Skip to main content
Log in

Nonlinear Elasticity in a Deforming Ambient Space

  • Published:
Journal of Nonlinear Science Aims and scope Submit manuscript

Abstract

In this paper, we formulate a nonlinear elasticity theory in which the ambient space is evolving. For a continuum moving in an evolving ambient space, we model time dependency of the metric by a time-dependent embedding of the ambient space in a larger manifold with a fixed background metric. We derive both the tangential and the normal governing equations. We then reduce the standard energy balance written in the larger ambient space to that in the evolving ambient space. We consider quasi-static deformations of the ambient space and show that a quasi-static deformation of the ambient space results in stresses, in general. We linearize the nonlinear theory about a reference motion and show that variation of the spatial metric corresponds to an effective field of body forces.

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
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. The generalization of a theory obtained by relaxing certain standard assumptions (in this case, the staticity of \({\varvec{g}_t}\)), commonly results in a deeper understanding of the original theory. Examples of this include the geometric notions of stress and traction obtained by allowing the spatial metric to be non-Euclidean.

  2. It may be helpful to imagine the strip as moving between two tori of infinitesimally different sizes, so that the strip is constrained from both sides.

  3. Note that for a given t, such an isometric embedding always exists for \(\dim \mathcal {Q}\) large enough, by Nash (1956)’s embedding theorem.

  4. Recall that the order matters since \(\varvec{\omega }_{ij}^t = - \varvec{\omega }_{ji}^t\,\). See Appendix 1 for more details and the definitions of both the second and the normal fundamental forms.

  5. Note that although the Lagrangian theory is formulated with respect to \(\mathcal {Q}\,\), the density is defined with respect to the volume element of \(\mathcal {B}\,\), i.e., \(\mathcal {L}\) is an n-dimensional density, not an m-dimensional one.

  6. Another way to see this is by looking at the elastic energy as a function of the right Cauchy–Green tensor, i.e., \(W=\tilde{W}(\varvec{X},\tilde{\varvec{C}},{\varvec{G}})\,\). First, we see that since \(\varvec{g}_t:=\psi _t^*\varvec{h}\,\), then \(\left( \psi _t\circ \varphi _t\right) ^*\varvec{h} = \varphi _t^*\psi _t^*\varvec{h} = \varphi _t^*\varvec{g}_t\,\), i.e., the right Cauchy–Green tensors \(\varvec{C}\) of \(\varphi _t\) and \(\tilde{\varvec{C}}\) of \(\tilde{\varphi }_t\) are equal. If we denote \(\varvec{f}:=T\psi _t\,\), we write in components \({\tilde{C}}_{AB} = f^\alpha {}_aF^a{}_Af^\beta {}_bF^b{}_Bh_{\alpha \beta } = F^a{}_AF^b{}_Bf^\alpha {}_af^\beta {}_bh_{\alpha \beta } = F^a{}_AF^b{}_B{g}_{ab}=C_{AB}.\) Therefore, \(W=\tilde{W}(\varvec{X},\tilde{\varvec{C}},{\varvec{G}})=\tilde{W}(\varvec{X},{\varvec{C}},{\varvec{G}})\,\), that is, the elastic energy does not depend on the embedding \(\psi _t\,\).

  7. This computation is a generalization for a higher codimension of the acceleration computation in Sadik et al. (2016). We should mention that for this calculation we benefited from a discussion with Fabio Sozio.

  8. The Lie derivative along the vector field \(\varvec{\mathcal {V}}\) is defined as \(\varvec{L}_{\varvec{\mathcal {V}}}\varvec{\mathcal {V}}_\parallel = \left. \frac{{\hbox {d}}}{{\hbox {d}}t}\right| _{t=s} \left[ \left( \tilde{\varphi }_t \circ \tilde{\varphi }_s^{-1}\right) ^*\varvec{\mathcal {V}}_\parallel \right] \,\), where \(\tilde{\varphi }_t \circ \tilde{\varphi }_s^{-1}\) is the flow of \(\varvec{\mathcal {V}}\,\).

  9. For a function f defined on \(\mathcal {S}_t\,\), we write \({\tilde{d}}f = \sum _{\alpha =1}^n \frac{\partial f}{\partial \chi ^\alpha } d\chi ^\alpha \).

  10. Recall that the normal fundamental 1-forms are defined for \(i,j\in \{1,\ldots ,k\}\) as \(\varvec{\omega }^t_{ij} \cdot \varvec{w}= \left\langle \left\langle \nabla ^{\varvec{h}}_{\varvec{w}} \varvec{\eta }^t_i, \varvec{\eta }^t_j \right\rangle \right\rangle _{\varvec{h}}\) for any vector \(\varvec{w}\) tangent to \(\mathcal {S}_t\,\).

  11. Note that since the vector \(\varvec{U}\) is tangent to \(\mathcal {S}_t\) at \(\tilde{\varphi }(X,t)\,\), the vector \(\left[ \varvec{\mathcal {V}}, \varvec{U} \right] =\varvec{L}_{\varvec{\mathcal {V}}}\varvec{U}\) is tangent to \(\mathcal {S}_t\,\) as well.

  12. For a function f defined on \(\mathcal {S}\,\), we write d\(f = \sum _{a=1}^n \frac{\partial f}{\partial x^a} {\hbox {d}}x^a\).

  13. An alternate proof of this result for the special case of a transversal embedding is given in Appendix 2.

  14. We communicated with A. DeSimone and M. Arroyo, and they kindly confirmed the mistake in their acceleration. They indicated that they followed the master balance law of Marsden and Hughes (1983, p. 129). In Appendix 2, we show this derivation by using the master balance law and demonstrate that the results are identical to those we obtain using Hamilton’s principle.

  15. Note that the Jacobian of the deformation \(\tilde{\varphi }\) is equal to that of \(\varphi \,\), i.e., \(\sqrt{\frac{\det {\varvec{h}}}{\det {\varvec{G}}}}\det \tilde{\varvec{F}} = \sqrt{\frac{\det {\varvec{g}_t}}{\det {\varvec{G}}}}\det \varvec{F}\,\), which follows from \(\varvec{g}_t:=\psi _t^*\varvec{h}\,\).

  16. Note that there is a typo in the corresponding equation in Marsden and Hughes (1983), p. 92.

  17. Similar to the discussion of Remark 2.1, we can conclude that \(E(\varvec{X},\textsf {N},\psi _*\varvec{F},{\varvec{G}},\varvec{h})={E}(\varvec{X},\textsf {N},\varvec{F},{\varvec{G}},{\varvec{g}_t})\,\).

  18. An alternate proof for this result can be found in Marsden and Hughes (1983), p. 101.

  19. In the local coordinate \(\{\chi ^\alpha \}_{\alpha =1,\ldots ,n+k}\,\), we denote \(u_\perp ^i = u^{n+i}\) for \(i\in \{1,\ldots ,k\}\,\).

  20. The proof given in Spivak (1999) still holds even when the embedding is time dependent. Note that \(\nabla ^{{\varvec{g}_t}}\) and \(\nabla ^{\varvec{h}}\) are the Levi–Civita connections corresponding to \({\varvec{g}_t}\) and \(\varvec{h}\), respectively.

  21. Recall that in the chosen coordinate chart \(\{\chi ^\alpha \}_{\alpha =1,\ldots ,n+k}\,\), one has \(h_{\alpha (n+i)} = \left\langle \!\left\langle \tilde{\partial }^t_\alpha ,\varvec{\eta }^t_i\right\rangle \!\right\rangle _{\varvec{h}} = \delta _{\alpha (n+i)}\,\).

  22. Note that \(D_t^{\varvec{g}_t}\varvec{g}_t = \frac{\partial \varvec{g}_t}{\partial t}\,\).

  23. It is also possible to define an alternative covariant time derivative, \(\tilde{D}_t^{{\varvec{g}_t}}\,\), so that an identity analogous to (4.4) holds. If we let

    one readily verifies that

    $$\begin{aligned} \frac{{\hbox {d}}}{{\hbox {dt}}} \left\langle \!\left\langle \varvec{u},\varvec{w} \right\rangle \!\right\rangle _{{\varvec{g}_t}}=\left\langle \!\left\langle \tilde{D}_{t}^{{\varvec{g}_t}} \varvec{u},\varvec{w} \right\rangle \!\right\rangle _{{\varvec{g}_t}} + \left\langle \!\left\langle \varvec{u},\tilde{D}_{t}^{{\varvec{g}_t}}\varvec{w}\right\rangle \!\right\rangle _{{\varvec{g}_t}}. \end{aligned}$$

    See Thiffeault (2001) for a discussion on this alternative covariant time derivative.

  24. Note that if we were to use the alternative covariant derivative (4.6), this formula would need to be modified.

  25. The autonomous Lie derivative \(\mathfrak {L}_{\varvec{Z}}{\varvec{g}_t}\) is defined by holding the explicit time dependence of \(\varvec{g}_t\) fixed, i.e., \(\mathfrak {L}_{\varvec{Z}}{\varvec{g}_t} = \left. \frac{{\hbox {d}}}{{\hbox {d}}s}\right| _{t=s}\left[ \left( \psi _t \circ \psi _s^{-1} \right) ^* \varvec{g}_s \right] \,\).

  26. Note that what Marsden and Hughes (1983) denote by \(\varvec{v}\) is the equivalent of \(\varvec{\Upsilon }\) in our notation, so that their \(\varvec{v}_\parallel \) corresponds to \(\varvec{v}\) (recall that \(\varvec{z} = \psi _t^*\varvec{\zeta }_\parallel = \varvec{0}\,\)) and their \(v_n\) would be \(\zeta _\perp \) in the particular case when \(\mathcal {S}_t\) is a hypersurface in \(\mathcal {Q}\,\).

  27. Note that (5.6) is equivalent (2.22) since by Lemma 3.2, we have \(\frac{\partial \varvec{g}_t}{\partial t} = 2\sum _{i=1}^{k}\zeta _\perp ^i\varvec{k}_i^t\,\).

References

  • Anderson, J.L.: Principles of Relativity Physics. Academic Press, New York (1967)

    Google Scholar 

  • Arroyo, M., DeSimone, A.: Relaxation dynamics of fluid membranes. Phys. Rev. E 79, 031915 (2009)

    Article  MathSciNet  Google Scholar 

  • Capovilla, R., Guven, J.: Geometry of deformations of relativistic membranes. Phys. Rev. D 51(12), 6736–6743 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  • Chow, B., Peng, L., Lei, N.: Hamilton’s Ricci flow. In: Graduate Studies in Mathematics, vol. 77. American Mathematical Society, Providence, RI (2006)

    Google Scholar 

  • Ciarlet, P.G.: An Introduction to Differential Geometry with Applications to Elasticity. Springer, Heidelberg (2005)

    MATH  Google Scholar 

  • do Carmo, M.: Riemannian geometry. In: Kadison, R.V., Singer, I.M. (eds.) Mathematics: Theory & Applications. Birkhäuser Boston (1992). ISBN 1584883553

  • Doyle, T., Ericksen, J.: Nonlinear elasticity. Adv. Appl. Mech. 4, 53–115 (1956)

    Article  MathSciNet  Google Scholar 

  • Eshelby, J.D.: The determination of the elastic field of an ellipsoidal inclusion, and related problems. In: Proceedings of the Royal Society of London. Series A: Mathematical, Physical and Engineering Sciences, vol. 241, pp. 376–396. The Royal Society (1957)

  • Kuchař, K.: Geometry of hyperspace. I. J. Math. Phys. 17(5), 777–791 (1976)

    Article  MathSciNet  Google Scholar 

  • Lee, J.M.: Riemannian Manifold: An Introduction to Curvature. Springer, New York (1997)

    Book  Google Scholar 

  • Marsden, J.E., Hughes, T.J.R.: Mathematical foundations of elasticity. In: Dover Civil and Mechanical Engineering Series. Dover, New York (1983). ISBN 9780486678658

  • Marsden, J.E., Ratiu, T.: Introduction to Mechanics and Symmetry. Springer, New York (2003)

    Google Scholar 

  • Mazzucato, A.L., Rachele, L.V.: Partial uniqueness and obstruction to uniqueness in inverse problems for anisotropic elastic media. J. Elast. 83, 205–245 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  • Nash, J.: The imbedding problem for Riemannian manifolds. Ann. Math. 63(1), 20–63 (1956)

  • Nishikawa, S.: Variational Problems in Geometry, volume 205 of Iwanami series in modern mathematics. American Mathematical Society, Providence (2002). ISBN 9780821813560

  • Ogden, R.: Non-linear elastic deformations. In: Dover Civil and Mechanical Engineering. Dover Publications (1997). ISBN 9780486678658

  • Ozakin, A., Yavari, A.: A geometric theory of thermal stresses. J. Math. Phys. 51, 032902 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  • Post, E.J.: Formal Structure of Electromagnetics General Covariance and Electromagnetics. Dover, New York (1997)

    MATH  Google Scholar 

  • Sadik, S., Yavari, A.: Geometric nonlinear thermoelasticity and the time evolution of thermal stresses. Math. Mech. Solids (2015). doi:10.1177/1081286515599458

    Google Scholar 

  • Sadik, S., Angoshtari, A., Goriely, A., Yavari, A.: A geometric theory of nonlinear morphoelastic shells. J. Nonlinear Sci. (2016). ISSN 1432-1467. doi:10.1007/s00332-016-9294-9

  • Scriven, L.E.: Dynamics of a fluid interface—equation of motion for newtonian surface fluids. Chem. Eng. Sci. 12(2), 98–108 (1960)

    Article  Google Scholar 

  • Simo, J., Marsden, J.: Stress tensors, Riemannian metrics and the alternative descriptions in elasticity. In: Ciarlet, P.G., Roseau, M. (eds.) Trends and Applications of Pure Mathematics to Mechanics: Invited and Contributed Papers presented at a Symposium at Ecole Polytechnique, Palaiseau, France November 28 – December 2, 1983, pp. 369–383. Springer, Berlin Heidelberg (1984a)

  • Simo, J.C., Marsden, J.E.: On the rotated stress tensor and the material version of the Doyle–Ericksen formula. Arch. Ration. Mech. Anal. 86, 213–231 (1984b)

    Article  MathSciNet  MATH  Google Scholar 

  • Spivak, M.: A Comprehensive Introduction to Differential Geometry, vol. III. Publish or Perish, Houston (1999)

    MATH  Google Scholar 

  • Thiffeault, J.L.: Covariant time derivatives for dynamical systems. J. Phys. A Math. Gen. 34, 5875–5885 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  • Truesdell, C., Noll, W.: The Non-Linear Field Theories of Mechanics. Springer, Berlin (2004)

    Book  MATH  Google Scholar 

  • Yavari, A.: A geometric theory of growth mechanics. J. Nonlinear Sci. 20(6), 781–830 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  • Yavari, A., Goriely, A.: Riemann–Cartan geometry of nonlinear dislocation mechanics. Arch. Ration. Mech. Anal. 205(1), 59–118 (2012a)

    Article  MathSciNet  MATH  Google Scholar 

  • Yavari, A., Goriely, A.: Riemann–Cartan geometry of nonlinear disclination mechanics. Math. Mech. Solids 18(1), 91–102 (2012b)

    Article  MathSciNet  MATH  Google Scholar 

  • Yavari, A., Goriely, A.: Weyl geometry and the nonlinear mechanics of distributed point defects. In: Proceedings of the Royal Society of London. Series A: Mathematical, Physical and Engineering Sciences, pp. 3902–3922. The Royal Society of London (2012c)

  • Yavari, A., Goriely, A.: Nonlinear elastic inclusions in isotropic solids. In: Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences, vol. 469, p. 20130415. The Royal Society (2013)

  • Yavari, A., Goriely, A.: The geometry of discombinations and its applications to semi-inverse problems in anelasticity. In: Proceedings of the Royal Society of London. Series A: Mathematical, Physical and Engineering Sciences, p. 20140403. The Royal Society of London (2014)

  • Yavari, A., Ozakin, A.: Covariance in linearized elasticity. J. Appl. Math. Phys. 59(6), 1081–1110 (2008)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgments

AY and AO started thinking about this problem in 2009 and benefited from a discussion with the late Professor Jerrold E. Marsden. AY was partially supported by AFOSR—Grant No. FA9550-12-1-0290, and NSF—Grant No. CMMI 1130856. SS was supported by a Fulbright Grant.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Arash Yavari.

Additional information

Communicated by Irene Fonseca.

Appendices

Appendix 1: Geometry of Riemannian Submanifolds

In the following, we tersely review a few elements of the geometry of embedded submanifolds. Here we mainly follow do Carmo (1992), Capovilla and Guven (1995), Spivak (1999), and Kuchař (1976). Let us consider a Riemannian manifold \(\mathcal {S}\) embedded in another Riemannian manifold \(\mathcal {Q}\) and assume that \(\dim \mathcal {S}<\dim \mathcal {Q}\,\). We consider a time-dependent embedding \(\psi _t:\mathcal {S}\rightarrow \mathcal {Q}\,\). The metric \(\varvec{h}\) on \(\mathcal {Q}\) induces a metric \({\varvec{g}_t}=\psi _t^*\varvec{h}\) on \(\mathcal {S}\) (the first fundamental form). At any given point p of \(\mathcal {S}\,\), the tangent space \(T_{p}\mathcal {S}_t\) has an orthogonal complement \(\left( T_{p}\mathcal {S}_t\right) ^{\perp }\subset T_{}\mathcal {Q}\) such that

$$\begin{aligned} T_{p}\mathcal {Q} = T_{p}\mathcal {S}_t \oplus \left( T_{p}\mathcal {S}_t\right) ^{\perp }. \end{aligned}$$
(4.1)

Note that such a decomposition is smooth in the sense that any smooth vector field \(\varvec{u}\) on \(\mathcal {S}_t\) can be smoothly decomposed into a vector field \(\varvec{u}_\parallel \) tangent to \(\mathcal {S}_t\) and a vector field \(\varvec{u}_\perp \) normal to \(\mathcal {S}_t\,\), so that \(p\rightarrow (\varvec{u}_\parallel )_p= (\varvec{u}_p)_\parallel \) and \(p\rightarrow (\varvec{u}_\perp )_p = (\varvec{u}_p)_\perp \) are smooth. We write \(\varvec{u} = \varvec{u}_\parallel + \varvec{u}_\perp \,\). The orientation of \(\varvec{\eta }^t_i\), for \(i\in \{1,\ldots ,k\}\,\), is chosen such that the orientations of \(\mathcal {S}_t\) and \(\mathcal {Q}\) are consistent in the sense that the orientation induced from \(\mathcal {S}_t\) along with the ordered sequence \(\{\varvec{\eta }^t_i\}_{i\in \{1,\ldots ,k\}}\,\) is equivalent to the orientation of \(\mathcal {Q}\,\). Let \(\dim \mathcal {S} = n\) and \(\dim \mathcal {Q} = n + k = m\,\). Following the smoothness of the decomposition (4.1), one can take a set of smooth vector fields \(\{\varvec{\eta }^t_i\}_{i=1,\ldots ,k}\) normal to \(\mathcal {S}_t\) such that they form an orthonormal basis for \(\mathfrak {X}^\perp (\mathcal {S}_t)\,\), the set of vector fields normal to \(\mathcal {S}_t\,\). Let \(\{\chi ^\alpha \}_{\alpha =1,\ldots ,n+k}\) be a local coordinate chart for \(\mathcal {Q}\) such that at any point of \(\mathcal {S}_t, \{\chi ^1,\ldots ,\chi ^{n}\}\) is a local coordinate chart for \(\mathcal {S}_t\,\), and such that the ith unit normal vector field \(\varvec{\eta }^t_i\) for \(i\in \{1,\ldots ,k\}\) is tangent to the coordinate curve \(\chi ^{n+i}\,\). Hence, every vector field \(\varvec{u}\) on \(\mathcal {Q}\) along \(\mathcal {S}_t\) can be written as \(\varvec{u} = \varvec{u}_\parallel + \sum _{i=1}^k u_\perp ^i \varvec{\eta }^t_i\,\).Footnote 19 Note that, for \(i,j\in \{1,\ldots ,k\}\,\), one has \(\left\langle \!\left\langle \varvec{\eta }^t_i,\varvec{\eta }^t_j \right\rangle \!\right\rangle _{\varvec{h}} = \delta _{ij}\) and \(\left\langle \!\left\langle \varvec{\eta }^t_i,\varvec{u}_\parallel \right\rangle \!\right\rangle _{\varvec{h}} = 0\,\), where the Kronecker delta symbol \(\delta _{ij}\) is defined as: \(\delta _{ij}=1\) if \(i=j\) and \(\delta _{ij}=0\) if \(i\ne j\). Note that at any point of \(\mathcal {S}_t\,\), one has \(h_{\alpha (n+i)} = \delta _{\alpha (n+i)}\,\), for \(i\in \{1,\ldots ,k\}\) and \(\alpha \in \{1,\ldots ,n+k\}\,\). We denote the connection coefficients for the Levi–Civita connections \(\nabla ^{\varvec{h}}\) and \(\nabla ^{{\varvec{g}_t}}\) corresponding to the metrics \(\varvec{h}\) and \(\varvec{g}_t\) by \(\tilde{\gamma }^{\alpha }_{\beta \gamma }\) and \(\gamma ^a_{bc}\,\), respectively. We denote by \(D^{\varvec{h}}_t\) and \(D^{\varvec{g}_t}_t\) the covariant derivatives along \(\tilde{\varphi }_X\) and \(\varphi _X\,\), respectively. For a vector field \(\varvec{u}\) on \(\mathcal {Q}\) along \(\mathcal {S}_t\,\), we write \(D^{\varvec{h}}_t \varvec{u} = \frac{\partial u^\alpha }{\partial t}\tilde{\partial }^t_\alpha + \frac{\partial u_\perp ^i}{\partial t}\varvec{\eta }^t_i+ \nabla ^{\varvec{h}}_{\varvec{\Upsilon }}\varvec{u}\,\) and for a vector field \(\varvec{w}\) on \(\mathcal {S}, D^{\varvec{g}_t}_t \varvec{w} = \frac{\partial w^a}{\partial t}\partial ^t_a + \nabla ^{\varvec{g}_t}_{\varvec{V}}\varvec{w}\,\), where \(\{\tilde{\partial }^t_\alpha \}_{\alpha =1,\ldots ,n}\) and \(\{\partial ^t_a\}_{a=1,\ldots ,n}\) denote local coordinate bases for \(\mathcal {S}_t\) and \(\mathcal {S}\,\), respectively.

Note that for vector fields \(\varvec{X}\) and \(\varvec{Y}\) defined on \(\mathcal {S}_t\) and \(\mathcal {Q}\,\), respectively, such that \(\varvec{Y}\) is everywhere tangent to \(\mathcal {S}_t, \nabla ^{{\varvec{g}_t}}_{\psi _t^*\varvec{X}}\psi _t^*\varvec{Y} =\psi _t^* (\nabla ^{\varvec{h}}_{\varvec{X}}\varvec{Y})_{\parallel }\).Footnote 20 As a corollary, given a curve c in \(\mathcal {S}_t\) and \(\varvec{X}\) a vector field along c tangent to \(\mathcal {S}_t\) everywhere, \(D^{{\varvec{g}_t}}_c\psi _t^*\varvec{X}=(D^{\varvec{h}}_c\varvec{X})_{\parallel }\,\). For \(i\in \{1,\ldots ,k\}\,\), the ith second fundamental form of \(\mathcal {S}_t\) along \(\varvec{\eta }^t_i\) is a \(\left( {\begin{array}{c}0\\ 2\end{array}}\right) \)-tensor \(\varvec{\kappa }_i^t\) on \(\mathcal {S}_t\) defined as (do Carmo 1992; Capovilla and Guven 1995)

$$\begin{aligned} \varvec{\kappa }_i^t(\varvec{u},\varvec{w}) =\left\langle \!\left\langle \nabla ^{\varvec{h}}_{\varvec{u}}\varvec{\eta }^t_i,\varvec{w} \right\rangle \!\right\rangle _{\varvec{h}},\quad \forall ~\varvec{u},\varvec{w}\in T_{\chi }\mathcal {S}_t. \end{aligned}$$
(4.2)

It is known that \(\varvec{\kappa }_i^t\) is a symmetric tensor and can equivalently be written as

$$\begin{aligned} \varvec{\kappa }_i^t=(\nabla ^{\varvec{h}}\varvec{\eta }^t_i)^{\flat }_\parallel ,\quad i=1,\ldots ,k. \end{aligned}$$

On \(\mathcal {S}\,\), we define, for \(i\in \{1,\ldots ,k\}\,\), the ith second fundamental form as \(\varvec{k}_i^t=\psi _t^*\varvec{\kappa }_i^t\,\). For vector fields \(\varvec{u},\varvec{w}\in T_{\varvec{x}}\mathcal {S}\), one can write \(\nabla _{\psi _*\varvec{u}}^{\varvec{h}}\psi _*\varvec{w} = \psi _*\nabla _{\varvec{u}}^{{\varvec{g}_t}}\varvec{w} + \sum _{i=1}^{k}h^i(\varvec{u},\varvec{w})\varvec{\eta }^t_i\,\), where \(h^i(.,.)\) is a bilinear form. Therefore

$$\begin{aligned} h^i(\varvec{u},\varvec{w})=\left\langle \!\left\langle \nabla _{\psi _*\varvec{u}}^{\varvec{h}}\psi _*\varvec{w}, \varvec{\eta }^t_i\right\rangle \!\right\rangle _{\varvec{h}},\quad i=1,\ldots ,k. \end{aligned}$$

Knowing that \(\left\langle \!\left\langle \psi _*\varvec{w},\varvec{\eta }^t_i\right\rangle \!\right\rangle _{\varvec{h}}=0\), one concludes that

$$\begin{aligned} \left\langle \!\left\langle \nabla _{\psi _*\varvec{u}}^{\varvec{h}}\psi _*\varvec{w}, \varvec{\eta }^t_i\right\rangle \!\right\rangle _{\varvec{h}} =-\left\langle \!\left\langle \nabla _{\psi _*\varvec{u}}^{\varvec{h}}\varvec{\eta }^t_i, \psi _*\varvec{w}, \right\rangle \!\right\rangle _{\varvec{h}},\quad i=1,\ldots ,k. \end{aligned}$$

Hence

$$\begin{aligned} h^i(\varvec{u},\varvec{w})= & {} -\left\langle \!\left\langle \nabla _{\psi _*\varvec{u}}^{\varvec{h}}\varvec{\eta }^t_i, \psi _*\varvec{w}, \right\rangle \!\right\rangle _{\varvec{h}}=-(\nabla ^{\varvec{h}}\varvec{\eta }^t_i)^{\flat }(\psi _*\varvec{u},\psi _*\varvec{w}) =-\varvec{k}_i^t(\varvec{u},\varvec{w}),\\ i= & {} 1,\ldots ,k. \end{aligned}$$

Therefore, we obtain Gauss’s equation

$$\begin{aligned} \nabla _{\psi _*\varvec{u}}^{\varvec{h}}\psi _*\varvec{w}=\psi _*\nabla _{\varvec{u}}^{{\varvec{g}_t}}\varvec{w} -\sum _{i=1}^{k}\varvec{k}_i^t(\varvec{u},\varvec{w})\varvec{\eta }^t_i. \end{aligned}$$

On the other hand, for \(i,j\in \{1,\ldots ,k\}\,\), the projection of \(\nabla ^{\varvec{h}}\varvec{\eta }^t_i\) along \(\varvec{\eta }^t_j\) defines \(\varvec{\omega }_{ij}^t\,\), the normal fundamental 1-form of \(\mathcal {S}_t\) relative to the unit normals \(\varvec{\eta }^t_i\) and \(\varvec{\eta }^t_j\,\). For any vector \(\varvec{w}\) tangent to \(\mathcal {S}_t\,\), the 1-form \(\varvec{\omega }_{ij}^t\) is defined by (Capovilla and Guven 1995)

$$\begin{aligned} \varvec{\omega }_{ij}^t \cdot \varvec{w}= \left\langle \!\left\langle \nabla ^{\varvec{h}}_{\varvec{w}} \varvec{\eta }^t_i, \varvec{\eta }^t_j \right\rangle \!\right\rangle _{\varvec{h}}. \end{aligned}$$

Note that, for \(i,j\in \{1,\ldots ,k\}\), the normal fundamental 1-form \(\varvec{\omega }_{ij}^t\) is such that \(\varvec{\omega }_{ij}^t = -\varvec{\omega }_{ji}^t\). On \(\mathcal {S}\,\), one defines the normal fundamental 1-forms, for \(i,j\in \{1,\ldots ,k\}\,\), as \(\varvec{o}_{ij}^t=\psi _t^*\varvec{\omega }_{ij}^t\,\). Note that, for a tangent vector field \(\varvec{w}\) on \(\mathcal {S}_t\), one can write the followingFootnote 21

$$\begin{aligned} \nabla ^{\varvec{h}}_{\varvec{w}}\varvec{\eta }^t_i = \varvec{h}^\sharp \cdot \varvec{\kappa }_i^t \cdot \varvec{w} + \sum _{j=1}^k \left( \varvec{\omega }_{ij}^t \cdot \varvec{w}\right) \varvec{\eta }^t_j. \end{aligned}$$
(4.3)

One needs to be careful in calculating time derivatives in \(\left( \mathcal {S},\varvec{g}_t\right) \,\), since the induced metric \({\varvec{g}_t}\) itself depends on time. In particular, when calculating the derivative of the inner product \(\left\langle \!\left\langle \varvec{u},\varvec{w}\right\rangle \!\right\rangle _{{\varvec{g}_t}}\) of two vector fields \(\varvec{u}\) and \(\varvec{w}\) along a time-parametrized curve \(c\,\), the usual formula

$$\begin{aligned} \frac{{\hbox {d}}}{{\hbox {d}}t} \left\langle \!\left\langle \varvec{u},\varvec{w} \right\rangle \!\right\rangle _{{\varvec{g}_t}} = \left\langle \!\left\langle D_{t}^{{\varvec{g}_t}} \varvec{u},\varvec{w} \right\rangle \!\right\rangle _{{\varvec{g}_t}} + \left\langle \!\left\langle \varvec{u},D_{t}^{{\varvec{g}_t}}\varvec{w}\right\rangle \!\right\rangle _{{\varvec{g}_t}}, \end{aligned}$$
(4.4)

is no longer valid when the metric \({\varvec{g}_t}\) is t dependent. One instead hasFootnote 22

$$\begin{aligned} \frac{{\hbox {d}}}{{\hbox {dt}}} \left\langle \!\left\langle \varvec{u},\varvec{w} \right\rangle \!\right\rangle _{{\varvec{g}_t}}=\left\langle \!\left\langle D_{t}^{{\varvec{g}_t}} \varvec{u},\varvec{w} \right\rangle \!\right\rangle _{{\varvec{g}_t}} + \left\langle \!\left\langle \varvec{u},D_{t}^{{\varvec{g}_t}}\varvec{w}\right\rangle \!\right\rangle _{{\varvec{g}_t}} + \left\langle \!\left\langle \varvec{u},\varvec{w} \right\rangle \!\right\rangle _{\frac{\partial \varvec{g}_t}{\partial t}}, \end{aligned}$$
(4.5)

where

$$\begin{aligned} \left\langle \!\left\langle \varvec{u},\varvec{w}\right\rangle \!\right\rangle _{\frac{\partial \varvec{g}_t}{\partial t}} = u^av^b \frac{\partial {g_t}_{ab}}{\partial t}. \end{aligned}$$

This can be written in terms of the inner product with respect to \({\varvec{g}_t}\) as

$$\begin{aligned} \left\langle \!\left\langle \varvec{u},\varvec{w}\right\rangle \!\right\rangle _{\frac{\partial \varvec{g}_t}{\partial t}} = \left\langle \!\left\langle \varvec{u}, {\varvec{g}_t^\sharp }\cdot \frac{\partial {\varvec{g}}}{\partial t}\cdot \varvec{w} \right\rangle \!\right\rangle _{{\varvec{g}_t}}, \end{aligned}$$

where \({\varvec{g}_t^\sharp }\) denotes the “inverse metric,” with components \(g_t^{ab}\,\). ThereforeFootnote 23

$$\begin{aligned} \frac{{\hbox {d}}}{{\hbox {d}}t} \left\langle \!\left\langle \varvec{u},\varvec{w} \right\rangle \!\right\rangle _{{\varvec{g}_t}}=\left\langle \!\left\langle D_{t}^{{\varvec{g}_t}} \varvec{u},\varvec{w} \right\rangle \!\right\rangle _{{\varvec{g}_t}} + \left\langle \!\left\langle \varvec{u},D_{t}^{{\varvec{g}_t}}\varvec{w}\right\rangle \!\right\rangle _{{\varvec{g}_t}} + \left\langle \!\left\langle \varvec{u},{\varvec{g}_t^\sharp }\cdot \frac{\partial \varvec{g}_t}{\partial t}\cdot \varvec{w} \right\rangle \!\right\rangle _{{\varvec{g}_t}}. \end{aligned}$$
(4.7)

Using the Levi–Civita connection for the metric \({\varvec{g}_t}\) to calculate covariant derivatives, the symmetry lemma of classical Riemann geometry (Lee 1997; Nishikawa 2002) still holds.Footnote 24

Lemma 3.1

For a Riemannian manifold with a time-dependent metric \({\varvec{g}_t}\),

$$\begin{aligned} D_{\epsilon }^{{\varvec{g}_t}} \frac{\partial c(t,\epsilon )}{\partial t} = D_{t}^{{\varvec{g}_t}} \frac{\partial c(t,\epsilon )}{\partial \epsilon }. \end{aligned}$$

The velocity of the time-dependent embedding \(\psi _t\) is defined as

$$\begin{aligned} \varvec{\zeta }=\frac{\partial \psi (t,\varvec{x})}{\partial t} = \varvec{\zeta }_{\parallel } + \sum _{i=1}^{k}\zeta _\perp ^i\varvec{\eta }^t_i, \end{aligned}$$

where \(\varvec{\zeta }_{\parallel }\) is the tangential velocity of the embedding. We also define \(\varvec{Z} := \psi _{t}^*\varvec{\zeta }_\parallel \circ \varphi _t\,\).

Lemma 3.2

For an arbitrary embedding \(\psi _t\,\), the following relation holds

$$\begin{aligned} \frac{\partial \varvec{g}_t}{\partial t} = \mathfrak {L}_{\varvec{Z}}{\varvec{g}_t} + 2\sum _{i=1}^{k}\zeta _\perp ^i\varvec{k}_i^t, \end{aligned}$$
(4.8)

where \(\mathfrak {L}\) denotes the autonomous Lie derivative.Footnote 25 For a transversal embedding, i.e., when \(\varvec{Z}=\varvec{0}\,\), (4.8) reduces to

$$\begin{aligned} \frac{\partial \varvec{g}_t}{\partial t} = 2\sum _{i=1}^{k}\zeta _\perp ^i\varvec{k}_i^t. \end{aligned}$$
(4.9)

Proof

First, we note that

$$\begin{aligned} \varvec{L}_{\varvec{\zeta }}\varvec{h}= & {} \left[ \frac{{\hbox {d}}}{{\hbox {d}}t}\left( \psi _{t} \circ \psi _{s}^{-1}\right) ^* \varvec{h} \right] _{s=t} = \left[ \frac{{\hbox {d}}}{{\hbox {dt}}} \psi _{s*}\psi _t^* \varvec{h} \right] _{s=t} = \left[ \frac{{\hbox {d}}}{{\hbox {dt}}} \psi _{s*}\varvec{g}_t \right] _{s=t}\nonumber \\= & {} \psi _{t*} \left[ D_t^{\varvec{g}_t} \varvec{g}_t \right] _{s=t} = \psi _{t*}\frac{\partial \varvec{g}_t}{\partial t}. \end{aligned}$$
(4.10)

On the other hand, we also have

$$\begin{aligned} \varvec{L}_{\varvec{\zeta }}\varvec{h} = \mathfrak {L}_{\varvec{\zeta }}\varvec{h} = \mathfrak {L}_{\varvec{\zeta }_\parallel }\varvec{h} + \sum _{i=1}^{k}\zeta _\perp ^i \mathfrak {L}_{\varvec{\eta }^t_i}\varvec{h}. \end{aligned}$$

However, for \(i\in \{1,\ldots ,k\}\) one has

$$\begin{aligned} \left( \mathfrak {L}_{\varvec{\eta }^t_i}\varvec{h} \right) _{\alpha \beta } = (\varvec{\eta }^t_i)_{\alpha |\beta } + (\varvec{\eta }^t_i)_{\beta |\alpha } = 2\kappa _{(i)\alpha \beta }. \end{aligned}$$
(4.11)

We observe that \(\mathfrak {L}_{\varvec{\zeta }_\parallel }\varvec{h} = \mathfrak {L}_{\psi _{t*}\varvec{Z}}\psi _{t*}\varvec{g}_t\,\), and following (Marsden and Hughes 1983, p. 98), we have \(\mathfrak {L}_{\psi _{t*}\varvec{Z}}\psi _{t*}\varvec{g}_t = \psi _{t*}\mathfrak {L}_{\varvec{Z}}\varvec{g}_t\,\). Thus

$$\begin{aligned} \varvec{L}_{\varvec{\zeta }}\varvec{h} = \psi _{t*}\left( \mathfrak {L}_{\varvec{Z}}\varvec{g}_t + 2\sum _{i=1}^{k}\zeta _\perp ^i\varvec{k}_i^t\right) . \end{aligned}$$
(4.12)

Finally, it follows from (4.10) and (4.12) that

$$\begin{aligned} \frac{\partial \varvec{g}_t}{\partial t} = \mathfrak {L}_{\varvec{Z}}{\varvec{g}_t} + 2\sum _{i=1}^{k}\zeta _\perp ^i\varvec{k}_i^t. \end{aligned}$$

\(\square \)

Appendix 2: An Alternative Derivation of the Tangent Balance of Linear Momentum

In this section, we provide an alternate proof for the tangential balance of linear momentum in the particular case of a transversal evolution of the ambient space. This derivation is a generalized version, for arbitrary co-dimension \(k = \dim \mathcal {Q} - \dim \mathcal {S}_t\), of a theorem appearing in Marsden and Hughes (1983), p. 129. The generalized version can be stated as follows (see § 2.1 and Fig. 2 to recall the notation):

Theorem 3.1

Assume that given scalar functions a and b, and a vector field \(\varvec{c}\) satisfy the following master balance law for any open set \(\mathcal {U}\) with \(C^1\) piecewise boundary:

$$\begin{aligned} \frac{{\hbox {d}}}{{\hbox {dt}}}\int _{\varphi _t(\mathcal {U})}a{\hbox {d}}v = \int _{\varphi _t(\mathcal {U})}b{\hbox {d}}v + \int _{\partial \varphi _t(\mathcal {U})}\left\langle \!\left\langle \varvec{c} , \varvec{\mathsf n} \right\rangle \!\right\rangle _{\varvec{g}_t} da, \end{aligned}$$
(5.1)

where \(\varvec{\mathsf n}\) is the unit normal vector to \(\partial \varphi _t\left( \mathcal {U}\right) \) in \(\mathcal {S}\,\). Localization of (5.1) gives one

$$\begin{aligned} \frac{{\hbox {d}}a}{{\hbox {d}}t} + a\, \mathrm{div}_{\varvec{g}_t}\varvec{v} + a \sum _{i=1}^{k}\zeta _\perp ^i \mathrm{tr}\left( \varvec{k}_i^t\right) = b + \mathrm{div}_{\varvec{g}_t} \varvec{c}, \end{aligned}$$
(5.2)

where we recall that \(\varvec{v}\) is the velocity field of \(\varphi _t\, and \varvec{\zeta } = \sum _{i=1}^{k}\zeta _\perp ^i\varvec{\eta }^t_i\) is the velocity field of \(\psi _t\,\) (we have \(\varvec{\zeta }_\parallel =\varvec{0}\) since we are assuming transversal evolution).Footnote 26

Proof

Note that

$$\begin{aligned} \frac{{\hbox {d}}}{{\hbox {d}}t}\int _{\varphi _t(\mathcal {U})}a{\hbox {d}}v = \frac{{\hbox {d}}}{{\hbox {d}}t}\int _{\mathcal {U}}aJ{\hbox {d}}V = \int _{\mathcal {U}}\frac{{\hbox {d}}}{{\hbox {d}}t}(aJ){\hbox {d}}V = \int _{\mathcal {U}}\left( \frac{{\hbox {d}}a}{{\hbox {dt}}}J+a\frac{{\hbox {d}}J}{{\hbox {d}}t}\right) {\hbox {d}}V. \end{aligned}$$

However

$$\begin{aligned} \frac{{\hbox {d}}J}{{\hbox {d}}t} =\left( {\text {div}}_{{\varvec{g}_t}}\varvec{v}\right) J + \frac{1}{2}J{\text {tr}}\left( \frac{\partial \varvec{g}_t}{\partial t}\right) , \end{aligned}$$

and following (4.8), one has

$$\begin{aligned} \frac{\partial \varvec{g}_t}{\partial t} = 2\sum _{i=1}^{k}\zeta _\perp ^i\varvec{k}_i^t. \end{aligned}$$

Therefore

$$\begin{aligned} \begin{aligned} \frac{{\hbox {d}}}{{\hbox {d}}t}\int _{\varphi _t(\mathcal {U})}a{\hbox {d}}v&= \int _{\mathcal {U}}\left( \frac{{\hbox {da}}}{{\hbox {d}}t}+a\, \mathrm{div}_{\varvec{g}_t}\varvec{v} + a \sum _{i=1}^{k}\zeta _\perp ^i \mathrm{tr}\left( \varvec{k}_i^t\right) \right) J{\hbox {d}}V \\ {}&= \int _{\varphi _t(\mathcal {U})}\left( \frac{{\hbox {da}}}{{\hbox {d}}t}+a\, \mathrm{div}_{\varvec{g}_t}\varvec{v} + a \sum _{i=1}^{k}\zeta _\perp ^i \mathrm{tr}\left( \varvec{k}_i^t\right) \right) {\hbox {d}}v. \end{aligned} \end{aligned}$$
(5.3)

On the other hand, by using Stokes’ theorem, one can write

$$\begin{aligned} \int _{\partial \varphi _t(\mathcal {U})}\left\langle \!\left\langle \varvec{c} , \varvec{\mathsf n} \right\rangle \!\right\rangle _{\varvec{g}_t} da = \int _{\varphi _t(\mathcal {U})}\mathrm{div}_{\varvec{g}_t} \varvec{c}\, {\hbox {d}}v. \end{aligned}$$
(5.4)

Therefore, by using (5.3) and (5.4), (5.1) transforms to

$$\begin{aligned} \int _{\varphi _t(\mathcal {U})}\left( \frac{{\hbox {d}}a}{{\hbox {d}}t}+a\, \mathrm{div}_{\varvec{g}_t}\varvec{v} + a \sum _{i=1}^{k}\zeta _\perp ^i \mathrm{tr}\left( \varvec{k}_i^t\right) \right) {\hbox {d}}v = \int _{\varphi _t(\mathcal {U})}\left( b + \mathrm{div}_{\varvec{g}_t} \varvec{c}\right) {\hbox {d}}v. \end{aligned}$$

Thus, by arbitrariness of the subset \(\mathcal {U}\,\), one finds that

$$\begin{aligned} \frac{{\hbox {d}}a}{{\hbox {d}}t} + a\, \mathrm{div}_{\varvec{g}_t}\varvec{v} + a \sum _{i=1}^{k}\zeta _\perp ^i \mathrm{tr}\left( \varvec{k}_i^t\right) = b + \mathrm{div}_{\varvec{g}_t} \varvec{c}. \end{aligned}$$
(5.5)

\(\square \)

First, the localized conservation of mass is derived. In integral form, conservation of mass reads

$$\begin{aligned} \frac{{\hbox {d}}}{{\hbox {d}}t}\int _{\varphi _t(\mathcal {U})} \rho {\hbox {d}}v = 0. \end{aligned}$$

Hence, using the above theorem (\(a=\rho , b=0\,\), and \(c=0\)), the conservation of mass in localized form readsFootnote 27

$$\begin{aligned} \frac{{\hbox {d}}\rho }{{\hbox {d}}t} + \rho \, \mathrm{div}_{\varvec{g}_t}\varvec{v} + \rho \sum _{i=1}^{k} \zeta _\perp ^i \mathrm{tr}\left( \varvec{k}_i^t\right) = 0. \end{aligned}$$
(5.6)

Next, we look at the balance of linear momentum, which in integral form reads

$$\begin{aligned} \frac{{\hbox {d}}}{{\hbox {d}}t}\int _{\varphi _t(\mathcal {U})}\rho \varvec{v} {\hbox {d}}v = \int _{\varphi _t(\mathcal {U})} \rho \varvec{B} {\hbox {d}}v + \int _{\partial \varphi _t(\mathcal {U})} \varvec{\sigma } \cdot \varvec{\mathsf n}{\hbox {d}}a, \end{aligned}$$
(5.7)

where \(\varvec{B}\) is the body force per unit mass, \(\varvec{\sigma }\) is the Cauchy stress tensor, and \(\varvec{\mathsf n}\) is the unit normal vector to \(\partial \varphi _t\left( \mathcal {U}\right) \,\).

Remark 3.3

Note that (5.7) makes sense only when the ambient space \(\mathcal {S}_t\) is endowed with a linear structure. In a general manifold, integrating a vector field does not make sense. Therefore, the proof given in this appendix is only valid for linear ambient spaces. However, the resulting localized tangential balance of linear momentum (5.8) still holds in the case of a general manifold as shown in § 2.1 using a Lagrangian field theory, see Eq. (2.17).

In order to use the above theorem, we contract the balance of linear momentum (5.7) with an arbitrary time-independent covariantly constant vector field \(\varvec{u}\) tangent to \(\varphi _t\left( \mathcal {U}\right) \,\), i.e.

$$\begin{aligned} \frac{{\hbox {d}}}{{\hbox {d}}t}\int _{\varphi _t(\mathcal {U})} \left\langle \!\left\langle \rho \varvec{v}, \varvec{u}\right\rangle \!\right\rangle _{\varvec{h}} {\hbox {d}}v = \int _{\varphi _t(\mathcal {U})} \left\langle \!\left\langle \rho \varvec{B}, \varvec{u}\right\rangle \!\right\rangle _{\varvec{h}} {\hbox {d}}v + \int _{\partial \varphi _t(\mathcal {U})} \left\langle \!\left\langle \varvec{\sigma } \cdot \varvec{\mathsf n}, \varvec{u}\right\rangle \!\right\rangle _{\varvec{h}}{\hbox {d}}a. \end{aligned}$$

We can then use the above theorem for \(a = \left\langle \!\left\langle \rho \varvec{v}, \varvec{u}\right\rangle \!\right\rangle _{\varvec{h}}, b = \left\langle \!\left\langle \rho \varvec{B}, \varvec{u}\right\rangle \!\right\rangle _{\varvec{h}}\,\), and \(\varvec{c} = \varvec{\sigma } \cdot \varvec{u}\,\). Hence, it follows that

$$\begin{aligned} \frac{{\hbox {d}}}{{\hbox {d}}t}\left\langle \!\left\langle \rho \varvec{v}, \varvec{u}\right\rangle \!\right\rangle _{\varvec{h}} + \left\langle \!\left\langle \rho \varvec{v}, \varvec{u}\right\rangle \!\right\rangle _{\varvec{h}} \mathrm{div}_{\varvec{g}_t}\varvec{v} + \left\langle \!\left\langle \rho \varvec{v}, \varvec{u}\right\rangle \!\right\rangle _{\varvec{h}} v_n \mathrm{tr}\varvec{k} = \left\langle \!\left\langle \rho \varvec{B}, \varvec{u}\right\rangle \!\right\rangle _{\varvec{h}} + \mathrm{div}_{\varvec{g}_t}\left( \varvec{\sigma } \cdot \varvec{u}\right) . \end{aligned}$$

Note that

$$\begin{aligned} \frac{{\hbox {d}}}{{\hbox {d}}t}\left\langle \!\left\langle \rho \varvec{v}, \varvec{u}\right\rangle \!\right\rangle _{\varvec{h}} = \frac{{\hbox {d}}\rho }{{\hbox {d}}t}\left\langle \!\left\langle \varvec{v}, \varvec{u}\right\rangle \!\right\rangle _{\varvec{h}} +\rho \left\langle \!\left\langle D_t^{\varvec{h}}\varvec{v}, \varvec{u}\right\rangle \!\right\rangle _{\varvec{h}}, \end{aligned}$$

where \(D_t^{\varvec{h}}\) denotes the covariant time derivative with respect to the metric \({\varvec{h}}\,\). Therefore, it follows that

$$\begin{aligned} \left( \frac{{\hbox {d}}\rho }{{\hbox {d}}t} + \rho \mathrm{div}_{\varvec{g}_t}\varvec{v} + \rho v_n \mathrm{tr}\varvec{k} \right) \left\langle \!\left\langle \varvec{v}, \varvec{u}\right\rangle \!\right\rangle _{\varvec{h}} +\rho \left\langle \!\left\langle D_t^{\varvec{h}}\varvec{v}, \varvec{u}\right\rangle \!\right\rangle _{\varvec{h}} = \left\langle \!\left\langle \rho \varvec{B}, \varvec{u}\right\rangle \!\right\rangle _{\varvec{h}} + \mathrm{div}_{\varvec{g}_t}\left( \varvec{\sigma } \cdot \varvec{u}\right) . \end{aligned}$$

The first term vanishes following the conservation of mass (5.6). Thus, by arbitrariness of \(\varvec{u}\) one concludes that

$$\begin{aligned} \rho \left( D_t^{\varvec{h}}\varvec{v}\right) _\parallel = \rho \varvec{B} + \mathrm{div}_{\varvec{g}_t}\varvec{\sigma }. \end{aligned}$$
(5.8)

Note that \(\left( D_t^{\varvec{h}}\varvec{v}\right) _\parallel \) is different from \(D_t^{\varvec{g}}\varvec{v}\,\). In fact, we can write following Proposition 2.1 that

$$\begin{aligned} \begin{aligned} \left( D_t^{\varvec{h}}\varvec{v}\right) _\parallel&= D^{\varvec{g}_t}_t \varvec{v} + 2\sum _{i=1}^k \zeta _\perp ^i \varvec{g}_t^\sharp \cdot \varvec{k}^t_i \cdot \varvec{v} - \sum _{i=1}^k \zeta _\perp ^i \left( {\hbox {d}}\zeta _\perp ^i\right) ^\sharp - \sum _{i,j=1}^k \zeta _\perp ^i\zeta _\perp ^j \varvec{o}^{t\sharp }_{ij} \\ {}&= D^{\varvec{g}_t}_t \varvec{v} + \varvec{g}_t^\sharp \cdot \frac{\partial \varvec{g}_t}{\partial t} \cdot \varvec{v} - \sum _{i=1}^k \zeta _\perp ^i \left( {\hbox {d}}\zeta _\perp ^i\right) ^\sharp - \sum _{i,j=1}^k \zeta _\perp ^i\zeta _\perp ^j \varvec{o}^{t\sharp }_{ij}. \end{aligned} \end{aligned}$$
(5.9)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yavari, A., Ozakin, A. & Sadik, S. Nonlinear Elasticity in a Deforming Ambient Space. J Nonlinear Sci 26, 1651–1692 (2016). https://doi.org/10.1007/s00332-016-9315-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00332-016-9315-8

Keywords

Mathematics Subject Classification

Navigation