Skip to main content
Log in

Crack Tip Equation of Motion in Dynamic Gradient Damage Models

  • Published:
Journal of Elasticity Aims and scope Submit manuscript

Abstract

We propose in this contribution to investigate the link between the dynamic gradient damage model and the classical Griffith’s theory of dynamic fracture during the crack propagation phase. To achieve this main objective, we first rigorously reformulate two-dimensional linear elastic dynamic fracture problems using variational methods and shape derivative techniques. The classical equation of motion governing a smoothly propagating crack tip follows by considering variations of a space-time action integral. We then give a variationally consistent framework of the dynamic gradient damage model. Owing to the analogies between the variational ingredients of these two models and under some basic assumptions concerning the damage band structuration, one obtains a generalized Griffith criterion which governs the crack tip evolution within the non-local damage model. Assuming further that the internal length is small compared to the dimension of the body, the previous criterion leads to the classical Griffith’s law through a separation of scales between the outer linear elastic domain and the inner damage process zone.

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
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

Notes

  1. But in practice it is the initial displacement \(\mathbf{u}_{0}\) and velocity \(\dot{\mathbf{u}}_{0}\) that are known and we will use the equations derived from the Hamiltonian principle to solve the physical Initial Value Problem.

References

  1. Abdelmoula, R., Debruyne, G.: Modal analysis of the dynamic crack growth and arrest in a DCB specimen. Int. J. Fract. 188(2), 187–202 (2014)

    Article  Google Scholar 

  2. Adda-Bedia, M., Arias, R., Amar, M.B., Lund, F.: Generalized Griffith criterion for dynamic fracture and the stability of crack motion at high velocities. Phys. Rev. E 60(2), 2366–2376 (1999)

    Article  ADS  Google Scholar 

  3. Alessi, R., Marigo, J.J., Vidoli, S.: Gradient damage models coupled with plasticity: variational formulation and main properties. Mech. Mater. 80, 351–367 (2015)

    Article  Google Scholar 

  4. Attigui, M., Petit, C.: Numerical path independent integral in dynamic fracture mechanics. In: ECF 11—Mechanisms and Mechanics of Damage and Failure (1996)

    Google Scholar 

  5. Ballarini, R., Royer-Carfagni, G.: Closed-path J-integral analysis of bridged and phase-field cracks. J. Appl. Mech. 83(6), 061,008 (2016)

    Article  Google Scholar 

  6. Borden, M.J., Verhoosel, C.V., Scott, M.A., Hughes, T.J., Landis, C.M.: A phase-field description of dynamic brittle fracture. Comput. Methods Appl. Mech. Eng. 217–220, 77–95 (2012). doi:10.1016/j.cma.2012.01.008

    Article  MathSciNet  MATH  Google Scholar 

  7. Bourdin, B., Francfort, G.A., Marigo, J.J.: The variational approach to fracture. J. Elast. 91(1–3), 5–148 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  8. Bourdin, B., Larsen, C.J., Richardson, C.L.: A time-discrete model for dynamic fracture based on crack regularization. Int. J. Fract. 168(2), 133–143 (2011)

    Article  MATH  Google Scholar 

  9. Destuynder, P., Djaoua, M.: Sur une interprétation mathématique de l’intégrale de Rice en théorie de la rupture fragile. Math. Methods Appl. Sci. 3(1), 70–87 (1981)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  10. Freddi, F., Royer-Carfagni, G.: Regularized variational theories of fracture: a unified approach. J. Mech. Phys. Solids 58(8), 1154–1174 (2010). doi:10.1016/j.jmps.2010.02.010

    Article  ADS  MathSciNet  MATH  Google Scholar 

  11. Freund, L.B.: Dynamic Fracture Mechanics. Cambridge University Press, Cambridge (1990). doi:10.1017/CBO9780511546761

    Book  MATH  Google Scholar 

  12. Griffith, A.A.: The phenomena of rupture and flow in solids. Philos. Trans. R. Soc. Lond. 221, 163–198 (1921)

    Article  ADS  Google Scholar 

  13. Hakim, V., Karma, A.: Laws of crack motion and phase-field models of fracture. J. Mech. Phys. Solids 57(2), 342–368 (2009). doi:10.1016/j.jmps.2008.10.012

    Article  ADS  MATH  Google Scholar 

  14. Hamilton, W.R.: On a general method in dynamics. Philos. Trans. R. Soc. Lond. 124, 247–308 (1834)

    Article  Google Scholar 

  15. Hintermüller, M., Kovtunenko, V.A.: From shape variation to topological changes in constrained minimization: a velocity method-based concept. Optim. Methods Softw. 26(4–5), 513–532 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  16. Khludnev, A., Sokołowski, J., Szulc, K.: Shape and topological sensitivity analysis in domains with cracks. Appl. Math. 55(6), 433–469 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  17. Li, T., Marigo, J.J., Guilbaud, D., Potapov, S.: Gradient damage modeling of brittle fracture in an explicit dynamics context. Int. J. Numer. Methods Eng. (2016). doi:10.1002/nme.5262

    Google Scholar 

  18. Li, T., Marigo, J.J., Guilbaud, D., Potapov, S.: Numerical investigation of dynamic brittle fracture via gradient damage models. Adv. Model. Simul. Eng. Sci. 3, 26 (2016). doi:10.1186/s40323-016-0080-x

    Article  Google Scholar 

  19. Lorentz, E., Andrieux, S.: A variational formulation for nonlocal damage models. Int. J. Plast. 15(2), 119–138 (1999). doi:10.1016/S0749-6419(98)00057-6

    Article  MATH  Google Scholar 

  20. Lorentz, E., Godard, V.: Gradient damage models: toward full-scale computations. Comput. Methods Appl. Mech. Eng. 200(21–22), 1927–1944 (2011). doi:10.1016/j.cma.2010.06.025

    Article  ADS  MathSciNet  MATH  Google Scholar 

  21. Maugin, G.: On the \(J\)-integral and energy-release rates in dynamical fracture. Acta Mech. 105(1–4), 33–47 (1994). doi:10.1007/BF01183940

    Article  MathSciNet  MATH  Google Scholar 

  22. Nishioka, T., Atluri, S.N.: Path-independent integrals, energy release rates, and general solutions of near-tip fields in mixed-mode dynamic fracture mechanics. Eng. Fract. Mech. 18(1), 1–22 (1983)

    Article  Google Scholar 

  23. Oleaga, G.E.: Remarks on a basic law for dynamic crack propagation. J. Mech. Phys. Solids 49(10), 2273–2306 (2001)

    Article  ADS  MATH  Google Scholar 

  24. Pham, K., Amor, H., Marigo, J.J., Maurini, C.: Gradient damage models and their use to approximate brittle fracture. Int. J. Damage Mech. 20(4), 618–652 (2011)

    Article  Google Scholar 

  25. Pham, K., Marigo, J.J.: Approche variationnelle de l’endommagement: II. Les modèles à gradient. C. R., Méc. 338(4), 199–206 (2010)

    Article  Google Scholar 

  26. Pham, K., Marigo, J.J.: From the onset of damage to rupture: construction of responses with damage localization for a general class of gradient damage models. Contin. Mech. Thermodyn. 25(2–4), 147–171 (2013)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  27. Pham, K., Marigo, J.J., Maurini, C.: The issues of the uniqueness and the stability of the homogeneous response in uniaxial tests with gradient damage models. J. Mech. Phys. Solids 59(6), 1163–1190 (2011)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  28. Sharon, E., Fineberg, J.: Microbranching instability and the dynamic fracture of brittle materials. Phys. Rev. B 54, 7128–7139 (1996). doi:10.1103/PhysRevB.54.7128

    Article  ADS  Google Scholar 

  29. Sicsic, P., Marigo, J.J.: From gradient damage laws to Griffith’s theory of crack propagation. J. Elast. 113(1), 55–74 (2013)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tianyi Li.

Appendices

Appendix A: Calculation of the First-Order Action Variation

We will carefully explore the first-order stability principle (15) by calculating the action variation with respect to arbitrary displacement and crack variations. The following easily established identities

$$\begin{aligned} \frac{\mathrm{d}}{\mathrm{d}l_{t}}\det\nabla\boldsymbol{\phi} \bigl(\mathbf{x}^{*} \bigr) &= \det \nabla\boldsymbol{\phi} \bigl(\mathbf{x}^{*} \bigr)\operatorname {tr} \bigl(\nabla \boldsymbol{\theta}^{*} \bigl(\mathbf{x}^{*} \bigr)\nabla\boldsymbol {\phi} \bigl( \mathbf{x}^{*} \bigr)^{-1} \bigr)=\operatorname{div} \boldsymbol{ \theta}_{t}(\mathbf{x})\det\nabla\boldsymbol{\phi} \bigl(\mathbf{x}^{*} \bigr), \\ \frac{\mathrm{d}}{\mathrm{d}l_{t}}\nabla\boldsymbol{\phi} \bigl(\mathbf{x}^{*} \bigr)^{-1} &= -\nabla\boldsymbol{\phi} \bigl(\mathbf{x}^{*} \bigr)^{-1}\nabla \boldsymbol{\theta}^{*} \bigl(\mathbf{x}^{*} \bigr)\nabla \boldsymbol {\phi} \bigl( \mathbf{x}^{*} \bigr)^{-1}=-\nabla\boldsymbol{ \phi} \bigl(\mathbf{x}^{*} \bigr)^{-1}\nabla\boldsymbol{ \theta}_{t}(\mathbf{x}), \end{aligned}$$

will be used for all subsequent calculations.

The classical wave equation can be obtained by calculating the action variation with zero crack advance \(\delta l=0\)

$$\begin{aligned} \mathcal{A}' \bigl(\mathbf{u}^{*},l \bigr) \bigl(\mathbf{w}^{*},0 \bigr)&= \int_{I}\,\mathrm{d}t \int_{\varOmega\setminus\varGamma_{0}} \biggl(\mathsf{A} \biggl({ \frac{1}{2}}\nabla \mathbf{u}_{t}^{*}\nabla\boldsymbol{\phi} ^{-1}+{ \frac{1}{2}}\nabla\boldsymbol{\phi}^{-\mathsf {T}} \bigl(\nabla \mathbf{u}_{t}^{*} \bigr)^{\mathsf{T}} \biggr)\\ &\quad{}\cdot \biggl({ \frac{1}{2}}\nabla\mathbf{w}_{t}^{*}\nabla \boldsymbol{ \phi}^{-1}+{\frac{1}{2}}\nabla\boldsymbol{ \phi}^{-\mathsf {T}} \bigl(\nabla\mathbf{w}_{t}^{*} \bigr)^{\mathsf{T}} \biggr)\det\nabla \boldsymbol{\phi} \\ &\quad{}-\rho \bigl(\dot{\mathbf{u}}_{t}^{*}-\dot{l}_{t}\nabla \mathbf{u}_{t}^{*}\nabla\boldsymbol{\phi} ^{-1}\boldsymbol{ \theta}^{*} \bigr)\cdot \bigl(\dot{\mathbf{w}}_{t}^{*}-\dot{l}_{t} \nabla\mathbf{w}_{t}^{*}\nabla\boldsymbol{\phi}^{-1}\boldsymbol{ \theta}^{*} \bigr)\det\nabla\boldsymbol{\phi} \biggr)\, \mathrm {d}\mathbf{x}^{*}\\ &\quad{}- \mathcal{W}_{t}^{*} \bigl(\mathbf{w}^{*}_{t} \bigr), \end{aligned}$$

which gives

$$\begin{aligned} \mathcal{A}' \bigl(\mathbf{u}^{*},l \bigr) \bigl( \mathbf{w}^{*},0 \bigr)&= \int_{I}\,\mathrm{d}t \int_{\varOmega\setminus\varGamma_{t}} \bigl(\boldsymbol{\sigma }_{t}\cdot \boldsymbol{\varepsilon}(\mathbf{w}_{t})+\rho\dot{l}_{t}\dot{ \mathbf{u}}_{t}\cdot\nabla\mathbf{w}_{t}\boldsymbol{ \theta}_{t} \bigr)\,\mathrm{d}\mathbf{x}-\mathcal{W}_{t}( \mathbf{w}_{t}) \\ &\quad{}+\underbrace{ \int_{I}\,\mathrm{d}t \int_{\varOmega\setminus\varGamma _{0}}\rho\frac{\mathrm{d}}{\mathrm{d}t} \bigl( \bigl(\dot{ \mathbf{u}}_{t}^{*}-\nabla\mathbf{u}_{t}^{*}\nabla\boldsymbol{ \phi}^{-1}\,\dot{l}_{t}\boldsymbol{\theta} ^{*} \bigr)\det\nabla \boldsymbol{\phi} \bigr)\cdot\mathbf{w}_{t}^{*}\,\mathrm{d} \mathbf{x}^{*}}_{R}, \end{aligned}$$
(88)

where \(\mathbf{w}\) denotes the pushforward of \(\mathbf{w}^{*}\) to the current cracked configuration via (6).

To proceed, we observe that the real acceleration \(\ddot{\mathbf{u}}_{t}\) can be obtained by differentiating (8)

$$ \ddot{\mathbf{u}}_{t}(\mathbf{x})=-\nabla\dot{ \mathbf{u}}_{t}(\mathbf{x})\dot{l}_{t}\boldsymbol{\theta}^{*} \bigl(\mathbf{x}^{*} \bigr)+\frac{\mathrm{d}}{\mathrm {d}t} \bigl(\dot{\mathbf{u}}_{t}^{*} \bigl(\mathbf{x}^{*} \bigr)-\nabla\mathbf{u}_{t}^{*} \bigl(\mathbf {x}^{*} \bigr) \nabla\boldsymbol{\phi} \bigl(\mathbf{x}^{*} \bigr)^{-1}\dot {l}_{t}\boldsymbol{\theta}^{*} \bigl(\mathbf{x}^{*} \bigr) \bigr), $$
(89)

where \(\nabla\dot{\mathbf{u}}_{t}\) is the (Eulerian) velocity gradient. Using (89), the last term above can be written

$$\begin{aligned} R &= \int_{I}\,\mathrm{d}t \int_{\varOmega\setminus\varGamma_{0}} \bigl(\rho \bigl(\ddot {\mathbf{u}}_{t} \circ \boldsymbol{\phi} +(\nabla\dot{\mathbf{u}}_{t}\circ\boldsymbol { \phi}) \dot{l}_{t}\boldsymbol{\theta}^{*} \bigr)\cdot\mathbf{w}_{t}^{*} \det\nabla\boldsymbol{\phi}+\rho\dot{l}_{t} \bigl(\dot{ \mathbf{u}}_{t}^{*}-\nabla\mathbf{u}_{t}^{*}\nabla\boldsymbol{ \phi}^{-1}\,\dot{l}_{t}\boldsymbol{\theta}^{*} \bigr) \\ &\quad{}\cdot \mathbf{w}_{t}^{*}\operatorname{tr} \bigl(\nabla\boldsymbol{ \phi}^{-1}\nabla\boldsymbol{\theta}^{*} \bigr)\det\nabla\boldsymbol {\phi} \bigr)\,\mathrm{d}\mathbf{x}^{*} \\ &= \int_{I}\,\mathrm{d}t \int_{\varOmega\setminus\varGamma_{t}}(\rho\ddot{\mathbf {u}}_{t}\cdot \mathbf{w}_{t}+\rho\dot{l}_{t}\nabla\dot{ \mathbf{u}}_{t}\boldsymbol{\theta}_{t}\cdot \mathbf{w}_{t}+\rho\dot{l}_{t}\dot{\mathbf{u}}_{t} \cdot\mathbf{w}_{t}\operatorname{div}\boldsymbol{\theta}_{t}) \,\mathrm{d}\mathbf{x} \\ &= \int_{I}\,\mathrm{d}t \int_{\varOmega\setminus\varGamma_{t}}(\rho\ddot{\mathbf {u}}_{t}\cdot \mathbf{w}_{t}-\rho\dot{l}_{t}\dot{\mathbf{u}}_{t} \cdot\nabla\mathbf{w}_{t}\boldsymbol{\theta}_{t})\,\mathrm{d} \mathbf{x}, \end{aligned}$$
(90)

where an integration by parts in \(\varOmega\setminus\varGamma_{t}\) has been used on establishing the last equality. Regrouping (88) and (90), we obtain thus the spatially weak dynamic equilibrium

$$ \mathcal{A}' \bigl(\mathbf{u}^{*},l \bigr) \bigl( \mathbf{w}^{*},0 \bigr)= \int_{I}\,\mathrm{d}t \int_{\varOmega\setminus\varGamma_{t}} \bigl(\boldsymbol{\sigma }_{t}\cdot \boldsymbol{\varepsilon}(\mathbf{w}_{t})+\rho\ddot{\mathbf{u}}_{t} \cdot\mathbf{w}_{t} \bigr)\,\mathrm{d}\mathbf{x} -\mathcal{W}_{t}( \mathbf{w}_{t}). $$
(91)

An integration by parts then gives the desired wave equation (18) for the displacement.

We then evaluate the action variation with respect to arbitrary crack increment \(\delta l\) but zero displacement variation

$$\begin{aligned} &\mathcal{A}' \bigl(\mathbf{u}^{*},l \bigr) ( \mathbf{0}, \delta l) \\ &\quad= \int_{I}G_{\mathrm{c}}\cdot\delta l_{t}\, \mathrm{d}t+ \int_{I}\delta l_{t}\,\mathrm{d}t \int_{\varOmega\setminus\varGamma_{t}} \bigl( \bigl(\psi \bigl(\boldsymbol{\varepsilon}( \mathbf{u}_{t}) \bigr)-\kappa(\dot{\mathbf{u}}_{t}) \bigr) \operatorname{div}\boldsymbol{\theta}_{t}-\boldsymbol{ \sigma}_{t}\cdot(\nabla\mathbf{u}_{t}\nabla\boldsymbol{ \theta}_{t})-\operatorname{div}(\mathbf{f}_{t}\otimes \boldsymbol{\theta}_{t})\cdot\mathbf{u}_{t} \bigr)\, \mathrm{d}\mathbf{x} \\ &\qquad{}-\underbrace{ \int_{I}\,\mathrm{d}t \int_{\varOmega\setminus\varGamma _{0}} \bigl(\rho \bigl(\dot{\mathbf{u}}^{*}_{t}- \dot {l}_{t}\nabla\mathbf{u}^{*}_{t}\nabla\boldsymbol{ \phi}^{-1}\boldsymbol{\theta}^{*} \bigr)\cdot \bigl(-\nabla \mathbf{u}_{t}^{*}\nabla\boldsymbol{\phi}^{-1}\boldsymbol{ \theta}^{*}\cdot\dot{\overline{\delta l}}_{t}+\dot{l}_{t} \nabla \mathbf{u}_{t}^{*}\nabla\boldsymbol{\phi}^{-1}\nabla \boldsymbol{ \theta}^{*}\nabla\boldsymbol{\phi} ^{-1}\boldsymbol {\theta}^{*} \cdot\delta l_{t} \bigr)\det\nabla\boldsymbol{\phi} \bigr)\, \mathrm{d} \mathbf{x}^{*}}_{R}. \end{aligned}$$
(92)

The last term can be written using integration by parts in the time domain

$$\begin{aligned} R =& \int_{I}\delta l_{t}\,\mathrm{d}t \int_{\varOmega\setminus\varGamma_{0}}\rho\frac{\mathrm {d}}{\mathrm{d}t} \bigl( \bigl(\dot{ \mathbf{u}}^{*}_{t}-\dot{l}_{t}\nabla\mathbf{u}^{*}_{t} \nabla\boldsymbol{\phi}^{-1}\boldsymbol{\theta}^{*} \bigr)\cdot \bigl( \nabla\mathbf{u}_{t}^{*}\nabla\boldsymbol{\phi}^{-1}\boldsymbol{ \theta}^{*} \bigr)\det\nabla\boldsymbol{\phi} \bigr)\,\mathrm {d}\mathbf{x}^{*} \\ &{}+ \int_{I}\delta l_{t}\,\mathrm{d}t \int_{\varOmega\setminus\varGamma_{t}}\rho\dot{l}_{t}\dot{\mathbf{u}}_{t} \cdot\nabla\mathbf{u}_{t}\nabla\boldsymbol{\theta}_{t} \boldsymbol{\theta}_{t}\,\mathrm{d}\mathbf{x}, \end{aligned}$$

which gives

$$\begin{aligned} R&= \int_{I}\delta l_{t}\,\mathrm{d}t \int_{\varOmega\setminus\varGamma_{0}} \bigl(\rho \bigl(\ddot {\mathbf{u}}_{t} \circ \boldsymbol{\phi}+(\nabla\dot{\mathbf{u}}_{t}\circ\boldsymbol{ \phi}) \dot{l}_{t}\boldsymbol{\theta}^{*} \bigr)\cdot \bigl(\nabla \mathbf{u}_{t}^{*}\nabla\boldsymbol{\phi}^{-1}\boldsymbol{ \theta}^{*} \bigr)\det\nabla\boldsymbol{\phi} \\ &\quad{}+\rho \bigl(\dot{\mathbf{u}}^{*}_{t}-\dot{l}_{t}\nabla \mathbf{u}^{*}_{t}\nabla\boldsymbol{\phi} ^{-1}\boldsymbol{ \theta}^{*} \bigr)\cdot \bigl(\nabla\dot{\mathbf{u}}^{*}_{t}\nabla \boldsymbol{ \phi}^{-1}\boldsymbol{\theta} ^{*}-\dot{l}_{t}\nabla \mathbf{u}^{*}_{t}\nabla\boldsymbol{\phi}^{-1}\nabla\boldsymbol{ \theta}^{*}\nabla\boldsymbol{\phi}^{-1}\boldsymbol{\theta}^{*} \bigr)\det \nabla\boldsymbol{\phi} \bigr)\, \mathrm{d}\mathbf{x}^{*} \\ &\quad{}+ \int_{I}\delta l_{t}\,\mathrm{d}t \int_{\varOmega\setminus\varGamma_{t}}(\rho\dot{l}_{t}\dot{\mathbf{u}}_{t} \cdot\nabla\mathbf{u}_{t}\boldsymbol{\theta}_{t} \operatorname{div}\boldsymbol{\theta}_{t}+\rho\dot{l}_{t} \dot{\mathbf{u}}_{t}\cdot\nabla\mathbf{u}_{t}\nabla \boldsymbol{\theta}_{t}\boldsymbol{\theta}_{t})\,\mathrm{d} \mathbf{x}. \end{aligned}$$

We obtain thus

$$\begin{aligned} R&= \int_{I}\delta l_{t}\,\mathrm{d}t \int_{\varOmega\setminus\varGamma _{t}}(\rho\ddot{\mathbf{u}}_{t}\cdot\nabla \mathbf{u}_{t}\boldsymbol{\theta}_{t}+\rho \dot{l}_{t}\nabla\dot{\mathbf{u}}_{t}\boldsymbol{ \theta}_{t}\cdot\nabla\mathbf{u}_{t}\boldsymbol{ \theta}_{t}+\rho\dot{l}_{t}\dot{\mathbf{u}}_{t}\cdot \nabla\mathbf{u}_{t}\boldsymbol{\theta}_{t}\operatorname{div} \boldsymbol{\theta}_{t})\,\mathrm{d}\mathbf{x} \\ &\quad{}+ \int_{I}\delta l_{t}\,\mathrm{d}t \int_{\varOmega\setminus\varGamma_{0}}\rho \bigl(\dot{\mathbf {u}}^{*}_{t}-\dot {l}_{t}\nabla\mathbf{u}^{*}_{t}\nabla\boldsymbol{ \phi}^{-1}\boldsymbol{\theta}^{*} \bigr)\cdot \bigl(\nabla\dot{ \mathbf{u}}^{*}_{t}\nabla\boldsymbol{\phi}^{-1}\boldsymbol{ \theta}^{*} \bigr)\det\nabla\boldsymbol{\phi}\,\mathrm{d}\mathbf{x}^{*}. \end{aligned}$$

Differentiating (7) to obtain the material time derivative of the deformation gradient \((\mathrm{d}/\mathrm{d}t)(\nabla\mathbf{u}_{t})\)

$$\frac{\mathrm{d}}{\mathrm{d}t} \bigl(\nabla\mathbf{u}_{t}(\mathbf {x}) \bigr)= \nabla \dot{\mathbf{u}}_{t}^{*} \bigl(\mathbf{x}^{*} \bigr)\nabla\boldsymbol {\phi} \bigl(\mathbf{x}^{*} \bigr)^{-1}-\dot{l}_{t}\nabla \mathbf{u}_{t}(\mathbf{x})\nabla\boldsymbol{\theta}_{t}( \mathbf{x}), $$

and with its definition

$$\frac{\mathrm{d}}{\mathrm{d}t} \bigl(\nabla\mathbf{u}_{t}(\mathbf {x}) \bigr)= \nabla \dot{\mathbf{u}}_{t}(\mathbf{x})+\nabla^{2} \mathbf{u}_{t}(\mathbf{x})\dot{l}_{t}\boldsymbol{\theta}^{*} \bigl(\mathbf{x}^{*} \bigr), $$

where \(\nabla^{2}\mathbf{u}_{t}\) is the second gradient of the displacement field (a third-order tensor), we obtain

$$\begin{aligned} R&= \int_{I}\delta l_{t}\,\mathrm{d}t \int_{\varOmega\setminus\varGamma_{t}} \bigl(\rho\ddot{\mathbf {u}}_{t}\cdot\nabla \mathbf{u}_{t}\boldsymbol{\theta}_{t}+\rho \dot{l}_{t}\nabla\dot{\mathbf{u}}_{t}\boldsymbol{ \theta}_{t}\cdot\nabla\mathbf{u}_{t}\boldsymbol{ \theta}_{t}+\rho\dot{\mathbf{u}}_{t}\cdot\nabla\dot{ \mathbf{u}}_{t}\boldsymbol{\theta}_{t}+\rho \dot{l}_{t}\dot{\mathbf{u}}_{t}\cdot \bigl(\nabla ^{2}\mathbf{u}_{t}\boldsymbol{\theta}_{t} \bigr) \boldsymbol{\theta}_{t} \\ &\quad{}+\rho\dot{l}_{t}\dot{\mathbf{u}}_{t}\cdot\nabla \mathbf{u}_{t}\nabla\boldsymbol{\theta}_{t}\boldsymbol{\theta} _{t}+\rho\dot{l}_{t}\dot{\mathbf{u}}_{t}\cdot\nabla \mathbf{u}_{t}\boldsymbol{\theta}_{t}\operatorname{div} \boldsymbol{\theta} _{t} \bigr)\,\mathrm{d}\mathbf{x}. \end{aligned}$$

Using an integration by parts in the domain \(\varOmega\setminus \varGamma_{t}\) knowing that \(\boldsymbol{\theta}_{t}=\mathbf{0}\) on \(\partial \varOmega\) and \(\boldsymbol{\theta}_{t}\cdot \mathbf{n}=0\) on \(\varGamma_{t}\) by definition

$$\begin{aligned} &\int_{\varOmega\setminus\varGamma_{t}}\rho\dot{l}_{t}\dot{\mathbf{u}}_{t} \cdot\nabla\mathbf{u}_{t}\nabla\boldsymbol{\theta}_{t} \boldsymbol{\theta}_{t}\,\mathrm{d}\mathbf{x}\\ &\quad=- \int_{\varOmega \setminus\varGamma_{t}} \bigl(\rho\dot{l}_{t}\dot{ \mathbf{u}}_{t}\cdot\nabla\mathbf{u}_{t}\boldsymbol{ \theta}_{t}\operatorname{div}\boldsymbol{\theta}_{t}+\rho \dot{l}_{t}\nabla\dot{\mathbf{u}}_{t}\boldsymbol{ \theta}_{t}\cdot\nabla\mathbf{u}_{t}\boldsymbol{\theta }_{t}+\rho\dot{l}_{t}\dot{\mathbf{u}}_{t}\cdot \bigl(\nabla^{2}\mathbf{u}_{t}\boldsymbol{ \theta}_{t} \bigr)\boldsymbol{\theta}_{t} \bigr)\,\mathrm{d} \mathbf{x}, \end{aligned}$$

we get finally

$$R= \int_{I}\delta l_{t}\,\mathrm{d}t \int_{\varOmega\setminus\varGamma _{t}}(\rho\ddot{\mathbf{u}}_{t}\cdot\nabla \mathbf{u}_{t}\boldsymbol{\theta}_{t}+\rho\dot{ \mathbf{u}}_{t}\cdot\nabla\dot{\mathbf{u}}_{t}\boldsymbol{ \theta} _{t})\,\mathrm{d}\mathbf{x} $$

which permits with (92) to deduce the desired equations (20) and (21).

Appendix B: Local Energy Balance Condition

In this section we will derive the equivalent local condition of the global energy balance (16), which gives the desired Griffith’s law of motion (22) when combined with the local stability condition (20). The Lagrangian defined in (14) is explicitly dependent on time solely through the external work potential (11). Its total derivative can thus be given by

$$ \frac{\mathrm{d}\mathcal {L}}{\mathrm{d}t}=\frac{\partial\mathcal {L}}{\partial\mathbf{u}_{t}^{*}}\dot{\mathbf{u}}_{t}^{*}+ \frac{\partial\mathcal{L}}{\partial\dot {\mathbf{u}}_{t}^{*}}\ddot{\mathbf{u}}_{t}^{*}+\frac{\partial\mathcal {L}}{\partial l_{t}}\dot {l}_{t}+\frac{\partial\mathcal{L}}{\partial\dot{l}_{t}}\ddot {l}_{t}+\frac {\partial\mathcal{L}}{\partial t}. $$
(93)

Using the weak dynamic equilibrium (91) and the fact that \(\dot{\mathbf{u}}_{t}^{*}-\dot{\mathbf{U}}_{t}\in\mathcal {C}_{0}\), we have

$$ \frac{\partial\mathcal{L}}{\partial \mathbf{u}_{t}^{*}} \bigl(\dot{\mathbf{u}}_{t}^{*}-\dot{ \mathbf{U}}_{t} \bigr)-\frac{\mathrm{d}}{\mathrm{d}t}\frac{\partial \mathcal {L}}{\partial\dot {\mathbf{u}}_{t}^{*}} \bigl(\dot{ \mathbf{u}}_{t}^{*}-\dot{\mathbf{U}}_{t} \bigr)=\mathbf{0}. $$
(94)

Plugging (94) into (93), we obtain

$$ \frac{\mathrm{d}\mathcal{L}}{\mathrm {d}t}=\frac{\mathrm {d}}{\mathrm{d}t} \biggl(\frac{\partial \mathcal{L}}{\partial\dot{\mathbf{u}}_{t}^{*}} \dot{\mathbf{u}}_{t}^{*} \biggr)+\frac {\partial\mathcal{L}}{\partial\mathbf{u}_{t}^{*}}\dot{ \mathbf{U}}_{t}-\frac {\mathrm{d} }{\mathrm{d}t}\frac{\partial\mathcal{L}}{\partial\dot{\mathbf {u}}_{t}^{*}}\dot{ \mathbf{U}}_{t}+\frac{\partial\mathcal{L}}{\partial l_{t}}\dot{l}_{t}+ \frac {\partial \mathcal{L}}{\partial\dot{l}_{t}}\ddot{l}_{t}+\frac{\partial\mathcal {L}}{\partial t}. $$
(95)

With all necessary temporal regularity, we note that the energy balance condition (16) can be equivalently written as

$$ \frac{\mathrm{d}\mathcal{H}}{\mathrm {d}t}=\frac{\mathrm {d}}{\mathrm{d}t} (\mathcal{L}+2\mathcal{K} )=\frac{\mathrm{d}}{\mathrm{d}t} \biggl(\mathcal{L}-\frac {\partial\mathcal{L}}{\partial\dot{\mathbf{u}}_{t}^{*}}\dot{ \mathbf{u}}_{t}^{*}-\frac {\partial\mathcal{L}}{\partial\dot{l}_{t}}\dot{l}_{t} \biggr)= \frac {\partial \mathcal{L}}{\partial\mathbf{u}_{t}^{*}}\dot{\mathbf{U}}_{t}-\frac {\mathrm {d}}{\mathrm{d}t} \frac {\partial\mathcal{L}}{\partial\dot{\mathbf{u}}_{t}^{*}}\dot{\mathbf {U}}_{t}+\frac {\partial\mathcal{L}}{\partial t}. $$
(96)

Comparing (95) and (96), we obtain the desired local energy balance condition

$$\biggl(\frac{\partial\mathcal{L}}{\partial l_{t}}-\frac{\mathrm {d}}{\mathrm{d}t}\frac {\partial\mathcal{L}}{\partial\dot{l}_{t}} \biggr)\cdot \dot{l}_{t}=0. $$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, T., Marigo, JJ. Crack Tip Equation of Motion in Dynamic Gradient Damage Models. J Elast 127, 25–57 (2017). https://doi.org/10.1007/s10659-016-9595-0

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10659-016-9595-0

Keywords

Mathematics Subject Classification (2010)

Navigation