Log in

Objective Rates as Covariant Derivatives on the Manifold of Riemannian Metrics

  • Published:
Archive for Rational Mechanics and Analysis Aims and scope Submit manuscript

Abstract

The subject of so-called objective derivatives in Continuum Mechanics has a long history and has generated varying views concerning their true mathematical interpretation. Several attempts have been made to provide a mathematical definition that would at least partially unify the existing notions. In this paper, we demonstrate that, under natural assumptions, all objective derivatives correspond to covariant derivatives on the infinite-dimensional manifold \(\textrm{Met}(\mathcal {B})\) of Riemannian metrics on the body. Furthermore, a natural Leibniz rule enables canonical extensions from covariant to contravariant tensor fields and vice versa. This makes the sometimes-used distinction between objective derivatives of “Lie type” and “co-rotational type” unnecessary. For an exhaustive list of objective derivatives found in the literature, we exhibit the corresponding covariant derivative on \(\textrm{Met}(\mathcal {B})\).

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

Access this article

Subscribe and save

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

Buy Now

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

Instant access to the full article PDF.

Notes

  1. To be more accurate, the manifold \(\mathcal {E}\) is not “the space” but the typical fiber of dimension 3, of a fibered manifold of dimension 4 with a Galilean structure [16, 49, 50].

  2. This formulation does not requires the introduction of the decomposition RU.

  3. Indeed, this choice seems to have been implicitly adopted by many authors [40, 94, 95].

  4. Note that what are denoted by \({\mathcal {L}}_{\varvec{v}}{\varvec{\sigma }}^{2}\) and \({\mathcal {L}}_{\varvec{v}}{\varvec{\sigma }}^{3}\) in [57, Chapter 1, Box 6.1] are not symmetric second-order tensors, and a mean of them is required in order to build an objective derivative on symmetric contravariant second-order tensors. This is our second objective derivative \(d^{2}\).

References

  1. Arnold, V.I.: Sur la géométrie différentielle des groupes de Lie de dimension infinie et ses applications à l’hydrodynamique des fluides parfaits. Ann. Inst. Fourier (Grenoble) 16(fasc. 1), 319–361, 1966

    Article  MathSciNet  Google Scholar 

  2. Arnold, V.I., Khesin, B.: Topological Methods in Hydrodynamics. Applied Mathematical Sciences, vol. 125. Springer, New York (1998)

    Book  Google Scholar 

  3. Bauer, M., Harms, P., Michor, P.W.: Sobolev metrics on the manifold of all Riemannian metrics. J. Differ. Geom. 94(2), 187–208, 2013

    Article  MathSciNet  Google Scholar 

  4. Becker, G.F.: The finite elastic stress–strain function. Am. J. Sci 3(275), 337–356, 1893

    Article  ADS  Google Scholar 

  5. Benzécri, J.P.: Un modèle géométrique de milieu continu déformable: la théorie des deux métriques. Annales de l’I.H.P. Probabilités et statistiques 3(3), 277–321, 1967

    MathSciNet  Google Scholar 

  6. Bernstein, B.: Hypo-elasticity and elasticity. Arch. Ration. Mech. Anal. 6(1), 89–104, 1960

    Article  MathSciNet  Google Scholar 

  7. Bertram, A.: Elasticity and Plasticity of Large Deformations, 3rd edn. Springer, Berlin (2012)

    Book  Google Scholar 

  8. Besson, J., Cailletaud, G., Chaboche, J.-L., Forest, S.: Mécanique non linéaire des matériaux. Hermès, Paris (2001)

    Google Scholar 

  9. Bhatia, R.: Positive Definite Matrices. Princeton University Press, Princeton (2009)

    Book  Google Scholar 

  10. Bruveris, M.: The \(L^2\)-metric on \(C^\infty (M,N)\). ar**v e-prints, ar**v:1804.00577 (2018)

  11. Clarke, B.: The metric geometry of the manifold of Riemannian metrics over a closed manifold. Calc. Var. Partial Differ. Equ. 39(3–4), 533–545, 2010

    Article  MathSciNet  Google Scholar 

  12. Clarke, B.: The Riemannian \(L^2\) topology on the manifold of Riemannian metrics. Ann. Global Anal. Geom. 39(2), 131–163, 2010

    Article  Google Scholar 

  13. Clarke, B.: The completion of the manifold of Riemannian metrics. J. Differ. Geom. 93(2), 203–268, 2013

    Article  MathSciNet  Google Scholar 

  14. Clarke, B.: Brian Clarke. The Completion of the Manifold of Riemannian Metrics with Respectto its L2 Metric. Ph.D. thesis, University of Leipzig (2018)

  15. Dimitrienko, Y.I.: Nonlinear Continuum Mechanics and Large Inelastic Deformations. Solid Mechanics and its Applications, vol. 174. Springer, Cham (2011)

    Google Scholar 

  16. Duval, C., Künzle, H.P.: Sur les connexions newtoniennes et l’extension non triviale du groupe de Galilée. C.R. Acad. Sc. Paris Série A 285, 813–816, 1977

    Google Scholar 

  17. Ebin, D. G.: On the space of Riemannian metrics. Ph.D. thesis, Massachusetts Institute of Technology, Cambridge (1967)

  18. Ebin, D.G.: On the space of Riemannian metrics. Bull. Am. Math. Soc. 74, 1001–1003, 1968

    Article  MathSciNet  Google Scholar 

  19. Ebin, D.G., Marsden, J.: Groups of diffeomorphisms and the motion of an incompressible fluid. Ann. Math. 92(1), 102, 1970

    Article  MathSciNet  Google Scholar 

  20. Epstein, M., Segev, R.: Differentiable manifolds and the principle of virtual work in continuum mechanics. J. Math. Phys. 21(5), 1243–1245, 1980

    Article  ADS  MathSciNet  Google Scholar 

  21. Eringen, A.C.: Nonlinear Theory of Continuous Media. McGraw-Hill Book Co., New York-Toronto-London (1962)

    Google Scholar 

  22. Euler, L.P.: Du mouvement de rotation des corps solides autour dun axe variable. Mémoires de l’académie des Sciences de Berlin 14, 154–193, 1765

    Google Scholar 

  23. Fiala, Z.: Time derivative obtained by applying the Riemannian manifold of Riemannian metrics to kinematics of continua. C. R. Mecanique 332, 97–102, 2004

    Article  ADS  Google Scholar 

  24. Fiala, Z.: Geometrical setting of solid mechanics. Ann. Phys. 326(8), 1983–1997, 2011

    Article  ADS  MathSciNet  Google Scholar 

  25. Fiala, Z.: Evolution equation of Lie-type for finite deformations, time-discrete integration, and incremental methods. Acta Mech. 226(1), 17–35, 2015

    Article  MathSciNet  Google Scholar 

  26. Fiala, Z.: Geometry of finite deformations and time-incremental analysis. Int. J. Non-Linear Mech. 81, 230–244, 2016

    Article  ADS  Google Scholar 

  27. Fiala, Z.: Objective time derivatives revised. Zeitschrift für angewandte Mathematik und Physik 71(1), 4, 2019

    Article  ADS  MathSciNet  Google Scholar 

  28. Freed, D.S., Groisser, D.: The basic geometry of the manifold of Riemannian metrics and of its quotient by the diffeomorphism group. Mich. Math. J. 36(3), 323–344, 1989

    Article  MathSciNet  Google Scholar 

  29. Frölicher, A., Kriegl, A.: Linear Spaces and Differentiation Theory. Pure and Applied Mathematics (New York). Wiley, Chichester (1988)

    Google Scholar 

  30. Gallot, S., Hulin, D., Lafontaine, J.: Riemannian Geometry. Universitext, 3rd edn. Springer, Berlin (2004)

    Book  Google Scholar 

  31. Giessen, E.V.D., Kollmann, F.G.: On mathematical aspects of dual variables in continuum mechanics. II. Applications in nonlinear solid mechanics. Z. Angew. Math. Mech. 76(9), 497–504, 1996

    Article  MathSciNet  Google Scholar 

  32. Gil-Medrano, O., Michor, P.W.: The Riemannian manifold of all Riemannian metrics. Q. J. Math. Oxford Ser. (2) 42(166), 183–202, 1991

    Article  MathSciNet  Google Scholar 

  33. Green, A., Naghdi, P.: A general theory of an elastic–plastic continuum. Arch. Ration. Mech. Anal. 18(4), 251–281, 1965

    Article  MathSciNet  Google Scholar 

  34. Green, A.E., Zerna, W.: Theoretical Elasticity, 2nd edn. Clarendon Press, Oxford (1968)

    Google Scholar 

  35. Hamilton, R.S.: The inverse function theorem of Nash and Moser. Bull. Am. Math. Soc. (N.S.) 7(1), 65–222, 1982

    Article  MathSciNet  Google Scholar 

  36. Hamilton, R.S.: Three-manifolds with positive Ricci curvature. J. Differ. Geom. 17(2), 255–306, 1982

    Article  MathSciNet  Google Scholar 

  37. Hamilton, R.S.: Four-manifolds with positive curvature operator. J. Differ. Geom. 24(2), 153–179, 1986

    Article  MathSciNet  Google Scholar 

  38. Haupt, P.: Mechanics, Continuum, Theory of Materials. 2nd edn. Springer, Berlin. Traduction de la quatrième édition allemande par G. Juvet et R, Leroy (2002)

  39. Hencky, H.: Über die Form des Elastizitätsgesetzes bei ideal elastischen Stoffen. Zeitschrift für technische Physik 9, 215–220, 1928

    Google Scholar 

  40. Hill, R.: Aspects of invariance in solid mechanics. Adv. Appl. Mech. 18, 1–75, 1978

    MathSciNet  Google Scholar 

  41. Husemoller, D.: Fibre Bundles. Springer, New York (1994)

    Book  Google Scholar 

  42. Iglesias-Zemmour, P.: Diffeology, volume 185 of Mathematical Surveys and Monographs. American Mathematical Society, Providence (2013)

    Google Scholar 

  43. Inci, H., Kappeler, T., Topalov, P.: On the Regularity of the Composition of Diffeomorphisms, Volume 226 of Memoirs of the American Mathematical Society, 1st edn. American Mathematical Society, Providence (2013)

    Google Scholar 

  44. Jaumann, C.: Geschlossenes System physikalischer und chemischer Differentialgesetze. Sitzber. Akad. Wiss. Wien (lIa) 120, 385–530, 1911

    Google Scholar 

  45. Kadianakis, N.: On the geometry of Lagrangian and Eulerian descriptions in continuum mechanics. ZAMM Z. Angew. Math. Mech. 79(2), 131–138, 1999

    Article  MathSciNet  Google Scholar 

  46. Kolev, B., Desmorat, R.: Éléments de géométrie pour la mécanique des milieux continus. Avaiable at : https://hal.archives-ouvertes.fr/hal-02343934 (in French), (2019)

  47. Kolev, B., Desmorat, R.: An intrinsic geometric formulation of hyper-elasticity, pressure potential and non-holonomic constraints. J. Elast. 146(1), 29–63, 2021

    Article  MathSciNet  Google Scholar 

  48. Kriegl, A., Michor, P.W.: The Convenient Setting of Global Analysis. Mathematical Surveys and Monographs, vol. 53. American Mathematical Society, Providence, RI (1997)

    Book  Google Scholar 

  49. Künzle, H.P.: Galilei and Lorentz structures on space-time: comparison of the corresponding geometry and physics. Ann. Inst. H. Poincaré Sect. A (N.S.) 17, 337–362, 1972

    MathSciNet  Google Scholar 

  50. Künzle, H.P.: Covariant Newtonian limit of Lorentz space-times. Gen. Relativ. Gravit. 7(5), 445–457, 1976

    Article  ADS  MathSciNet  Google Scholar 

  51. Ladevèze, P.: Sur la théorie de la plasticité en grandes déformations. Internal Report Nb 9 of LMT-Cachan (1980)

  52. Ladevèze, P.: Nonlinear Computational Structural Mechanics: New Approaches and Non-Incremental Methods of Calculation (translated by J. G. Simmonds from the French edition “Mecanique non linéaire des structures”, Hermès, Paris 1996. Springer, Mechanical Engineering Series (1999)

  53. Lang, S.: Fundamentals of Differential Geometry. Graduate Texts in Mathematics, vol. 191. Springer, New York (1999)

    Book  Google Scholar 

  54. Lichnerowicz, A.: Tensor-Distributions. In: Magnetohydrodynamics: Waves and Shock Waves in Curved Space-Time, pp 1–17. Springer, Cham (1994)

  55. Lubliner, J.: Plasticity Theory. Macmillan, New York (Maxwell Macmillan International Editions) (1990)

  56. Lützen, J.: De Rham’s Currents. In: Studies in the History of Mathematics and Physical Sciences, pp. 144–147. Springer, New York (1982).

  57. Marsden, J.E., Hughes, T.J.R.: Mathematical Foundations of Elasticity. Dover Publications Inc, New York (1994). Corrected reprint of the 1983 original

  58. Martin, R.J., Münch, I., Eidel, B., Neff, P.: A brief history of logarithmic strain measures in nonlinear elasticity. PAMM 18(1), e201800366, 2018

    Article  Google Scholar 

  59. Michor, P.W.: Manifolds of Differentiable Map**s. Shiva Mathematics Series, vol. 3. Shiva Publishing Ltd., Nantwich (1980)

    Google Scholar 

  60. Michor, P.W.: Topics in Differential Geometry. Graduate Studies in Mathematics, vol. 93. American Mathematical Society, Providence, RI (2008)

    Google Scholar 

  61. Miehe, C., Apel, N., Lambrecht, M.: Anisotropic additive plasticity in the logarithmic strain space: modular kinematic formulation and implementation based on incremental minimization principles for standard materials. Comput. Methods Appl. Mech. Eng. 191(47–48), 5383–5425, 2002

    Article  ADS  MathSciNet  Google Scholar 

  62. Milnor, J.: Remarks on infinite-dimensional Lie groups. In: Relativity, groups and topology, II (Les Houches, 1983), pp. 1007–1057. North-Holland, Amsterdam (1984)

  63. Noll, W.: A mathematical theory of the mechanical behavior of continuous media. Arch. Ration. Mech. Anal. 2(1), 197–226, 1958

    Article  MathSciNet  Google Scholar 

  64. Noll, W.: The Foundations of Classical Mechanics in the Light of Recent Advances in Continuum Mechanics, pp 266–281 (1959)

  65. Noll, W.: A new mathematical theory of simple materials. Arch. Ration. Mech. Anal. 48(1), 1–50, 1972

    Article  MathSciNet  Google Scholar 

  66. Noll, W.: La mécanique classique, basée sur un axiome d’objectivité. In: The Foundations of Mechanics and Thermodynamics, pp. 135–144. Springer, Berlin (1974)

  67. Noll, W.: A general framework for problems in the statics of finite elasticity. In: Contemporary Developments in Continuum Mechanics and Partial Differential Equations. Proceedings of the International Symposium on Continuum Mechanics and Partial Differential Equations, pp. 363–387. Elsevier (1978)

  68. Noll, W., Seguin, B.: Basic concepts of thermomechanics. J. Elast. 101(2), 121–151, 2010

    Article  MathSciNet  Google Scholar 

  69. Oldroyd, J.G.: On the formulation of rheological equations of state. Proc. Roy. Soc. Lond. A200, 523–541, 1950

    ADS  MathSciNet  Google Scholar 

  70. Panicaud, B., Rouhaud, E., Altmeyer, G., Wang, M., Kerner, R., Roos, A., Ameline, O.: Consistent hypo-elastic behavior using the four-dimensional formalism of differential geometry. Acta Mech. 227(3), 651–675, 2015

    Article  MathSciNet  Google Scholar 

  71. Perelman, G.: The entropy formula for the Ricci flow and its geometric applications. Available at ar**v:math/0211159 (2002)

  72. Perelman, G.: Finite extinction time for the solutions to the Ricci flow on certain three-manifolds. Available at ar**v:math/0307245 (2003)

  73. Perelman, G.: Ricci flow with surgery on three-manifolds. Available at ar**v:math/0303109 (2003)

  74. Rougée, P.: Formulation lagrangienne intrinsèque en mécanique des milieux continus. J. Mécanique 19, 7–32, 1980

    MathSciNet  Google Scholar 

  75. Rougée, P.: The intrinsic Lagrangian metric and stress variables. In: Finite Inelastic Deformations - Theory and Applications, IUTAM Symposium Hannover/Germany, vol. 199, pp. 217–226 (1991)

  76. Rougée, P.: A new Lagrangian intrinsic approach of continuous media in large deformation. Eur. J. Mech. A/Solids 10, 15–39, 1991

    MathSciNet  Google Scholar 

  77. Rougée, P.: Mécanique des grandes transformations, vol. 25. Mathématiques & Applications (Berlin) [Mathematics & Applications]. Springer, Berlin (1997)

  78. Rougée, P.: An intrinsic Lagrangian statement of constitutive laws in large strain. Comput. Struct. 84(17–18), 1125–1133, 2006

    Article  MathSciNet  Google Scholar 

  79. Rouhaud, E., Panicaud, B., Kerner, R.: Canonical frame-indifferent transport operators with the four-dimensional formalism of differential geometry. Comput. Mater. Sci. 77, 120–130, 2013

    Article  Google Scholar 

  80. Sansour, C.: On the geometric structure of the stress and strain tensors, dual variables and objective rates in continuum mechanics. Arch. Mech. (Arch. Mech. Stos.) 44(5–6), 527–556, 1992

    MathSciNet  Google Scholar 

  81. Segev, R.: Forces and the existence of stresses in invariant continuum mechanics. J. Math. Phys. 27(1), 163–170, 1986

    Article  ADS  MathSciNet  Google Scholar 

  82. Segev, R., Epstein, M.: Geometric Continuum Mechanics, vol. 42. ACM, Basel (2020)

    Book  Google Scholar 

  83. Segev, R., Rodnay, G.: Cauchy’s theorem on manifolds. J. Elast. 56(2), 129–144, 1999

    Article  MathSciNet  Google Scholar 

  84. Simo, J.C., Marsden, J.E.: Stress tensors, Riemannian metrics and the alternative descriptions in elasticity. In: Trends and Applications of Pure Mathematics to Mechanics (Palaiseau, 1983), volume 195 of Lecture Notes in Physics, pp. 369–383. Springer, Berlin (1984)

  85. Simo, J.C., Pister, K.S.: Remarks on rate constitutive equations for finite deformation problems: computational implications. Comput. Methods Appl. Mech. Eng. 46(2), 201–215, 1984

    Article  ADS  Google Scholar 

  86. Souriau, J.-M.: Structure of Dynamical Systems, volume 149 of Progress in Mathematics. Birkhäuser Boston Inc., Boston, MA, 1997. A symplectic view of physics, Translated from the French by C. H. Cushman-de Vries, Translation edited and with a preface by R. H. Cushman and G. M. Tuynman.

  87. Steinmann, P.: Geometrical Foundations of Continuum Mechanics, volume 2 of Lecture Notes in Applied Mathematics and Mechanics. Springer, Heidelberg (2015). An application to first- and second- order elasticity and elasto-plasticity.

  88. Stumpf, H., Hoppe, U.: The application of tensor algebra on manifolds to nonlinear continuum mechanics-invited survey article. Z. Angew. Math. Mech. 77(5), 327–339, 1997

    Article  MathSciNet  Google Scholar 

  89. Svendsen, B.: A local frame formulation of dual stress-strain pairs and time derivatives. Acta Mech. 111(1–2), 13–40, 1995

    Article  MathSciNet  Google Scholar 

  90. Svendsen, B., Tsakmakis, C.: A local differential geometric formulation of dual stress-strain pairs and time derivatives. Arch. Mech. (Arch. Mech. Stos.) 46(1–2), 49–91, 1994

    MathSciNet  Google Scholar 

  91. Truesdell, C.: Hypo-elasticity. J. Ration. Mech. Anal. 4, 83–133, 1955

    MathSciNet  Google Scholar 

  92. Truesdell, C., Noll, W.: The non-linear field theories of mechanics. In: Handbuch der Physik, pp. 1–602. III(3). Springer, Berlin (1965)

  93. Wang, C.C., Truesdell, C.: Introduction to Rational Elasticity. Noordhoff International Publishing, Leyden (1973). Monographs and Textbooks on Mechanics of Solids and Fluids: Mechanics of Continua

  94. **ao, H., Bruhns, O., Meyers, A.: On objective corotational rates and their defining spin tensors. Int. J. Solids Struct. 35(30), 4001–4014, 1998

    Article  MathSciNet  Google Scholar 

  95. **ao, H., Bruhns, O.T., Meyers, A.T.M.: Strain rates and material spins. J. Elast. 52(1), 1–41, 1998

    Article  MathSciNet  Google Scholar 

  96. Yavari, A., Marsden, J.E.: Covariantization of nonlinear elasticity. Zeitschrift für angewandte Mathematik und Physik 63(5), 921–927, 2012

    Article  ADS  MathSciNet  Google Scholar 

  97. Yavari, A., Marsden, J.E., Ortiz, M.: On spatial and material covariant balance laws in elasticity. J. Math. Phys. 47(4), 042903, 2006

    Article  ADS  MathSciNet  Google Scholar 

  98. Zaremba, S.: Sur une forme perfectionnée de la theorie de la relaxation. Bull. Int. Acad. Sci. Cracovie, 534–614 (1903)

Download references

Acknowledgements

It is a pleasure to thank Emmanuelle Rouhaud for stimulating discussions concerning the concept of objective derivative and related to her own work [70, 79]. We would also like to thank Boris Desmorat for the day-long scientific discussions that helped initiate the present work.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to B. Kolev.

Additional information

Communicated by V. Šverák

Publisher's Note

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

This work is dedicated to Professor Paul Rougée.

Appendices

Appendix A. Second-Order Tensors and Their Interpretations

Given a real (finite dimensional) vector space E and denoting its dual by \(E^{\star }\), we can define four types of second-order tensors on E, which we interpret as bilinear map**s.

  1. (1)

    \(\textbf{b}: E \times E \rightarrow \mathbb {R}\), \((\textbf{x},\textbf{y}) \mapsto \textbf{b}(\textbf{x},\textbf{y})\),

  2. (2)

    \(\textbf{b}: E \times E^{\star } \rightarrow \mathbb {R}\), \((\textbf{x},\beta ) \mapsto \textbf{b}(\textbf{x},\beta )\),

  3. (3)

    \(\textbf{b}: E^{\star } \times E \rightarrow \mathbb {R}\), \((\alpha ,\textbf{y}) \mapsto \textbf{b}(\alpha ,\textbf{y})\),

  4. (4)

    \(\textbf{b}: E^{\star } \times E^{\star } \rightarrow \mathbb {R}\), \((\alpha ,\beta ) \mapsto \textbf{b}(\alpha ,\beta )\).

To each of these tensors, we associate, using the convention, of that we call the second argument, a linear map**

  1. (1)

    \(\tilde{\textbf{b}}: E \rightarrow E^{\star }\), \(\textbf{y}\mapsto \textbf{b}(\cdot ,\textbf{y})\),

  2. (2)

    \(\tilde{\textbf{b}}: E^{\star } \rightarrow E^{\star }\), \(\beta \mapsto \textbf{b}(\cdot ,\beta )\),

  3. (3)

    \(\tilde{\textbf{b}}: E \rightarrow (E^{\star })^{\star } = E\), \(\textbf{y}\mapsto \textbf{b}(\cdot ,\textbf{y})\),

  4. (4)

    \(\tilde{\textbf{b}}: E^{\star } \rightarrow (E^{\star })^{\star } = E\), \(\beta \mapsto \textbf{b}(\cdot ,\beta )\),

and if there is no ambiguity, we will not distinguish between \(\textbf{b}\) and \(\tilde{\textbf{b}}\) and only use the notation \(\textbf{b}\). Given a basis \((\varvec{e}_{i})\) of E, and denoting its dual basis by \((\varvec{e}^{i})\), their respective components write , , and .

Given two vector spaces E and F, and a linear map** \(L: E \rightarrow F\), its adjoint (or dual linear map**) is defined as

$$\begin{aligned} L^{\star }: F^{\star } \rightarrow E^{\star }, \qquad \alpha \mapsto \alpha \circ L. \end{aligned}$$

Note that \((L^{-1})^{\star } = (L^{\star })^{-1}\) and \((L_{1} \, L_{2})^{\star } = L_{2}^{\star }\, L_{1}^{\star }\).

Remark A.1

A second-order covariant, or contravariant, tensor \(\textbf{b}\) is symmetric if and only if \(\textbf{b}^{\star } = \textbf{b}\).

When the spaces E and F are respectively equipped with scalar products, noted \(\textbf{q}_{E}\) and \(\textbf{q}_{F}\) respectively, the transpose \(L^{t}: F \rightarrow E\) of a linear map** \(L: E \rightarrow F\) is defined implicitly by the relation

$$\begin{aligned} \langle L\textbf{x}, \textbf{y}\rangle _{F} = \langle \textbf{x}, L^{t}\textbf{y}\rangle _{E}. \end{aligned}$$

The following diagram makes clear the relation between \(L^{\star }\) and \(L^{t}\)

figure c

and leads to

$$\begin{aligned} \textbf{q}_{E}\,L^{t} = L^{\star }\,\textbf{q}_{F}. \end{aligned}$$

Appendix B. Pull-Back and Push-Forward

The fundamental concept of differential geometry that allows to pass from material variables to spatial variables (and vice versa) are the operations of pull-back and push-forward. For functions, these operations are defined by

$$\begin{aligned} p^{*}f = f \circ p\quad \text {(pull-back)}, \qquad p_{*}F = F \circ p^{-1} \quad \text {(push-forward)}, \end{aligned}$$

where \(f \in \textrm{C}^{\infty }(\Omega ,\mathbb {R})\) and \(F \in \textrm{C}^{\infty }(\mathcal {B},\mathbb {R})\). For vector fields, the following diagram

figure d

leads immediately to the natural definitions

$$\begin{aligned} p^{*}\varvec{u}= Tp^{-1} \circ \varvec{u}\circ p\quad \text {(pull-back)}, \qquad p_{*}\varvec{U}= Tp\circ \varvec{U}\circ p^{-1} \quad \text {(push-forward)}. \end{aligned}$$

For covector fields, the following diagram

figure e

leads to the following definitions

$$\begin{aligned} p^{*}\beta = Tp^{\star } \circ \beta \circ p\quad \text {(pull-back)}, \qquad p_{*}\alpha = (Tp^{\star })^{-1} \circ \alpha \circ p^{-1} \quad \text {(push-forward)}. \end{aligned}$$

The pull-back and push-forward operations are inverse to each other, meaning that \(p^{*} = (p_{*})^{-1} = (p^{-1})_{*}\). They are easily extended to higher-order contravariant, covariant or mixed tensor fields.

In local coordinate systems, \((X^{I})\) on \(\mathcal {B}\) and \((x^{i})\) on \(\mathcal {E}\), where we have set \(p(\textbf{X}) = \textbf{x}\), the linear tangent map \(Tp: T\mathcal {B}\rightarrow T\mathcal {E}\) is represented by the square matrix \(\textbf{F}\) defined by

and its dual tangent linear map \(Tp^{\star }: T^{\star }_{\textbf{x}}\mathcal {E}\rightarrow T^{\star }_{\textbf{X}}\mathcal {B}\), by

We will write \(\textbf{F}^{-\star }:= (\textbf{F}^{\star })^{-1}=\left( \textbf{F}^{-1}\right) ^{\star }\).

Proposition B.1

For tensor fields of order one or two, we have the following expressions.

  1. (1)

    For contravariant vector fields \(\varvec{W}=(W^{I})\), \(\varvec{w}=(w^{i})\), we have

    $$\begin{aligned} p_{*} \varvec{W}= \textbf{F}\left( \varvec{W}\circ p^{-1}\right) , \qquad p^{*} \varvec{w}= \textbf{F}^{-1} \left( \varvec{w}\circ p\right) . \end{aligned}$$
  2. (2)

    For covariant vector fields \(\alpha =(\alpha _{I})\), \(\beta =(\beta _{i})\), we have

    $$\begin{aligned} p_{*} \alpha = \textbf{F}^{-\star }\left( \alpha \circ p^{-1}\right) , \qquad p^{*} \beta = \textbf{F}^{\star }\left( \beta \circ p\right) . \end{aligned}$$
  3. (3)

    For second-order covariant tensor fields \({\varvec{\varepsilon }}= (\varepsilon _{IJ})\), \(\textbf{k}= (k_{ij})\), we have

    $$\begin{aligned} p_{*} {\varvec{\varepsilon }}= \textbf{F}^{-\star } \left( {\varvec{\varepsilon }}\circ p^{-1}\right) \textbf{F}^{-1}, \qquad p^{*} \textbf{k}= \textbf{F}^{\star } (\textbf{k}\circ p) \textbf{F}. \end{aligned}$$
  4. (4)

    For second-order contravariant tensor fields \({\varvec{\theta }}= ({\varvec{\theta }}^{IJ})\) and \({\varvec{\tau }}= ({\varvec{\tau }}^{ij})\), we have

    $$\begin{aligned} p_{*} {\varvec{\theta }}= \textbf{F}\left( {\varvec{\theta }}\circ p^{-1}\right) \textbf{F}^{\star }, \qquad p^{*} {\varvec{\tau }}= \textbf{F}^{-1} \left( {\varvec{\tau }}\circ p\right) \textbf{F}^{-\star }. \end{aligned}$$
  5. (5)

    For second-order mixed tensor fields \(\hat{\textbf{T}} = ({\hat{T}^{I}}_{\; J})\) and \(\hat{\textbf{t}} = ({\hat{t}^{i}}_{\, j})\), we have

    $$\begin{aligned} p_{*} \hat{\textbf{T}} = \textbf{F}\left( \hat{\textbf{T}} \circ p^{-1}\right) \textbf{F}^{-1}, \qquad p^{*} \hat{\textbf{t}} = \textbf{F}^{-1} \left( \hat{\textbf{t}} \circ p\right) \textbf{F}. \end{aligned}$$
  6. (6)

    For second-order mixed tensor fields \(\check{\textbf{T}} = ({\check{T}_{I}}^{\,J})\) and \(\check{\textbf{t}} = ({\check{t}_{i}}^{\; j})\), we have

    $$\begin{aligned} p_{*} \check{\textbf{T}} = \textbf{F}^{-\star } (\check{\textbf{T}} \circ p^{-1}) \textbf{F}^{\star }, \qquad p^{*} \check{\textbf{t}} = \textbf{F}^{\star } (\check{\textbf{t}} \circ p)\textbf{F}^{-\star }. \end{aligned}$$

Remark B.2

The push-forward and pull-back operations commute with the contraction between covariant and contravariant tensors. So in particular, we get

$$\begin{aligned} (p_{*}\alpha ) \cdot (p_{*}\varvec{W}) = p_{*}(\alpha \cdot \varvec{W}), \qquad (p_{*} {\varvec{\theta }}): (p_{*} {\varvec{\varepsilon }}) = p_{*} ({\varvec{\theta }}: {\varvec{\varepsilon }}), \qquad {{\,\textrm{tr}\,}}(p_{*} \hat{\textbf{T}}) = p_{*}({{\,\textrm{tr}\,}}\hat{\textbf{T}}). \end{aligned}$$

Appendix C. Lie Derivative

Given a manifold M, the Lie derivative of a (time-independent) tensor field \(\textbf{t}\in \mathbb {T}(M)\) corresponds to the infinitesimal version of the pull-back operation. More precisely, let \(\varvec{u}\) be a vector field on M, \(\phi (t)\) its flow and \(\textbf{t}\) be a tensor field on M. The Lie derivative \({{\,\textrm{L}\,}}_{\varvec{u}} \textbf{t}\) of the tensor field \(\textbf{t}\) with respect to \(\varvec{u}\) is defined by

$$\begin{aligned} {{\,\textrm{L}\,}}_{\varvec{u}} \textbf{t}:= \left. \frac{\partial }{\partial t}\right| _{t=0} \phi (t)^{*} \textbf{t}. \end{aligned}$$

The Lie derivative has the following properties.

  1. (1)

    When \(\textbf{t}= \varvec{v}\) is a vector field, \({{\,\textrm{L}\,}}_{\varvec{u}} \varvec{v}= [\varvec{u},\varvec{v}]\), where \([\varvec{u},\varvec{v}]\) is the Lie bracket of the two vector fields \(\varvec{u}\) and \(\varvec{v}\).

  2. (2)

    Given a diffeomorphism \(\varphi \) of M, we have

    $$\begin{aligned} \varphi ^{*} {{\,\textrm{L}\,}}_{\varvec{u}} \textbf{t}= {{\,\textrm{L}\,}}_{\varphi ^{*}\varvec{u}} \varphi ^{*}\textbf{t}. \end{aligned}$$
    (C.1)
  3. (3)

    At any time t where \(\phi (t)\) is defined, we have (note that \(\phi (t)^{*} \varvec{u}=\varvec{u}\))

    $$\begin{aligned} \frac{\partial }{\partial t}\left( \phi (t)^{*} \textbf{t}\right) = \phi (t)^{*} {{\,\textrm{L}\,}}_{\varvec{u}} \textbf{t}= {{\,\textrm{L}\,}}_{\varvec{u}} \phi (t)^{*} \textbf{t}. \end{aligned}$$
    (C.2)
  4. (4)

    Given two vector fields \(\varvec{u},\varvec{v}\) on M, we have

    $$\begin{aligned} {{\,\textrm{L}\,}}_{[\varvec{u},\varvec{v}]} \textbf{t}= {{\,\textrm{L}\,}}_{\varvec{u}} {{\,\textrm{L}\,}}_{\varvec{v}} \textbf{t}- {{\,\textrm{L}\,}}_{\varvec{v}} {{\,\textrm{L}\,}}_{\varvec{u}} \textbf{t}. \end{aligned}$$

Consider now a time dependent vector field \(\varvec{u}(t)\) on M. Its flow \(\phi (t,s)\) is defined as the solution a time t of the initial value problem

$$\begin{aligned} \dot{c}(t) = \varvec{u}(t,c(t)), \qquad c(s)=\textbf{x}. \end{aligned}$$
(C.3)

Then, \(\phi (t,s)\) is a local diffeomorphism with inverse \(\phi (s,t)\) and we have

$$\begin{aligned} \phi (t,s) = \phi (t,\tau ) \circ \phi (\tau ,s), \end{aligned}$$

as soon as the three map**s are defined. The Lie derivative can be extended to time dependent vector fields as follows.

$$\begin{aligned} {{\,\textrm{L}\,}}_{\varvec{u}(t)} \textbf{t}:= \left. \frac{\partial }{\partial \tau }\right| _{\tau =t} \phi (\tau ,t)^{*} \textbf{t}. \end{aligned}$$

Remark C.1

Let \(\widetilde{\varphi } = (\varphi (t))\) be a path of diffeomorphism. Its Eulerian velocity is defined as the time dependent vector field \(\varvec{u}(t):= \partial _{t}\varphi \circ \varphi (t)^{-1}\). Given a (time-independent) tensor field \(\textbf{t}\), we have

$$\begin{aligned} \frac{\partial }{\partial t} \left( \varphi (t)^{*} \textbf{t}\right) =\varphi (t)^{*} ({{\,\textrm{L}\,}}_{\varvec{u}(t)} \textbf{t}) \qquad \text {and} \qquad \frac{\partial }{\partial t} \left( \varphi (t)_{*} \textbf{t}\right) =-{{\,\textrm{L}\,}}_{\varvec{u}(t)} (\varphi (t)_{*} \textbf{t}). \end{aligned}$$

The preceding results extend to paths of embeddings between two manifolds \(\mathcal {B}\) and M and will be summarized by the following lemma.

Lemma C.2

Let \(\widetilde{p} = (p(t))\) be a path of embeddings, \(\varvec{u}(t):= (\partial _{t}p)\circ p(t)^{-1}\) be its (right) Eulerian velocity and \(\textbf{t}(t)\) be a tensor field defined along \(p(t)\) (i.e. on \(\Omega _{p(t)} = p(t)(\mathcal {B})\) and possibly time-dependent). Then

$$\begin{aligned} \partial _{t}(p(t)^{*}\textbf{t}(t)) = p(t)^{*} \left( \partial _{t}\textbf{t}+ {{\,\textrm{L}\,}}_{\varvec{u}(t)}\textbf{t}(t) \right) . \end{aligned}$$

Proof

Let \(\phi (t,s)\) be the flow of \(\varvec{u}(t)\). Then we get

$$\begin{aligned} p(s) = \phi (s,t) \circ p(t). \end{aligned}$$

We have thus

$$\begin{aligned} \partial _{t}(p(t)^{*}\textbf{t}(t))&= \left. \frac{\partial }{\partial s}\right| _{s=t} (\phi (s,t) \circ p(t))^{*} \textbf{t}(s)\\&= p(t)^{*} \left. \frac{\partial }{\partial s} \right| _{s=t} \phi (s,t)^{*} \textbf{t}(s)\\&= p(t)^{*}\left( \partial _{t} \textbf{t}(t) + {{\,\textrm{L}\,}}_{\varvec{u}(t)}\textbf{t}(t)\right) . \end{aligned}$$

\(\square \)

Appendix D. Vector Bundles and Covariant Derivatives

The interested reader may consult [30, 53, 57, 60, 82] for complementary points of view on differential geometry. Let \(\mathbb {E}\) be a vector bundle over a manifold M. We denote by \(\Gamma (\mathbb {E})\) the space of smooth sections of \(\mathbb {E}\) and by \(\Omega ^{k}(M,\mathbb {E})\) the space of k-forms with values in \(\mathbb {E}\), in other words, the space of sections of the vector bundle \(\Lambda ^{k}T^{\star }M \otimes \mathbb {E}\) (in particular, \(\Omega ^{0}(M,\mathbb {E}) = \Gamma (\mathbb {E})\)).

Definition D.1

(Covariant derivative) A covariant derivative on vector bundle \(\mathbb {E}\) over M is a linear operator

$$\begin{aligned} \nabla : \Gamma (\mathbb {E}) \rightarrow \Omega ^{1}(M,\mathbb {E}), \qquad \textbf{s}\mapsto \nabla \textbf{s}, \end{aligned}$$

which satisfies the Leibniz identity.

$$\begin{aligned} \nabla (f\textbf{s}) = df \otimes \textbf{s}+ f \, \nabla \textbf{s}, \end{aligned}$$

for any function \(f \in \textrm{C}^{\infty }(M)\) and any section \(\textbf{s}\in \Gamma (\mathbb {E})\).

Remark D.2

The set of all covariant derivatives defined on a given vector bundle \(\mathbb {E}\) has an affine structure. Indeed, given two covariant derivatives \(\nabla ^{1}\) and \(\nabla ^{2}\), the difference \(\nabla ^{2}-\nabla ^{1}\) is a section of the vector bundle

$$\begin{aligned} (T^{\star }M \otimes \mathbb {E}^{\star }) \otimes \mathbb {E}. \end{aligned}$$

Hence, this set is well an affine space with associated vector space \(\Gamma ((T^{\star }M \otimes \mathbb {E}^{\star }) \otimes \mathbb {E})\).

Remark D.3

If a vector bundle \(\mathbb {E}\) with base manifold M and fiber type E (a vector space) is trivializable, meaning that there exists a vector bundle isomorphism

$$\begin{aligned} \Psi : \mathbb {E}\rightarrow M \times E, \qquad \varvec{v}_{x} \mapsto (x,\varvec{v}), \end{aligned}$$

then, each section \(\textbf{s}\) of \(\mathbb {E}\) corresponds bijectively to a vector valued function S defined by

$$\begin{aligned} S: M \rightarrow E, \qquad x \mapsto p_{2} \circ \Psi (\textbf{s}(x)), \end{aligned}$$

where \(p_{2}: M \times E\) is the projection onto the second factor. Therefore, there is a canonical covariant derivative associated with this trivialization which is given by

$$\begin{aligned} (\nabla _{X} \textbf{s})(x):= \Psi ^{-1}(x, d_{x}S.X). \end{aligned}$$

Definition D.4

(Curvature) Given a covariant derivative \(\nabla \) on a vector bundle \(\mathbb {E}\), its curvature is the map**

$$\begin{aligned} R: \Gamma (\mathbb {E}) \rightarrow \Omega ^{2}(M,\mathbb {E}) \end{aligned}$$

defined by

$$\begin{aligned} R(X,Y)\textbf{s}:= \nabla _{X}\nabla _{Y}\textbf{s}- \nabla _{Y}\nabla _{X}\textbf{s}- \nabla _{[X,Y]}\textbf{s}. \end{aligned}$$

When \(\mathbb {E}= TM\) is the tangent bundle of a manifold M, we define the torsion of this covariant derivative, by the formula

$$\begin{aligned} T(X,Y):= \nabla _{X} Y - \nabla _{Y} X - [X, Y], \qquad X, Y \in \textrm{Vect}(M), \end{aligned}$$

which is a mixed tensor field of type (1, 2). The curvature tensor of \(\nabla \) is a mixed tensor field of type (1, 3), which writes

$$\begin{aligned} R(X,Y)Z = \nabla _{X}\nabla _{Y}Z - \nabla _{Y} \nabla _{X}Z - \nabla _{[X,Y]}Z, \qquad X, Y, Z \in \textrm{Vect}(M). \end{aligned}$$

Definition D.5

A covariant derivative on the tangent bundle TM of a manifold M is symmetric if its torsion is zero, that is if

$$\begin{aligned} \nabla _{\varvec{v}} \varvec{w}- \nabla _{\varvec{w}} \varvec{v}= [\varvec{v}, \varvec{w}], \qquad \forall \varvec{v}, \varvec{w}\in \textrm{Vect}(M), \end{aligned}$$

where \([\varvec{v}, \varvec{w}]:={{\,\textrm{L}\,}}_{\varvec{v}} \varvec{w}\) is the Lie bracket of the vector fields \(\varvec{v}\) and \(\varvec{w}\).

Remark D.6

One can show the existence on any differential manifold M of a covariant derivative. However, there are an infinite number of such derivatives and none of them play a particular role. On the other hand, if a manifold M has a Riemannian metric g, then there is a unique symmetric covariant derivative \(\nabla \) such that \(\nabla g = 0\) (see for example [30, Theorem 2.51]), this is the Riemannian covariant derivative.

Any covariant derivative on TM induces by the Leibniz rule a covariant derivative on all tensor bundles of M. The link between the Lie derivative and a symmetric covariant derivative is then recalled in the following theorem.

Proposition D.7

Let M be a differential manifold with a symmetric covariant derivative \(\nabla \). Then we have the following relations:

  1. (1)

    Lie derivative of a function f:

    $$\begin{aligned} {{\,\textrm{L}\,}}_{\varvec{u}} f = \nabla _{\varvec{u}} f = df.\varvec{u}; \end{aligned}$$
  2. (2)

    Lie derivative of a vector field \(\varvec{w}= (w^{i})\):

    $$\begin{aligned} {{\,\textrm{L}\,}}_{\varvec{u}} \varvec{w}= [\varvec{u}, \varvec{w}] = \nabla _{\varvec{u}} \varvec{w}- \nabla _{\varvec{w}} \varvec{u}; \end{aligned}$$
  3. (3)

    Lie derivative of a covector field (1-form) \(\alpha = (\alpha _{i})\):

    $$\begin{aligned} {{\,\textrm{L}\,}}_{\varvec{u}} \alpha = \nabla _{\varvec{u}} \alpha + (\nabla \varvec{u})^{\star } \alpha ; \end{aligned}$$
  4. (4)

    Lie derivative of a second-order covariant tensor field, \(\textbf{k}= (k_{ij})\):

    $$\begin{aligned} {{\,\textrm{L}\,}}_{\varvec{u}} \textbf{k}= \nabla _{\varvec{u}} \textbf{k}+ (\nabla \varvec{u})^{\star } \textbf{k}+ \textbf{k}(\nabla \varvec{u}); \end{aligned}$$
  5. (5)

    Lie derivative of a second-order contravariant tensor field, \({\varvec{\tau }}= (\tau ^{ij})\):

    $$\begin{aligned} {{\,\textrm{L}\,}}_{\varvec{u}} {\varvec{\tau }}= \nabla _{\varvec{u}} {\varvec{\tau }}- (\nabla \varvec{u}) {\varvec{\tau }}- {\varvec{\tau }}(\nabla \varvec{u})^{\star }; \end{aligned}$$
  6. (6)

    Lie derivative of a second-order mixed tensor field, :

    $$\begin{aligned} {{\,\textrm{L}\,}}_{\varvec{u}} \hat{\textbf{t}} = \nabla _{\varvec{u}} \hat{\textbf{t}} - (\nabla \varvec{u}) \hat{\textbf{t}} + \hat{\textbf{t}} (\nabla \varvec{u}); \end{aligned}$$
  7. (7)

    Lie derivative of a second-order mixed tensor field, :

    $$\begin{aligned} {{\,\textrm{L}\,}}_{\varvec{u}} \check{\textbf{t}} = \nabla _{\varvec{u}} \check{\textbf{t}} + (\nabla \varvec{u})^{\star } \check{\textbf{t}} - \check{\textbf{t}} (\nabla \varvec{u})^{\star }. \end{aligned}$$

In order to intrinsically define the geodesic equation of a Riemannian manifold, but also the covariant derivative of the Lagrangian velocity, it is necessary to extend the notion of covariant derivative to vector fields (and more generally tensor fields) which are only defined along a map** \(f: K \rightarrow M\), between two manifolds K and M. The rigorous formulation of such a definition requires first to introduce the notion of pull-back of a vector bundle [41].

Definition D.8

Let \(\pi : \mathbb {E}\rightarrow M\) be a vector bundle and \(f: K \rightarrow M\) be a smooth map**. Then the set

$$\begin{aligned} f^{*}\mathbb {E}:= \bigsqcup _{k \in K} E_{f(k)} \subset \mathbb {E}\end{aligned}$$

is a vector bundle above K, referred to as the pull-back by f of the vector bundle \(\mathbb {E}\). A section of this bundle is therefore a map** \(s: K \rightarrow \mathbb {E}\), such as \(\pi (s(k)) = f(k)\).

Example D.9

A vector field defined along a curve \(c: I \rightarrow M\) is a curve \(X: I \rightarrow TM\), such that \(X(t) \in T_{c(t)}M\), for any \(t \in I\).

Example D.10

The Lagrangian velocity \(\varvec{V}(t): \mathcal {B}\rightarrow T\mathcal {E}\), at time t, is a section of the pullback bundle \(p(t)^{*}T\mathcal {E}\).

Proposition D.11

Let \(\pi : \mathbb {E}\rightarrow M\) be a vector bundle, equipped with a covariant derivative \(\nabla \) and \(f: K \rightarrow M\), a smooth map**. Then, there exists a unique covariant derivative, denoted \(f^{*}\nabla \), on the vector bundle \(f^{*}\mathbb {E}\), called the pull-back of \(\nabla \), and such that

$$\begin{aligned} (f^{*}\nabla )_{X} (f \circ s) = f^{*}\left( \nabla _{Tf.X} s\right) , \end{aligned}$$

for any section s of \(\mathbb {E}\) and any vector field X on K.

Example D.12

Consider the case where \(K = \mathcal {B}\), \(M=\mathcal {E}\), \(\mathbb {E}= T\mathcal {E}\) is the tangent vector bundle to \(\mathcal {E}\) and

$$\begin{aligned} f = p: \mathcal {B}\rightarrow \mathcal {E}, \end{aligned}$$

is an embedding of \(\mathcal {B}\) in \(\mathcal {E}\). Consider a local coordinate system \((X^{I})\) on \(\mathcal {B}\) and a local coordinate system \((x^{i})\) on \(\mathcal {E}\), where Christoffel’s symbols are written \(\Gamma _{ij}^{k}\). Let \(\varvec{V}\) be a section of \(p^{*}T\mathcal {E}\), i.e. a map**

$$\begin{aligned} \varvec{V}: \mathcal {B}\rightarrow T\mathcal {E}, \quad \text {such that} \quad \pi (X(X)) = p(X), \end{aligned}$$

then, we have

$$\begin{aligned} {[}(p^{*}\nabla )_{\partial _{X^{I}}} V]^{k} = \partial _{X^{I}} V^{k} + \Gamma _{ij}^{k} \frac{\partial p^{i}}{\partial _{X^{I}}} V^{j}. \end{aligned}$$

Example D.13

Let \((M,{\varvec{\gamma }})\) be a Riemannian manifold and \(\nabla \) the associated covariant derivative. Let \(c: I \rightarrow M\) be curve. Then, the pull-back of \(\nabla \) by c, usually noted \(D_{t}\) is defined on the vector space of vector fields defined along c. It is characterized by the following properties:

  1. (1)

    for any function \(f: I \rightarrow \mathbb {R}\),

    $$\begin{aligned} D_{t}(fX)(t) = f^{\prime }(t)\, X(t) + f(t)\, D_{t}(X)(t); \end{aligned}$$
  2. (2)

    If \(X(t) = \tilde{X}(c(t))\) where \(\tilde{X}\) is a vector field on M, then

    $$\begin{aligned} (D_{t}X)(t) = (\nabla _{c^{\prime }(t)}\tilde{X})((t)). \end{aligned}$$

In a local coordinate system \((x^{i})\) of M, we have:

$$\begin{aligned} (D_{t}X)^{k} = \partial _{t}X^{k} + \Gamma _{ij}^{k} (\partial _{t}x^{i})X^{j}. \end{aligned}$$

Remark D.14

On an infinite dimensional manifold, equipped with a covariant derivative, the curvature is defined along a parameterized surface c(st), and writes

$$\begin{aligned} R(\partial _s, \partial _{t})X = D_{s}D_{t}X - D_{t}D_{s}X. \end{aligned}$$

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

Kolev, B., Desmorat, R. Objective Rates as Covariant Derivatives on the Manifold of Riemannian Metrics. Arch Rational Mech Anal 248, 66 (2024). https://doi.org/10.1007/s00205-024-02010-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00205-024-02010-x

Mathematics Subject Classification

Navigation