Skip to main content
Log in

Phase field approximation of dynamic brittle fracture

  • Original Paper
  • Published:
Computational Mechanics Aims and scope Submit manuscript

Abstract

Numerical methods that are able to predict the failure of technical structures due to fracture are important in many engineering applications. One of these approaches, the so-called phase field method, represents cracks by means of an additional continuous field variable. This strategy avoids some of the main drawbacks of a sharp interface description of cracks. For example, it is not necessary to track or model crack faces explicitly, which allows a simple algorithmic treatment. The phase field model for brittle fracture presented in Kuhn and Müller (Eng Fract Mech 77(18):3625–3634, 2010) assumes quasi-static loading conditions. However dynamic effects have a great impact on the crack growth in many practical applications. Therefore this investigation presents an extension of the quasi-static phase field model for fracture from Kuhn and Müller (Eng Fract Mech 77(18):3625–3634, 2010) to the dynamic case. First of all Hamilton’s principle is applied to derive a coupled set of Euler-Lagrange equations that govern the mechanical behaviour of the body as well as the crack growth. Subsequently the model is implemented in a finite element scheme which allows to solve several test problems numerically. The numerical examples illustrate the capabilities of the developed approach to dynamic fracture in brittle materials.

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
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23

Similar content being viewed by others

References

  1. Ambrosio L, Tortorelli VM (1990) Approximation of functional depending on jumps by elliptic functional via \(\gamma \)-convergence. Commun Pure Appl Math 43(8):999–1036

    Article  MathSciNet  MATH  Google Scholar 

  2. Amor H, Marigo JJ, Maurini C (2009) Regularized formulation of the variational brittle fracture with unilateral contact: numerical experiments. J Mech Phys Solid 57(8):1209–1229

    Article  MATH  Google Scholar 

  3. Bertram A, Kalthoff JF (2003) Crack propagation toughness of rock for the range of low to very high crack speeds. Key Eng Mater 251–252(1):423–430

    Article  Google Scholar 

  4. Borden MJ, Verhoosel CV, Scott MA, Hughes TJR, Landis CM (2012) A phase-field description of dynamic brittle fracture. Comput Meth Appl Mech Eng 217–220:77–95

    Article  MathSciNet  Google Scholar 

  5. Bourdin B (2007) Numerical implementation of the variational formulation of quasi-static brittle fracture. Interfaces Free Boundaries 9:411–430

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MATH  Google Scholar 

  7. Braides A (2002) \(\Gamma \)-convergence for beginners. In: Oxford Lecture Series in Mathematics and its Applications, vol. 22. Oxford University Press, Oxford

  8. Chambolle A (2004) An approximation result for special functions with bounded deformation. J Math Pure Appl 83(7):929–954

    Article  MathSciNet  Google Scholar 

  9. Das S (2003) Dynamic fracture mechanics in the study of the earthquake rupturing process: theory and observation. J Mech Phys Solid 51(11–12):1939–1955

    Article  MATH  Google Scholar 

  10. Fineberg J, Gross SP, Marder M, Swinney HL (1992) Instability in the propagation of fast cracks. Phys Rev B 45:5146–5154

    Article  Google Scholar 

  11. Francfort GA, Marigo JJ (1998) Revisiting brittle fracture as an energy minimization problem. J Mech Phys Solid 46(8):1319–1342

    Article  MathSciNet  MATH  Google Scholar 

  12. Freund LB (1990) Dynamic Fracture Mechanics. Cambridge University Press, Cambridge

    Book  MATH  Google Scholar 

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

    Article  Google Scholar 

  14. Gross D, Seelig T (2010) Fracture mechanics: with an introduction to micromechanics. Mechanical engineering series. Springer, Berlin

    Google Scholar 

  15. Gross SP, Fineberg J, Marder M, McCormick WD, Swinney HL (1993) Acoustic emissions from rapidly moving cracks. Phys Rev Lett 71:3162–3165

    Article  Google Scholar 

  16. Gurtin ME (1996) Generalized ginzburg-landau and cahn-hilliard equations based on a microforce balance. Phys D 92(3–4):178–192

    Article  MathSciNet  MATH  Google Scholar 

  17. Hakim V, Karma A (2009) Laws of crack motion and phase-field models of fracture. J Mech Phys Solid 57(2):342–368

    Article  MATH  Google Scholar 

  18. Hofacker M, Miehe C (2012) Continuum phase field modeling of dynamic fracture: variational principles and staggered fe implementation. Int J Fract 178(1–2):113–129

    Article  Google Scholar 

  19. Hofacker M, Miehe C (2013) A phase field model of dynamic fracture: robust field updates for the analysis of complex crack patterns. Int J Numer Methods Eng 93(3):276–301

    Article  MathSciNet  Google Scholar 

  20. Hopkinson B (1921) The pressure of a blow. The Scientific Papers of Bertram Hopkinson pp 423–437

  21. Hopkinson J (1872) On the rupture of iron wire by a blow. Proc Manch Lit Philos Soc 11:40–45

    Google Scholar 

  22. Hughes TJR (2000) The finite element method: linear static and dynamic finite element analysis. Dover Publications Inc, Mineola

    Google Scholar 

  23. Irwin GR (1957) Analysis of stresses and strains near the end of a crack traversing a plate. J Appl Mech 24:361–364

    Google Scholar 

  24. Kalthoff J (2000) Modes of dynamic shear failure in solids. Int J Fract 101:1–31

    Article  Google Scholar 

  25. Kalthoff JF, Winkler S (1987) Failure mode transition of high rates of shear loading. Proc. I. Con. Imp. Load. Dyn. Beh. Mat. 1:185–195

    Google Scholar 

  26. Katzav E, Adda-Bedia M, Arias R (2007) Theory of dynamic crack branching in brittle materials. Int. J. Fract. 143:245–271

    Article  MATH  Google Scholar 

  27. Ravi-Chandar K, Knauss WG (1984) An experimental investigation into dynamic fracture: Iii on steady-state crack propagation and crack branching. Int J Fract 26:33–72

    Article  Google Scholar 

  28. Krueger R (2004) Virtual crack closure technique: history, approach, and applications. Appl Mech Rev 57(2):109

    Article  MathSciNet  Google Scholar 

  29. Kuhn C (2013) Numerical and analytical investigation of a phase field model for fracture. Phd thesis, Technische Universität Kaiserslautern

  30. Kuhn C, Müller R (2010) A continuum phase field model for fracture. Eng Fract Mech 77(18):3625–3634. Computational mechanics in fracture and damage: a special issue in honor of Prof. Gross

  31. Maso GD (1993) Introduction to \(\Gamma \)-convergence. Progress in nonlinear differential equations and their applications, Birkhäuser

  32. Mello M, Bhat HS, Rosakis AJ, Kanamori H (2014) Reproducing the supershear portion of the 2002 denali earthquake rupture in laboratory. Earth Planet Sci Lett 387:89–96

    Article  Google Scholar 

  33. Moës N, Dolbow J, Belytschko T (1999) A finite element method for crack growth without remeshing. Int J Numer Meth Eng 46(1):131–150

  34. Mumford D, Shah J (1989) Optimal approximations by piecewise smooth functions and associated variational problems. Commun Pure Appl Math 42(5):577–685

    Article  MathSciNet  MATH  Google Scholar 

  35. Ravi-Chandar K, Knauss W (1984) An experimental investigation into dynamic fracture: I. crack initiation and arrest. Int J Fract 25(4):247–262

    Article  Google Scholar 

  36. Ravi-Chandar K, Knauss W (1984) An experimental investigation into dynamic fracture: Ii. microstructural aspects. Int J Fract 26(1):65–80

    Article  Google Scholar 

  37. Ravi-Chandar K, Knauss W (1984) An experimental investigation into dynamic fracture: Iii. on steady-state crack propagation and crack branching. Int J Fract 26(2):141–154

    Article  Google Scholar 

  38. Ravi-Chandar K, Knauss W (1984) An experimental investigation into dynamic fracture: Iv. on the interaction of stress waves with propagating cracks. Int J Fract 26(3):189–200

    Article  Google Scholar 

  39. Rosakis AJ, Samudrala O, Coker D (2000) Intersonic shear crack growth along weak planes. Mater Res Innov 3(4):236–243

    Article  Google Scholar 

  40. Schlueter A, Willenbuecher A, Kuhn C (2013) GPU-accelerated crack path computation based on a phase field approach for brittle fracture. In: Proceedings of the 2nd young researcher symposium (YRS) 2013, pp 60–65. Fraunhofer Verlag, Kaiserslautern, Germany

  41. Seelig T, Gross D (1999) On the interaction and branching of fast running cracksâa numerical investigation. J Mech Phys Solids 47(4):935–952

    Article  MATH  Google Scholar 

  42. Sharon E, Gross SP, Fineberg J (1995) Local crack branching as a mechanism for instability in dynamic fracture. Phys Rev Lett 74:5096–5099

    Article  Google Scholar 

Download references

Acknowledgments

A. Schlüter, C. Kuhn, and R. Müller acknowledge the support from the German Research Foundation (DFG) within the CRC 926 “Microscale Morphology of Component Surfaces”

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alexander Schlüter.

Appendix: Finite element implementation

Appendix: Finite element implementation

The set of coupled partial differential Eqs. (29) and (30) can be solved by standard multi-field finite element solvers. This section outlines the implementation of the present dynamic phase field model with the finite element code FEAP. The initial boundary value problem is solved simultaneously for the displacements \(\varvec{u}\) and the crack field \(s\). To start with, the weak forms of (29) and (30) are formulated

$$\begin{aligned}&\int \limits _{\varOmega }{\left( -\rho \ddot{\varvec{u}}\cdot \delta {\varvec{u}}-\varvec{\sigma }:\delta {{\varvec{\varepsilon }}}\right) }\ \mathrm{d}V\nonumber \\&\quad +{\int \limits _{\partial {\varOmega _t}}{\varvec{t}^*\cdot \delta \varvec{u}}\ \mathrm{d}A}=0 \end{aligned}$$
(80)

and

$$\begin{aligned}&\int \limits _{\varOmega }{\left( \left[ 2s\psi _{e,zs}-\mathcal{G}_c\frac{1-s}{2\epsilon }\right] \delta {s}\right. }\nonumber \\&\quad +{\left. 2\epsilon \mathcal{G}_c\nabla {s}\cdot \nabla {\delta {s}}\right) \mathrm{d}V}=0. \end{aligned}$$
(81)

Standard isoparametric, four node in 2D and eight node in 3D finite elements are used to discretize the body \(\varOmega \). Therefore, the primary fields can be interpolated in the interior of an element by means of the nodal values \(\underline{\varvec{u}}_I, {s}_I\) and the scalar bilinear or trilinear shape functions \(N_I\)

$$\begin{aligned} \begin{array}{ll} \underline{\varvec{u}}_h= \displaystyle {\sum _{I=1}^{N_e}}{N_I\underline{\varvec{u}}_I}&\text { and }{s}_h= \displaystyle {\sum _{I=1}^{N_e}}{N_I{{s}}_I} \end{array} \end{aligned}$$
(82)

where \(N_e\) is the total number of nodes per element. The 2D, plane strain implementation is discussed in more detail in the following. However a 3D finite implementation is conceptually similar and straightforward. For further information about the finite element method the reader is referred to textbooks like Hughes [22]. In a two dimensional setting the spatial derivatives in (80) and (81) can be expressed by means of the matrices

$$\begin{aligned} \begin{array}{ll} \underline{\varvec{B}}_I^{\varvec{u}}= \left[ \begin{array}{ll} N_{I,1} &{} 0 \\ 0 &{} N_{I,2} \\ N_{I,2} &{} N_{I,1} \end{array} \right] \text { and } &{} \underline{\varvec{B}}_I^{s}= \left[ \begin{array}{l} N_{I,1} \\ N_{I,2} \end{array} \right] . \end{array} \end{aligned}$$
(83)

In these matrices \(N_{I,1}\) and \(N_{I,2}\) denote the differentiation of the shape functions with respect to the coordinates \(x_1\) and \(x_2\). This way, it is possible to express the gradients

$$\begin{aligned} \underline{{\varvec{\varepsilon }}}_h= \left[ \begin{array}{c} \varepsilon _{11_h}\\ \varepsilon _{22_h}\\ 2\varepsilon _{12_h} \end{array} \right] = \displaystyle {\sum _{I=1}^{N_e}}{\underline{\varvec{B}}_I^{\varvec{u}}\underline{\varvec{u}}_I} \end{aligned}$$
(84)

and

$$\begin{aligned} \underline{\nabla {s}}_h= \left[ \begin{array}{l} s_h,_1\\ s_h,_2 \end{array}\right] = \displaystyle {\sum _{I=1}^{N_e}}{\underline{\varvec{B}}_I^{s}{s}_I} \end{aligned}$$
(85)

in matrix notation. Furthermore, two matrices are introduced that represent the deviatoric part of \(\underline{{\varvec{\varepsilon }}}_h\)

$$\begin{aligned} \underline{{\varvec{\varepsilon }}_D}_h^*= \left[ \begin{array}{c} \frac{1}{2}(\varepsilon _{11_h}-\varepsilon _{22_h})\\ \frac{1}{2}(\varepsilon _{22_h}-\varepsilon _{11_h})\\ 2\varepsilon _{12_h} \end{array} \right] \end{aligned}$$
(86)

and

$$\begin{aligned} \underline{{\varvec{\varepsilon }}_{D}}_h= \left[ \begin{array}{c} \frac{1}{2}(\varepsilon _{11_h}-\varepsilon _{22_h})\\ \frac{1}{2}(\varepsilon _{22_h}-\varepsilon _{11_h})\\ \varepsilon _{12_h} \end{array} \right] . \end{aligned}$$
(87)

By means of the above terms, it is possible to express the elastic energy density as

$$\begin{aligned}&\psi _{\text {e}_h} = \frac{K_n}{2}{\langle \varepsilon _{11_h}+\varepsilon _{22_h}}\rangle _{-}^2\nonumber \\&\quad +({s}_h^2+\eta )\underbrace{\left[ \frac{K_n}{2}{\langle \varepsilon _{11_h} +\varepsilon _{22_h}}\rangle _{+}^2+\mu \left( {\underline{{\varvec{\varepsilon }}_D}_h^*}^T \underline{{\varvec{\varepsilon }}_{D}}_h\right) \right] }_{\displaystyle \psi _{\text {e,ts}_{h}}}, \end{aligned}$$
(88)

where the bracket terms \(\langle \cdot \rangle _{+/-}\) are treated according to (12) and (13). Consequently the stress tensor is

$$\begin{aligned} \underline{\varvec{\sigma }}_h=&\dfrac{\partial \psi _{e_h}}{\partial {\varvec{\varepsilon }}_h}\nonumber \\ =&K_n\text {tr}^{-}(\underline{{\varvec{\varepsilon }}}_h)\underline{\varvec{1}}\nonumber \\&+(s_h^2+\eta )\left[ K_n\text {tr}^{+}(\underline{{\varvec{\varepsilon }}}_h)\underline{\varvec{1}}+2\mu \underline{{\varvec{\varepsilon }}_{D}}_h\right] , \end{aligned}$$
(89)

where

$$\begin{aligned} \underline{\varvec{1}}&= \left[ \begin{array}{l@{\quad }l@{\quad }l} 1&1&0 \end{array} \right] ^T \end{aligned}$$
(90)

is the second-order identity tensor in matrix notation. By means of these expressions one can reformulate the governing equations (80) and (81) in discretized form

$$\begin{aligned} \displaystyle \bigcup _{e=1}^{n_e}\displaystyle \sum _{I=1}^{{N_e}} {\delta \underline{\varvec{u}}_I^T\underline{\varvec{R}}_{I,e}^{\varvec{u}}}={{0}},\end{aligned}$$
(91)
$$\begin{aligned} \displaystyle \bigcup _{e=1}^{n_e}\sum _{I=1}^{N_e}\delta {s}_I{R}_{I,e}^{s}={{0}}, \end{aligned}$$
(92)

where \(n_e\) and \(N_e\) are the total number of elements and the number of nodes per element, respectively. The symbol \(\bigcup \) denotes the assembly operator. The contribution of element \(e\) at node \(I\) to the global residuum are

$$\begin{aligned} \begin{array}{ll} \underline{\varvec{R}}_{I,e}^{\varvec{u}}&{}=\underbrace{\int \limits _{\varOmega _e} {(-{\left[ \underline{\varvec{B}}_I^{\varvec{u}}\right] }^T\underline{\varvec{\sigma }}_h-\rho {N_I}\underline{\ddot{\varvec{u}}}_h)\mathrm{d}V}}_{\displaystyle -\underline{\varvec{P}}_{I,e}^{\varvec{u}}}\\ &{}\quad +\underbrace{\int \limits _{\partial {\varOmega }_e^t\cap \partial {\varOmega ^t}} {N_I\underline{\varvec{t}}_h^*\mathrm{d}A}}_{\displaystyle \underline{\varvec{F}}_{I,e}^{\varvec{u}}} \end{array} \end{aligned}$$
(93)

and

$$\begin{aligned} \begin{array}{ll} {R}_{I,e}^{s}=&{}-{P}_{I,e}^s=\displaystyle \int \limits _{\varOmega _e}\left( -[\underline{\varvec{B}}_I^{s}]^T2\epsilon \mathcal{G}_c\underline{\nabla {s}}_h\right. \\ &{}\left. -N_I\left( 2{s}_h{\psi _{e,ts}}_h -\frac{\mathcal{G}_c}{2\epsilon }(1-{s}_h)\right) \right) \mathrm{d}V. \end{array} \end{aligned}$$
(94)

Terms that can be attributed to inner forces are denoted \(\underline{\varvec{P}}_{I,e}^{\varvec{u}}\) and \({P}_{I,e}^{s}\) respectively, whereas external mechanical loads are labelled \(\underline{\varvec{F}}_{I,e}^{\varvec{u}}\). The residuals \(\underline{\varvec{R}}_{I,e}^{\varvec{u}}\) and \({R}_{I,e}^{s}\) have to be computed on element level and are subsequently assembled to a global residuum \(\underline{\varvec{R}}\). The discretized equations (91) and (92) define a global system of equations

$$\begin{aligned} \delta \underline{\varvec{d}}^T\underline{\varvec{R}}=0\quad \Leftrightarrow \quad \underline{\varvec{R}}=\underline{\varvec{0}} \end{aligned}$$
(95)

where

$$\begin{aligned} \underline{\varvec{R}} = \bigcup _{e=1}^{n_e}{\underline{\varvec{R}}_{e}}= \left[ \begin{array}{lll} \left( \underline{\varvec{R}}_{1}\right) ^T&...&\left( \underline{\varvec{R}}_{N}\right) ^T \end{array} \right] ^T, \end{aligned}$$
(96)

and

$$\begin{aligned} \begin{array}{ll} \delta \underline{\varvec{d}}= \left[ \begin{array}{lll} \left( \delta \underline{\varvec{d}}_1\right) ^T &{}...&{}\left( \delta \underline{\varvec{d}}_N\right) ^T \end{array} \right] ^T,&{} \\ \delta \underline{\varvec{d}}_I= \left[ \begin{array}{ll} \left( \delta \underline{\varvec{u}}_I\right) ^T &{}\delta {s}_I \end{array} \right] ^T \end{array} \end{aligned}$$
(97)

are the global vectors of nodal residuals and nodal values of the test functions respectively.

1.1 Appendix B: Time discretization and linearization

By means of the internal and external forces \(\underline{\varvec{P}}\) and \(\underline{\varvec{F}}\) the global system of Eq. (95) reads

$$\begin{aligned} \underline{\varvec{R}}=\underline{\varvec{F}}-\underline{\varvec{P}}\left( {{\underline{\varvec{d}}}}, \dot{{{\underline{\varvec{d}}}}},\ddot{{{\underline{\varvec{d}}}}}\right) =\underline{\varvec{0}} \end{aligned}$$
(98)

or in time discretized form (\(t_{n+1}=t_n+\Delta {t}\))

$$\begin{aligned} \underline{\varvec{R}}_{n+1}=\underline{\varvec{F}}_{n+1}-\underline{\varvec{P}} \left( {{\underline{\varvec{d}}}}_{n+1},\dot{{{\underline{\varvec{d}}}}}_{n+1},\ddot{{{\underline{\varvec{d}}}}}_{n+1}\right) =\underline{\varvec{0}}. \end{aligned}$$
(99)

Herein, the subscript \(n+1\) denotes terms that are evaluated at time \(t_{n+1}\). In order to approximate \(\dot{{{\underline{\varvec{d}}}}}_{n+1}\) and \(\ddot{{{\underline{\varvec{d}}}}}_{n+1}\) the implicit Newmark method

$$\begin{aligned} \dot{{\underline{\varvec{d}}}}_{n+1}&= \dot{{\underline{\varvec{d}}}}_n + \Delta t\ \ddot{{\underline{\varvec{d}}}}_\gamma , \text { where }\nonumber \\ \ddot{{\underline{\varvec{d}}}}_\gamma&= (1-\gamma )\ \ddot{{\underline{\varvec{d}}}}_n+\gamma \ \ddot{{\underline{\varvec{d}}}}_{n+1}\,,\end{aligned}$$
(100)
$$\begin{aligned} {{\underline{\varvec{d}}}}_{n+1}&= {{\underline{\varvec{d}}}}_n+\Delta t\ \dot{{\underline{\varvec{d}}}}_n+\frac{1}{2}\Delta t^2\ \ddot{{\underline{\varvec{d}}}}_\beta , \text { where }\nonumber \\ \ddot{{\underline{\varvec{d}}}}_\beta&= (1-2\beta )\ \ddot{{\underline{\varvec{d}}}}_n+2\beta \ \ddot{{\underline{\varvec{d}}}}_{n+1}\, \end{aligned}$$
(101)

with parameters \(\gamma =\beta =0.5\) is utilized. The time step size \(\Delta t\) is chosen by an automatic time step size control that adapts the time step size according to the number of Newton iterations needed for convergence in the previous time step. Substituting the relations (100) and (101) in (99) yields a system of nonlinear equations for the unknowns \(\varvec{\underline{d}}_{n+1}\) that has to be solved in every time step

$$\begin{aligned} \underline{\varvec{R}}_{n+1}&=\widehat{\underline{\varvec{R}}}\left( \underline{\varvec{F}}_{n+1}, {{\underline{\varvec{d}}}}_{n},\dot{{{\underline{\varvec{d}}}}}_{n},\ddot{{{\underline{\varvec{d}}}}}_{n},{{\underline{\varvec{d}}}}_{n+1}\right) \nonumber \\&=\underline{\varvec{F}}_{n+1}-\widehat{\underline{\varvec{P}}} \left( {{\underline{\varvec{d}}}}_{n},\dot{{{\underline{\varvec{d}}}}}_{n},\ddot{{{\underline{\varvec{d}}}}}_{n},{{\underline{\varvec{d}}}}_{n+1}\right) =\underline{\varvec{0}}. \end{aligned}$$
(102)

For this purpose, Newton’s method is applied

$$\begin{aligned} \underline{\varvec{R}}_{n+1}^{(k+1)}&=\underline{\varvec{R}}_{n+1}^{(k)}+\Delta {\underline{\varvec{R}}_{n+1}^{(k)}}\nonumber \\&\approx \underline{\varvec{R}}_{n+1}^{(k)}+\underbrace{D{\underline{\varvec{R}}_{n+1}^{(k)}}}_{=-\underline{\varvec{S}}_{n+1}^{(k)}} \Delta {{{\underline{\varvec{d}}}}_{n+1}}\nonumber \\&= \underline{\varvec{R}}_{n+1}^{(k)}-\underline{\varvec{S}}_{n+1}^{(k)}\Delta {{{\underline{\varvec{d}}}}_{n+1}^{(k)}}=\underline{\varvec{0}} \end{aligned}$$
(103)

where the index \(k\) indicates the current Newton iteration and \(\underline{\varvec{S}}_{n+1}^{(k)}\) is the overall tangent matrix of the problem. Equation (103) defines an iteration scheme, in which a system of linear equations (SLE) has to be solved for \(\Delta {{{\underline{\varvec{d}}}}_{n+1}^{(k)}}\) in each iteration. Subsequently the vector of unknowns is updated according to

$$\begin{aligned} {{{\underline{\varvec{d}}}}_{n+1}^{(k+1)}} = {{{\underline{\varvec{d}}}}_{n+1}^{(k)}} +\Delta {{{\underline{\varvec{d}}}}_{n+1}^{(k)}} \end{aligned}$$
(104)

and the norm \(\left\| \underline{\varvec{R}}\left( {{\underline{\varvec{d}}}}_{n+1}^{(k+1)}\right) \right\| \) is compared to a predefined tolerance \(\delta \) to check for convergence. The overall tangent matrix is computed from

$$\begin{aligned} \underline{\varvec{S}}_{n+1}^{(k)}&= \frac{\partial \widehat{\underline{\varvec{P}}}}{\partial {{\underline{\varvec{d}}}}_{n+1}^{(k)}}\left( {{\underline{\varvec{d}}}}_{n},\dot{{{\underline{\varvec{d}}}}}_{n},\ddot{{{\underline{\varvec{d}}}}}_{n},{{\underline{\varvec{d}}}}_{n+1}^{(k)}\right) \nonumber \\&= \frac{\mathrm{d}\underline{\varvec{P}}}{\mathrm{d}{{\underline{\varvec{d}}}}_{n+1}^{(k)}}\left( {{\underline{\varvec{d}}}}_{n+1}^{(k)},\dot{{{\underline{\varvec{d}}}}}_{n+1}^{(k)},\ddot{{{\underline{\varvec{d}}}}}_{n+1}^{(k)}\right) \nonumber \\&= \underline{\varvec{K}}^{(k)}+\underline{\varvec{D}}^{(k)}\frac{\partial \dot{{{\underline{\varvec{d}}}}}_{n+1}^{(k)}}{\partial {{\underline{\varvec{d}}}}_{n+1}^{(k)}} +\underline{\varvec{M}}^{(k)}\frac{\partial \ddot{{{\underline{\varvec{d}}}}}_{n+1}^{(k)}}{\partial {{\underline{\varvec{d}}}}_{n+1}^{(k)}}\nonumber \\&= \underline{\varvec{K}}^{(k)}+\frac{\gamma }{\beta \Delta {t}}\underline{\varvec{D}}^{(k)}+\frac{1}{\beta \Delta {t}^2}\underline{\varvec{M}}^{(k)}. \end{aligned}$$
(105)

Similar to the residuals, the stiffness matrix

$$\begin{aligned} \underline{\varvec{K}}^{(k)}=\frac{\partial \underline{\varvec{P}}}{\partial {{\underline{\varvec{d}}}}_{n+1}^{(k)}}, \end{aligned}$$
(106)

the damping matrix

$$\begin{aligned} \underline{\varvec{D}}^{(k)}=\frac{\partial \underline{\varvec{P}}}{\partial \dot{{{\underline{\varvec{d}}}}}_{n+1}^{(k)}}=\varvec{\underline{0}}, \end{aligned}$$
(107)

and the mass matrix

$$\begin{aligned} \underline{\varvec{M}}^{(k)}=\frac{\partial \underline{\varvec{P}}}{\partial \ddot{{{\underline{\varvec{d}}}}}_{n+1}^{(k)}}. \end{aligned}$$
(108)

are computed for each finite element separately and are assembled to the global matrices. The stiffness matrix of each four node element can be written as

$$\begin{aligned} \underline{\varvec{K}}_e^{(k)}= \left[ \begin{array}{l@{\quad }l@{\quad }l@{\quad }l} \underline{\varvec{K}}_{11,e}^{(k)} &{}\underline{\varvec{K}}_{12,e}^{(k)} &{}\underline{\varvec{K}}_{13,e}^{(k)} &{}\underline{\varvec{K}}_{14,e}^{(k)}\\ \underline{\varvec{K}}_{21,e}^{(k)} &{}\underline{\varvec{K}}_{22,e}^{(k)} &{}\underline{\varvec{K}}_{23,e}^{(k)} &{}\underline{\varvec{K}}_{24,e}^{(k)}\\ \underline{\varvec{K}}_{31,e}^{(k)} &{}\underline{\varvec{K}}_{32,e}^{(k)} &{}\underline{\varvec{K}}_{33,e}^{(k)} &{}\underline{\varvec{K}}_{34,e}^{(k)}\\ \underline{\varvec{K}}_{41,e}^{(k)} &{}\underline{\varvec{K}}_{42,e}^{(k)} &{}\underline{\varvec{K}}_{43,e}^{(k)} &{}\underline{\varvec{K}}_{44,e}^{(k)} \end{array} \right] , \end{aligned}$$
(109)

where the submatrices \(\underline{\varvec{K}}_{IJ,e}^{(k)}\) have the structure

$$\begin{aligned} \underline{\varvec{K}}_{IJ,e}^{(k)}= \left[ \begin{array}{ll} \underline{\varvec{K}}_{IJ,e}^{\varvec{u}\varvec{u}} &{}\underline{\varvec{K}}_{IJ,e}^{\varvec{u}{s}}\\ \underline{\varvec{K}}_{IJ,e}^{s\varvec{u}} &{} {K}_{IJ,e}^{ss} \end{array} \right] . \end{aligned}$$
(110)

Note that the iteration index is omitted for readability. The submatrices follow from

$$\begin{aligned} \underline{\varvec{K}}_{IJ,e}^{\varvec{u}\varvec{u}}&=\frac{\partial \underline{\varvec{P}}_{I,e}^{\varvec{u}}}{\partial \underline{\varvec{u}}_J}\nonumber \\&=\int \limits _{\varOmega _e}{\left[ \underline{\varvec{B}}_I^{\varvec{u}}\right] ^T\frac{\partial \underline{\varvec{\sigma }}_h}{\partial \underline{{\varvec{\varepsilon }}}_h}\left[ \frac{\partial \underline{{\varvec{\varepsilon }}}_h}{\partial \underline{\varvec{u}}_J}\right] \mathrm{d}V}\nonumber \\&=\int \limits _{\varOmega _e}{\left[ \underline{\varvec{B}}_I^{\varvec{u}}\right] ^T\frac{\partial \underline{\varvec{\sigma }}_h}{\partial \underline{{\varvec{\varepsilon }}}_h}\left[ \underline{\varvec{B}}_J^{\varvec{u}}\right] \mathrm{d}V},\end{aligned}$$
(111)
$$\begin{aligned} \underline{\varvec{K}}_{IJ,e}^{\varvec{u}{s}}&=\frac{\partial \underline{\varvec{P}}_{I,e}^{\varvec{u}}}{\partial {s}_J}\nonumber \\&=\int \limits _{\varOmega _e}{\left[ \underline{\varvec{B}}_I^{\varvec{u}}\right] ^T\frac{\partial \underline{\varvec{\sigma }}_h}{\partial {s_h}}\frac{\partial {s_h}}{\partial {s_J}}\mathrm{d}V}\nonumber \\&=\int \limits _{\varOmega _e}{\left[ \underline{\varvec{B}}_I^{\varvec{u}}\right] ^T\frac{\partial \underline{\varvec{\sigma }}_h}{\partial {s_h}}N_J^s\mathrm{d}V},\end{aligned}$$
(112)
$$\begin{aligned} \underline{\varvec{K}}_{IJ,e}^{{s}\varvec{u}}&=\frac{\partial {{P}}_{I,e}^{s}}{\partial {\underline{\varvec{u}}}_J}\nonumber \\&=\int \limits _{\varOmega _e}{N_I^{{s}}2\underline{s}_h\left[ \frac{\partial {\psi _{\text {e,ts}}}_h}{\partial {\underline{{\varvec{\varepsilon }}}_h}}\right] ^T\left[ \frac{\partial \underline{{\varvec{\varepsilon }}}_h}{\partial \underline{\varvec{u}}_J}\right] \mathrm{d}V}\nonumber \\&=\int \limits _{\varOmega _e}{N_I^{{s}}2\underline{s}_h\left[ \frac{\partial {\psi _{\text {e,ts}}}_h}{\partial {\underline{{\varvec{\varepsilon }}}_h}}\right] ^T\left[ \underline{\varvec{B}}_J^{\varvec{u}}\right] \mathrm{d}V},\end{aligned}$$
(113)
$$\begin{aligned} {K}_{IJ,e}^{{ss}} =&\,\,\frac{\partial {{P}}_{I,e}^{s}}{\partial {s}_J}\nonumber \\ =&\int \limits _{\varOmega _e}\left( \left[ \underline{\varvec{B}}_I^{s}\right] ^T\left[ 2\mathcal{G}_c\epsilon \right] \frac{\partial \underline{\nabla {s}}_h}{\partial {s_J}}\right. \nonumber \\&+\left. N_I\left( 2{\psi _{\text {e,ts}}}_h+\frac{\mathcal{G}_c}{2\epsilon }\right) \frac{\partial \underline{s}_h}{\partial {s_J}}\right) \mathrm{d}V\nonumber \\ =&\int \limits _{\varOmega _e}\left( \left[ \underline{\varvec{B}}_I^{s}\right] ^T\left[ 2\mathcal{G}_c\epsilon \right] \underline{\varvec{B}}_J^{s}\right. \nonumber \\&+\left. N_I\left( 2{\psi _{\text {e,ts}}}_h+\frac{\mathcal{G}_c}{2\epsilon }\right) N_J^s\right) \mathrm{d}V. \end{aligned}$$
(114)

The expressions

$$\begin{aligned} \frac{\partial \underline{\varvec{\sigma }}_h}{\partial \underline{{\varvec{\varepsilon }}}_h}={\varvec{\underline{\mathbb {C}}}}= \left[ \begin{array}{l@{\quad }l@{\quad }l} \mathbb {C}_{11} &{}\mathbb {C}_{12} &{}\mathbb {C}_{13}\\ \mathbb {C}_{21} &{}\mathbb {C}_{22} &{}\mathbb {C}_{23}\\ \mathbb {C}_{31} &{}\mathbb {C}_{32} &{}\mathbb {C}_{33} \end{array} \right] \end{aligned}$$
(115)

with

$$\begin{aligned} \mathbb {C}_{11}&=\mathbb {C}_{22}=\frac{\partial \text {tr}^{-}(\underline{{\varvec{\varepsilon }}}_h)}{\partial \text {tr}(\underline{{\varvec{\varepsilon }}}_h)}K_n\nonumber \\&\quad +(s_h^2+\eta )\left( \frac{\partial \text {tr}^{+}(\underline{{\varvec{\varepsilon }}}_h)}{\partial \text {tr}(\underline{{\varvec{\varepsilon }}}_h)}K_n+\mu \right) ,\nonumber \\ \mathbb {C}_{12}&=\mathbb {C}_{21}=\frac{\partial \text {tr}^{-}(\underline{{\varvec{\varepsilon }}}_h)}{\partial \text {tr}(\underline{{\varvec{\varepsilon }}}_h)}K_n\nonumber \\&\quad +(s_h^2+\eta )\left( \frac{\partial \text {tr}^{+}(\underline{{\varvec{\varepsilon }}}_h)}{\partial \text {tr}(\underline{{\varvec{\varepsilon }}}_h)}K_n-\mu \right) ,\nonumber \\ \mathbb {C}_{33}&=(s_h^2+\eta )\mu ,\nonumber \\ \mathbb {C}_{23}&=\mathbb {C}_{32}=\mathbb {C}_{13}=\mathbb {C}_{31}=0 \end{aligned}$$
(116)

is the material tangent for plane strain, whereas

$$\begin{aligned} \frac{\partial {\psi _{\text {e,ts}}}_h}{\partial {\underline{{\varvec{\varepsilon }}}_h}} =&K_n\text {tr}^{+}(\underline{{\varvec{\varepsilon }}}_h)\frac{\partial \text {tr}^{+}(\underline{{\varvec{\varepsilon }}}_h)}{\partial \underline{{\varvec{\varepsilon }}}_h}\nonumber \\&+\mu \frac{\partial \left( \underline{{\varvec{\varepsilon }}_D}_h^*\cdot \underline{{\varvec{\varepsilon }}_{D}}_h\right) }{\partial \underline{{\varvec{\varepsilon }}}_h}\nonumber \\ =&K_n\text {tr}^{+}(\underline{{\varvec{\varepsilon }}}_h)\underline{\varvec{1}}+2\mu \underline{{\varvec{\varepsilon }}_{D}}_h\end{aligned}$$
(117)

is the tensile/deviatoric part of the stress tensor. The derivative of the stresses with respect to the phase field reads

$$\begin{aligned} \frac{\partial \underline{\varvec{\sigma }}_h}{\partial {s_h}}=2s_h\left( K_n\text {tr}^{+}(\underline{{\varvec{\varepsilon }}}_h)\underline{\varvec{1}}+2\mu \underline{{\varvec{\varepsilon }}_{D}}_h\right) . \end{aligned}$$
(118)

Similar to the element stiffness matrix it is possible to write the mass matrix as

$$\begin{aligned} \underline{\varvec{M}}_e= \left[ \begin{array}{l@{\quad }l@{\quad }l@{\quad }l} \underline{\varvec{M}}_{11,e} &{}\underline{\varvec{M}}_{12,e} &{}\underline{\varvec{M}}_{13,e} &{}\underline{\varvec{M}}_{14,e}\\ \underline{\varvec{M}}_{21,e} &{}\underline{\varvec{M}}_{22,e} &{}\underline{\varvec{M}}_{23,e} &{}\underline{\varvec{M}}_{24,e}\\ \underline{\varvec{M}}_{31,e} &{}\underline{\varvec{M}}_{32,e} &{}\underline{\varvec{M}}_{33,e} &{}\underline{\varvec{M}}_{34,e}\\ \underline{\varvec{M}}_{41,e} &{}\underline{\varvec{M}}_{42,e} &{}\underline{\varvec{M}}_{43,e} &{}\underline{\varvec{M}}_{44,e} \end{array} \right] , \end{aligned}$$
(119)

with submatrices

$$\begin{aligned} \underline{\varvec{M}}_{IJ,e}= \left[ \begin{array}{l@{\quad }l} \underline{\varvec{M}}_{IJ,e}^{\varvec{u}\varvec{u}} &{}\underline{\varvec{M}}_{IJ,e}^{\varvec{u}{s}} \\ \underline{\varvec{M}}_{IJ,e}^{s\varvec{u}} &{}\underline{M}_{IJ,e}^{ss} \end{array} \right] . \end{aligned}$$
(120)

The components of the submatrices follow from

$$\begin{aligned} \underline{\varvec{M}}_{IJ,e}^{\varvec{u}\varvec{u}}&= \frac{\partial \underline{\varvec{P}}_{I,e}^{\varvec{u}}}{\partial \underline{\ddot{\varvec{u}}}_J}\nonumber \\&=\int \limits _{\varOmega _e}{N_I\rho \frac{\partial \ddot{\underline{\varvec{u}}}_h}{\partial \underline{\ddot{\varvec{u}}}_J}\mathrm{d}V\varvec{1}}\nonumber \\&=\int \limits _{\varOmega _e}{N_I\rho {N_J}\mathrm{d}V}\varvec{1},\end{aligned}$$
(121)
$$\begin{aligned} \underline{\varvec{M}}_{IJ,e}^{\varvec{u}{s}}&=\frac{\partial \underline{\varvec{P}}_{I,e}^{\varvec{u}}}{\partial {\ddot{s}_J}}=\underline{\varvec{0}},\end{aligned}$$
(122)
$$\begin{aligned} \underline{\varvec{M}}_{IJ,e}^{s\varvec{u}}&=\frac{\partial {{P}}_{I,e}^{s}}{\partial \underline{\ddot{\varvec{u}}}_J}=\varvec{\underline{0}},\end{aligned}$$
(123)
$$\begin{aligned} \underline{\varvec{M}}_{IJ,e}^{ss}&=\frac{\partial {{P}}_{I,e}^{s}}{\partial \ddot{s}_J}=0. \end{aligned}$$
(124)

Finally the element tangent matrix can be formulated

$$\begin{aligned} \underline{\varvec{S}}_e^{(k)}=\underline{\varvec{K}}_e^{(k)}+\frac{\gamma }{\beta \Delta {t}}\underline{\varvec{D}}_e^{(k)}+\frac{1}{\beta \Delta {t}^2}\underline{\varvec{M}}_e^{(k)} \end{aligned}$$
(125)

and the assembly of all element tangent matrix yields the global tangent matrix

$$\begin{aligned} \underline{\varvec{S}}_{n+1}^{(k)} = \bigcup _{e=1}^{n_e}{\underline{\varvec{S}}_e^{(k)}}. \end{aligned}$$
(126)

1.2 Irreversibility

As proposed in Kuhn [29], the irreversibility of the crack growth is modelled by defining homogeneous Dirichlet boundary conditions for the phase field \(s(\varvec{x},t)\) once the value \(s=0\) is reached for the first time i.e.

$$\begin{aligned} s_{I,n+1}=0 \ \ \text {if} \ \ s_{I,n}=0. \end{aligned}$$
(127)

This is accomplished by a reformulation of the residual and the tangent matrix on element level rather than changing global boundary conditions in order to achieve

$$\begin{aligned} \Delta {s}_{I,n+1}^{(k)}=-{s}_{I,n+1}^{(k)}. \end{aligned}$$
(128)

This causes the phase field to remain zero in subsequent time steps

$$\begin{aligned} s_{I,n+1}^{(k+1)}=s_{I,n+1}^{(k)}+\Delta {s}_{I,n+1}^{(k)} =s_{I,n+1}^{(k)}-{s}_{I,n+1}^{(k)}=0. \end{aligned}$$
(129)

The result (128) is enforced by reformulation of (103)

$$\begin{aligned} \underline{\varvec{R}}_{e,n+1}^{(k)}=\underline{\varvec{S}}_{e,n+1}^{(k)}\Delta \underline{\varvec{d}}_{e,n+1}^{(k)} \end{aligned}$$
(130)

on the element level.

  • In a first step, the column \({\underline{\varvec{S}}_{e,n+1}^{(k)}}^{[:,s_I]}\) of the element tangent matrix is multiplied by \({s}_{I,n+1}^{(k)}\) and added to the residual \(\underline{\varvec{R}}_{e,n+1}^{(k)}\leftarrow \underline{\varvec{R}}_{e,n+1}^{(k)}+{s}_{I,n+1}^{(k)}{\underline{\varvec{S}}_{e,n+1}^{(k)}}^{[:,s_I]}.\)

  • Then the corresponding residual entry to \(s_I\) is replaced by \(-{s}_{I,n+1}^{(k)}\), i.e. \({{R}_{e,n+1}^{(k)}}^{[s_I]}~=~-~{s}_{I,n+1}^{(k)}\).

  • Finally, the row \({\underline{\varvec{S}}_{e,n+1}^{(k)}}^{[s_I,:]}\) and the column \({\underline{\varvec{S}}_{e,n+1}^{(k)}}^{[:,s_I]}\) are set to zero and the entry \({\underline{\varvec{S}}_{e,n+1}^{(k)}}^{[s_I,s_I]}\) is set to \({\underline{\varvec{S}}_{e,n+1}^{(k)}}^{[s_I,s_I]} = 1.\)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Schlüter, A., Willenbücher, A., Kuhn, C. et al. Phase field approximation of dynamic brittle fracture. Comput Mech 54, 1141–1161 (2014). https://doi.org/10.1007/s00466-014-1045-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00466-014-1045-x

Keywords

Navigation