{{\mathfrak {c}}}}(e_1)\cap {{{\mathfrak {t}}}}(\infty )^\perp \cap {{{\mathfrak {t}}}}'. \end{aligned}$$
(7.12)

This is seen as follows: By (7.9) we have \(E_j^1(\alpha )+X_j^1(\alpha ) \in {{{\mathfrak {t}}}}(\infty )^\perp \). Because we know \(\lim _{\alpha \rightarrow \infty }X_j^1(\alpha )=0\) this implies \(E_j(\infty ) \in {{{\mathfrak {t}}}}(\infty )^\perp \). Moreover, we have \(E_j^1(\alpha )\in V_1(\alpha )\) by definition. Since \(V_1(\alpha )\rightarrow {{{\mathfrak {t}}}}'\) as \(\alpha \rightarrow \infty \), we deduce \(E_j(\infty )\in {{{\mathfrak {t}}}}'\), so it follows that \(E_j(\infty ) \in {{{\mathfrak {c}}}}(e_1)\) by (7.11).

An important consequence is that

$$\begin{aligned} {{{\mathfrak {t}}}}(\infty ) \subsetneq {{{\mathfrak {t}}}}(\infty ) \oplus {{{\mathfrak {t}}}}' \subset {{{\mathfrak {c}}}}(e_1). \end{aligned}$$
(7.13)

In fact we have that

$$\begin{aligned} e_2,\ldots ,e_{\vert I\vert }, E_2(\infty ),\ldots ,E_{\vert I\vert }(\infty ) \in {{{\mathfrak {c}}}}(e_1). \end{aligned}$$

Step 3: We show by induction on p, for \(1 \le p \le \dim {{{\mathfrak {t}}}}(\infty )\), that \({{{\mathfrak {t}}}}(\infty )\subsetneq {{{\mathfrak {c}}}}({{{\mathfrak {t}}}}(\infty ))\), which is clearly impossible. This will prove our contradiction for the specific choice of Q-orthonormal basis \(b \in Z_{\vert I\vert }\) and sequence \((b_\alpha )_{\alpha \in {{{\mathbb {N}}}}}:=(e_1(\alpha ),\ldots ,e_{\dim {{{{\mathfrak {g}}}}}}(\alpha ))_{\alpha \in {{{\mathbb {N}}}}}\) of Q-orthonormal bases for \({{{\mathfrak {g}}}}\) with \(\lim _{\alpha \rightarrow \infty }b_\alpha =b\). Our induction hypotheses for each p, denoted by \((A1)_p\),...,\((A4)_p\), are defined here:

Assumption \((A1)_p\): There exist a Q-orthonormal frame \((e_1^1:=e_1,\ldots ,e_1^p)\) in \({{{\mathfrak {t}}}}(\infty )\), positive numbers \(d_1(\alpha ),\ldots ,d_p(\alpha )>0\), and \(r_1^{p+1}(\alpha )\in {{{\mathfrak {t}}}}(\infty )\), \(r_1^{p+1}(\alpha ) \perp e_1^1,\ldots ,e_1^p\), such that

$$\begin{aligned} e_1(\alpha )=d_1(\alpha )\cdot e_1^1+\cdots +d_p(\alpha )\cdot e_1^p +r_1^{p+1}(\alpha ). \end{aligned}$$
(7.14)

Before we define \((A2)_p\), we set \({{{\mathfrak {c}}}}(0):={{{\mathfrak {g}}}}\) and

$$\begin{aligned} {{{\mathfrak {c}}}}(p):= \bigcap _{q=1}^p {{{\mathfrak {c}}}}(e_1^q). \end{aligned}$$

For \(j \in I'\) let

$$\begin{aligned} e_j^1=e_j \quad \text { and } \quad e_j^{p}(\alpha )= \textrm{pr}_{{{{\mathfrak {c}}}}(p-1)}(e_j(\alpha )). \end{aligned}$$

Let

$$\begin{aligned} V_p(\alpha ):=\langle e_j^p(\alpha ),\ldots , e_{\vert I\vert }^p(\alpha )\rangle _{{{{{\mathbb {R}}}}}} \end{aligned}$$

and let \(V_p(\alpha )^\perp \) denote the Q-orthogonal complement of \(V_p(\alpha )\) in \({{{\mathfrak {c}}}}(p-1)\). Let

$$\begin{aligned} E_j^p(\alpha ):=\tfrac{ \textrm{pr}_{V_p(\alpha )}([e_1(\alpha ),e_j^p(\alpha )])}{ s_{jk(j)}(\alpha )} \quad \text { and } \quad X_j^p(\alpha ):=\tfrac{ \textrm{pr}_{V_p(\alpha )^\perp }([e_1(\alpha ),e_j^p(\alpha )])}{ s_{jk(j)}(\alpha )}, \end{aligned}$$

recalling, as we showed in Step 2, that \(s_{jk(j)}(\alpha )>0\) for all sufficiently large \(\alpha \in {{{\mathbb {N}}}}\).

Assumption \((A2)_p\): We have

$$\begin{aligned} {{{\mathfrak {t}}}}(\infty )\subsetneq {{{\mathfrak {t}}}}(\infty ) \oplus {{{\mathfrak {t}}}}' \subset {{{\mathfrak {c}}}}(p). \end{aligned}$$
(7.15)

Notice that because \(e_1(\alpha )\in {{{\mathfrak {t}}}}(\infty )\subset {{{\mathfrak {c}}}}(p)\subset {{{\mathfrak {c}}}}(p-1)\), we have

$$\begin{aligned} \text {pr} _{{{{\mathfrak {c}}}}(p-1)}([e_1(\alpha ),e_j(\alpha )])= [e_1(\alpha ),e_j^{p}(\alpha )]=s_{jk(j)}\cdot ( E^{p}_j(\alpha )+ X^{p}_j(\alpha )) \in {{{\mathfrak {c}}}}(p-1) \cap {{{\mathfrak {t}}}}(\infty )^\perp . \end{aligned}$$

Assumption \((A3)_p\): For every \(j \in I'\) we know \({\displaystyle \lim _{\alpha \rightarrow \infty }X_j^p(\alpha )=0}\) and

$$\begin{aligned} \lim _{\alpha \rightarrow \infty } (E^{p}_j(\alpha )+X_j^p(\alpha ))=E_j(\infty ) \in ({{{\mathfrak {c}}}}(p)\cap {{{\mathfrak {t}}}}'\cap {{{\mathfrak {t}}}}(\infty )^\perp )\backslash \{0\}. \end{aligned}$$
(7.16)

Assumption \((A4)_p\): For every \(j \in I'\), we have

$$\begin{aligned} \lim _{\alpha \rightarrow \infty } e_j^{p}(\alpha )=e_j \in {{{\mathfrak {c}}}}(p) \cap {{{\mathfrak {t}}}}'. \end{aligned}$$
(7.17)

The base case \(p=1\) of our induction was essentially established in Step 2: By definition we have \({\displaystyle \lim _{\alpha \rightarrow \infty }e_1(\alpha )=e_1}\), that is, we can write \(e_1(\alpha )=d_1(\alpha )\cdot e_1 + r_1^2(\alpha )\) with \(d_1(\alpha )>0\), \(r_1^2(\alpha )\in {{{\mathfrak {t}}}}(\infty )\) and \(e_1 \perp r_1^2(\alpha )\). Thus \((A1)_1\) holds. Next, \((A2)_1\) is established in (7.13). Equations (7.10) and (7.12) give \((A3)_1\). And then, because \(e_j^1(\alpha )=e_j(\alpha ) \rightarrow e_j \in {{{\mathfrak {t}}}}' \subset {{{\mathfrak {c}}}}(e_1)\), \((A4)_1\) is also clear.

We now prove the induction step \(p \rightarrow p+1\le \dim {{{\mathfrak {t}}}}(\infty )\): By the induction hypothesis \((A2)_p\) we have \({{{\mathfrak {t}}}}(\infty ) \subset {{{\mathfrak {c}}}}(p)\), thus \([{{{\mathfrak {t}}}}_\infty ,{{{\mathfrak {c}}}}(p)^\perp ]\subset {{{\mathfrak {c}}}}(p)^\perp \). For \(j \in I'\) we define \(e_j^{p+1}(\alpha ):=\textrm{pr}_{{{{{\mathfrak {c}}}}}(p)}(e_j(\alpha ))\). Then by \((A1)_p\) we have

$$\begin{aligned} \text {pr} _{{{{{\mathfrak {c}}}}}(p)}( [e_1(\alpha ),e_j(\alpha ) ] ) = [e_1(\alpha ),e_j^{p+1}(\alpha ) ] = [ r_1^{p+1}(\alpha ),e_j^{p+1}(\alpha ) ] \in {{{\mathfrak {c}}}}(p) \cap {{{\mathfrak {t}}}}(\infty )^\perp . \end{aligned}$$

Using \((A2)_p\), the equation after (7.15), and \({{{\mathfrak {c}}}}(p-1) \supset {{{\mathfrak {c}}}}(p)\), we see

$$\begin{aligned} {[}e_1(\alpha ),e_j^{p+1}(\alpha )]= & {} s_{jk(j)}(\alpha )\cdot \underbrace{ \text {pr} _{{{{{\mathfrak {c}}}}}(p)} (E_j^p(\alpha )+X_j^p(\alpha ))}_{=:\tilde{E}^{p+1}_j(\alpha )} \in {{{\mathfrak {c}}}}(p)\cap {{{\mathfrak {t}}}}(\infty )^\perp . \end{aligned}$$
(7.18)

By \((A3)_p\) we know

$$\begin{aligned} \lim _{\alpha \rightarrow \infty }X_j^p(\alpha )=0 \text { and } \lim _{\alpha \rightarrow \infty }\tilde{E}^{p+1}_j(\alpha )=E_j(\infty ) \in ({{{\mathfrak {c}}}}(p)\cap {{{\mathfrak {t}}}}' \cap {{{\mathfrak {t}}}}(\infty )^\perp )\backslash \{0\}. \end{aligned}$$
(7.19)

Since \(s_{jk(j)}(\alpha )>0\) we must have \(r_1^{p+1}(\alpha )\ne 0\) for all sufficiently large \(\alpha \in {{{\mathbb {N}}}}\). Set

$$\begin{aligned} e_1^{p+1}(\alpha ):=\tfrac{ r_1^{p+1}(\alpha ) }{\Vert r_1^{p+1}(\alpha )\Vert }. \end{aligned}$$

Then, by passing to a further subsequence, we may assume \({\displaystyle \lim _{\alpha \rightarrow \infty } e_1^{p+1}(\alpha )=e_1^{p+1} \in {{{\mathfrak {t}}}}(\infty )}\), with \(e_1^{p+1}\perp e_1,\ldots ,e_1^p\) and \(\Vert e_1^{p+1} \Vert =1\). This shows \((A1)_{p+1}\).

The induction hypothesis \((A4)_p\) guarantees for all \(j \in I'\),

$$\begin{aligned} \lim _{\alpha \rightarrow \infty } e_j^{p+1}(\alpha )=e_j \in {{{\mathfrak {c}}}}(p) \cap {{{\mathfrak {t}}}}' . \end{aligned}$$
(7.20)

Thus for all sufficiently large \(\alpha \in {{{\mathbb {N}}}}\),

$$\begin{aligned} V_{p+1}(\alpha ):=\langle e_2^{p+1}(\alpha ),\ldots , e_{\vert I\vert }^{p+1}(\alpha )\rangle _{{{{{\mathbb {R}}}}}} \end{aligned}$$

is a subspace of \({{{\mathfrak {c}}}}(p)\) of dimension \(\vert I'\vert \), converging to \({{{\mathfrak {t}}}}'\) as \(\alpha \rightarrow \infty \). Using (7.14), the definition of \({{{\mathfrak {c}}}}(p)\), and \(e_1^{p+1}(\alpha )\) we deduce from (7.18) for all \(j \in I'\)

$$\begin{aligned} {[}e_1^{p+1}(\alpha ),e_j^{p+1}(\alpha )]= \tfrac{s_{jk(j)}(\alpha )}{\Vert r_1^{p+1}(\alpha ) \Vert }\cdot ( E^{p+1}_j(\alpha )+ X^{p+1}_j(\alpha )) \in {{{\mathfrak {c}}}}(p) \cap {{{\mathfrak {t}}}}(\infty )^\perp , \end{aligned}$$
(7.21)

where \({\displaystyle E^{p+1}_j(\alpha ):= \text {pr} _{V_{p+1}(\alpha ) }(\tilde{E}_j^{p+1}(\alpha ))\in {{{\mathfrak {c}}}}(p)}\) and \({\displaystyle X^{p+1}_j(\alpha ):= \textrm{pr}_{V_{p+1}(\alpha )^\perp }(\tilde{E}_j^{p+1}(\alpha ))\in {{{\mathfrak {c}}}}(p)}\).

Notice, by Eq. (7.21),

$$\begin{aligned} {[}e_1(\alpha ),e_j^{p+1}(\alpha )]=s_{jk(j)}\cdot ( E^{p+1}_j(\alpha )+ X^{p+1}_j(\alpha )) \in {{{\mathfrak {c}}}}(p) \cap {{{\mathfrak {t}}}}(\infty )^\perp . \end{aligned}$$

Since by (7.19) \({\displaystyle \lim _{\alpha \rightarrow \infty }\tilde{E}^{p+1}_j(\alpha )=E_j(\infty )\in {{{\mathfrak {t}}}}'}\) and \(V_{p+1}(\alpha )\rightarrow {{{\mathfrak {t}}}}'\) as \(\alpha \rightarrow \infty \), we get

$$\begin{aligned} \lim _{\alpha \rightarrow \infty } X^{p+1}_j(\alpha )=0 \quad \text { and } \quad \lim _{\alpha \rightarrow \infty } E^{p+1}_j(\alpha )=E_j(\infty ). \end{aligned}$$

This shows the first part of \((A3)_{p+1}\). Next, we claim \({\displaystyle \lim _{\alpha \rightarrow \infty }\tfrac{s_{jk(j)}(\alpha )}{\Vert r_1^{p+1}(\alpha ) \Vert } =0}\). To see this, notice that \(\lim _{\alpha \rightarrow \infty }e_1^{p+1}(\alpha )=e_1^{p+1}\in {{{\mathfrak {t}}}}(\infty )\), that \(\lim _{\alpha \rightarrow \infty } e_j^{p+1}(\alpha ) =e_j \in {{{\mathfrak {t}}}}'\) by (7.20) and that \(\lim _{\alpha \rightarrow \infty } (E_j^{p+1}(\alpha )+X_j^{p+1}(\alpha ))=E_j(\infty ) \in {{{\mathfrak {t}}}}'\) for all \(j \in I'\). Now if \({\displaystyle \tfrac{s_{jk(j)}(\alpha )}{\Vert r_1^{p+1}(\alpha ) \Vert } }\) had a uniform positive lower bound, then by passing to a subsequence if necessary, then we deduce \(Q([e_1^{p+1},e_j],E_j(\infty ))>0\) by (7.21), because \(E_j(\infty )\ne 0\). But since \(e_j,E_j(\infty )\in {{{\mathfrak {t}}}}'\) and since \({{{\mathfrak {t}}}}'\) is abelian, the left-hand side equals zero. This shows the above claim.

By (7.21) we now deduce \([e_1^{p+1},e_j]=0\) for all \(j \in I'\), that is,

$$\begin{aligned} {{{\mathfrak {t}}}}(\infty ) \subset {{{\mathfrak {t}}}}(\infty ) \oplus {{{\mathfrak {t}}}}' \subset {{{\mathfrak {c}}}}(e_1^{p+1}). \end{aligned}$$

Using this and \(E_j(\infty )\in {{{\mathfrak {c}}}}(p)\cap {{{\mathfrak {t}}}}'\) we deduce

$$\begin{aligned} E_j(\infty ) \in {{{\mathfrak {c}}}}(p+1) \cap {{{\mathfrak {t}}}}(\infty )^\perp \cap {{{\mathfrak {t}}}}', \end{aligned}$$

where \({{{\mathfrak {c}}}}(p+1)={{{\mathfrak {c}}}}(p)\cap {{{\mathfrak {c}}}}(e_1^{p+1})\). By (7.20), this yields \((A2)_{p+1}\):

$$\begin{aligned} {{{\mathfrak {t}}}}(\infty )\subsetneq {{{\mathfrak {t}}}}(\infty ) \oplus {{{\mathfrak {t}}}}' \subset {{{\mathfrak {c}}}}(p+1). \end{aligned}$$

Notice that this implies

$$\begin{aligned} e_2,\ldots ,e_{\vert I\vert }, E_2(\infty ),\ldots ,E_{\vert I\vert }(\infty ) \in {{{\mathfrak {c}}}}(p+1), \end{aligned}$$

implying both the second part of \((A3)_{p+1}\) and also \((A4)_{p+1}\).

So far we have proved that for a given \(b\in Z_{\vert I\vert }\) and a given sequence \((b_\alpha )_{\alpha \in {{{\mathbb {N}}}}}\) converging to b, such a constant C exists for all sufficiently large \(\alpha \). We next show that C can be chosen independently of both \(b\in Z_{\vert I\vert }\) and the sequence \((b_\alpha )_{\alpha \in {{{\mathbb {N}}}}}\).

To this end, suppose that there exists a sequence \((b_\alpha )_{\alpha \in {{{\mathbb {N}}}}}\) in \(Z_{\vert I\vert }\) and corresponding sequences \((b_{\alpha ,\beta })_{\beta \in {{{\mathbb {N}}}}}\) with \(\lim _{\beta \rightarrow \infty }b_{\alpha ,\beta }=b_\alpha \) for all \(\alpha \in {{{\mathbb {N}}}}\), such that for the corresponding optimal constants, as \(\alpha \rightarrow \infty \), \(C_\alpha \rightarrow +\infty \). Here by optimal we mean that the constants \(C_\alpha -1\) would not work. Since the constants \(C_\alpha \) are chosen to be optimal, for each \(\alpha \in {{{\mathbb {N}}}}\) there exists \(N(\alpha ) \in {{{\mathbb {N}}}}\), such that we may assume (after passing to a subsequence) that

$$\begin{aligned} \sum _{j,k \in I'} [1jk]_{\alpha ,\beta }\ge & {} (C_\alpha -1) \cdot \sum _{j\in I',k \in I^{\textsf {c}}} [1jk]_{\alpha ,\beta } \end{aligned}$$
(7.22)

for all \(\beta \ge N(\alpha )\). By again passing to a subsequence, we may assume that this holds for all \(\alpha ,\beta \ge 1\). Since \(Z_{\vert I\vert }\) is compact, we may assume that \(\lim _{\alpha \rightarrow \infty }b_\alpha =b\). Moreover, it is clear that we can construct a sequence \(({\tilde{b}}_\alpha )_{\alpha \in {\textbf{N}} }\) converging to b and satisfying (7.22). Since \(C_\alpha \rightarrow + \infty \) as \(\alpha \rightarrow \infty \), we obtain a contradiction. This shows that C can be chosen independently.

To prove the theorem, suppose now that no such constant \(C=C(G,Q)\) and no such open neighborhood \(U_{\vert I\vert }\) of \(Z_{\vert I\vert }\) exist. Then there exists a sequence of bases \((b_\alpha )_{\alpha \in {{{\mathbb {N}}}}}\) converging to \(b \in Z_{\vert I\vert }\) for which (7.22) holds. Contradiction. \(\square \)

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

Böhm, C., Kerr, M.M. Homogeneous Einstein metrics and butterflies. Ann Glob Anal Geom 63, 29 (2023). https://doi.org/10.1007/s10455-023-09905-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10455-023-09905-0

Keywords

Mathematics Subject Classification

Navigation