Skip to main content

Warped Riemannian Metrics for Location-Scale Models

  • Chapter
  • First Online:
Geometric Structures of Information

Abstract

The present contribution shows that warped Riemannian metrics, a class of Riemannian metrics which play a prominent role in Riemannian geometry, are also of fundamental importance in information geometry. Precisely, the starting point is a new theorem, which states that the Rao–Fisher information metric of any location-scale model, defined on a Riemannian manifold, is a warped Riemannian metric, whenever this model is invariant under the action of some Lie group. This theorem is a valuable tool in finding the expression of the Rao–Fisher information metric of location-scale models defined on high-dimensional Riemannian manifolds. Indeed, a warped Riemannian metric is fully determined by only two functions of a single variable, irrespective of the dimension of the underlying Riemannian manifold. Starting from this theorem, several original results are obtained. The expression of the Rao–Fisher information metric of the Riemannian Gaussian model is provided, for the first time in the literature. A generalised definition of the Mahalanobis distance is introduced, which is applicable to any location-scale model defined on a Riemannian manifold. The solution of the geodesic equation, as well as an explicit construction of Riemannian Brownian motion, are obtained, for any Rao–Fisher information metric defined in terms of warped Riemannian metrics. Finally, using a mixture of analytical and numerical computations, it is shown that the parameter space of the von Mises–Fisher model of n-dimensional directional data, when equipped with its Rao–Fisher information metric, becomes a Hadamard manifold, a simply-connected complete Riemannian manifold of negative sectional curvature, for \(n = 2,\ldots ,8\). Hopefully, in upcoming work, this will be proved for any value of n.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 129.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 169.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Similar content being viewed by others

References

  1. Absil, P.A., Mahony, R., Sepulchre, R.: Optimization Algorithms on Matrix Manifolds. Princeton University Press, Princeton (2008)

    Book  Google Scholar 

  2. Amari, S.I.: Natural gradient works efficiently in learning. Neural Comput. 10(2), 251–276 (1998)

    Article  Google Scholar 

  3. Amari, S., Nagaoka, H.: Methods of Information Geometry. American Mathematical Society, Providence (2000)

    MATH  Google Scholar 

  4. Arnaudon, M.: Semi-martingales dans les espaces homogènes. Annales de l’I.H.P 29(3), 269–288 (1993)

    MATH  Google Scholar 

  5. Arnaudon, M., Miclo, L.: A stochastic algorithm finding generalized means on compact manifolds. Stochastic. Process. Appl. 124(10), 3463–3479 (2014)

    Article  MathSciNet  Google Scholar 

  6. Atkinson, C., Mitchell, A.: Rao’s distance measure. Sankhya Ser. A 43, 345–365 (1981)

    MathSciNet  MATH  Google Scholar 

  7. Ay, N., Jost, J., Lê, H.V., Schwachhöfer, L.: Information geometry and sufficient statistics. Probab. Theory Relat. Fields 162(1), 327–364 (2015)

    Article  MathSciNet  Google Scholar 

  8. Bensadon, J.: Black-box optimization using geodesics in statistical manifolds. Entropy 17(1), 304–345 (2015)

    Article  MathSciNet  Google Scholar 

  9. Bensadon, J.: Applications of information theory to machine learning. Ph.D. thesis, Université Paris-Saclay (2016). https://tel.archives-ouvertes.fr/tel-01297163

  10. Berthoumieu, Y., Bombrun, L., Said, S.: Classification approach based on the product of Riemannian manifolds from Gaussian parameterization space. In: International Conference on Image Processing (ICIP) (2017)

    Google Scholar 

  11. Bishop, R.L., O’NEILL, B.: Manifolds of negative curvature. Trans. Amer. Math. Soc. 145, 1–49 (1969)

    Article  MathSciNet  Google Scholar 

  12. Bottou, L., Curtis, F.E., Nocedal, J.: Optimization methods for large-scale machine learning (2017). arXiv:1606.04838v2

  13. Chavel, I.: Riemannian Geometry, A Modern Introduction. Cambridge University Press, Cambridge (2006)

    Book  Google Scholar 

  14. Cheng, G., Vemuri, B.C.: A novel dynamic system in the space of SPD matrices with applications to appearance tracking. SIAM J. Imaging Sci. 6(1), 592–615 (2013)

    Article  MathSciNet  Google Scholar 

  15. Chentsov, N.N.: Statistical Decision Rules and Optimal Inference. American Mathematical Society, Providence (1982)

    MATH  Google Scholar 

  16. Chikuse, Y.: Statistics on Special Manifolds. Lecture Notes in Statistics, 174. Springer Science+Business Media, LLC (2003)

    Book  Google Scholar 

  17. Dobarro, F., Unal, B.: Curvature of multiply warped products. J. Geom. Phys. 55, 75–106 (2005)

    Article  MathSciNet  Google Scholar 

  18. Do Carmo, M.P.: Riemannian Geometry, 2nd edn. Birkhauser, Boston (1992)

    Book  Google Scholar 

  19. Emery, M.: Stochastic Calculus in Manifolds. Springer, Berlin (1980)

    Google Scholar 

  20. Estrade, A.: Exponentielle stochastique et intégrale multiplicative discontinues. Annales de l’I.H.P. 28(1), 107–129 (1992)

    Google Scholar 

  21. Gallavotti, G.: The Elements of Mechanics. Springer Science+Business Media, LLC, New York (1983)

    Book  Google Scholar 

  22. Helgason, S.: Differential Geometry and Symmetric Spaces. Academic Press, New York (1962)

    MATH  Google Scholar 

  23. Hsu, E.P.: Stochastic Analysis on Manifolds. American Mathematical Society, Providence (2002)

    Book  Google Scholar 

  24. Ikeda, N., Watanabe, S.: Stochastic Differential Equations and Diffusion Processes. North-Holland Publishing Company, Amsterdam (1981)

    MATH  Google Scholar 

  25. Kallenberg, O.: Foundations of Modern Probability, 2nd edn. Springer, New York (2001)

    MATH  Google Scholar 

  26. Kloeden, P.E., Platen, E.: Numerical Solution of Stochastic Differential Equations. Springer, New York (2011)

    MATH  Google Scholar 

  27. Knapp, A.W.: Lie Groups Beyond an Introduction, 2nd edn. Birkhauser, Boston (2002)

    MATH  Google Scholar 

  28. Lang, S.: Introduction to Linear Algebra, 2nd edn. Springer, New York (1986)

    Book  Google Scholar 

  29. Lee, J.M.: Introduction to Smooth Manifolds, 2nd edn. Springer Science+Business Media, New York (2012)

    Book  Google Scholar 

  30. Lehmann, E.L., Romano, J.P.: Testing Statistical Hypotheses, 3rd edn. Springer Science+Business Media Inc, New York (2005)

    MATH  Google Scholar 

  31. Liao, E.: Lévy Processes on Lie Groups. Cambridge University Press, Cambridge (2004)

    Book  Google Scholar 

  32. Luenbeger, D.G.: Linear and Nonlinear Programming, 2nd edn. Addison-Wesley publishing company, Reading (1973)

    Google Scholar 

  33. Mardia, K.V., Jupp, P.E.: Directional Statistics. Wiley, Chichester (2002)

    MATH  Google Scholar 

  34. Marsden, J.E., Ratiu, T.: Manifolds, Tensor Analysis, and Applications. Springer Publishing Company Inc., New York (2001)

    MATH  Google Scholar 

  35. Martens, J.: New insights and perspectives on the natural gradient method (2017). arXiv:1412.1193v8

  36. McLachlan, G.J.: Discriminant Analysis and Statistical Pattern Recognition. Wiley, Hoboken (2004)

    MATH  Google Scholar 

  37. Ollivier, Y., Arnold, L., Auger, A., Hansen, N.: Information geometric optimization algorithms: a unifying picture via invariance principles. J. Mach. Learn. Res. 18(18), 1–65 (2017)

    MathSciNet  MATH  Google Scholar 

  38. O’Neill, B.: Semi-Riemannian Geometry with Applications to Relativity. Academic Press, San Diego (1983)

    MATH  Google Scholar 

  39. Pennec, X., Fillard, P., Ayache, N.: A Riemannian framework for tensor computing. Int. J. Comput. Vision 66(1), 41–66 (2006)

    Article  Google Scholar 

  40. Petersen, P.: Riemannian Geometry, 2nd edn. Springer Science+Business Media, LLC, New York (2006)

    MATH  Google Scholar 

  41. Said, S., Bombrun, L., Berthoumieu, Y., Manton, J.H.: Riemannian Gaussian distributions on the space of symmetric positive definite matrices. IEEE. Trans. Inf. Theory 63(4), 2153–2170 (2017)

    Article  MathSciNet  Google Scholar 

  42. Said, S., Hajri, H., Bombrun, L., Vemuri, B.C.: Gaussian distributions on Riemannian symmetric spaces: statistical learning with structured covariance matrices. IEEE. Trans. Inf. Theory 64(2), 752–772 (2018)

    Article  MathSciNet  Google Scholar 

  43. Spanier, E.H.: Algebraic Topology. Springer, New York (1966)

    MATH  Google Scholar 

  44. Terras, A.: Harmonic Analysis on Symmetric Spaces and Applications, vol. II. Springer, New York (1988)

    Book  Google Scholar 

  45. Unal, B.: Doubly warped products. Differ. Geom. Appl. 15, 253–263 (2001)

    Article  MathSciNet  Google Scholar 

  46. Watson, G.N.: A Treatise on the Theory of Bessel Functions. Cambridge University Press, Cambridge (1922)

    MATH  Google Scholar 

  47. Wierstra, D., Schaul, T., Glasmachers, T., Sun, Y., Peters, J., Schmidhuber, J.: Natural evolution strategies. J. Mach. Learn. Res. 15(1), 949–980 (2014)

    MathSciNet  MATH  Google Scholar 

  48. Young, G.A., Smith, R.L.: Essentials of Statistical Inference. Cambridge University Press, Cambridge (2005)

    Book  Google Scholar 

  49. Zanini, P., Said, S., Congedo, M., Berthoumieu, Y., Jutten, C.: Parameter estimates of Riemannian Gaussian distributions in the manifold of covariance matrices. In: Sensor Array and Multichannel Signal Processsing Workshop (SAM) (2016)

    Google Scholar 

  50. Zanini, P., Said, S., Berthoumieu, Y., Congedo, M., Jutten, C.: Riemannian online algorithm for estimating mixture model parameters. In: Geometric Science of Information (GSI) (2017)

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Salem Said .

Editor information

Editors and Affiliations

Appendices

Appendix A – Proof of Theorem 1

In order to complete the proof of Theorem 1, the following proposition is needed. The notation is that of Remark 4 and of (11).

Proposition 9

Assume condition (7) holds. Then,

$$\begin{aligned} \partial _\sigma \ell (z)\circ g \,=\, \partial _\sigma \ell (g^{-1}\!\cdot z) \quad \nabla _{\bar{x}}\,\ell (z)\circ g \,=\, dg_{\bar{x}}\,\nabla _{\bar{x}}\,\ell (g^{-1}\!\cdot z)\, \end{aligned}$$
(49a)

In particular, if \(g = s_{\bar{x}}\),

$$\begin{aligned} \partial _\sigma \ell (z)\circ s_{\bar{x}} \,=\, \partial _\sigma \ell (z) \quad \nabla _{\bar{x}}\,\ell (z)\circ s_{\bar{x}} \,=\, -\,\nabla _{\bar{x}}\,\ell (z) \end{aligned}$$
(49b)

Proof

Note that (49b) follows from (49a), by the definition of the geodesic-reversing isometry \(s_{\bar{x}}\) [22]. Indeed, \(s_{\bar{x}}\cdot \bar{x} = \bar{x}\) so \(s^{-1}_{\bar{x}}\cdot z = z\). Moreover, \(ds_{\bar{x}} = -\mathrm {Id}\), as a linear mapping of \(T_{\bar{x}}M\), where \(\mathrm {Id}\) denotes the identity. To prove (49a), note that

$$\begin{aligned} \left( \partial _\sigma \ell (z)\circ g\right) (x) \,=\, \partial _\sigma \log p(g\cdot x|z) \,=\, \partial _\sigma \log p(x|g^{-1}\!\cdot z) \end{aligned}$$
(50a)

where the second equality follows from condition (7). However,

$$\begin{aligned} \partial _\sigma \log p(x|g^{-1}\!\cdot z) \,=\, \partial _\sigma \ell (g^{-1}\!\cdot z)(x) \end{aligned}$$
(50b)

Replacing (50b) in (50a) gives,

$$ \left( \partial _\sigma \ell (z)\circ g\right) (x) \,=\, \partial _\sigma \ell (g^{-1}\!\cdot z)(x) $$

which is the first part of (49a). For the second part, a similar reasoning can be applied. Precisely, using condition (7), it follows,

$$\begin{aligned} \left( d\ell (z)\circ g\right) (x) \,=\, d\log p(g\cdot x|z) = d\log p(x|g^{-1}\!\cdot z) \,=\, d\ell ^{(g)}(z)(x) \end{aligned}$$
(51a)

where \(d\ell (z)\) denotes the derivative of \(\ell (z)\) with respect to \(\bar{x}\), and \(\ell ^{(g)}(z) = \ell (g^{-1}\!\cdot z)\), so (51a) implies that,

$$\begin{aligned} d\ell (z)\circ g \,=\, d\ell ^{(g)}(z) \end{aligned}$$
(51b)

By the chain rule [29], for \(u \in T_{\bar{x}}M\),

$$ \left. d\ell ^{(g)}(z)\right| _{\bar{x}}\, u \,=\, d\ell (g^{-1}\!\cdot z) \, dg^{-1}_{\bar{x}}\, u $$

Replacing in (51b),

$$\begin{aligned} \left. d\ell (z)\circ g\right| _{\bar{x}} \,=\, d\ell (g^{-1}\!\cdot z) \, dg^{-1}_{\bar{x}} \end{aligned}$$
(51c)

The second part of (49a) can now be obtained as follows. By the definition of the Riemannian gradient [40],

$$\begin{aligned} Q\left( \nabla _{\bar{x}}\,\ell (z)\circ g\,,u\right) \,=\, \left. d\ell (z)\circ g\right| _{\bar{x}}\,u \,=\, d\ell (g^{-1}\!\cdot z) \, dg^{-1}_{\bar{x}}\,u \end{aligned}$$
(52a)

where the second equality follows from (51c). However,

$$ d\ell (g^{-1}\!\cdot z) \, dg^{-1}_{\bar{x}}\,u \,=\, Q\left( \nabla _{\bar{x}}\,\ell (g^{-1}\!\cdot z), dg^{-1}_{\bar{x}}\,u\right) $$

Since g is an isometry of M, its derivative \(dg_{\bar{x}}\) preserves the Riemannian metric Q. Therefore,

$$\begin{aligned} Q\left( \nabla _{\bar{x}}\,\ell (g^{-1}\!\cdot z), dg^{-1}_{\bar{x}}\,u\right) \,=\, Q\left( dg_{\bar{x}}\,\nabla _{\bar{x}}\,\ell (g^{-1}\!\cdot z),\,u\right) \end{aligned}$$
(52b)

Replacing (52b) in (52a) gives,

$$ Q\left( \nabla _{\bar{x}}\,\ell (z)\circ g\,,u\right) \,=\, Q\left( dg_{\bar{x}}\,\nabla _{\bar{x}}\,\ell (g^{-1}\!\cdot z),\,u\right) $$

To finish the proof, it is enough to note that the vector u is arbitrary. \(\blacksquare \)

Proof of (10b): recall the polarisation identity, from elementary linear algebra [28], (see p. 29),

$$ I_z(\partial _\sigma ,u) = \frac{1}{4}\,I_z(\partial _\sigma +u,\partial _\sigma +u) - \frac{1}{4}\,I_z(\partial _\sigma -u,\partial _\sigma -u) $$

by replacing (8) into this identity, it can be seen that,

$$ I_z(\partial _\sigma ,u)\,=\,\mathbb {E}_z\left( \left( \partial _\sigma \ell (z)\right) \left( d\ell (z)\,u\right) \right) $$

Then, by recalling the definition of the Riemannian gradient [40],

$$\begin{aligned} I_z(\partial _\sigma ,u)\,=\,\mathbb {E}_z\left( \left( \partial _\sigma \ell (z)\right) Q\left( \nabla _{\bar{x}}\,\ell (z),u\right) \right) \end{aligned}$$
(53a)

Denote the function under the expectation by f, and apply (11) with \(g = s_{\bar{x}}\). Then,

$$\begin{aligned} \mathbb {E}_z\,f \,=\, \mathbb {E}_{s_{\bar{x}}\cdot z}\,f \,=\, \mathbb {E}_z\left( f\circ s_{\bar{x}}\right) \end{aligned}$$
(53b)

since \(s_{\bar{x}}\cdot z = z\). Note that (10b) amounts to saying that \(\mathbb {E}_z\,f = 0\). To prove this, note that

$$\begin{aligned} f\circ s_{\bar{x}} = \left( \partial _\sigma \ell (z)\circ s_{\bar{x}}\right) Q\left( \nabla _{\bar{x}}\ell (z)\circ s_{\bar{x}}\,,u\right) = -\partial _\sigma \ell (z) Q\left( \nabla _{\bar{x}}\ell (z),u\right) = - f \end{aligned}$$
(53c)

where the second equality follows from (49b). Replacing in (53b) shows that \(\mathbb {E}_z\,f = 0\). \(\blacksquare \)

Proof of (10c): the idea is to apply Schur’s lemma to \(I_z(u,u)\), considered as a symmetric bilinear form on \(T_{\bar{x}}M\). First, it is shown that this symmetric bilinear form is invariant under the isotropy representation. That is,

$$\begin{aligned} I_z(u,u) \,=\, I_z\left( dk_{\bar{x}}\,u\,,dk_{\bar{x}}\,u\right) \quad \text {for all } k \in K_{\bar{x}} \end{aligned}$$
(54a)

This is done using (11). Note from (8),

$$\begin{aligned} I_z(u,u) = \mathbb {E}_z\left[ \left( \,Q\left( \nabla _{\bar{x}}\,\ell (z),u\right) \right) ^2\right] \end{aligned}$$
(54b)

Denote the function under the expectation by f. By (11),

$$\begin{aligned} \mathbb {E}_z\,f \,=\, \mathbb {E}_{k^{-1}\cdot z}\,f \,=\, \mathbb {E}_z\left( f\circ k^{-1}\right) \end{aligned}$$
(54c)

since \(k^{-1}\cdot z = z\) for \(k \in K_{\bar{x}}\). To find \(f\circ k^{-1}\), note that,

$$ Q\left( \nabla _{\bar{x}}\,\ell (z)\circ k^{-1},u\right) \,= \,Q\left( dk^{-1}_{\bar{x}}\nabla _{\bar{x}}\,\ell (z),u\right) \,= \,Q\left( \nabla _{\bar{x}}\,\ell (z),dk_{\bar{x}}\,u\right) $$

where the first equality follows from (49a) and the fact that \(k\cdot \bar{x} = \bar{x}\), and the second equality from the fact that \(dk_{\bar{x}}\) preserves the Riemannian metric Q. Now, by (54b) and (54c),

$$ I_z(u,u) = \mathbb {E}_z\,f = \mathbb {E}_z\left( f\circ k^{-1}\right) = \mathbb {E}_z\left( Q\left( \nabla _{\bar{x}}\ell (z),dk_{\bar{x}}\,u\right) \right) ^2 = I_z\left( dk_{\bar{x}}\,u\,,dk_{\bar{x}}\,u\right) $$

and this proves (54a).

Recall Schur’s lemma, ([27], p. 240). Applied to (54a), this lemma implies that there exists some multiplicative factor \(\beta ^2\), such that

$$\begin{aligned} I_z(u,u) \,=\, \beta ^2\,Q_{\bar{x}}(u,u) \end{aligned}$$
(55a)

It remains to show that \(\beta ^2\) is given by (9). Taking the trace of (55a),

$$\begin{aligned} \mathrm {tr}\, I_z = \beta ^2\, \mathrm {tr}\,Q_{\bar{x}} = \beta ^2\, \mathrm {dim}\,M \end{aligned}$$
(55b)

If \(e_1\,,\ldots ,\,e_d\) is an orthonormal basis of \(T_{\bar{x}}M\), then by (54b),

$$\begin{aligned} \mathrm {tr}\, I_z = \mathbb {E}_z\sum ^d_{i=1} \left( Q\left( \nabla _{\bar{x}}\,\ell (z),e_i\right) \right) ^2 = \mathbb {E}_z\,Q\left( \nabla _{\bar{x}}\ell (z),\nabla _{\bar{x}}\ell (z)\right) \end{aligned}$$
(55c)

Thus, (55b) and (55c) show that \(\beta ^2\) is given by (9). \(\blacksquare \)

To complete the proof of Theorem 1, it remains to show that the expectations appearing in (9) do not depend on \(\bar{x}\).

For the first expectation, giving \(\alpha ^2(\sigma )\), note that,

$$\begin{aligned} \mathbb {E}_{g\cdot z}\left( \partial _\sigma \ell (g\cdot z)\right) ^2 \,=\, \mathbb {E}_z\left( \partial _\sigma \ell (g\cdot z)\circ g\right) ^2 \,=\, \mathbb {E}_z\left( \partial _\sigma \ell (z)\right) ^2 \end{aligned}$$
(56)

where the first equality follows from (11) and the second equality follows from (49a). Thus, this expectation has the same value, whether computed at \(g\cdot z = (g\cdot \bar{x},\sigma )\), or at \(z = (\bar{x},\sigma )\). Therefore, it does not depend on \(\bar{x}\), since the action of G on M is transitive.

For the second expectation, giving \(\beta ^2(\sigma )\), note that by (11),

$$\begin{aligned} \mathbb {E}_{g\cdot z} \,Q\left( \nabla _{\bar{x}}\,\ell (g\cdot z)\,,\nabla _{\bar{x}}\,\ell (g\cdot z)\,\right) = \mathbb {E}_{z} \,Q\left( \nabla _{\bar{x}}\,\ell (g\cdot z) \circ g\,,\nabla _{\bar{x}}\,\ell (g\cdot z) \circ g\,\right) \end{aligned}$$
(57a)

On the other hand, by (49a),

$$ \nabla _{\bar{x}}\,\ell (g\cdot z) \circ g \,=\, dg_{\bar{x}}\,\nabla _{\bar{x}}\,\ell (z) $$

Moreover, since \(dg_{\bar{x}}\) preserves the Riemannian metric Q,

$$ \begin{array}{rl} Q(\nabla _{\bar{x}}\,\ell (g\cdot z) \circ g\,,\nabla _{\bar{x}}\,\ell (g\cdot z) \circ g) &{} = Q(dg_{\bar{x}}\,\nabla _{\bar{x}}\,\ell (z)\,,dg_{\bar{x}} \nabla _{\bar{x}}\,\ell (z) ) \\ &{}= Q(\nabla _{\bar{x}}\,\ell (z)\,,\nabla _{\bar{x}}\,\ell (z) ) \end{array} $$

Replacing in (57a) gives

$$\begin{aligned} \mathbb {E}_{g\cdot z} \,Q\left( \nabla _{\bar{x}}\,\ell (g\cdot z)\,,\nabla _{\bar{x}}\,\ell (g\cdot z)\,\right) \,=\, \mathbb {E}_z\,Q(\nabla _{\bar{x}}\,\ell (z)\,,\nabla _{\bar{x}}\,\ell (z) ) \end{aligned}$$
(57b)

so this expectation has the same value, at \(g\cdot z\) and at z. By the same argument made after (56), it does not depend on \(\bar{x}\). \(\blacksquare \)

Proof of (11): let dv denote the invariant Riemannian volume element of M, and note that,

$$\begin{aligned} \mathbb {E}_{g\cdot \, z}\,f \,=\, \int _{M}\, f(x)\,p(x|g\cdot z) \,dv(x) \,=\, \int _{M}\, f(x)\,p(g^{-1}\cdot \!x| z) \,dv(x) \end{aligned}$$
(58a)

where the second equality follows from (7). Introduce the variable \(y = g^{-1}\cdot x\). Since the volume element dv is invariant,

$$\begin{aligned} \int _{M}\, f(x)\,p(g^{-1}\cdot x|z) \,dv(x) \,=\, \int _{M}\, f(g\cdot y)\,p(y|z) \,dv(y) \end{aligned}$$
(58b)

The last integral is the same as \(\mathbb {E}_z\left( f\circ g\right) \). Therefore, (11) follows from (58a) and (58b). \(\blacksquare \)

Appendix B – Proof of Proposition 2

It remains to prove (17a) and (17b). To do so, introduce the following notation, using (15),

$$\begin{aligned} t = \langle x,\bar{x}\rangle \quad Z(\eta ) = e^{\psi (\eta )} = (2\pi )^\nu \,\eta ^{1-\nu }I_{\nu -1}(\eta ) \end{aligned}$$
(59a)

Then, \(Z(\eta )\) is the moment generating function of t, so

$$\begin{aligned} \mathbb {E}_z(t) = \frac{Z^\prime (\eta )}{Z(\eta )} \quad \text {and}\quad \mathbb {E}_z\left( t^2\right) = \frac{Z^{\prime \prime }(\eta )}{Z(\eta )} \end{aligned}$$
(59b)

where the prime denotes differentiation with respect to \(\eta \). Recall the derivative and recurrence relations of modified Bessel functions [46],

$$\begin{aligned} \left( \eta ^{-a}I_{a}(\eta )\right) ^\prime = \eta ^{-a}I_{a+1}(\eta ) \quad I_{a-1}(\eta ) - I_{a+1}(\eta ) \,=\, \frac{2a}{\eta }\,I_a(\eta ) \end{aligned}$$
(60)

where a is any complex number. By applying these relations to (59b), it is possible to show, through a direct calculation,

$$\begin{aligned} \mathbb {E}_z(t) = \frac{I_\nu (\eta )}{I_{\nu -1}(\eta )}\quad \end{aligned}$$
(61a)
$$\begin{aligned} \mathbb {E}_z\left( t^2\right) = \frac{1}{n} + \frac{n-1}{n}\, \frac{I_{\nu +1}(\eta )}{I_{\nu -1}(\eta )} \end{aligned}$$
(61b)

Formulae (61) will provide the proof of (17a) and (17b).

Proof of (17a): since \(\psi (\eta )\) is the cumulant generating function of t,

$$\begin{aligned} \psi ^{\prime \prime }(\eta ) \,=\, \mathrm {Var}_z(t) \,=\, \mathbb {E}_z\left( t^2\right) - \mathbb {E}_z\left( t\right) ^2 \end{aligned}$$
(62a)

where \(\mathrm {Var}\) denotes the variance. Now, (17a) follows immediately by replacing from (61) into the right-hand side. \(\blacksquare \)

Proof of (17b): recall from (18a),

$$\begin{aligned} \beta ^2(\eta ) \,=\, \frac{\eta ^2}{n-1}\,\mathbb {E}_z\left( \, 1 - t^2\,\right) \end{aligned}$$
(62b)

However, from (61b),

$$\begin{aligned} \mathbb {E}_z\left( \, 1 - t^2\,\right) = \frac{n-1}{n}\,\left( \, 1 + \frac{I_{\nu +1}(\eta )}{I_{\nu -1}(\eta )}\,\right) \end{aligned}$$
(62c)

Now, (17b) follows by replacing (62c) into (62b). \(\blacksquare \)

Proof of (61a): using the derivative relation of modified Bessel functions, which is the first relation in (60), with \(a = \nu - 1\), it follows that

$$\begin{aligned} Z^\prime (\eta ) \,=\, (2\pi )^\nu \,\eta ^{1-\nu }I_\nu (\eta ) \end{aligned}$$
(63a)

Now, Formula (61a) follows by replacing this into (59b) and using (59a). \(\blacksquare \)

Proof of (61b): write (63a) in the form

$$ Z^\prime (\eta ) \,=\, (2\pi )^\nu \,\eta \,\left( \eta ^{-\nu }I_\nu (\eta )\right) $$

By the product rule

$$ Z^{\prime \prime }(\eta ) \,=\, (2\pi )^\nu \, \eta ^{-\nu }I_\nu (\eta ) + (2\pi )^\nu \,\eta \,\left( \eta ^{-\nu }I_\nu (\eta )\right) ^\prime $$

The derivative in the second term can be evaluated from the derivative relation of modified Bessel functions, with \(a = \nu \). Then,

$$ Z^{\prime \prime }(\eta ) \,=\, (2\pi )^\nu \, \eta ^{-\nu }I_\nu (\eta ) + (2\pi )^\nu \,\eta ^{1-\nu }I_{\nu +1}(\eta ) $$

Rearrange this formula as

$$ Z^{\prime \prime }(\eta ) \,=\, (2\pi )^\nu \,\eta ^{1-\nu }\,\left( \eta ^{-1}I_\nu (\eta ) + I_{\nu +1}(\eta )\right) $$

By the recurrence relation of modified Bessel functions, which is the second relation in (60), with \(a = \nu \), it then follows

$$ Z^{\prime \prime }(\eta ) \,=\, (2\pi )^\nu \,\eta ^{1-\nu }\,\left( \frac{1}{2\nu }\,I_{\nu -1}(\eta ) - \frac{1}{2\nu }\,I_{\nu +1}(\eta ) + I_{\nu +1}(\eta )\right) $$

Recalling that \(2\nu = n\), this can be written,

$$\begin{aligned} Z^{\prime \prime }(\eta ) \,=\, (2\pi )^\nu \,\eta ^{1-\nu }\,\left( \frac{1}{n}\,I_{\nu -1}(\eta ) + \frac{n-1}{n}\, I_{\nu +1}(\eta )\right) \end{aligned}$$
(63b)

Now, Formula (61b) follows by replacing this into (59b) and using (59a). \(\blacksquare \)

Appendix C – Proof of Proposition 6

The setting and notations are the same as in Sect. 7, except for the fact that \(\bar{x}\) is written as x, without the bar, in order to avoid notations such as \(\dot{\bar{x}}\) or \(\ddot{\bar{x}}\). This being said, let \(\tilde{\nabla }\) and \(\nabla \) denote the Levi-Civita connections of the Riemannian metrics I and Q, respectively. Thus, \(\tilde{\nabla }\) is a connection on the tangent bundle of the manifold \(\mathcal {M}\), and \(\nabla \) is a connection on the tangent bundle of the manifold M [13, 40]. Introduce the shape operator \(S:T_{x}M\rightarrow T_{x}M\), which is given as in [40],

$$\begin{aligned} S(u) = \tilde{\nabla }_u\,\partial _r \quad u \in T_{x}M \end{aligned}$$
(64)

for any \(x \in M\). The following identities can be found in [40] (Sect. 2.4, p. 41),

$$\begin{aligned} \tilde{\nabla }_{\partial _r}\,\partial _r&= 0 \quad \end{aligned}$$
(65a)
$$\begin{aligned} \tilde{\nabla }_{\partial _r}\,X&= S(X) \quad \end{aligned}$$
(65b)
$$\begin{aligned} \tilde{\nabla }_{X}\,Y&= \nabla _X\,Y - I(S(X),Y)\,\partial _r \, \end{aligned}$$
(65c)

for any vector fields X and Y on M. Using these identities, it is possible to write the geodesic equation of the Riemannian metric I, in terms of the shape operator S. This is given in the following proposition.

Proposition 10

Let \(\gamma (t)\) be a curve in \(\mathcal {M}\) , with \(\gamma (t) = (x(t),\sigma (t))\) and let \(r(t) = r(\sigma (t))\). The curve \(\gamma (t)\) is a geodesic of the Riemannian metric I if and only if it satisfies the geodesic equation

$$\begin{aligned} \ddot{r}&= I(S(\dot{x}),\dot{x}) \end{aligned}$$
(66a)
$$\begin{aligned} \,\ddot{x}&= -2\,\dot{r}S(\dot{x}) \quad \end{aligned}$$
(66b)

where \(\ddot{x} = \nabla _{\dot{x}}\,\dot{x}\) is the acceleration of the curve x(t) in M.

The shape operator S moreover admits a simple expression, which can be derived from expression (40) of the Riemannian metric I, using the fact that \(\tilde{\nabla }\) is a metric connection [13] (Theorem I.5.1, p. 16).

Proposition 11

In the notation of (40), the shape operator S is given by

$$\begin{aligned} S(u) = \sum ^r_{q=1}\, \frac{\partial _r\,\beta _q(r)}{\beta _q(r)}\,u_q \end{aligned}$$
(67)

In other words, the decomposition \(u = u_1\,+\cdots +\,u_r\) provides a block-diagonalisation of S, where each block is a multiple of identity.

Combining Propositions 10 and 11, the geodesic equation (66) takes on a new form. Precisely, replacing (67) into (66) gives the following equations

$$\begin{aligned} \ddot{r}&= \sum ^r_{q=1}\, \beta _q(r)\partial _r\beta _q(r)\,Q(\dot{x}_q,\dot{x}_q) \end{aligned}$$
(68a)
$$\begin{aligned} \ddot{x}_q&= -2\,\dot{r}\,\frac{\partial _r\,\beta _q(r)}{\beta _q(r)}\,\dot{x}_q \end{aligned}$$
(68b)

where \(\dot{x} = \dot{x}_1\,+\cdots +\,\dot{x}_r\) and \(\ddot{x} = \ddot{x}_1\,+\cdots +\,\ddot{x}_{r\,}\). The proof of Proposition 6 can be obtained directly from Eq. (68), using the following conservation laws.

Proposition 12

Each one of the following quantities \(C_q\) is a conserved quantity,

$$\begin{aligned} C_q \,=\, \beta ^4_q(r)\,Q(\dot{x}_{q\,},\dot{x}_q) \quad \text {for }\,\, q = 1\,,\ldots ,\,r \end{aligned}$$
(69)

In other words, \(C_q\) remains constant when evaluated along any geodesic \(\gamma (t)\) of the Riemannian metric I.

For now, assume that Propositions 1012 are true. To prove Proposition 6, note the following.

Proof of (41a): it is enough to show that the right-hand side of (41a) is the same as the right-hand side of (68a). To do so, note from (41a) and (69) that

$$\begin{aligned} V(r) \,=\, \sum ^r_{q=1}\,\frac{\beta ^2_q(r(0))}{\beta ^2_q(r)}\,I_z(u_{q\,},u_q) \,=\, \sum ^r_{q=1}\,\frac{C_q}{\beta ^2_q(r)} \end{aligned}$$
(70a)

Indeed, since \(\dot{x}_q(0) = u_q\) and since \(C_q\) is a conserved quantity

$$ \beta ^2_q(r(0))\,I_z(u_{q\,},u_q) \,=\, \left. \beta ^4_q(r)\,Q(\dot{x}_{q\,},\dot{x}_q)\right| _{t=0} \,=\,C_q $$

Now, replacing the derivative of (70a) into the right-hand side of (41a) directly leads to the right-hand side of (68a). \(\blacksquare \)

Proof of (41b): recall from Remark 7 that M is the Riemannian product of the \(M_q\). Therefore, the Riemannian exponential mapping of M is also the product of the Riemannian exponential mappings of the \(M_q\). Precisely, (41b) is equivalent to

$$\begin{aligned} x_q(t) \,=\, \exp _{{x_q(0)}}\,\left[ \,\left( \,\int ^t_0 \frac{\beta ^2_q(r(0))}{\beta ^2_q(r(s))}ds\right) \,u_q\,\right] \quad \text {for }\,\, q = 1\,,\ldots ,\,r \end{aligned}$$
(70b)

This means that the curve \(x_q(t)\) in \(M_q\) is a reparameterised geodesic \(\left( \delta _q\circ F\right) (t)\) where \(\delta _q(t)\) is the geodesic given by \(\delta _q(t)= \exp (t\,u_q)\) and F(t) is the integral inside the parentheses in (70b). To prove (41b), it is sufficient to prove that (70b) solves Eq. (68b). Using the chain rule, (70b) implies that

$$\begin{aligned} \ddot{x}_q \,=\, \dot{F}^{\,2}\left( \ddot{\delta }_q\circ F\right) \,+\, F^{\,\prime \prime }\left( \dot{\delta }_q\circ F\right) \,=\, \dot{F}^{\,2}\left( \ddot{\delta }_q\circ F\right) \,+\, \frac{F^{\,\prime \prime }}{F^\prime }\,\dot{x}_q \,=\, \frac{F^{\,\prime \prime }}{F^\prime }\,\dot{x}_q \end{aligned}$$
(70c)

where the third equality follows because \(\delta _q\) is a geodesic, and therefore its acceleration \(\ddot{\delta }_q\) is zero. By replacing the definition of the function F(t), it is seen that (70c) is the same as (68b). It follows that (70b) solves (68b), as required. \(\blacksquare \)

Proof of Proposition 10: recall the geodesic equation is \(\tilde{\nabla }_{\dot{\gamma }}\,\dot{\gamma } \,=0\), which means that the velocity \(\dot{\gamma }(t)\) is self-parallel [13, 40]. Here, the velocity \(\dot{\gamma }(t)\) is given by \(\dot{\gamma }(t) \,=\, \dot{r}\,\partial _r \,+\, \dot{x}\). Accordingly, the left-hand side of the geodesic equation is

$$\begin{aligned} \tilde{\nabla }_{\dot{\gamma }}\,\dot{\gamma } \,=\, \tilde{\nabla }_{\dot{\gamma }}\,\dot{r}\,\partial _r \,+\, \tilde{\nabla }_{\dot{\gamma }}\,\dot{x} \,=\, \ddot{r}\,\partial _r \,+\,\dot{r}\,\tilde{\nabla }_{\dot{\gamma }}\,\partial _r\,+\, \tilde{\nabla }_{\dot{\gamma }}\,\dot{x} \end{aligned}$$
(71a)

where the second equality follows by the product rule for the covariant derivative [13, 40]. The second and third terms on the right-hand side of (71a) can be written in terms of the shape operator S. Precisely, for the second term,

$$\begin{aligned} \tilde{\nabla }_{\dot{\gamma }}\,\partial _r \,=\, \dot{r}\,\tilde{\nabla }_{\partial _r}\,\partial _r\,+\, \tilde{\nabla }_{\dot{x}}\,\partial _r \,=\, S(\dot{x}) \end{aligned}$$
(71b)

where the second equality follows from (64) and (65a). Moreover, for the third term,

$$\begin{aligned} \tilde{\nabla }_{\dot{\gamma }}\,\dot{x} \,=\, \dot{r}\,\tilde{\nabla }_{\partial _r}\,\dot{x}\,+\, \tilde{\nabla }_{\dot{x}}\,\dot{x} \,=\, \dot{r}\,S(\dot{x}) \,+\,\ddot{x} \,-\, I(S(\dot{x}),\dot{x})\,\partial _r \end{aligned}$$
(71c)

where the second equality follows from (65b) and (65c). Replacing (71b) and (71c) into (71a), the left-hand side of the geodesic equation becomes

$$ \tilde{\nabla }_{\dot{\gamma }}\,\dot{\gamma } \,=\, \left( \,\ddot{r} - I(S(\dot{x}),\dot{x}) \,\right) \,\partial _r \,+\, \left( \, \ddot{x} + 2\dot{r}S(\dot{x})\,\right) $$

Setting this equal to zero immediately gives Eq. (66). \(\blacksquare \)

Proof of Proposition 11 : recall the shape operator S is symmetric, since it is essentially the Riemannian Hessian of r [40] (Sect. 2.4, p. 41). Therefore, it is enough to evaluate I(S(u), u) for \(u \in T_xM\). Let X by a vector field on M, with \(X(x) = u\). Then,

$$\begin{aligned} I(S(u),u) \,=\, I\left( S(X),X\right) \,=\,I\left( \,\tilde{\nabla }_{\partial _r}\,X\,,X\right) \end{aligned}$$
(72a)

where the second equality follows from (65b). Using the fact that \(\tilde{\nabla }\) is a metric connection [13] (Theorem I.5.1, p. 16), the right-hand side can be written as

$$\begin{aligned} I\left( \,\tilde{\nabla }_{\partial _r}\,X\,,X\right) \,=\, \frac{1}{2}\, \partial _r\,I(X,X) \,=\, \frac{1}{2}\, \partial _r\,\sum ^r_{q=1}\,\beta ^2_q(r)\,Q_x(u_{q\,},u_q) \end{aligned}$$
(72b)

where the second equality follows from (40). It remains to note that

$$ \frac{1}{2}\, \partial _r\,\beta ^2_q(r)\,Q_x(u_{q\,},u_q) \,=\, \frac{\partial _r\,\beta _q(r)}{\beta _q(r)}\,I_z(u_{q\,},u_q) $$

Accordingly, (72a) and (72b) imply

$$\begin{aligned} I(S(u),u) \,=\, I\left( \, \sum ^r_{q=1}\, \frac{\partial _r\,\beta _q(r)}{\beta _q(r)}\,u_{q\,},u_q \,\right) \,=\, I\left( \, \sum ^r_{q=1}\, \frac{\partial _r\,\beta _q(r)}{\beta _q(r)}\,u_{q\,},u \,\right) \end{aligned}$$
(72c)

and (67) follows from the fact that S is symmetric. \(\blacksquare \)

Proof of Proposition 12: to say that \(C_q\) is a conserved quantity means that \(\dot{C}_q = 0\). From (69),

$$\begin{aligned} \dot{C}_q \,=\, 4\dot{r}\,\beta ^3_q(r)\,\partial _r\beta _q(r)\,Q(\dot{x}_{q\,}\,\dot{x}_{q}) \,+\,\beta ^4_q(r)\,\frac{d}{dt}\,Q(\dot{x}_{q\,}\,\dot{x}_{q}) \end{aligned}$$
(73a)

The last derivative can be expressed as

$$\begin{aligned} \frac{d}{dt}\,Q(\dot{x}_{q\,}\,\dot{x}_{q}) \,=\, 2\,Q(\ddot{x}_{q\,},\dot{x}_q) \,=\, -4\dot{r}\,\frac{\partial _r\beta _q(r)}{\beta _q(r)}\,Q(\dot{x}_{q\,},\dot{x}_q) \end{aligned}$$
(73b)

where the second equality follows from (68b). By replacing (73b) into (73a), it follows immediately that \(\dot{C}_q = 0\). \(\blacksquare \)

Appendix D – Proof of Proposition 7

The proof of Formula (44b), for the Laplace–Beltrami operator \(\varDelta _{\mathcal {M}}\) , will introduce some useful notation.

Proof of Formula (44b): for any smooth function f on \(\mathcal {M}\), it follows from (40) that the Riemannian gradient \(\tilde{\nabla }f\) of f, with respect to the multiply-warped Riemannian metric I, is given by

$$\begin{aligned} \tilde{\nabla }f \,=\, \left( \partial _rf\right) \,\partial _r \,+\, \sum ^r_{q=1}\,\beta ^{-2}_q(r)\,\nabla _{\bar{x}_q}\,f \end{aligned}$$
(74a)

where \(\nabla _{\bar{x}_q}\,f\) is the Riemannian gradient of f with respect to \(\bar{x}_q \in M_q\,\), computed with all other arguments of f being fixed. Expression (74a) can be verified by checking that

$$ df\,U \,=\, I(\tilde{\nabla }f,U) $$

for any tangent vector U to \(\mathcal {M}\). This follows directly from (40) and (74a). Now, by definition of the Laplace–Beltrami operator [34], (see p. 443),

$$\begin{aligned} \varDelta _{\mathcal {M}}f \,=\, \mathrm {div}\,\tilde{\nabla }f \end{aligned}$$
(74b)

where the divergence \(\mathrm {div}\, V\) of a vector field V on \(\mathcal {M}\) is found from

$$\begin{aligned} \mathcal {L}_V\,\mathrm {vol}\,=\,\left( \mathrm {div}\,V\right) \,\mathrm {vol} \end{aligned}$$
(74c)

with the notation \(\mathcal {L}\) for the Lie derivative, and \(\mathrm {vol}\) for the Riemannian volume element of the metric I. This last formula can be applied along with the following expression of \(\mathrm {vol}\), which follows from (40),

$$\begin{aligned} \mathrm {vol}\,=\, G(r)\,dr\bigwedge _{q}\,\mathrm {vol}_q(\bar{x}_q) \end{aligned}$$
(74d)

where the function G(r) was defined after (44b), and where \(\wedge \) denotes the exterior product, and \(\mathrm {vol}_q\) is the Riemannian volume of \(M_q\,\). From (74c) and (74d), applying the product formula of the Lie derivative [34] (Theorem 7.4.8, p. 414),

$$\begin{aligned} \mathrm {div}\, V = \frac{1}{G}\, \partial _r\left( G\,V_r\right) \,+\, \sum ^r_{q=1}\, \mathrm {div}_{M_q}\, V_q \end{aligned}$$
(74e)

where \(V = V_r\,\partial _r\,+\,\sum _q\,V_q\) with each \(V_q\) tangent to \(M_q\) , and where \(\mathrm {div}_{M_q}\, V_q\) denotes the divergence of \(V_q\) with respect to \(\bar{x}_q \in M_q\,\). Formula (44b) follows directly from (74a), (74b) and (74e). Indeed, for the vector field \(V = \tilde{\nabla }f\),

$$ V_r \,=\,\partial _rf \quad V_q \,=\, \beta ^{-2}_q(r)\,\nabla _{\bar{x}_q}\,f $$

as can be seen from (74a). \(\blacksquare \)

For the proof of Proposition 7, assume the process z is a Riemannian Brownian motion associated to I. Write \(z(t) = (\bar{x}(t),\sigma (t))\) where \(\bar{x}(t) = \left( \bar{x}_1(t)\,,\ldots ,\,\bar{x}_r(t)\right) \) and each \(\bar{x}_q(t)\) belongs to \(M_q\,\). The proof consists in showing that the joint distribution of the processes \(r(t) = r(\sigma (t))\) and \(\bar{x}_q(t)\) is the same as described in Proposition 7. This is done through the following steps.

Step 1r(t) verifies (45a): recall that, for any smooth function f on \(\mathcal {M}\), the process z verifies (44a). If \(f = r\), then by (44a) and (44b)

$$\begin{aligned} dr(t) \,=\, \frac{1}{2}\,\varDelta _{\mathcal {M}}r(t)dt + dm^r(t) \,=\, \frac{1}{2}\,\frac{\partial _rG}{G}(r(t))dt\,+\,dm^r(t) \end{aligned}$$
(75)

To prove that r(t) verifies (45a), it is enough to prove that \(dm^r(t) = dw(t)\) where w is a standard Brownian motion. Note that \(dr^2(t)\) can be computed in two ways. The first way, by Itô’s formula and (75)

$$ dr^2 = r\varDelta _{\mathcal {M}}r(t)dt \,+\, 2r(t)dm^r(t)\,+\,d[m^r](t) $$

where \([m^r](t)\) is the quadratic variation process of the local martingale \(m^r(t)\) [25] (Theorem 17.16, p. 339). The second way, by (44a) with \(f = r^2\),

$$ \begin{array}{rl} dr^2 &{}= \left( I(\tilde{\nabla }r,\tilde{\nabla }r)+r\varDelta _{\mathcal {M}}r(t)\right) \,dt \,+\,dm^{r^2}(t)dt \\ &{}= \left( 1+r\varDelta _{\mathcal {M}}r(t)\right) \,dt \,+\,dm^{r^2}(t)dt \end{array} $$

where the second line follows from (40) because \(\tilde{\nabla }r = \partial _r\) . By equating these two expressions of \(dr^2(t)\) , it follows that \(dt - d[m^r](t)\) is the differential of a continuous local martingale of finite variation, and therefore identically zero [25] (Proposition 17.2, p. 330). In other words, \(d[m^r](t) = dt\) and Lévy’s characterisation implies that \(dm^r(t) = dw(t)\), where w is a standard Brownian motion [25] (Theorem 18.3, p. 352). \(\blacksquare \)

Step 2\(\bar{x}_q(t)\) verifies (45b): if f is a smooth function on \(\mathcal {M}\), such that \(f(z) = f(\bar{x}_q)\), then by (44a) and (44b)

$$\begin{aligned} df(\bar{x}_q(t)) \,=\, \frac{1}{2}\beta ^{-2}_q(r(t))\,\varDelta _{M_q}\,f(\bar{x}_q(t))\,dt\,+\,dm^f(t) \end{aligned}$$
(76)

Define \(l_q(t)\) to be the inverse of the time change process \(\tau _q(t)\) defined in (45b), and let \(\theta _q(t) = \left( \bar{x}_q\circ l_q\right) (t)\). By applying the time change \(l_q(t)\) to (76)

$$ df(\theta _q(t)) \,=\, \frac{1}{2}\,\varDelta _{M_q}\,f(\theta _q(t))\,dt\,+\,d(m^f\circ l_q)(t) $$

where the first term on the right-hand side is obtained by replacing from the definition of \(\tau _q(t)\) in (45b). Recall that a time change of a local martingale is a local martingale [25] (Theorem 17.24, p. 344). Therefore, \(m^f\circ l_q\) is a local martingale, so \(\theta _q\) solves the martingale problem associated to \((1/2)\varDelta _{M_q}\). This means that \(\theta _q\) is a Riemannian Brownian motion in \(M_q\) . \(\blacksquare \)

Step 3r(t) and \(\theta _q(t)\) are independent: it is required to prove that the processes r(t) and \(\theta _1(t)\,,\ldots ,\,\theta _r(t)\) are jointly independent. A detailed proof is given only of the fact that r(t) and \(\theta _q(t)\) are independent, for any fixed q. The complete proof is obtained by repeating similar arguments.

The following proof is modeled on [23] (Example 3.3.3, p. 84). Let \(\left( \xi ^i(t)\,;\right. \left. i = 1\,,\ldots ,\,\mathrm {dim}\,M_q\right) \) be the stochastic anti-development of \(\bar{x}_q(t)\). Precisely [19] (Definition 8.23, p. 119)

$$\begin{aligned} d\xi ^i(t) \,=\, Q\left( e^i\,,d\bar{x}_q(t)\right) \,=\,\beta ^{-2}_q(r(t))\,I\left( e^i\,,dz(t)\right) \end{aligned}$$
(77a)

where the \(e^i\) form a parallel orthonormal moving frame above the stochastic process \(\bar{x}_q(t)\) in \(M_q\,\). On the other hand, it is possible to write, using Itô’s formula [19] (Proposition 7.34, p. 109)

$$\begin{aligned} dr(t) \,=\, I\left( \partial _r\,,dz(t)\right) \,+\, \frac{1}{2}\,\varDelta _{\mathcal {M}}r(t)\,dt \end{aligned}$$
(77b)

Let \([\xi ^i,r]\) denote the quadratic covariation of \(\xi ^i\) and r. From [19] (Proposition 5.18, p. 63), since z is a Riemannian Brownian motion, it follows from (77a) and (77b) that

$$\begin{aligned} d[\xi ^i,r](t) \,=\, \beta ^{-2}_q(r(t))\,I(e^i,\partial _r)(z(t))\,dt \,=\, 0 \end{aligned}$$
(77c)

as \(e^i\) and \(\partial _r\) are orthogonal with respect to I. This means that \((\xi ^i(t))\) and r(t) have zero quadratic covariation, and therefore \((\xi ^i(t))\) and w(t) have zero quadratic covariation. It follows as in [23] (Lemma 3.3.4, p. 85), that r(t) and \(\theta _q(t)\) are independent. \(\blacksquare \)

Rights and permissions

Reprints and permissions

Copyright information

© 2019 Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Said, S., Bombrun, L., Berthoumieu, Y. (2019). Warped Riemannian Metrics for Location-Scale Models. In: Nielsen, F. (eds) Geometric Structures of Information. Signals and Communication Technology. Springer, Cham. https://doi.org/10.1007/978-3-030-02520-5_10

Download citation

  • DOI: https://doi.org/10.1007/978-3-030-02520-5_10

  • Published:

  • Publisher Name: Springer, Cham

  • Print ISBN: 978-3-030-02519-9

  • Online ISBN: 978-3-030-02520-5

  • eBook Packages: EngineeringEngineering (R0)

Publish with us

Policies and ethics