Skip to main content
Log in

Thermodynamically consistent variational theory of porous media with a breaking component

  • Original Article
  • Published:
Continuum Mechanics and Thermodynamics Aims and scope Submit manuscript

Abstract

If a porous media is being damaged by excessive stress, the elastic matrix at every infinitesimal volume separates into a ‘solid’ and a ‘broken’ component. The ‘solid’ part is the one that is capable of transferring stress, whereas the ‘broken’ part is advecting passively and is not able to transfer the stress. In previous works, damage mechanics was addressed by introducing the damage parameter affecting the elastic properties of the material. In this work, we take a more microscopic point of view, by considering the transition from the ‘solid’ part, which can transfer mechanical stress, to the ‘broken’ part, which consists of microscopic solid particles and does not transfer mechanical stress. Based on this approach, we develop a thermodynamically consistent dynamical theory for porous media including the transfer between the ‘broken’ and ‘solid’ components, by using a variational principle recently proposed in thermodynamics. This setting allows us to derive an explicit formula for the breaking rate, i.e., the transition from the ‘solid’ to the ‘broken’ phase, dependent on the Gibbs’ free energy of each phase. Using that expression, we derive a reduced variational model for material breaking under one-dimensional deformations. We show that the material is destroyed in finite time, and that the number of ‘solid’ strands vanishing at the singularity follows a power law. We also discuss connections with existing experiments on material breaking and extensions to multi-phase porous media.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

References

  1. Abali, B.E., Klunker, A., Barchiesi, E., Placidi, L.: A novel phase-field approach to brittle damage mechanics of gradient metamaterials combining action formalism and history variable. ZAMM Z. Angew. Math. Mech. 101(9), e202000289 (2021)

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  3. Athesian, G., Warden, W., Kim, J.: Finite deformation biphasic material properties of bovine articular cartilage from confined compression experiments. J. Biomech. 30, 1157–1164 (1997)

    Article  Google Scholar 

  4. Atkin, R.J., Craine, R.E.: Continuum theories of mixtures: basic theory and historical development. Q. J. Mech. Appl. Math. 29(2), 209–244 (1976)

    Article  MathSciNet  Google Scholar 

  5. Auffray, N., dell’Isola, F., Eremeyev, V.A., Madeo, A., Rosi, G.: Analytical continuum mechanics à la Hamilton-Piola least action principle for second gradient continua and capillary fluids. Math. Mech. Solids 20(4), 375–417 (2015)

    Article  MathSciNet  Google Scholar 

  6. Aulisa, E., Cakmak, A., Ibragimov, A., Solynin, A.: Variational principle and steady state invariants for non-linear hydrodynamic interactions in porous media. Dyn. Contin. Discret. Impuls. Syst. (Series A) (2007)

  7. Aulisa, E., Ibragimov, A., Toda, M.: Geometric framework for modeling nonlinear flows in porous media, and its applications in engineering. Nonlinear Anal. Real World Appl. 11(3), 1734–1751 (2010)

    Article  MathSciNet  Google Scholar 

  8. Bauer, W., Gay-Balmaz, F.: Variational integrators for anelastic and pseudo-incompressible flows. J. Geom. Mech. 11(4), 511–537 (2019)

    Article  MathSciNet  Google Scholar 

  9. Bažant, Z.P., Jirásek, M.: Nonlocal integral formulations of plasticity and damage: survey of progress. J. Eng. Mech. 128(11), 1119–1149 (2002)

    Article  Google Scholar 

  10. Bedford, A., Drumheller, D.S.: A variational theory of porous media. Int. J. Solids Struct. 15(12), 967–980 (1979)

    Article  MathSciNet  Google Scholar 

  11. Biot, M.A.: General theory of three-dimensional consolidation. J. Appl. Phys. 12(2), 155–164 (1941)

    Article  ADS  Google Scholar 

  12. Brecht, R., Bauer, W., Bihlo, A., Gay-Balmaz, F., MacLachlan, S.: Variational integrator for the rotating shallow-water equations on the sphere. Q. J. R. Meteorol. Soc. 145(720), 1070–1088 (2019)

    Article  ADS  Google Scholar 

  13. Brinkman, H.C.: A calculation of the viscous force exerted by a flowing fluid on a dense swarm of particles. Flow Turbul. Combust. 1(1), 27–34 (1949)

    Article  Google Scholar 

  14. Brinkman, H.C.: On the permeability of media consisting of closely packed porous particles. Flow Turbul. Combust. 1(1), 81–86 (1949)

    Article  Google Scholar 

  15. Brinkman, H.C.: The viscosity of concentrated suspensions and solutions. J. Chem. Phys. 20(4), 571–571 (1952)

    Article  ADS  Google Scholar 

  16. Bucchi, A., De Tommasi, D., Puglisi, G., Saccomandi, G.: Damage as a material phase transition. J. Elast. (2023). https://doi.org/10.1007/s10659-023-10014-z

    Article  MathSciNet  Google Scholar 

  17. Cohen, B., Lai, W., Mow, V.: A transversely isotropic biphasic model for unconfined compression of growth plate and chondroepiphysis. J. Biomech. Eng. 120(4), 491–496 (1998)

    Article  Google Scholar 

  18. Coussy, O.: Mechanics of Porous Continua. Wiley, New York (1995)

    Google Scholar 

  19. de Boer, R.: Contemporary progress in porous media theory. Appl. Mech. Rev. 53(12), 323–370 (2000)

    Article  ADS  Google Scholar 

  20. de Boer, R.: Trends in Continuum Mechanics of Porous Media, vol. 18. Springer Science & Business Media, Berlin (2005)

    Book  Google Scholar 

  21. de Boer, R.: Theory of Porous Media: Highlights in Historical Development and Current State. Springer Science & Business Media, Berlin (2012)

    Google Scholar 

  22. de Groot, S.R., Mazur, P.: Nonequilibrium Thermodynamics. North-Holland, Amsterdam (1969)

    Google Scholar 

  23. dell’Isola, F., Guarascio, M., Hutter, K.: A variational approach for the deformation of a saturated porous solid. A second-gradient theory extending Terzaghi’s effective stress principle. Arch. Appl. Mech. 70(5), 323–337 (2000)

    Article  ADS  Google Scholar 

  24. dell’Isola, F., Madeo, A., Seppecher, P.: Boundary conditions at fluid-permeable interfaces in porous media: a variational approach. Int. J. Solids Struct. 46(17), 3150–3164 (2009)

    Article  MathSciNet  Google Scholar 

  25. Eldred, C., Bauer, W.: Variational and Hamiltonian formulations of geophysical fluids using split exterior calculus. HAL-INRIA preprint, (2018)

  26. Farkhutdinov, T., Gay-Balmaz, F., Putkaradze, V.: Geometric variational approach to the dynamics of porous media filled with incompressible fluid. Acta Mech. 231, 3897–3924 (2020)

    Article  MathSciNet  Google Scholar 

  27. Farkhutdinov, T., Gay-Balmaz, F., Putkaradze, V.: Actively deforming porous media in an incompressible fluid: a variational approach. Phys. D Nonlinear Phenom. 426, 132984 (2020)

    Article  MathSciNet  Google Scholar 

  28. Gastaldi, D., Sassi, V., Petrini, L., Vedani, M., Trasatti, S., Migliavacca, F.: Continuum damage model for bioresorbable magnesium alloy devices-application to coronary stents. J. Mech. Behav. Biomed. Mate. 4(3), 352–365 (2011)

    Article  Google Scholar 

  29. Gawlik, E.S., Gay-Balmaz, F.: A variational finite element discretization of compressible flow. Found. Comput. Math. 21, 961–1001 (2020)

    Article  MathSciNet  Google Scholar 

  30. Gawlik, E. S., Gay-Balmaz, F.: Variational and thermodynamically consistent finite element discretization for heat conducting viscous fluids. preprint arXiv:2211.08745, (2023)

  31. Gay-Balmaz, F.: A variational derivation of the nonequilibrium thermodynamics of a moist atmosphere with rain process and its pseudoincompressible approximation. Geophys. Astrophys. Fluid Dyn. 113, 428–465 (2019)

    Article  ADS  MathSciNet  Google Scholar 

  32. Gay-Balmaz, F., Putkaradze, V.: Variational geometric approach to the thermodynamics of porous media. ZAMM Z. Angew. Math. Mech. 102(11), e202100198 (2022)

    Article  MathSciNet  Google Scholar 

  33. Gay-Balmaz, F., Yoshimura, H.: A Lagrangian variational formulation for nonequilibrium thermodynamics. Part I: discrete systems. J. Geom. Phys. 111, 169–193 (2017)

    Article  ADS  MathSciNet  Google Scholar 

  34. Gay-Balmaz, F., Yoshimura, H.: A Lagrangian variational formulation for nonequilibrium thermodynamics. Part II: continuum systems. J. Geom. Phys. 111, 194–212 (2017)

    Article  ADS  MathSciNet  Google Scholar 

  35. Gay-Balmaz, F., Yoshimura, H.: From Lagrangian mechanics to nonequilibrium thermodynamics: a variational perspective. Entropy 21, 8 (2019)

    Article  ADS  Google Scholar 

  36. Gay-Balmaz, F., Marsden, J.E., Ratiu, T.S.: Reduced variational formulations in free boundary continuum mechanics. J. Nonlinear Sci. 22(4), 463–497 (2012)

    Article  ADS  MathSciNet  Google Scholar 

  37. Giorgio, I., dell’Isola, F., Andreaus, U., Alzahrani, F., Hayat, T., Lekszycki, T.: On mechanically driven biological stimulus for bone remodeling as a diffusive phenomenon. Biomech. Model. Mechanobiol. 18, 1639–1663 (2019)

    Article  Google Scholar 

  38. Hamiel, Y., Lyakhovsky, V., Agnon, A.: Coupled evolution of damage and porosity in poroelastic media: theory and applications to deformation of porous rocks. Geophys. J. Int. 156(3), 701–713 (2004)

    Article  ADS  Google Scholar 

  39. Holmes, M., Mow, V.C.: The nonlinear characteristics of soft gels and hydrated connective tissues in ultrafiltration. J. Biomech. 23(11), 1145–1156 (1990)

    Article  Google Scholar 

  40. Homand-Etienne, F., Hoxha, D., Shao, J.-F.: A continuum damage constitutive law for brittle rocks. Comput. Geotech. 22(2), 135–151 (1998)

    Article  Google Scholar 

  41. Jeffers, J.R., Browne, M., Taylor, M.: Damage accumulation, fatigue and creep behaviour of vacuum mixed bone cement. Biomaterials 26(27), 5532–5541 (2005)

    Article  Google Scholar 

  42. Kachanov, L.: Introduction to Continuum Damage Mechanics, vol. 10. Springer Science & Business Media, Berlin (1986)

    Google Scholar 

  43. Kachanov, M.: Elastic solids with many cracks and related problems. Adv. Appl. Mech. 30, 259–445 (1993)

    Article  Google Scholar 

  44. Kannan, K., Rajagopal, K.R.: Flow through porous media due to high pressure gradients. Appl. Math. Comput. 199(2), 748–759 (2008)

    MathSciNet  Google Scholar 

  45. Lemaitre, J.: A Course on Damage Mechanics. Springer Science & Business Media, Berlin (2012)

    Google Scholar 

  46. Lopatnikov, S.L., Cheng, A.H.-D.: Macroscopic Lagrangian formulation of poroelasticity with porosity dynamics. J. Mech. Phys. Solids 52(12), 2801–2839 (2004)

    Article  ADS  MathSciNet  Google Scholar 

  47. Lopatnikov, S.L., Gillespie, J.W.: Poroelasticity-I: governing equations of the mechanics of fluid-saturated porous materials. Transp. Porous Med. 84(2), 471–492 (2010)

    Article  MathSciNet  Google Scholar 

  48. Lyakhovsky, V., Hamiel, Y.: Damage evolution and fluid flow in poroelastic rock. Izvestiya Phys. Solid Earth 43(1), 13–23 (2007)

    Article  ADS  Google Scholar 

  49. Marsden, J.E., Hughes, T.J.R.: Mathematical Foundations of Elasticity. Courier Corporation, Chelmsford (1994)

    Google Scholar 

  50. Mihai, L.A., Chin, L., Janmey, P.A., Goriely, A.: A comparison of hyperelastic constitutive models applicable to brain and fat tissues. J. R. Soc. Interface 12(110), 20150486 (2015)

    Article  Google Scholar 

  51. Mihai, L.A., Budday, S., Holzapfel, G.A., Kuhl, E., Goriely, A.: A family of hyperelastic models for human brain tissue. J. Mech. Phys. Solids 106, 60–79 (2017)

    Article  ADS  MathSciNet  Google Scholar 

  52. Mobasher, M.E., Berger-Vergiat, L., Waisman, H.: Non-local formulation for transport and damage in porous media. Comput. Methods Appl. Mech. Eng. 324, 654–688 (2017)

    Article  ADS  MathSciNet  Google Scholar 

  53. Murphy, B.P., Prendergast, P.J.: On the magnitude and variability of the fatigue strength of acrylic bone cement. Int. J. Fatigue 22(10), 855–864 (2000)

    Article  Google Scholar 

  54. Ogden, R.W.: Large deformation isotropic elasticity-on the correlation of theory and experiment for incompressible rubberlike solids. Proc. R. Soc. London A Math. Phys. Sci. 326(1567), 565–584 (1972)

    ADS  Google Scholar 

  55. Placidi, L., dell’Isola, F., Ianiro, N., Sciarra, G.: Variational formulation of pre-stressed solid-fluid mixture theory, with an application to wave phenomena. Eur. J. Mech. A Solids 27(4), 582–606 (2008)

    Article  ADS  MathSciNet  Google Scholar 

  56. Placidi, L., Barchiesi, E., Misra, A.: A strain gradient variational approach to damage: a comparison with damage gradient models and numerical results. Math. Mech. Complex Syst. 6(2), 77–100 (2018)

    Article  MathSciNet  Google Scholar 

  57. Placidi, L., Misra, A., Barchiesi, E.: Two-dimensional strain gradient damage modeling: a variational approach. ZAMM Z. Angew. Math. Mech. 69, 1–19 (2018)

    MathSciNet  Google Scholar 

  58. Placidi, L., Barchiesi, E., Misra, A., Andreaus, U.: Variational methods in continuum damage and fracture mechanics. Encycl. Contin. Mech. 2634–2643 (2020)

  59. Rhodes, M.E., Hillen, T., Putkaradze, V.: Comparing the effects of linear and one-term Ogden elasticity in a model of glioblastoma invasion. Brain Multiphys. 3, 100050 (2022)

    Article  Google Scholar 

  60. Sciarra, G., dell’Isola, F., Ianiro, N., Madeo, A.: A variational deduction of second gradient poroelasticity I: general theory. J. Mech. Mater. Struct. 3(3), 507–526 (2008)

    Article  Google Scholar 

  61. Sciarra, G., dell’Isola, F., Ianiro, N., Sciarra, G.: A variational deduction of second gradient poroelasticity II: an application to the consolidation problem. J. Mech. Mater. Struct. 3(4), 607–625 (2008)

    Article  Google Scholar 

  62. Serpieri, R., Rosati, L.: Formulation of a finite deformation model for the dynamic response of open cell biphasic media. J. Mech. Phys. Solids 59(4), 841–862 (2011)

    Article  ADS  MathSciNet  Google Scholar 

  63. Serpieri, R., Travascio, F.: General quantitative analysis of stress partitioning and boundary conditions in undrained biphasic porous media via a purely macroscopic and purely variational approach. Contin. Mech. Thermodyn. 28(1–2), 235–261 (2016)

    Article  ADS  MathSciNet  Google Scholar 

  64. Serpieri, R., Travascio, F.: Variational Continuum Multiphase Poroelasticity. Springer, Berlin (2017)

    Book  Google Scholar 

  65. Serpieri, R., Travascio, F., Asfour, S., Rosati, L.: Variationally consistent derivation of the stress partitioning law in saturated porous media. Int. J. Solids Struct. 56, 235–247 (2015)

    Article  Google Scholar 

  66. Serpieri, R., Della Corte, A., Travascio, F., Rosati, L.: Variational theories of two-phase continuum poroelastic mixtures: a short survey. In: Generalized Continua as Models for Classical and Advanced Materials, pp. 377–394. Springer, Berlin (2016)

    Chapter  Google Scholar 

  67. Srinivasan, S., Rajagopal, K.R.: A thermodynamic basis for the derivation of the Darcy, Forchheimer and Brinkman models for flows through porous media and their generalizations. Int. J. Non-Linear Mech. 58, 162–166 (2014)

    Article  ADS  Google Scholar 

  68. Travascio, F., Asfour, S., Serpieri, R., Rosati, L.: Analysis of the consolidation problem of compressible porous media by a macroscopic variational continuum approach. Math. Mech. Solids 22(5), 952–968 (2017)

    Article  MathSciNet  Google Scholar 

  69. Weibull, W.: A statistical distribution function of wide applicability. J. Appl. Mech. (1951)

  70. Weymouth, G.D., Yue, D.K.-P.: Conservative volume-of-fluid method for free-surface simulations on cartesian-grids. J. Comput. Phys. 229(8), 2853–2865 (2010)

    Article  ADS  MathSciNet  Google Scholar 

Download references

Acknowledgements

We are grateful for fruitful discussions to Profs. D. D. Holm and D. V. Zenkov, and Drs. C. Eldred and A. Lokhov, elucidating the applicability of our theory and the extent of development of damage mechanics theories. VP was partially supported by NSERC Discovery Grant 2023-03590.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Vakhtang Putkaradze.

Ethics declarations

Conflict of interest

Authors declare no competing interests.

Additional information

Publisher's Note

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

Appendices

Variational formulation for a viscous medium with heat conduction

In this Appendix, we review the variational theory of a single-component fluid or elastic medium with heat conduction and viscosity, based on the variational thermodynamics approach developed in [33,34,35]. This work was generalized for porous media in [32], in the case where the elastic medium consists of a single component. We use this variational thermodynamics framework in Sect. 3 to treat the case of a two-component (‘solid’ and ‘broken’) elastic media which includes the additional irreversible processes of transitions between the components, where both components are embedded in the fluid.

1.1 Lagrangian (material) description

In the material description, the variational formulation of thermodynamics is an extension of the Hamilton principle of continuum mechanics. We recall the material description here for pedagogical reasons since the variational formulation takes its simplest form. The variational formulation in the spatial (Eulerian) description is then derived from it, see the next section, and turns out to be a useful tool for porous media modeling.

For a single-component fluid or elastic medium, we take the Lagrangian L to be a function of the configuration diffeomorphism \(\varvec{\varphi }(t, {\textbf{X}} )\), see Sect. 2.1.2, the material velocity \( {\varvec{V}}(t, {\textbf{X}} )= \dot{\varvec{\varphi }}(t, {\textbf{X}} )\), and the entropy density \(S(t, {\varvec{X}})\). It also depends parametrically on the density \(\varrho _0({\varvec{X}})\) and on the Riemannian metric \(G_0({\varvec{X}})\) on the reference configuration which can be chosen as the Euclidean one when \( {\mathcal {B}} \subset {\mathbb {R}} ^3\). The metric \(G_0({\varvec{X}})\) is needed to introduce the Eulerian deformation tensors.

For each fixed \( \varrho _0( {\textbf{X}} ) \) and \(G_0( {\textbf{X}} )\), we can write the Lagrangian as a function \(L: T {\text {Diff}}_0( {\mathcal {B}}) \times {\text {Den}}( {\mathcal {B}} ) \rightarrow {\mathbb {R}}\), where \( {\text {Diff}}_0( {\mathcal {B}} )\ni \varvec{\varphi } \) is the group of diffeomorphisms of \( {\mathcal {B}} \) keeping \( \partial {\mathcal {B}} \) pointwise fixed, corresponding to no-slip boundary conditions, \(T{\text {Diff}}_0( {\mathcal {B}} )\ni (\varvec{\varphi }, \dot{\varvec{\varphi }})\) is its tangent bundle, and \({\text {Den}}( {\mathcal {B}} )\ni S\) is the space of densities on \( {\mathcal {B}} \). In the material description, a standard expression is given by

$$\begin{aligned} L( \varvec{\varphi }, \dot{ \varvec{\varphi } }, S)= \int _ {\mathcal {B}} \left[ \frac{1}{2} \varrho _0 | \dot{\varvec{\varphi } }|- \varrho _0 \,{\mathcal {E}} ( \nabla \varvec{\varphi }, \varrho _0, S, G_0) \right] \textrm{d}^3 {\varvec{X}}, \end{aligned}$$
(97)

with \({\mathcal {E}}\) the internal energy in the material description, subject to the usual covariance assumptions, [36, 49]. In particular, due to the material covariance assumption, the internal energy can be written in terms of the Eulerian variables \( \rho , s, b\) see Sect. A.2, as

$$\begin{aligned} {\mathcal {E}} ( \nabla \varvec{\varphi }, \varrho _0, S, G_0)= e( \rho , s, b) \circ \varvec{\varphi }, \end{aligned}$$
(98)

with e the specific internal energy in the Eulerian description.

The critical action principle for thermodynamics needs the introduction of two additional variables: \( \Sigma (t,{\varvec{X}}) \) which is identified with the entropy generated by the irreversible processes (unlike \(S(t, {\textbf{X}} )\), which is the total entropy), and \( \Gamma (t,{\varvec{X}}) \) identified with the thermal displacement [34]. With that notation, the critical action principle for a heat conducting viscous continuum reads [33, 34]:

$$\begin{aligned} \delta \int _0^ T \left[ L (\varvec{\varphi }, \dot{ \varvec{\varphi } }, S) + \int _ {\mathcal {B}} (S- \Sigma ) {\dot{\Gamma }} \,\textrm{d}^3 {\varvec{X}} \right] \textrm{d}t=0\,, \end{aligned}$$
(99)

subject to the phenomenological constraint

$$\begin{aligned} \frac{\delta L}{ \delta S}{\dot{\Sigma }} = - {\varvec{P}}: \nabla \dot{\varvec{\varphi }} + {\varvec{J}}_S \cdot \nabla \dot{ \Gamma } \end{aligned}$$
(100)

and with respect to variations \( \delta \varvec{\varphi } \), \( \delta S\), \( \delta \Sigma \), \( \delta \Gamma \) subject to the variational constraint

$$\begin{aligned} \frac{\delta L}{ \delta S}\delta \Sigma = - {\varvec{P}}: \nabla \delta \varvec{\varphi } + {\varvec{J}}_S \cdot \nabla \delta \Gamma . \end{aligned}$$
(101)

The tensor \({\varvec{P}}\) is the Piola–Kirchhoff viscous stress tensor and \( {\varvec{J}}_S\) is the entropy flux density in Lagrangian representation, see [34]. They are the Lagrangian objects corresponding to the more widely used Eulerian viscous stress tensor \( \varvec{\sigma }\) and Eulerian entropy flux density \( {\varvec{j}}_s\), see below.

An application of (99)–(101) yields, in addition to the equations of motion in the Lagrangian frame that we do not present here, the conditions

$$\begin{aligned} {\dot{\Sigma }} = \dot{S} + {\text {DIV}} {\varvec{J}}_S \;\; \text{ and } \;\;{\dot{\Gamma }} = - \frac{\delta L}{\delta S} = T \end{aligned}$$
(102)

with T being the temperature, see [34], with the notation \({\text {DIV}}\) indicating the divergence operator in the reference frame, as opposed to the divergence operator \({\text {div}}\) in the spatial frame. These conditions attribute to \( \Sigma \) and \(\Gamma \) their physical meaning mentioned above.

Since it is \(\Sigma \) that describes the entropy of irreversible processes, from the second law of thermodynamics, we must have

$$\begin{aligned} {\dot{\Sigma }} \ge 0 \,, \end{aligned}$$
(103)

whereas \(\dot{S}\) does not necessarily has to have a particular sign.

Remark A.1

(Structure of the variational formulation) Observe that (99)–(101) is an extension of the Hamilton principle \( \delta \int _0^T L( \varvec{\varphi }, \dot{\varvec{\varphi }} ) \textrm{d} t=0\) of continuum mechanics which involves two types of constraints: the constraint (100) on the critical curve and the constraint (101) on the variations to be used when computing this critical curve. One passes from (100) to (101) by formally replacing the time rate of changes by \(\delta \)-variations for each irreversible processes, such as \(F_i \dot{x}^i \leadsto F_i \delta x^i\) in the case of an irreversible process due to a friction force for a finite dimensional system, see [33,34,35]. This variational formulation is reminiscent from the Lagrange-d’Alembert principle used in nonholonomic mechanics. A finite dimensional version of the variational formulation (99)–(101) will be used as a modeling tool in Sect. 5.

1.2 Eulerian (spatial) description

The Eulerian variables are the velocity \({\varvec{u}}(t, {\varvec{x}})\), mass density \( \rho (t, {\varvec{x}})\), entropy density \( s(t, {\varvec{x}})\), and Finger deformation tensor \(b(t, {\varvec{x}})\) defined from the material variables \( \varvec{\varphi } (t, {\varvec{X}})\), \(\dot{ \varvec{\varphi }} (t, {\varvec{X}})\), \(\varrho _0 (t, {\varvec{X}})\), \(S (t, {\varvec{X}})\), and \(G_0 (t, {\varvec{X}})\) in the usual way. In particular, the Finger deformation tensor is the push-forward of the inverted metric by the configuration diffeomorphism \( \varvec{\varphi } \):

$$\begin{aligned} b= \varvec{\varphi } _* G_0 ^{-1}, \end{aligned}$$
(104)

see [36, 49]. In the case when \({\mathcal {B}}\) is a subset of \({\mathbb {R}}^3\) and \(G_0\) is the identity tensor, the Finger deformation tensor b is written in terms of the deformation gradient tensor \({\mathbb {F}}= \nabla \varvec{\varphi } \) as

$$\begin{aligned} b = {\mathbb {F}} \cdot {\mathbb {F}}^T \,, \quad b^{ij}(t, {\varvec{x}}) = \frac{\partial x^i}{\partial X^k}\frac{\partial x^j}{\partial X^k}( \varvec{\varphi }^{-1}(t, {\varvec{x}})). \end{aligned}$$
(105)

Note that from \( \varvec{\varphi } \in {\text {Diff}}_0( {\mathcal {B}})\), we have the no-slip boundary conditions \( {\varvec{u}}=0\) on \( \partial {\mathcal {B}} \). We also consider the thermal displacement \( \gamma (t, {\varvec{x}}) \), internal entropy density \( \sigma (t, {\varvec{x}}) \), Cauchy stress \( \varvec{\sigma } (t, {\varvec{x}}) \) and entropy flux \( {\varvec{j}}_s (t, {\varvec{x}}) \) defined as the Eulerian quantities associated to \( \Gamma (t, {\varvec{X}}) \), \( \Sigma (t, {\varvec{X}}) \), \( {\varvec{P}} (t, {\varvec{X}}) \), and \( {\varvec{J}}_S (t, {\varvec{X}}) \), see [34].

With these definitions, the spatial version of (99)–(101) gives the variational formulation

$$\begin{aligned} \delta \int _0^ T \left[ \ell ({\varvec{u}}, \rho , s, b) + \int _ {\mathcal {B}} (s - \sigma ) D_t \gamma \,\textrm{d}^3 {\varvec{x}} \right] \textrm{d}t=0\,, \end{aligned}$$
(106)

subject to the phenomenological constraint

$$\begin{aligned} \frac{\delta \ell }{ \delta s} {\bar{D}}_t \sigma = - \varvec{\sigma }: \nabla {\varvec{u}} + {\varvec{j}}_s \cdot \nabla D_t \gamma \end{aligned}$$
(107)

and with respect to variations

$$\begin{aligned} \delta {\varvec{u}}= \partial _t \varvec{\eta } + {\varvec{u}} \cdot \nabla \varvec{\eta } - \varvec{\eta } \cdot \nabla {\varvec{u}}, \quad \delta \rho = - {\text {div}}( \rho \varvec{\eta } ), \quad \delta b= - \pounds _ { \varvec{\eta } }b, \end{aligned}$$
(108)

\(\delta s\), \(\delta \sigma \), and \(\delta \gamma \) subject to the variational constraint

$$\begin{aligned} \frac{\delta \ell }{ \delta s}{\bar{D}}_ \delta \sigma = - \varvec{ \sigma }: \nabla \varvec{\eta } + {\varvec{j}}_s \cdot \nabla D_ \delta \gamma . \end{aligned}$$
(109)

We recall that \( \varvec{\eta }(t, {\varvec{x}})= \delta \varvec{\varphi } (t, \varvec{\varphi } ^{-1} (t, {\varvec{x}}))\) denotes the variation of the fluid trajectories in the Eulerian frame. In the expression of \( \delta b\), \(\pounds _{ \varvec{\eta }} b\) denotes the Lie derivative of the symmetric contravariant tensor b in the direction \(\varvec{\eta }\), which follows from (104). We have introduced the notations

$$\begin{aligned} \begin{aligned}&D_t f= \partial _t f + {\varvec{u}} \cdot \nabla f&\qquad&D_ \delta f= \delta f + \varvec{\eta } \cdot \nabla f \\&{\bar{D}}_t f = \partial _t f + {\text {div}}(f {\varvec{u}})&\qquad&{\bar{D}}_ \delta f = \delta f + {\text {div}}(f \varvec{\eta }) \end{aligned} \end{aligned}$$
(110)

for the Lagrangian time derivative and Lagrangian variations of a scalar function and a density.

A direct application of the variational principle (106)–(109) yields the general equations of motion for a heat conducting viscous continuum with Lagrangian \(\ell ( {\varvec{u}}, \rho , s, b)\) in Eulerian coordinates as

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle \partial _t \frac{\delta \ell }{\delta {\varvec{u}} }+\pounds _{{\varvec{u}}}\frac{\delta \ell }{\delta {\varvec{u}}} =\rho \nabla \frac{\delta \ell }{\delta \rho } +s\nabla \frac{\delta \ell }{\delta s}- \frac{\delta \ell }{\delta b}:\nabla b- 2{\text {div}} \left( \frac{\delta \ell }{\delta b}\cdot b \right) + {\text {div}} \varvec{\sigma } \\ \displaystyle \frac{\delta \ell }{\delta s}( {\bar{D}}_t s + {\text {div}} {\varvec{j}}_s) = - \varvec{\sigma }: \nabla {\varvec{u}}- {\varvec{j}}_s \cdot \nabla \frac{\delta \ell }{\delta s} \end{array} \right. \end{aligned}$$
(111)

together with the conditions

$$\begin{aligned} {\bar{D}}_t \sigma = {\bar{D}}_t s+ {\text {div}} {\varvec{j}}_s \qquad \text {and}\qquad {\bar{D}}_t \gamma = - \frac{\delta \ell }{\delta s}. \end{aligned}$$
(112)

From these two conditions, \({\bar{D}}_t \sigma \) is interpreted as the total entropy generation rate density and \(D_t \gamma \) is the temperature, hence \( \gamma \) is the thermal displacement. The equations for \( \rho \) and b are

$$\begin{aligned} \partial _t \rho + {\text {div}}( \rho {\varvec{u}} )=0 \quad \text {and}\quad \partial _t b + \pounds _{ {\varvec{u}}} b=0, \end{aligned}$$

which follow from the definition of \( \rho \) and b in terms of \( \varrho _0\) and \(G_0\). Also, the variations \( \delta \gamma \) at the boundary yields the insulated boundary conditions \( {\varvec{j}} _s \cdot {\varvec{n}}=0\) on \( \partial {\mathcal {B}}\). In the fluid momentum equations above, \( \pounds _ {{\varvec{u}}} \frac{\delta \ell }{\delta {\varvec{u}}}\) denotes the Lie derivative of the fluid momentum density \(\frac{\delta \ell }{\delta {\varvec{u}}}\) in the direction \({\varvec{u}}\), explicitly given as \(\pounds _ {{\varvec{u}}} {\varvec{m}} = {\varvec{u}} \cdot \nabla {\varvec{m}} + \nabla {\varvec{u}}^{\textsf{T}} {\varvec{m}} + {\varvec{m}} {\text {div}} {\varvec{u}}\).

By using the standard expression of the Lagrangian

$$\begin{aligned} \ell ({\varvec{u}}, \rho , s, b) = \int _{{{\mathcal {B}}}} \left[ \frac{1}{2} \rho |{\varvec{u}}|^2 - \rho e(\rho ,s/ \rho ,b)\right] \textrm{d}^3 {\varvec{x}}\,, \end{aligned}$$
(113)

which is the Eulerian form of (97), see [36], the equations of motion (114) give the following system of equations for a viscous and heat conducting continuum

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle \rho (\partial _t {\varvec{u}} + {\varvec{u}} \cdot \nabla {\varvec{u}}) = - \nabla p + {\text {div}} \varvec{\sigma }_{\textrm{el}}+ {\text {div}} \varvec{\sigma }\\ \displaystyle T({\bar{D}}_t s + {\text {div}} {\varvec{j}}_s) = \varvec{\sigma }: \nabla {\varvec{u}}- {\varvec{j}}_s \cdot \nabla T, \end{array} \right. \end{aligned}$$
(114)

with \(p= \rho ^2 \frac{\partial e}{\partial \rho } \) the pressure, \(T= \frac{\partial e}{\partial \eta } \) the temperature, and \( \varvec{\sigma } _{\textrm{el}}=2 \rho \frac{\partial e}{\partial b}\cdot b\) the elastic stress. We refer to [31, 33,34,35] for the statement of the variational formulation in both the material and spatial description as well as the detailed computations and several applications and extensions.

Remark A.2

(Structure of the variational formulation) The variational formulation (106)–(109) inherits from its Lagrangian counterpart (99)–(101) the same structure, in which the variational constraint (109) follows from the phenomenological constraint (107) by formally replacing the time rate of changes by \(\delta \)-variation, now in the Eulerian setting, see [33,34,35]. The same structure underlies the variational approach to porous media developed in Sect. 3.1.

Alternative expressions for heat and mass transfer coefficients involving cross-phenomena

We describe here possible cross-effects in the reversible processes, based on the form of the entropy production equation (30).

Regarding mass and heat transfer, cross-effects can occur for \(k=s,\ell =b\), while still preserving (38), which are the discrete analogues to the Soret and Dufour effects. We can thus consider

$$\begin{aligned} \begin{bmatrix} J_{sb} \frac{T_s-T_b}{T_s T_b} \\ q_{sb} \end{bmatrix} = {\mathcal {L}} _{sb} \begin{bmatrix} T_b - T_s \\ \frac{1}{T_s}\big ( \frac{1}{2} | {\varvec{u}}_s| ^2 - g_s \big )- \frac{1}{T_b}\big ( \frac{1}{2} | {\varvec{u}}_b| ^2 - g_b\big ) \end{bmatrix} \end{aligned}$$
(115)

with the symmetric part of the \(2 \times 2\) matrix \({\mathcal {L}} _{sb}\) positive, and with \(J_{sf}\) and \(J_{bf}\) satisfying (39) or, equivalently (41). We shall note that the conditions (42) and (115) for the mass and heat transfer coefficients seem to be novel and not encountered in previous literature on the subject. On the other hand, since \(q_{sb}\) can be interpreted physically as the rate of damage to the elastic matrix, these equations do have some aspects of the phenomenological equations for the evolution of damage parameters used previously in the literature, see, e.g., [48]. The equations for transfer rate q described there, in our notation, would be an affine function of the internal pressure p. Our expression has these pressure terms, but also involves additional terms related to the elastic energy, as we show below.

Besides the cross-phenomena between the scalar processes associated to mass and heat transfer, another cross-effect can be postulated, mathematically similar to the cross-phenomena between bulk viscosity and chemistry, see [22]. Ignoring the friction forces and assuming that the stresses \( \varvec{\sigma } _{k\ell }\) are symmetric, we can rewrite the entropy production expression (30) as

$$\begin{aligned}&\sum _{k<\ell }\Bigg [ \varvec{\sigma } _{k\ell }^\circ : \left( \frac{{\mathbb {F}} _k^\circ }{T_k}- \frac{{\mathbb {F}} _\ell ^\circ }{T_\ell } \right) + \frac{1}{3} {\text {Tr}}( \varvec{\sigma } _{k\ell }) \left( \frac{{\text {div}} {\textbf{u}} _k}{T_k} - \frac{{\text {div}} {\textbf{u}} _\ell }{T_\ell } \right) \\&\qquad + q_{k\ell }\left( \frac{\frac{1}{2} | {\varvec{u}}_k| ^2 - g_k}{T_k}-\frac{\frac{1}{2} | {\varvec{u}}_\ell | ^2 - g_\ell }{T_\ell } \right) + J_{k\ell }\left( \frac{1}{T_\ell } - \frac{1}{T_k}\right) (T_\ell -T_k)\Bigg ]\,. \end{aligned}$$

Therefore, one is naturally led to postulate the following cross-effects of scalar processes:

$$\begin{aligned} \left[ \begin{array}{c} \frac{1}{3} {\text {Tr}}( \varvec{\sigma } _{k\ell })\\ \displaystyle q_{k\ell }\\ J_{k\ell } \frac{T_k-T_\ell }{T_kT_\ell } \end{array} \right] = {\mathcal {L}}_{k\ell } \left[ \begin{array}{c} \frac{1}{T_k}{\text {div}} {\textbf{u}} _k - \frac{1}{T_\ell }{\text {div}} {\textbf{u}} _\ell \\ \frac{1}{T_k} (\frac{1}{2} | {\varvec{u}}_k| ^2 - g_k)-\frac{1}{T_\ell }(\frac{1}{2} | {\varvec{u}}_\ell | ^2 - g_\ell ) \\ \displaystyle T_\ell -T_k \end{array} \right] , \end{aligned}$$
(116)

where the symmetric part of the \(3 \times 3\) matrices \( {\mathcal {L}}_{k\ell } \) must be positive from the second law. We shall not endeavor to consider the more general conditions (115) and (116) involving the cross-effects, and only concentrate on the simpler conditions (40) as the one leading to the most simple mathematical expressions treatable analytically. In spite of the relative mathematical simplicity compared to (115) and (116), expressions (40) lead to physically relevant systems providing physically valid quantitative predictions.

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

Gay-Balmaz, F., Putkaradze, V. Thermodynamically consistent variational theory of porous media with a breaking component. Continuum Mech. Thermodyn. 36, 75–105 (2024). https://doi.org/10.1007/s00161-023-01262-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00161-023-01262-4

Keywords

Navigation