Skip to main content
Log in

Incompatible Deformations in Relativistic Elasticity

  • Published:
Lobachevskii Journal of Mathematics Aims and scope Submit manuscript

Abstract

The work develops differential-geometric methods for modeling incompatible deformations and stresses caused by them in hyperelastic supermassive bodies within the framework of special and general relativity. Particular attention is paid to a unified geometric language used both for modeling distributed defects and gravitational interaction. It is shown that finite incompatible deformations that do not evolve in time can be formalized in terms of the curvature or torsion of the connection on the 3D material manifold obtained as the image of the projection of the body world tube, while the gravitational interaction is formalized in terms of the curvature and torsion of 4D space-time manifold. To describe the evolution of the material composition of a body, which occurs, for example, as a result of accretion, instead of a 3D material manifold, one should consider a body-tube in a 4D material space. In this case, from a formal mathematical point of view, the description of distributed defects and gravity becomes equivalent, and both, material manifold and space-time manifold are derived from a single fundamental object, which is a four-dimensional topological vector space. The developed approach can be used to model the processes of evolution of the stress-strain state of neutron star crusts and their local spontaneous destruction.

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

Notes

  1. These points can be treated as positions in the framework of conventional continuum mechanics.

  2. Here \(I_{m}\) denotes totally ordered \(m\)-element index set.

  3. This follows from the corresponding property of geometric tangent vectors.

  4. Here we put \(I_{3}=\{1,\,2,\,3\}\) for the index set of the preferred basis.

  5. Which is assumed to be open interval.

  6. In this formula \(dt\otimes\boldsymbol{\epsilon}\) denotes tensor product of Euclidean measures.

  7. Symbol \({\mathop{D}\limits_{{{4}}}}\cdot\) means divergence operation.

  8. Within the framework of Relativity an observer is required to be a massive particle with moving orthonormal basis. Dealing with continuous matter, we restrict ourselves with internal observer only, i.e., every particle is observer on its own right. By this reason, one needs to consider a matter that consists of massive particles only. The general case is beyond the scope of the present work.

  9. Explicitly, if \(4\)-acceleration \(\mathbf{a}\) of the observer vanishes (the world-line is a straight line), this neighborhood coincides with the whole Minkowski space-time, while if this acceleration doesn’t vanish, then the neighborhood is formed by points at distances less than \(\frac{1}{\sqrt{\mathbf{g}(\mathbf{a},\,\mathbf{a})}}\) from the world-line ([51], p. 92).

  10. With translation vector space, that equals to spatial platform \(\mathcal{V}^{3}_{\mathbf{c}_{0}}\), and metric, induced by \(\mathbf{g}\).

  11. The vector and affine structures are now in shadow and we are not able to extract them explicitly. Note, that the underlying manifold can change only in special case, when the space-time singularity occurs. This case is beyond the scope of present work.

  12. That is, Einstein gravitational constant related with Newtonian gravitational constant \(G\) as \(\varkappa=\frac{8\pi G}{c^{4}}\).

  13. Inverse deformations were considered by Schield in [88].

  14. In the general case, we consider solids with memory and take into account the history of events in order to evaluate the reaction of the material in a given event. In this case, the lower part of the world-line is considered, and the upper part is not taken into account due to the principle of definiteness, that is typically adopted in classical mechanics [62].

  15. Formally, one should use another designation for Lagrangian density, but for the economical writing we here and in what follows use the same symbol \(L\).

  16. The comprehensive treatment of such space-time foliations can be found in [92].

  17. According to Whitney’s theorem, such an embedding can be always performed [96].

  18. This problem is especially significant when constructing the geometry space-time. If the observer is inside it, then he cannot fully identify the geometry of the space in which the space-time is embedded. It is only possible to construct the space-time geometry by means that are available to the observer.

  19. It should be noted, however, that we do not completely break with the <<external>> geometry; it is used as a support for thought.

  20. It is worth to mention that at least one affine connection always exists on manifold [15].

  21. This feature is due to the transformation law (A1).

  22. Within interpretation of vector fields as derivations of algebra \(C^{\infty}(M)\).

  23. Here it was taken into account that \(ddg_{jk}=\mathsf{0}\).

REFERENCES

  1. G. Maugin, Continuum Mechanics through the Eighteenth and Nineteenth Centuries: Historical Perspectives from John Bernoulli (1727) to Ernst Hellinger (1914) (Springer, Cham, 2014).

    Book  MATH  Google Scholar 

  2. S. Hawking and G. Ellis, The Large Scale Structure of Space-Time (Cambridge Univ. Press, Cambridge, 1973).

    Book  MATH  Google Scholar 

  3. G. Maugin, Continuum Mechanics through the Twentieth Century: A Concise Historical Perspective (Springer, Dordrecht, 2013).

    Book  MATH  Google Scholar 

  4. B. Bilby, R. Bullough, and E. Smith, ‘‘Continuous distributions of dislocations: A new application of the methods of non-Riemannian geometry,’’ Proc. R. Soc. London, Ser. A 231, 263–273 (1955).

    Article  MathSciNet  Google Scholar 

  5. K. Kondo, ‘‘Non-Riemannian geometry of imperfect crystals from a macroscopic viewpoint,’’ in Memoirs of the Unifying Study of the Basic Problems in Engineering Science by Means of Geometry (Gakujutsu Bunken Fukyu-Kai, Tokyo, 1955), Vol. 1, pp. 6–17.

    MATH  Google Scholar 

  6. K. Kondo, ‘‘Non-Riemannian and Finslerian approaches to the theory of yielding,’’ Int. J. Eng. Sci. 1, 71–88 (1963).

    Article  MATH  Google Scholar 

  7. E. Kröner, ‘‘Allgemeine Kontinuumstheorie der Versetzungen und Eigenspannungen,’’ Arch. Ration. Mech. Anal. 4(273) (1959).

  8. W. Noll, ‘‘Materially uniform simple bodies with inhomogeneities,’’ Arch. Ration. Mech. Anal. 27 (1), 1–32 (1967).

    Article  MathSciNet  MATH  Google Scholar 

  9. C.-C. Wang, ‘‘On the geometric structures of simple bodies, a mathematical foundation for the theory of continuous distributions of dislocations,’’ Arch. Ration. Mech. Anal. 27, 33–94 (1967).

    Article  MathSciNet  MATH  Google Scholar 

  10. L. Rakotomanana, A Geometric Approach to Thermomechanics of Dissipating Continua (Birkhäuser, Boston, 2004).

    Book  MATH  Google Scholar 

  11. P. Steinmann, Geometrical Foundations of Continuum Mechanics: An Application to First- and Second-Order Elasticity and Elasto-Plasticity (Springer, Berlin, 2015).

    Book  MATH  Google Scholar 

  12. S. Lychev and K. Koifman, Geometry of Incompatible Deformations: Differential Geometry in Continuum Mechanics (De Gruyter, Berlin, 2018).

    Book  MATH  Google Scholar 

  13. A. Yavari and A. Goriely, ‘‘Riemann–Cartan geometry of nonlinear dislocation mechanics,’’ Arch. Ration. Mech. Anal. 205, 59–118 (2012).

    Article  MathSciNet  MATH  Google Scholar 

  14. A. Yavari and A. Goriely, ‘‘Weyl geometry and the nonlinear mechanics of distributed point defects,’’ Proc. R. Soc. London, Ser. A 468, 3902–3922 (2012).

    MathSciNet  MATH  Google Scholar 

  15. M. Postnikov, Geometry VI: Riemannian Geometry (Springer, Berlin, 2001).

    Book  MATH  Google Scholar 

  16. G. Ferrarese and D. Bini, Introduction to Relativistic Continuum Mechanics (Springer, Berlin, 2008).

    Book  MATH  Google Scholar 

  17. A. Romano and M. Furnari, The Physical and Mathematical Foundations of the Theory of Relativity: A Critical Analysis (Birkhäuser, Cham, 2019).

    Book  MATH  Google Scholar 

  18. M. Karlovini and L. Samuelsson, ‘‘Elastic stars in general relativity: I. Foundations and equilibrium models,’’ Class. Quantum Grav. 20, 3613 (2003).

    Article  MathSciNet  MATH  Google Scholar 

  19. M. Karlovini, L. Samuelsson, and M. Zarroug, ‘‘Elastic stars in general relativity: II. Radial perturbations,’’ Class. Quantum Grav. 21, 1559 (2004).

    Article  MathSciNet  MATH  Google Scholar 

  20. M. Karlovini and L. Samuelsson, ‘‘Elastic stars in general relativity: III. Stiff ultrarigid exact solutions,’’ Class. Quantum Grav. 21, 4531 (2004).

    Article  MathSciNet  MATH  Google Scholar 

  21. M. Karlovini and L. Samuelsson, ‘‘Elastic stars in general relativity: IV. Axial perturbations,’’ Class. Quantum Grav. 24, 3171 (2007).

    Article  MathSciNet  MATH  Google Scholar 

  22. C. Brown, L. Goodman, and H. Jeffreys, ‘‘Gravitational stresses in accreted bodies,’’ Proc. R. Soc. London, Ser. A 276 (1367), 571–576 (1963).

    Article  MATH  Google Scholar 

  23. J. Kadish, J. Barber, and P. Washabaugh, ‘‘Stresses in rotating spheres grown by accretion,’’ Int. J. Solids Struct. 42, 5322–5334 (2005).

    Article  MATH  Google Scholar 

  24. J. Kadish, J. Barber, P. Washabaugh, and D. Scheeres, ‘‘Stresses in accreted planetary bodies,’’ Int. J. Solids Struct. 45, 540–550 (2008).

    Article  MATH  Google Scholar 

  25. A. Manzhirov and D. Parshin, ‘‘Accretion of a viscoelastic ball in a centrally symmetric force field,’’ Mech. Solids 41, 51–64 (2006).

    Google Scholar 

  26. S. Shapiro and S. Teukolsky, Black Holes, White Dwarfs, and Neutron Stars: The Physics of Compact Objects (Wiley-VCH, Weinheim, 1983).

    Book  Google Scholar 

  27. N. Andersson and G. L. Comer, ‘‘Relativistic fluid dynamics: Physics for many different scales,’’ Living Rev Relativ. 24 (3) (2021).

  28. J. Frank, A. King, and D. Raine, Accretion Power in Astrophysics (Cambridge Univ. Press, Cambridge, 2002).

    Book  Google Scholar 

  29. H. Silva and N. Yunes, ‘‘Neutron star pulse profiles in scalar-tensor theories of gravity,’’ Phys. Rev. D 99, 044034 (2019).

    Article  MathSciNet  Google Scholar 

  30. M. Rigoselli, S. Mereghetti, R. Turolla, R. Taverna, V. Suleimanov, and A. Potekhin, ‘‘Thermal emission and magnetic beaming in the radio and X-ray mode-switching PSR B0943+10,’’ Astrophys. J. 872, 15 (2019).

    Article  Google Scholar 

  31. A. Danilenko, A. Karpova, D. Ofengeim, Y. Shibanov, and D. Zyuzin, ‘‘XMM-Newton observations of a gamma-ray pulsar J0633+0632: Pulsations, cooling and large-scale emission,’’ Mon. Not. R. Astron. Soc. 493, 1874–1887 (2020).

    Article  Google Scholar 

  32. S. Sengupta, ‘‘General relativistic effects on the induced electric field exterior to pulsars,’’ Astrophys. J. 449, 224 (1995).

    Article  Google Scholar 

  33. A. Chugunov and C. Horowitz, ‘‘Breaking stress of neutron star crust,’’ Mon. Not. R. Astron. Soc. Lett. 407, L54–L58 (2010).

    Article  Google Scholar 

  34. S. Lander, N. Andersson, D. Antonopoulou, and A. Watts, ‘‘Magnetically driven crustquakes in neutron stars,’’ Mon. Not. R. Astron. Soc. 449, 2047–2058 (2015).

    Article  Google Scholar 

  35. T. Wood and R. Hollerbach, ‘‘Three dimensional simulation of the magnetic stress in a neutron star crust,’’ Phys. Rev. Lett. 114, 191101 (2015).

    Article  Google Scholar 

  36. B. Carter and H. Quintana, ‘‘Foundations of general relativistic high-pressure elasticity theory,’’ Proc. R. Soc. London, Ser. A 331, 57–83 (1972).

    Article  MathSciNet  MATH  Google Scholar 

  37. M. Ruggiero and A. Tartaglia, ‘‘Einstein–Cartan theory as a theory of defects in space-time,’’ Am. J. Phys. 71, 1303–1313 (2003).

    Article  MathSciNet  MATH  Google Scholar 

  38. D. Bennett, C. Das, L. Laperashvili, and H. Nielsen, ‘‘The relation between the model of a crystal with defects and Plebanski’s theory of gravity,’’ Int. J. Mod. Phys. A 28, 1350044 (2013).

    Article  MathSciNet  Google Scholar 

  39. J. Synge, ‘‘A theory of elasticity in general relativity,’’ Math. Z. 72, 82–87 (1959).

    Article  MathSciNet  MATH  Google Scholar 

  40. C. Rayner, ‘‘Elasticity in general relativity,’’ Proc. R. Soc. London, Ser. A 272, 44–53 (1963).

    Article  MathSciNet  MATH  Google Scholar 

  41. M. Epstein, D. Burton, and R. Tucker, ‘‘Relativistic anelasticity,’’ Class. Quantum Grav. 23, 3545–3571 (2006).

    Article  MathSciNet  MATH  Google Scholar 

  42. L. Andersson, T. Oliynyk, and B. Schmidt, ‘‘Dynamical compact elastic bodies in general relativity,’’ Arch. Ration. Mech. Anal. 220, 849–887 (2016).

    Article  MathSciNet  MATH  Google Scholar 

  43. H. Weyl, Space, Time, Matter (Dover, New York, 1952).

    Google Scholar 

  44. R. Penrose and W. Rindler, Spinors and Space-Time, Vol. 1: Two-Spinor Calculus and Relativistic Fields (Cambridge Univ. Press, Cambridge, 1984).

  45. M. Postnikov, Lectures in Geometry: Analytic Geometry (URSS, Moscow, 1994) [in Russian].

    MATH  Google Scholar 

  46. S. Lychev, K. Koifman, and D. Bout, ‘‘Finite incompatible deformations in elastic solids: Relativistic approach,’’ Lobachevskii J. Math. 43, 1908–1933 (2022).

    Article  MathSciNet  MATH  Google Scholar 

  47. M. Postnikov, Lectures in Geometry: Linear Algebra and Differential Geometry (URSS, Moscow, 1994) [in Russian].

    MATH  Google Scholar 

  48. A. Norden, Spaces of an Affine Connection (URSS, Moscow, 2018) [in Russian].

    Google Scholar 

  49. B. Cooperstein, Advanced Linear Algebra (CRC, Boca Raton, FL, 2015).

    Book  MATH  Google Scholar 

  50. M. Gurtin, E. Fried, and L. Anand, The Mechanics and Thermodynamics of Continua (Cambridge Univ. Press, Cambridge, 2010).

    Book  Google Scholar 

  51. E. Gourgoulhon, Special Relativity in General Frames (Springer, Berlin, 2013).

    Book  MATH  Google Scholar 

  52. J. Lee, Introduction to Smooth Manifolds (Springer, New York, 2012).

    Book  Google Scholar 

  53. R. Abraham, J. Marsden, and T. Ratiu, Manifolds, Tensor Analysis, and Applications (Springer Science, New York, 2012).

    MATH  Google Scholar 

  54. M. Modugno and R. Vitolo, ‘‘The geometry of Newton’s law and rigid systems,’’ Arch. Math. 043, 197–229 (2007).

    MathSciNet  MATH  Google Scholar 

  55. E. Kanso, M. Arroyo, Y. Tong, A. Yavari, J. Marsden, and M. Desbrun, ‘‘On the geometric character of stress in continuum mechanics,’’ Zeitschr. Angew. Math. Phys. 58, 843–856 (2007).

    Article  MathSciNet  MATH  Google Scholar 

  56. H. Coxeter, Regular Polytopes (Dover, New York, 2012).

    MATH  Google Scholar 

  57. O. Kellogg, Foundations of Potential Theory (Springer Nature, Switzerland, 1967).

    Book  MATH  Google Scholar 

  58. J. Marsden and T. Hughes, Mathematical Foundations of Elasticity (Courier, New York, 1994).

    MATH  Google Scholar 

  59. C. Truesdell and R. Toupin, ‘‘The classical field theories,’’ in Principles of Classical Mechanics and Field Theory. Encyclopedia of Physics, Ed. by S. Flügge (Springer, Berlin, 1960), pp. 226–858.

    Google Scholar 

  60. A. Gusev and S. Lurie, ‘‘Theory of spacetime elasticity,’’ Int. J. Mod. Phys. B 26, 1250032 (2012).

    Article  MathSciNet  MATH  Google Scholar 

  61. S. Lurie and P. Belov, ‘‘On the nature of the relaxation time, the Maxwell–Cattaneo and Fourier law in the thermodynamics of a continuous medium, and the scale effects in thermal conductivity,’’ Continuum Mech. Thermodyn. 32, 709–728 (2020).

    Article  MathSciNet  MATH  Google Scholar 

  62. C. Truesdell and W. Noll, The Non-Linear Field Theories of Mechanics (Springer Science, New York, 2013).

    MATH  Google Scholar 

  63. S. Lychev, ‘‘On conservation laws of micromorphic nondissipative thermoelasticity,’’ Vestn. Samara Univ., Ser.: Estestv. Nauki 4, 225–262 (2007).

    MathSciNet  MATH  Google Scholar 

  64. M. Epstein and M. Elzanowski, Material Inhomogeneities and their Evolution: A Geometric Approach (Springer Science, New York, 2007).

    MATH  Google Scholar 

  65. S. Lychev and K. Koifman, ‘‘Contorsion of material connection in growing solids,’’ Lobachevskii J. Math. 42, 1852–1875 (2021).

    Article  MathSciNet  MATH  Google Scholar 

  66. T. Levi-Civita, ‘‘Nozione di parallelismo in una varietà qualunque e conseguente specificazione geometrica della curvatura riemanniana,’’ Rend. Circ. Mat. Palermo 42, 173–204 (1916).

    Article  Google Scholar 

  67. O. Fernandez and A. Bloch, ‘‘The Weitzenböck connection and time reparameterization in nonholonomic mechanics,’’ J. Math. Phys. 52, 012901 (2011).

    Article  MathSciNet  MATH  Google Scholar 

  68. E. Zeeman, ‘‘The topology of Minkowski space,’’ Topology 6, 161–170 (1967).

    Article  MathSciNet  MATH  Google Scholar 

  69. C. Rovelli, Quantum Gravity (Cambridge Univ. Press, Cambridge, 2004).

    Book  MATH  Google Scholar 

  70. E. Poisson, A Relativist’s Toolkit: The Mathematics of Black-Hole Mechanics (Cambridge Univ. Press, Cambridge, 2004).

    Book  MATH  Google Scholar 

  71. A. Palatini, ‘‘Deduzione invariantiva delle equazioni gravitazionali dal principio di Hamilton,’’ Rend. Circ. Mat. Palermo 43, 203–212 (1919).

    Article  MATH  Google Scholar 

  72. R. Aldrovandi and J. Pereira, Teleparallel Gravity: An Introduction (Springer, Dordrecht, 2013).

    Book  MATH  Google Scholar 

  73. M. Gasperini, Theory of Gravitational Interactions (Springer, Cham, 2017).

    Book  MATH  Google Scholar 

  74. C. Heinicke and F. Hehl, ‘‘Schwarzschild and Kerr solutions of Einstein’s field equation: An introduction,’’ Int. J. Mod. Phys. D 24, 1530006 (2015).

    Article  MathSciNet  MATH  Google Scholar 

  75. S. Chandrasekhar, The Mathematical Theory of Black Holes (Oxford Univ. Press, New York, 1983).

    MATH  Google Scholar 

  76. M. Born, ‘‘Die Theorie des starren Elektrons in der Kinematik des Relativitätsprinzips,’’ Ann. Phys. 335 (11), 1–56 (1909).

    Article  MATH  Google Scholar 

  77. M. Born, ‘‘Zur Kinematik des starren Körpers im System des Relativitätsprinzips,’’ Göttinger Nachr. 2, 161–179 (1910).

    MATH  Google Scholar 

  78. K. Schwarzschild, ‘‘Über das Gravitationsfeld eines Massenpunktes nach der Einsteinschen Theorie,’’ Sitzungsber. Königl. Preuss. Akad. Wissensch. 7, 189–196 (1916).

    MATH  Google Scholar 

  79. K. Schwarzschild, ‘‘Über das Gravitationsfeld einer Kugel aus Inkompressibler Flüssigkeit nach der Einsteinschen Theorie,’’ Sitzungsber. König. Preuss. Akad. Wissensch., 424–434 (1916).

  80. R. Tolman, ‘‘Static solutions of Einstein’s field equations for spheres of fluid,’’ Phys. Rev. 55, 364–373 (1939).

    Article  MATH  Google Scholar 

  81. H. Reissner, ‘‘Über die Eigengravitation des elektrischen Feldes nach der Einsteinschen Theorie,’’ Ann. Phys. 50 (9), 106–120 (1916).

    Article  Google Scholar 

  82. G. Nordström, ‘‘On the energy of the gravitational field in Einstein’s theory,’’ Verhandl. Koninkl. Ned. Akad. Wetenschap. 26, 1201–1208 (1918).

    MATH  Google Scholar 

  83. L. Patel and B. Pandya, ‘‘A Reissner–Nordström interior solution,’’ Acta Phys. Hung. 60, 57–65 (1986).

    Article  Google Scholar 

  84. R. Rivlin, ‘‘Large elastic deformations of isotropic materials, I. Fundamental concepts,’’ Phil. Trans. R. Soc. London, Ser. A 240, 459–490 (1948).

    Article  MathSciNet  MATH  Google Scholar 

  85. S. Lychev, K. Koifman, and N. Djuzhev, ‘‘Incompatible deformations in additively fabricated solids: Discrete and continuous approaches,’’ Symmetry 13, 2331 (2021).

    Article  Google Scholar 

  86. S. Lychev, G. Kostin, T. Lycheva, and K. Koifman, ‘‘Non-Euclidean geometry and defected structure for bodies with variable material composition,’’ J. Phys.: Conf. Ser. 1250, 012035 (2019).

    Google Scholar 

  87. M. Gurtin, An Introduction to Continuum Mechanics (Academic, New York, 1981).

    MATH  Google Scholar 

  88. R. Schield, ‘‘Inverse deformation results in finite elasticity,’’ J. Appl. Math. Phys. 18, 490–500 (1967).

    Google Scholar 

  89. A. Bressan, ‘‘Una teoria di relatività generale includente, oltre all’elettromagnetismo e alla termodinamica, le equazioni costitutive dei materiali ereditari. Sistemazione assiomatica,’’ Rend. Semin. Mat. Univ. Padova 34, 74–109 (1964).

    MATH  Google Scholar 

  90. L. Bragg, ‘‘On relativistic worldlines and motions, and on non-sentient response,’’ Arch. Ration. Mech. Anal. 18, 127–166 (1965).

    Article  MATH  Google Scholar 

  91. L. Söderholm, ‘‘A principle of objectivity for relativistic continuum mechanics,’’ Arch. Ration. Mech. Anal. 39, 89–107 (1970).

    Article  MathSciNet  MATH  Google Scholar 

  92. D. Delphenich, ‘‘Proper time foliations of Lorentz manifolds,’’ arXiv: gr-qc/0211066 (2002).

  93. C. Gundlach, I. Hawke, and S. Erickson, ‘‘A conservation law formulation of nonlinear elasticity in general relativity,’’ Class. Quantum Grav. 29, 015005 (2011).

    Article  MathSciNet  MATH  Google Scholar 

  94. R. Beig and B. Schmidt, ‘‘Relativistic elasticity,’’ Class. Quantum Grav. 20, 889 (2003).

    Article  MathSciNet  MATH  Google Scholar 

  95. G. Maugin and A. Eringen, ‘‘Relativistic continua with directors,’’ J. Math. Phys. 13, 1788–1797 (1972).

    Article  Google Scholar 

  96. M. Hirsch, Differential Topology (Springer, New York, 1976).

    Book  MATH  Google Scholar 

  97. S. Chern, W. Chen, and K. Lam, Lectures on Differential Geometry (World Scientific, Singapore, 1999).

    Book  MATH  Google Scholar 

  98. S. Sternberg, Lectures on Differential Geometry (Prentice-Hall, Englewood Cliffs, NJ, 1964).

    MATH  Google Scholar 

  99. J. Lee, Introduction to Riemannian Manifolds (Springer, Cham, 2018).

    Book  MATH  Google Scholar 

  100. D. Lovelock and H. Rund, Tensors, Differential Forms, and Variational Principles (Dover, New York, 1989).

    MATH  Google Scholar 

Download references

Funding

The study was supported by the grant of Russian Science Foundation no. 22-21-00457. https://rscf.ru/en/project/22-21-00457/.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to S. A. Lychev, K. G. Koifman or N. A. Pivovaroff.

Additional information

(Submitted by A. M. Elizarov)

Appendix

Appendix

1.1 REVIEW OF GENERAL AFFINE CONNECTIONS

1.2 1. Covariant Derivative and its Coordinate Representation

There are two approaches to the construction of geometric structures on manifolds. The first approach uses the ‘‘outer’’ geometry methodology, when a manifold is embedded in a Euclidean space of possibly large dimensionFootnote 17 and the Euclidean parallel transport rule is induced on this manifold, similar to how it is done in classical theory of surfaces. At the same time, despite the seeming simplicity, the implementation of this approach has at least two problems: 1) due to the high dimension, the formulas are cumbersome, and 2) the physical interpretation of the surrounding space is unknown.Footnote 18 In this regard, as a rule, the second approach is used, based on the ‘‘internal’’ geometry methodology. Namely, the parallel transfer procedure is defined axiomatically, using only the structure of the considered manifold.Footnote 19

Let \(M\) be a smooth manifold of dimension \(n\). By \(\textrm{Vec}(M)\) we denote the \(C^{\infty}(M)\)-module of smooth vector fields on \(M\).

Definition 1. Affine connection on manifold \(M\) is a mapping \(\nabla:\>\textrm{Vec}(M)\times\textrm{Vec}(M)\rightarrow\textrm{Vec}(M)\), \((\mathsf{u},\>\mathsf{v})\mapsto\nabla_{\mathsf{u}}\mathsf{v}\), that satisfies the following axioms [15, 97]:

  1. 1.

    \(\forall\mathsf{u},\>\mathsf{v},\>\mathsf{w}\in\textrm{Vec}(M):\;\nabla_{\mathsf{u}+\mathsf{v}}\mathsf{w}=\nabla_{\mathsf{u}}\mathsf{w}+\nabla_{\mathsf{v}}\mathsf{w}\);

  2. 2.

    \(\forall\mathsf{u},\>\mathsf{v}\in\textrm{Vec}(M)\>\forall f\in C^{\infty}(M):\;\nabla_{f\mathsf{u}}\mathsf{v}=f\nabla_{\mathsf{u}}\mathsf{v}\);

  3. 3.

    \(\forall\mathsf{u},\>\mathsf{v},\>\mathsf{w}\in\textrm{Vec}(M):\;\nabla_{u}(\mathsf{v}+\mathsf{w})=\nabla_{\mathsf{u}}\mathsf{v}+\nabla_{\mathsf{u}}\mathsf{w}\);

  4. 4.

    \(\forall\mathsf{u},\>\mathsf{v}\in\textrm{Vec}(M)\>\forall f\in C^{\infty}(M):\;\nabla_{\mathsf{u}}(f\mathsf{v})=f\nabla_{\mathsf{u}}\mathsf{v}+(\mathsf{u}f)\mathsf{v}\).

In Definition 1 the symbol \(\mathsf{u}f\) denotes the action of vector field \(\mathsf{u}\), interpreted as derivation, onto scalar field \(f\). In coordinate neighborhood \(U\) one thus has relation \(\mathsf{u}f=u^{i}\,\frac{\partial f}{\partial x^{i}}\). The vector field \(\nabla_{\mathsf{u}}\mathsf{v}\) is usually referred to as covariant derivative of \(\mathsf{v}\) along vector field \(\mathsf{u}\).

In arbitrary chart \((U,\,\varphi)=(U,\,x^{1},\,\ldots,\,x^{n})\) on manifold \(M\) the restriction of connection \(\nabla\) onto \(U\) can be expressed in coordinate form. Indeed, let \((\partial_{i})_{i=1}^{n}\) be coordinate frame on \(U\), and let \((dx^{i})_{i=1}^{n}\) be the corresponding coordinate coframe, i.e., \(\left\langle dx^{i},\>\partial_{j}\right\rangle=\delta^{i}_{j}\), \(i,\,j=1,\,\ldots,\,n\). Then, taking \(\Gamma^{i}_{jk}:=\left\langle dx^{i},\>\nabla_{\partial_{j}}\partial_{k}\right\rangle\), one obtains relation \(\nabla_{\partial_{j}}\partial_{k}=\Gamma^{i}_{jk}\partial_{i}\) on \(U\). Scalar functions \(\Gamma^{i}_{jk}\in C^{\infty}(U)\), \(i,\>j,\>k=1,\>\ldots,\>n\), would be referred to as connection coefficients in chart \((U,\,\varphi)\). Furthermore, utilizing the axioms of affine connection, one obtains that for arbitrary vector fields \(\mathsf{u},\,\mathsf{v}\in\textrm{Vec}(U)\), the following formula holds

$$\nabla_{\mathsf{u}}\mathsf{v}=u^{i}\left\{\partial_{i}v^{k}+v^{j}\Gamma^{k}_{ij}\right\}\partial_{k}.$$

Suppose that \((V,\,\psi)=(V,\,\widetilde{x}^{1},\,\ldots,\,\widetilde{x}^{n})\) is another chart on \(M\), that overlaps with the chart \((U,\,\varphi)\), i.e., \(U\cap V\neq\varnothing\). Denoting connection coefficients that correspond to the chart \((V,\>\psi)\) by \(\widetilde{\Gamma}^{i}_{jk}\), one can show directly that on \(U\cap V\) the following transformation law holds

$$\widetilde{\Gamma}^{i}_{jk}=\dfrac{\partial\widetilde{x}^{i}}{\partial x^{p}}\dfrac{\partial x^{l}}{\partial\widetilde{x}^{j}}\dfrac{\partial x^{m}}{\partial\widetilde{x}^{k}}\Gamma^{p}_{lm}+\dfrac{\partial\widetilde{x}^{i}}{\partial x^{p}}\dfrac{\partial^{2}x^{p}}{\partial\widetilde{x}^{j}\partial\widetilde{x}^{k}}.$$
(A1)

Formula (A1) shows, in particular, that quantities \(\Gamma^{i}_{jk}\) do not form tensor field on \(M\).

1.3 2. Parallel Transport

Let \(\gamma:\>\mathbb{I}\rightarrow M\) be some smooth curve on \(M\).

Definition 2. A vector field \(\mathsf{u}\in\textrm{Vec}(M)\) is called parallel along \(\gamma\), if for every value \(t\in\mathbb{I}\) the following relation holds

$$\nabla_{\dot{\gamma}(t)}\mathsf{u}=\mathsf{0},$$
(A2)

where \(\dot{\gamma}(t)\) is velocity vector of \(\gamma\).

Choosing chart \((U,\,x^{1},\,\ldots,\,x^{n})\) on \(M\), and supposing that the image of the curve \(\gamma\) completely lies in \(U\), one can express equation (A2) in components:

$$\dot{u}^{k}+\dot{x}^{i}u^{j}\Gamma^{k}_{ij}=0,\quad k=1,\>\ldots,\>n.$$

Here \(\mathsf{u}=u^{i}\partial_{i}\), \(\dot{\gamma}=\dot{x}^{i}\partial_{i}\), and the dot above denotes the operation of taking derivative \(d/dt\).

If \(p=\gamma(t_{0})\), then one can formulate initial condition \(\mathsf{u}_{p}=\mathsf{u}_{0}\). Thus, in coordinate neighborhood \(U\) there is a unique solution of the following Cauchy problem

$$\dot{u}^{k}+\dot{x}^{i}u^{j}\Gamma^{k}_{ij}=0,\quad k=1,\,\ldots,\,n,$$
$$u_{p}^{k}=u_{0}^{k},\quad k=1,\,\ldots,\,n.$$
(A3)

By this reason, one can define a mapping \(P_{t_{0}}^{t}:\>T_{\gamma(t_{0})}M\rightarrow T_{\gamma(t)}M\), that with each vector \(\mathsf{u}_{0}\in T_{\gamma(t_{0})}M\) associates value \(\mathsf{u}(t)\) of the solution \(\mathsf{u}\) of the Cauchy problem (A3) \(P_{t_{0}}^{t}:\>\mathsf{u}_{0}\mapsto\mathsf{u}(t)\). This mapping is linear and characterizes parallel transport [98].

If \(\mathsf{u},\>\mathsf{v}\in\textrm{Vec}(M)\) are two vector fields and if \(\gamma:\>\mathbb{I}\rightarrow M\) is integral curve of \(\mathsf{u}\), passing through point \(p=\gamma(t_{0})\), then the value of covariant derivative \(\nabla_{\mathsf{u}}\mathsf{v}|_{p}\) at \(p\) can be calculated via the following limit

$$\nabla_{\mathsf{u}}\mathsf{v}|_{p}=\lim\limits_{h\to 0}\dfrac{P_{t_{0}+h}^{t_{0}}\mathsf{v}_{\gamma(t_{0}+h)}-\mathsf{v}_{\gamma(t_{0})}}{h}=\left.\dfrac{d}{dt}P_{t}^{t_{0}}\mathsf{v}_{\gamma(t)}\right|_{t=t_{0}}.$$

Note that vector \(P_{t_{0}+h}^{t_{0}}\mathsf{v}_{\gamma(t_{0}+h)}\) belongs to \(T_{\gamma(t_{0})}M\), so, for sufficiently small \(h\) the difference \(P_{t_{0}+h}^{t_{0}}\mathsf{v}_{\gamma(t_{0}+h)}-\mathsf{v}_{\gamma(t_{0})}\) is well-defined and belongs to \(T_{\gamma(t_{0})}M\).

1.4 3. Torsion, Curvature and Nonmetricity

If manifold \(M\) is endowed with some affine connection \(\nabla\), then we will refer to the structure \((M,\,\nabla)\) as space of affine connectionFootnote 20 [48]. In such the space one can define the notion of parallel transport, so, in particular, there is an opportunity to differentiate tensor fields, that is necessary for formulating of balance equations. If, moreover, the manifold \(M\) is endowed with metric tensor \(\mathsf{g}\), then one can consider a structure \((M,\,\mathsf{g},\,\nabla)\). We will refer to it as affine-metric manifold.

The parallel transport induced by the connection defines some geometry on the manifold, which, in the general case, is different from the Euclidean one. The connection coefficients are not suitable for describing this geometry: already in the Euclidean space, the connection coefficients are equal to zero in Cartesian coordinates, and are non-zero in curvilinear coordinates.Footnote 21 In this regard, additional tensor fields characterizing the degree of non-Euclideness are introduced.

For affine-metric manifold \((M,\,\mathsf{g},\,\nabla)\) the connection \(\nabla\) defines tensor fields of torsion \(\mathfrak{T}:\>\textrm{Vec}(M)\times\textrm{Vec}(M)\rightarrow\textrm{Vec}(M)\), curvature \(\mathfrak{R}:\>\textrm{Vec}(M)\times\textrm{Vec}(M)\times\textrm{Vec}(M)\rightarrow\textrm{Vec}(M)\), and nonmetricity \(\mathfrak{Q}:\>\textrm{Vec}(M)\times\textrm{Vec}(M)\times\textrm{Vec}(M)\rightarrow C^{\infty}(M)\) upon formulae [15, 99]

$$\mathfrak{T}(\mathsf{u},\>\mathsf{v}):=\nabla_{\mathsf{u}}\mathsf{v}-\nabla_{\mathsf{v}}\mathsf{u}-[\mathsf{u},\>\mathsf{v}];$$
(A4)
$$\mathfrak{R}(\mathsf{u},\>\mathsf{v},\>\mathsf{w}):=[\nabla_{\mathsf{u}},\>\nabla_{\mathsf{v}}]\mathsf{w}-\nabla_{[\mathsf{u},\>\mathsf{v}]}\mathsf{w};$$
(A5)
$$\mathfrak{Q}(\mathsf{u},\>\mathsf{v},\>\mathsf{w}):=\mathsf{g}(\nabla_{\mathsf{u}}\mathsf{v},\>\mathsf{w})+\mathsf{g}(\mathsf{v},\>\nabla_{\mathsf{u}}\mathsf{w})-\mathsf{u}[\mathsf{g}(\mathsf{v},\>\mathsf{w})].$$
(A6)

HereFootnote 22\([\mathsf{u},\>\mathsf{v}]\) and \([\nabla_{\mathsf{u}},\>\nabla_{\mathsf{v}}]\) are commutators: \([\mathsf{u},\,\mathsf{v}]=\mathsf{u}\circ\mathsf{v}-\mathsf{v}\circ\mathsf{u}\), \([\nabla_{\mathsf{u}},\,\nabla_{\mathsf{v}}]=\nabla_{\mathsf{u}}\circ\nabla_{\mathsf{v}}-\nabla_{\mathsf{v}}\circ\nabla_{\mathsf{u}}\), while \(\mathsf{u}[\mathsf{g}(\mathsf{v},\,\mathsf{w})]\) is action of vector field \(\mathsf{u}\) onto scalar field \(\mathsf{g}(\mathsf{v},\,\mathsf{w})\).

In chart \((U,\,\varphi)\) the tensor fields \(\mathfrak{T}\), \(\mathfrak{R}\) and \(\mathfrak{Q}\) (rather, their restrictions on \(U\)) have the following components

$$\mathfrak{T}^{k}_{ij}=\Gamma^{k}_{ij}-\Gamma^{k}_{ji};$$
$$\mathfrak{R}^{t}_{ijk}=\partial_{i}\Gamma^{t}_{jk}-\partial_{j}\Gamma^{t}_{ik}+\Gamma^{l}_{jk}\Gamma^{t}_{il}-\Gamma^{l}_{ik}\Gamma^{t}_{jl};$$
$$-\mathfrak{Q}_{ijk}=\partial_{i}g_{jk}-\Gamma^{m}_{ij}g_{mk}-\Gamma^{m}_{ik}g_{mj},$$

where \(i,\,j,\,k,\,t=1,\,\ldots,\,n\).

1.5 4. Connection in Non-holonomic Frame

Suppose now, that in the coordinate neighborhood \(U\) of a chart \((U,\,\varphi)=(U,\,x^{1},\,\ldots,\,x^{n})\) there is a field of non-degenerate \(n\times n\)-matrices \(\Omega=[\Omega^{i}_{j}]:\>U\rightarrow\textrm{GL}(n;\>\mathbb{R})\). With respect to coordinate frame \((\partial_{i})_{i=1}^{n}\) this field defines a new frame \((\mathsf{z}_{i})_{i=1}^{n}\) on the same set \(U\) as \(\mathsf{z}_{i}=\Omega^{j}_{i}\partial_{j}\). In contrast to the coordinate frame, the latter frame may be non-holonomic, i.e., \([\mathsf{z}_{i},\>\mathsf{z}_{j}]\neq\mathsf{0}\). From physical viewpoint the frame \((\mathsf{z}_{i})_{i=1}^{n}\) may represent the collection of distorted crystal axes, while field \(\Omega\) is the distortion field.

The coframe \((\vartheta^{i})_{i=1}^{n}\), dual to \((\mathsf{z}_{i})_{i=1}^{n}\), is determined upon the system of \(n^{2}\) equalities: \(\left\langle\vartheta^{i},\>\mathsf{z}_{j}\right\rangle=\delta^{i}_{j}\), \(i,\,j=1,\,\ldots,\,n\). Explicitly, \(\vartheta^{i}=\mho^{i}_{j}dx^{j}\). Here \((dx^{i})_{i=1}^{n}\) is the coordinate coframe, and \(\mho=\Omega^{-1}\) is the field of inverse matrices, i.e., \(\mho^{i}_{k}\,\Omega^{k}_{j}=\Omega^{i}_{k}\,\mho^{k}_{j}=\delta^{i}_{j}\) on \(U\).

Since for all \(i,\,j=1,\,\ldots,\,n\), the commutator \([\mathsf{z}_{i},\,\mathsf{z}_{j}]\) is a vector field on \(U\), one can decompose it with respect to the frame \((\mathsf{z}_{i})_{i=1}^{n}\): \([\mathsf{z}_{i},\,\mathsf{z}_{j}]=-c^{k}_{ij}\mathsf{z}_{k}\). The coefficients \(c^{k}_{ij}\in C^{\infty}(U)\) of this expansion are referred to as the objects of anholonomity and can be expressed explicitly in terms of matrix fields \(\Omega\) and \(\mho\):

Theorem 1. The fields \(c^{k}_{ij}\) can be determined upon equality \(c^{k}_{ij}=\mho^{k}_{m}\left(\Omega^{s}_{j}\,\partial_{s}\Omega^{m}_{i}-\Omega^{s}_{i}\,\partial_{s}\Omega^{m}_{j}\right),\) or, equivalently, from \(c^{k}_{ij}=\Omega^{m}_{i}\Omega^{q}_{j}\left(\partial_{m}\mho^{k}_{q}-\partial_{q}\mho^{k}_{m}\right).\)

The proof of this claim can be found in ([12], p. 222).

The objects of anholonomity are components of exterior differential \(d\vartheta^{k}\) with respect to coframe \((\vartheta^{i})_{i=1}^{n}\).

Theorem 2. The following relation holds

$$d\vartheta^{k}=\displaystyle\sum\limits_{p<q}c^{k}_{pq}\vartheta^{p}\wedge\vartheta^{q}.$$
(A7)

The proof can be found in ([12], p. 223).

Now, let \(\gamma^{i}_{jk}=\left\langle\vartheta^{i},\>\nabla_{\mathsf{z}_{j}}\mathsf{z}_{k}\right\rangle\) and \(\Gamma^{i}_{jk}=\left\langle dx^{i},\>\nabla_{\partial_{j}}\partial_{k}\right\rangle\) be coefficients of connection \(\nabla\) with respect to frames \((\mathsf{z}_{i})_{i=1}^{n}\) and \((\partial_{i})_{i=1}^{n}\):

$$\nabla_{\mathsf{z}_{j}}\mathsf{z}_{k}=\gamma^{i}_{jk}\mathsf{z}_{i}\quad\text{and}\quad\nabla_{\partial_{j}}\partial_{k}=\Gamma^{i}_{jk}\partial_{i}.$$
(A8)

The next claim expresses relation between \(\gamma^{i}_{jk}\) and \(\Gamma^{i}_{jk}\).

Theorem 3. Fields \(\gamma^{i}_{jk}\) and \(\Gamma^{i}_{jk}\) on \(U\), defined by relations (A8), are related as

$$\gamma^{i}_{jk}=\Gamma^{m}_{sq}\mho^{i}_{m}\Omega^{s}_{j}\Omega^{q}_{k}+\mho^{i}_{m}\Omega^{s}_{j}\partial_{x^{s}}\Omega^{m}_{k}.$$
(A9)

For the proof see ([12], p. 213, Section 9.2.1).

From equality (A9), as particular case (when \((\mathsf{z}_{i})_{i=1}^{n}\) is coordinate frame, generated by another chart), the equality (A1) follows.

Define components of torsion, curvature and nonmetricity in non-holonomic frame. Let

$$\mathfrak{T}=\mathfrak{T}^{k}_{ij}\,\mathsf{z}_{k}\otimes\vartheta^{i}\otimes\vartheta^{j},\quad\mathfrak{R}=\mathfrak{R}^{m}_{ijk}\,\mathsf{z}_{m}\otimes\vartheta^{i}\otimes\vartheta^{j}\otimes\vartheta^{k},\quad\mathfrak{Q}=\mathfrak{Q}_{ijk}\,\vartheta^{i}\otimes\vartheta^{j}\otimes\vartheta^{k}$$

be polyadic decompositions of fields \(\mathfrak{T}\), \(\mathfrak{R}\) and \(\mathfrak{Q}\) with respect to the frame \((\mathsf{z}_{i})_{i=1}^{n}\) and coframe \((\vartheta^{i})_{i=1}^{n}\). Expressions of components of these expansions are stated in the following claim.

Theorem 4. With respect to the frame \((\mathsf{z}_{i})_{i=1}^{n}\) , related with the coordinate frame \((\partial_{i})_{i=1}^{n}\) by equalities \(\mathsf{z}_{i}=\Omega^{j}_{i}\partial_{j}\) , \(i=1,\,\ldots,\,n\) , one has the following formulas

$$\mathfrak{T}^{k}_{ij}=\gamma^{k}_{ij}-\gamma^{k}_{ji}+c^{k}_{ij},$$
(A10)
$$\mathfrak{R}^{m}_{ijk}=\Omega^{s}_{i}\,\partial_{s}\gamma^{m}_{jk}-\Omega^{s}_{j}\,\partial_{s}\gamma^{m}_{ik}+\gamma^{s}_{jk}\gamma^{m}_{is}-\gamma^{s}_{ik}\gamma^{m}_{js}+c^{s}_{ij}\gamma^{m}_{sk},$$
(A11)
$$-\mathfrak{Q}_{ijk}=\Omega^{s}_{i}\,\partial_{s}g_{jk}-g_{mk}\gamma^{m}_{ij}-g_{jm}\gamma^{m}_{ik}.$$
(A12)

Here \(\gamma^{i}_{jk}\) are coefficients of connection \(\nabla\) in frame \((\mathsf{z}_{i})_{i=1}^{n}\), while \(g_{ij}\) are metric tensor components; \(g_{ij}=\mathsf{g}(\mathsf{z}_{i},\,\mathsf{z}_{j})\).

Proof. Due to formula (A4),

$$\mathfrak{T}(\mathsf{z}_{i},\>\mathsf{z}_{j})=\nabla_{\mathsf{z}_{i}}\mathsf{z}_{j}-\nabla_{\mathsf{z}_{j}}\mathsf{z}_{i}-[\mathsf{z}_{i},\>\mathsf{z}_{j}]=\gamma^{k}_{ij}\mathsf{z}_{k}-\gamma^{k}_{ji}\mathsf{z}_{k}+c^{k}_{ij}\mathsf{z}_{k},$$

that with respect to equality \(\mathfrak{T}^{k}_{ij}=\left\langle\vartheta^{k},\>\mathfrak{T}(\mathsf{z}_{i},\>\mathsf{z}_{j})\right\rangle\) leads to (A10).

Using definition of curvature (formula (A5)), one gets

$$\mathfrak{R}(\mathsf{z}_{i},\>\mathsf{z}_{j},\>\mathsf{z}_{k})=\nabla_{\mathsf{z}_{i}}\nabla_{\mathsf{z}_{j}}\mathsf{z}_{k}-\nabla_{\mathsf{z}_{j}}\nabla_{\mathsf{z}_{i}}\mathsf{z}_{k}-\nabla_{[\mathsf{z}_{i},\>\mathsf{z}_{j}]}\mathsf{z}_{k}$$
$${}=\nabla_{\mathsf{z}_{i}}\left(\gamma^{m}_{jk}\mathsf{z}_{m}\right)-\nabla_{\mathsf{z}_{j}}\left(\gamma^{m}_{ik}\mathsf{z}_{m}\right)-\nabla_{(-c^{s}_{ij}\mathsf{z}_{s})}\mathsf{z}_{k}$$
$${}=\gamma^{m}_{jk}\nabla_{\mathsf{z}_{i}}\mathsf{z}_{m}+\mathsf{z}_{i}[\gamma^{m}_{jk}]\mathsf{z}_{m}-\gamma^{m}_{ik}\nabla_{\mathsf{z}_{j}}\mathsf{z}_{m}-\mathsf{z}_{j}[\gamma^{m}_{ik}]\mathsf{z}_{m}+c^{s}_{ij}\nabla_{\mathsf{z}_{s}}\mathsf{z}_{k}$$
$${}=\gamma^{m}_{jk}\gamma^{s}_{im}\mathsf{z}_{s}+\Omega^{s}_{i}\partial_{s}\gamma^{m}_{jk}\mathsf{z}_{m}-\gamma^{m}_{ik}\gamma^{s}_{jm}\mathsf{z}_{s}-\Omega^{s}_{j}\partial_{s}\gamma^{m}_{ik}\mathsf{z}_{m}+c^{s}_{ij}\gamma^{m}_{sk}\mathsf{z}_{m}.$$

From the latter equality, according to relation \(\mathfrak{R}^{m}_{ijk}=\left\langle\vartheta^{m},\>\mathfrak{R}(\mathsf{z}_{i},\>\mathsf{z}_{j},\>\mathsf{z}_{k})\right\rangle\), the formula (A11) follows. Finally, for nonmetricity (formula (A6)) one gets

$$\mathfrak{Q}_{ijk}=\mathfrak{Q}(\mathsf{z}_{i},\>\mathsf{z}_{j},\>\mathsf{z}_{k})=\mathsf{g}(\nabla_{\mathsf{z}_{i}}\mathsf{z}_{j},\>\mathsf{z}_{k})+\mathsf{g}(\mathsf{z}_{j},\>\nabla_{\mathsf{z}_{i}}\mathsf{z}_{k})-\mathsf{z}_{i}[\mathsf{g}(\mathsf{z}_{j},\>\mathsf{z}_{k})]$$
$${}=\mathsf{g}(\gamma^{m}_{ij}\mathsf{z}_{m},\>\mathsf{z}_{k})+\mathsf{g}(\mathsf{z}_{j},\>\gamma^{m}_{ik}\mathsf{z}_{m})-\Omega^{s}_{i}\partial_{s}g_{jk}=\gamma^{m}_{ij}g_{mk}+\gamma^{m}_{ik}g_{jm}-\Omega^{s}_{i}\partial_{s}g_{jk},$$

and then the equality (A12) directly follows. \(\Box\)

Using formulae (A10) and (A12), one can obtain direct expression for connection coefficients \(\gamma^{i}_{jk}\) in terms of metric, torsion, nonmetricity and objects of anholonomity.

Theorem 5. The following relation holds

$$\gamma^{i}_{jk}=\frac{g^{is}}{2}\left[\mathsf{z}_{j}(g_{sk})+\mathsf{z}_{k}(g_{js})-\mathsf{z}_{s}(g_{jk})\right]+\frac{g^{is}}{2}\left(g_{jm}c^{m}_{ks}-g_{ms}c^{m}_{jk}-g_{mk}c^{m}_{sj}\right)$$
$${}+\frac{g^{is}}{2}\left(g_{ms}\mathfrak{T}^{m}_{jk}+g_{mk}\mathfrak{T}^{m}_{sj}-g_{mj}\mathfrak{T}^{m}_{ks}\right)+\frac{g^{is}}{2}\left(\mathfrak{Q}_{jks}+\mathfrak{Q}_{ksj}-\mathfrak{Q}_{sjk}\right).$$
(A13)

Proof. Under cyclic permutation \((i,\,j,\,k)\mapsto(k,\,i,\,j)\mapsto(j,\,k,\,i)\) one gets three counterparts of formula (A12):

$$\mathsf{z}_{i}(g_{jk})-g_{mk}\gamma^{m}_{ij}-g_{jm}\gamma^{m}_{ik}=-\mathfrak{Q}_{ijk},$$
$$\mathsf{z}_{k}(g_{ij})-g_{mj}\gamma^{m}_{ki}-g_{im}\gamma^{m}_{kj}=-\mathfrak{Q}_{kij},\quad\mathsf{z}_{j}(g_{ki})-g_{mi}\gamma^{m}_{jk}-g_{km}\gamma^{m}_{ji}=-\mathfrak{Q}_{jki}$$

(here we take into account that \(\mathsf{z}_{i}(g_{jk})=\Omega^{s}_{i}\,\partial_{s}g_{jk}\)). Adding two latter formulae with each other and subtracting the first formula from the result, one gets

$$\mathsf{z}_{j}(g_{ki})+\mathsf{z}_{k}(g_{ij})-\mathsf{z}_{i}(g_{jk})-g_{mi}\gamma^{m}_{jk}-g_{im}\gamma^{m}_{kj}+g_{mk}(\gamma^{m}_{ij}-\gamma^{m}_{ji})-g_{jm}(\gamma^{m}_{ki}-\gamma^{m}_{ik})$$
$${}=\mathfrak{Q}_{ijk}-\mathfrak{Q}_{kij}-\mathfrak{Q}_{jki}.$$

Expressing the fourth term in the left hand side as \(g_{mi}\gamma^{m}_{jk}=2g_{mi}\gamma^{m}_{jk}-g_{mi}\gamma^{m}_{jk}\), and then replacing the differences \(\gamma^{m}_{jk}-\gamma^{m}_{kj}\), \(\gamma^{m}_{ij}-\gamma^{m}_{ji}\), \(\gamma^{m}_{ki}-\gamma^{m}_{ik}\), according to formula (A10), one gets

$$\mathsf{z}_{j}(g_{ki})+\mathsf{z}_{k}(g_{ij})-\mathsf{z}_{i}(g_{jk})-2g_{mi}\gamma^{m}_{jk}+g_{mi}(\mathfrak{T}^{m}_{jk}-c^{m}_{jk})+g_{mk}(\mathfrak{T}^{m}_{ij}-c^{m}_{ij})$$
$${}-g_{jm}(\mathfrak{T}^{m}_{ki}-c^{m}_{ki})=\mathfrak{Q}_{ijk}-\mathfrak{Q}_{kij}-\mathfrak{Q}_{jki}.$$

Replacing all occurrences of index \(i\) by \(s\) and performing simple algebraic transformations, one arrives at the formula

$$g_{ms}\gamma^{m}_{jk}=\frac{1}{2}\left[\mathsf{z}_{j}(g_{sk})+\mathsf{z}_{k}(g_{js})-\mathsf{z}_{s}(g_{jk})\right]+\frac{1}{2}\left(g_{jm}c^{m}_{ks}-g_{ms}c^{m}_{jk}-g_{mk}c^{m}_{sj}\right)$$
$${}+\frac{1}{2}\left(g_{ms}\mathfrak{T}^{m}_{jk}+g_{mk}\mathfrak{T}^{m}_{sj}-g_{mj}\mathfrak{T}^{m}_{ks}\right)+\frac{1}{2}\left(\mathfrak{Q}_{jks}+\mathfrak{Q}_{ksj}-\mathfrak{Q}_{sjk}\right).$$

Finally, multiplying both sides of the latter relation by \(g^{is}\) and performing summation over \(s\), one obtains (A13). \(\Box\)

In particular case, when the frame \((\mathsf{z}_{i})_{i=1}^{n}\) is \(\mathsf{g}\)-orthonormal, one gets that \(g_{ij}=\delta_{ij}\), \(g^{ij}=\delta^{ij}\), and the equality (A13) takes the form

$$\gamma^{i}_{jk}=\frac{1}{2}\left(c^{j}_{ki}-c^{i}_{jk}-c^{k}_{ij}\right)+\frac{1}{2}\left(\mathfrak{T}^{i}_{jk}+\mathfrak{T}^{k}_{ij}-\mathfrak{T}^{j}_{ki}\right)+\frac{1}{2}\left(\mathfrak{Q}_{jki}+\mathfrak{Q}_{kij}-\mathfrak{Q}_{ijk}\right).$$

1.6 5. Cartan Structural Equations and Bianchi Identities

It is convenient to represent affine connection in terms of family of \(1\)-forms, since within such representation one can obtain relations, that define the pair \((\nabla,\,(\vartheta_{i})_{i=1}^{n})\) as solution of system of equations, which right hand sides are torsion, curvature and nonmetricity, expressed in terms of \(2\)-forms (for torsion and curvature) and in terms of \(1\)-forms (for curvature). These relations are referred to as Cartan equations, since Cartan was first who derived them.

Suppose that on some coordinate neighborhood \(U\subset M\) the frame \((\mathsf{z}_{i})_{i=1}^{n}\) is defined. This field is related with the original coordinate frame \((\partial_{i})_{i=1}^{n}\) as \(\mathsf{z}_{i}=\Omega^{j}_{i}\partial_{j}\), where \(\Omega=[\Omega^{i}_{j}]\) is smooth field of invertible matrices on \(U\).

Define \(1\)-forms \(\omega^{i}_{k}:=\gamma^{i}_{jk}\vartheta^{j}\), where \(\gamma^{i}_{jk}\) are coefficients of connection \(\nabla\) with respect to the frame \((\mathsf{z}_{i})_{i=1}^{n}\), while \((\vartheta^{i})_{i=1}^{n}\) is coframe, dual to \((\mathsf{z}_{i})_{i=1}^{n}\). We will refer to each \(1\)-form \(\omega^{i}_{k}\) as connection form.

Introduce families of forms \(Q_{jk}\), \(T^{i}\) and \(R^{i}_{j}\) as

$$Q_{jk}=\mathfrak{Q}_{ijk}\vartheta^{i},\quad T^{i}=\dfrac{1}{2}\mathfrak{T}^{i}_{ms}\vartheta^{m}\wedge\vartheta^{s},\quad R^{i}_{j}=\dfrac{1}{2}\mathfrak{R}^{i}_{klj}\vartheta^{k}\wedge\vartheta^{l}.$$

Here \(\mathfrak{Q}_{ijk}\), \(\mathfrak{T}^{i}_{ms}\) and \(\mathfrak{R}^{i}_{klj}\) are components of the corresponding fields with respect to \((\mathsf{z}_{i})\). The field \(Q_{jk}\) is referred to as nonmetricity form, the quantity \(T^{i}\) is torsion form, while \(R^{i}_{j}\) is curvature form.

The following claim is central for this exposition.

Theorem 6 (Cartan Structural Equations). The following relations take place:

$$g_{mk}\omega^{m}_{j}+g_{mj}\omega^{m}_{k}-dg_{jk}=Q_{jk}$$
(A14)

(\(0th\) Cartan structural equation; here \(g_{jk}=\mathsf{g}(\mathsf{z}_{j},\,\mathsf{z}_{k})\)),

$$d\vartheta^{i}+\omega^{i}_{j}\wedge\vartheta^{j}=T^{i}$$
(A15)

(\(1\) st Cartan structural equation) and

$$d\omega^{i}_{j}+\omega^{i}_{k}\wedge\omega^{k}_{j}=R^{i}_{j}$$
(A16)

( \(2\) nd Cartan structural equation).

Proof. Cartan structural equations are nothing but the reformulations of equalities (A10)–(A12) in terms of differential forms.

Indeed, multiply both sides of equality (A12) on \(\vartheta^{i}\) from the right and sum over \(i\):

$$-\Omega^{s}_{i}\partial_{s}g_{jk}\vartheta^{i}+g_{mk}\gamma^{m}_{ij}\vartheta^{i}+g_{jm}\gamma^{m}_{ik}\vartheta^{i}=\mathfrak{Q}_{ijk}\vartheta^{i}.$$

The right hand side of the obtained relation is just \(Q_{jk}\). As for the left hand side, its second and third terms are equal, respectively, to \(g_{mk}\omega^{m}_{j}\) and \(g_{jm}\omega^{m}_{k}\). Finally, the first term is \(dg_{jk}\):

$$dg_{jk}=\partial_{s}g_{jk}dx^{s}=\Omega^{s}_{i}\partial_{s}g_{jk}\vartheta^{i}.$$

All this implies (A14).

Multiplying both sides of the equality (A10) on \(2\)-covector \(\dfrac{1}{2}\vartheta^{i}\wedge\vartheta^{j}\) from the right and summing over all values \(i,\,j\), one has

$$\dfrac{1}{2}\gamma^{k}_{ij}\vartheta^{i}\wedge\vartheta^{j}-\dfrac{1}{2}\gamma^{k}_{ji}\vartheta^{i}\wedge\vartheta^{j}+\dfrac{1}{2}c^{k}_{ij}\vartheta^{i}\wedge\vartheta^{j}=\dfrac{1}{2}\mathfrak{T}^{k}_{ij}\vartheta^{i}\wedge\vartheta^{j}.$$

The right hand side of the obtained equality is torsion form \(T^{k}\). The third term of the left hand side, according to (A7), is equal to differential \(d\vartheta^{k}\). As for the first two terms on the left hand side, they give

$$\dfrac{1}{2}\gamma^{k}_{ij}\vartheta^{i}\wedge\vartheta^{j}-\dfrac{1}{2}\gamma^{k}_{ji}\vartheta^{i}\wedge\vartheta^{j}=\dfrac{1}{2}\omega^{k}_{j}\wedge\vartheta^{j}+\dfrac{1}{2}\omega^{k}_{i}\wedge\vartheta^{i}=\omega^{k}_{i}\wedge\vartheta^{i}$$

(in the second term, the factors of the \(2\)-covector are interchanged). Thus, up to index notation, one has (A15).

Finally, multiply both sides of the equality (A11) on \(\dfrac{1}{2}\vartheta^{i}\wedge\vartheta^{j}\) from the right and sum over all values \(i,\>j\). This results in the following formula

$$\dfrac{1}{2}\Omega^{s}_{i}\,\partial_{s}\gamma^{m}_{jk}\vartheta^{i}\wedge\vartheta^{j}-\dfrac{1}{2}\Omega^{s}_{j}\,\partial_{s}\gamma^{m}_{ik}\vartheta^{i}\wedge\vartheta^{j}+\dfrac{1}{2}\gamma^{s}_{jk}\gamma^{m}_{is}\vartheta^{i}\wedge\vartheta^{j}-\dfrac{1}{2}\gamma^{s}_{ik}\gamma^{m}_{js}\vartheta^{i}\wedge\vartheta^{j}$$
$${}+\dfrac{1}{2}c^{s}_{ij}\gamma^{m}_{sk}\vartheta^{i}\wedge\vartheta^{j}=\dfrac{1}{2}\mathfrak{R}^{m}_{ijk}\vartheta^{i}\wedge\vartheta^{j},$$

the right hand side of which is curvature form \(R^{m}_{k}\). The third and fourth terms give

$$\dfrac{1}{2}\gamma^{s}_{jk}\gamma^{m}_{is}\vartheta^{i}\wedge\vartheta^{j}-\dfrac{1}{2}\gamma^{s}_{ik}\gamma^{m}_{js}\vartheta^{i}\wedge\vartheta^{j}=\dfrac{1}{2}\omega^{m}_{s}\wedge\omega^{s}_{k}-\dfrac{1}{2}\omega^{s}_{k}\wedge\omega^{m}_{s}=\omega^{m}_{s}\wedge\omega^{s}_{k}.$$

It remains to consider the sum

$$\dfrac{1}{2}\Omega^{s}_{i}\,\partial_{s}\gamma^{m}_{jk}\vartheta^{i}\wedge\vartheta^{j}-\dfrac{1}{2}\Omega^{s}_{j}\,\partial_{s}\gamma^{m}_{ik}\vartheta^{i}\wedge\vartheta^{j}+\dfrac{1}{2}c^{s}_{ij}\gamma^{m}_{sk}\vartheta^{i}\wedge\vartheta^{j}.$$
(A17)

Interchanging indices \(i\) and \(j\) in the second term and then interchanging factors of \(2\)-covector, one gets

$$\dfrac{1}{2}\Omega^{s}_{i}\,\partial_{s}\gamma^{m}_{jk}\vartheta^{i}\wedge\vartheta^{j}-\dfrac{1}{2}\Omega^{s}_{i}\,\partial_{s}\gamma^{m}_{jk}\vartheta^{j}\wedge\vartheta^{i}=\Omega^{s}_{i}\,\partial_{s}\gamma^{m}_{jk}\vartheta^{i}\wedge\vartheta^{j}.$$

Thus, (A17) has the form

$$\dfrac{1}{2}\Omega^{s}_{i}\,\partial_{s}\gamma^{m}_{jk}\vartheta^{i}\wedge\vartheta^{j}-\dfrac{1}{2}\Omega^{s}_{j}\,\partial_{s}\gamma^{m}_{ik}\vartheta^{i}\wedge\vartheta^{j}+\dfrac{1}{2}c^{s}_{ij}\gamma^{m}_{sk}\vartheta^{i}\wedge\vartheta^{j}=\Omega^{s}_{i}\,\partial_{s}\gamma^{m}_{jk}\vartheta^{i}\wedge\vartheta^{j}+\gamma^{m}_{sk}d\vartheta^{s},$$

where the equality (A7) was utilized as well. The first term of the resulting expression can also be transformed

$$\Omega^{s}_{i}\,\partial_{s}\gamma^{m}_{jk}\vartheta^{i}\wedge\vartheta^{j}=\partial_{s}\gamma^{m}_{jk}dx^{s}\wedge\vartheta^{j}=d\gamma^{m}_{sk}\wedge\vartheta^{s}.$$

Therefore, in fact, the expression (A17) is equal to

$$\dfrac{1}{2}\Omega^{s}_{i}\,\partial_{s}\gamma^{m}_{jk}\vartheta^{i}\wedge\vartheta^{j}-\dfrac{1}{2}\Omega^{s}_{j}\,\partial_{s}\gamma^{m}_{ik}\vartheta^{i}\wedge\vartheta^{j}+\dfrac{1}{2}c^{s}_{ij}\gamma^{m}_{sk}\vartheta^{i}\wedge\vartheta^{j}$$
$${}=d\gamma^{m}_{sk}\wedge\vartheta^{s}+\gamma^{m}_{sk}d\vartheta^{s}=d(\gamma^{m}_{sk}\wedge\vartheta^{s})=d\omega^{m}_{k}.$$

Collecting the obtained results, one arrives at the relation (A16). \(\Box\)

Cartan structural equations make it possible to derive relations for the differentials of torsion, curvature, and non-metricity.

Theorem 7 (Bianchi Identities). The following identities hold

$$dQ_{jk}=g_{mk}R^{m}_{j}+g_{mj}R^{m}_{k}-Q_{mk}\wedge\omega^{m}_{j}-Q_{mj}\wedge\omega^{m}_{k},$$
(A18)
$$dT^{i}=R^{i}_{j}\wedge\vartheta^{j}-\omega^{i}_{j}\wedge T^{j},$$
(A19)
$$dR^{i}_{j}=R^{i}_{k}\wedge\omega^{k}_{j}-\omega^{i}_{k}\wedge R^{k}_{j}.$$
(A20)

Proof. Each of the relations (A18)–(A20) can be obtained just with Cartan equations (A14)–(A16) only. Indeed, applying the exterior differential to both sides of \(0\)th Cartan equation (A14), one getsFootnote 23

$$dQ_{jk}=dg_{mk}\wedge\omega^{m}_{j}+g_{mk}d\omega^{m}_{j}+dg_{mj}\wedge\omega^{m}_{k}+g_{mj}d\omega^{m}_{k}.$$
(A21)

From another perspective, \(0\)th and \(2\)nd Cartan equations (A14) and (A16) imply

$$dg_{mk}=g_{sk}\omega^{s}_{m}+g_{sm}\omega^{s}_{k}-Q_{mk},\quad dg_{mj}=g_{sj}\omega^{s}_{m}+g_{sm}\omega^{s}_{j}-Q_{mj},$$
$$d\omega^{m}_{j}=R^{m}_{j}-\omega^{m}_{s}\wedge\omega^{s}_{j},\quad d\omega^{m}_{k}=R^{m}_{k}-\omega^{m}_{s}\wedge\omega^{s}_{k}.$$

Substituting the obtained expressions into (A21) and collecting similar terms, one arrives at (A18).

Acting by exterior derivative \(d\) onto both sides of the \(1\)st Cartan equation (A15) and taking into account that \(d\circ d=0\), one gets

$$dT^{i}=d\omega^{i}_{j}\wedge\vartheta^{j}-\omega^{i}_{j}\wedge d\vartheta^{j}.$$
(A22)

Further, from the first and second Cartan equations (A15) and (A16) one can obtain expressions for \(d\omega^{i}_{j}\) and \(d\vartheta^{j}\):

$$d\omega^{i}_{j}=R^{i}_{j}-\omega^{i}_{k}\wedge\omega^{k}_{j},\quad d\vartheta^{j}=T^{j}-\omega^{j}_{k}\wedge\vartheta^{k}.$$

Their substitution into (A22) gives (A19).

Finally, let us act by exterior derivative \(d\) onto \(2\)nd Cartan equation (A16) (and, like in previous cases, take into account that \(d\circ d=0\)):

$$dR^{i}_{j}=d\omega^{i}_{k}\wedge\omega^{k}_{j}-\omega^{i}_{k}\wedge d\omega^{k}_{j}.$$
(A23)

On the other hand, from the second Cartan equation one can obtain representations for \(d\omega^{i}_{k}\) and \(d\omega^{k}_{j}\):

$$d\omega^{i}_{k}=R^{i}_{k}-\omega^{i}_{s}\wedge\omega^{s}_{k},\quad d\omega^{k}_{j}=R^{k}_{j}-\omega^{k}_{s}\wedge\omega^{s}_{j}.$$

Substituting them into (A23), one obtains (A20). \(\Box\)

We will call the equalities (A18)–(A20) as Bianchi identities.

The Cartan equations and the Bianchi identities can be written in a more concise form. This can be achieved if one introduces the generalization of exterior derivative, namely, covariant exterior derivative ([100], p. 139). Consider family of \(p\)-forms

$$\Pi^{i_{1}\ldots i_{k}}{}_{j_{1}\ldots j_{l}}=\frac{1}{p!}A^{i_{1}\ldots i_{k}}{}_{j_{1}\ldots j_{l}j_{l+1}j_{l+p}}\,\vartheta^{j_{l+1}}\wedge\cdots\wedge\vartheta^{j_{l+p}},\quad i_{1},\,\ldots,\,i_{k},\,j_{1},\,\ldots,\,j_{l}=1,\,\ldots,\,n,$$

where \(A^{i_{1}\ldots i_{k}}{}_{j_{1}\ldots j_{l}j_{l+1}j_{l+p}}\) are components of \((k,\,l+p)\)-tensor, that are antisymmetric with respect to the latter \(p\) lower indices. Then, covariant exterior derivative \(D\Pi^{i_{1}\ldots i_{k}}{}_{j_{1}\ldots j_{l}}\) is defined as

$$D\Pi^{i_{1}\ldots i_{k}}{}_{j_{1}\ldots j_{l}}=d\Pi^{i_{1}\ldots i_{k}}{}_{j_{1}\ldots j_{l}}+\omega^{i_{1}}_{s}\wedge\Pi^{s\ldots i_{k}}{}_{j_{1}\ldots j_{l}}+\cdots+\omega^{i_{k}}_{s}\wedge\Pi^{i_{1}\ldots s}{}_{j_{1}\ldots j_{l}}$$
$${}-\omega^{s}_{j_{1}}\wedge\Pi^{i_{1}\ldots i_{k}}{}_{s\ldots j_{l}}-\cdots-\omega^{s}_{j_{l}}\wedge\Pi^{i_{1}\ldots i_{k}}{}_{j_{1}\ldots s},$$
(A24)

where \(d\) is exterior differential. In particular,

$$D\Pi^{i}=d\Pi^{i}+\omega^{i}_{k}\wedge\Pi^{k}\quad\text{and}\quad D\Pi^{i}_{j}=d\Pi^{i}_{j}+\omega^{i}_{k}\wedge\Pi^{k}_{j}-\omega^{k}_{j}\wedge\Pi^{i}_{k}.$$

Thus, \(D\) increases the rank of the ‘‘outer’’ part by \(1\).

Using operation \(D\), one can rewrite Cartan equations (A14)–(A16) as

$$Dg_{ij}=-Q_{ij},\quad D\vartheta^{i}=T^{i},\quad D\omega^{i}_{j}=R^{i}_{j}.$$
(A25)

The Bianchi identities (A18)–(A20), in turn, are represented as

$$DQ_{jk}=g_{mk}R^{m}_{j}+g_{mj}R^{m}_{k},\quad DT^{i}=R^{i}_{j}\wedge\vartheta^{j},\quad DR^{i}_{j}=\mathsf{0}.$$
(A26)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lychev, S.A., Koifman, K.G. & Pivovaroff, N.A. Incompatible Deformations in Relativistic Elasticity. Lobachevskii J Math 44, 2352–2397 (2023). https://doi.org/10.1134/S1995080223060343

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1134/S1995080223060343

Keywords:

Navigation