Skip to main content
Log in

Phase-field modeling of ductile fracture

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

Abstract

Phase-field modeling of brittle fracture in elastic solids is a well-established framework that overcomes the limitations of the classical Griffith theory in the prediction of crack nucleation and in the identification of complicated crack paths including branching and merging. We propose a novel phase-field model for ductile fracture of elasto-plastic solids in the quasi-static kinematically linear regime. The formulation is shown to capture the entire range of behavior of a ductile material exhibiting \(J_2\)-plasticity, encompassing plasticization, crack initiation, propagation and failure. Several examples demonstrate the ability of the model to reproduce some important phenomenological features of ductile fracture as reported in the experimental literature.

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
Fig. 24
Fig. 25
Fig. 26
Fig. 27
Fig. 28
Fig. 29
Fig. 30
Fig. 31
Fig. 32
Fig. 33
Fig. 34

Similar content being viewed by others

References

  1. Besson J (2010) Continuum models of ductile fracture: a review. Int J Damage Mech 19(1):3–52

    Article  Google Scholar 

  2. Sumpter JDG, Forbes AT (1992) Constraint based analysis of shallow cracks in mild steel. In: TWI/EWI/IS international conference on shallow crack fracture mechanics, toughness tests and applications, Cambridge

  3. Sumpter JDG (1993) An experimental investigation of the T stress approach. In: Constraint effects in fracture, ASTM STP 1171

  4. ODowd NP, Shih CF (1991) Family of crack-tip fields characterized by a triaxiality parameter I. Structure of fields. J Mech Phys Solids 39(8):9891015

    Google Scholar 

  5. Dawicke DS, Piascik RS, Jr Newman JC (1997) Prediction of stable tearing and fracture of a 2000 series aluminium alloy plate using a CTOA criterion. In: Piascik RS, Newman JC, Dowlings NE (eds) Fatigue and fracture mechanics, ASTM STP 1296, 27:90–104

  6. James MA, Newman JC (2003) The effect of crack tunneling on crack growth: experiments and CTOA analyses. Eng Fract Mech 70:457–468

    Article  Google Scholar 

  7. Mahmoud S, Lease K (2003) The effect of specimen thickness on the experimental characterisation of critical crack tip opening angle in 2024–T351 aluminum alloy. Eng Fract Mech 70:443–456

    Article  Google Scholar 

  8. Trädegard A, Nilsson F, Östlund S (1998) FEM-remeshing technique applied to crack growth problems. Comput Methods Appl Mech Eng 160:115–131

    Article  Google Scholar 

  9. Pineau A (1980) Review of fracture micromechanisms and a local approach to predicting crack resistance in low strength steels. In: Proceedings of ICF5 conference, vol 2, Cannes

  10. Berdin C, Besson J, Bugat S, Desmorat R, Feyel F, Forest S, Lorentz E, Maire E, Pardoen T, Pineau A, Tanguy B (2004) Local approach to fracture. Presses de l’Ecole des Mines, Paris

    Google Scholar 

  11. Gurson AL (1977) Continuum theory of ductile rupture by void nucleation and growth: part I yield criteria and flow rules for porous ductile media. J Eng Mater Technol 99:215

    Article  Google Scholar 

  12. Tvergaard V, Needleman A (1984) Analysis of the cup–cone fracture in a round tensile bar. Acta Metall 32:157169

    Google Scholar 

  13. Lemaitre J (1985) A continuous damage mechanics model for ductile fracture. J Eng Mater Technol 107:8389

    Article  Google Scholar 

  14. Lemaitre J (1996) A course on damage mechanics. Springer, New York

    Book  Google Scholar 

  15. Bazant ZP, Pijaudier-Cabot G (1988) Non local continuum damage. Localization, instability and convergence. J Appl Mech 55:287294

    Google Scholar 

  16. Peerlings RHJ, De Borst R, Brekelmans WAM, De Vree JHP, Spee I (1996) Some Observations on localisation in non-local and gradient damage models. Eur J Mech 15A(6):937–953

    Google Scholar 

  17. Enakoutsa K, Leblond JB, Perrin G (2007) Numerical implementation and assessment of a phenomenological nonlocal model of ductile rupture. Comput Methods Appl Mech Eng 196(1316):19461957

    Google Scholar 

  18. Reusch F, Svendsen B, Klingbeil D (2003) Local and non-local Gurson-based ductile damage and failure modelling at large deformation. Eur J Mech 22A:779792

    Google Scholar 

  19. Moës N, Dolbow J, Belytschko T (1999) A finite element method for crack growth without remeshing. Int J Numer Methods Eng 46(1):131150

    Article  Google Scholar 

  20. Mediavilla J, Peerlings RHJ, Geers MGD (2006) Discrete crack modelling of ductile fracture driven by non-local softening plasticity. Int J Numer Methods Eng 66(4):661688

    Article  Google Scholar 

  21. Ortiz M, Pandolfi A (1999) Finite-deformation irreversible cohesive elements for three-dimensional crack-propagation analysis. Int J Numer Methods Eng 44(9):1267–1282

    Article  Google Scholar 

  22. Alfano G, Crisfield MA (2001) Finite element interface models for the delamination analysis of laminated composites: mechanical and computational issues. Int J Numer Methods Eng 50:17011736

    Google Scholar 

  23. De Lorenzis L, Fernando D, Teng JG (2013) Coupled mixed-mode cohesive zone modeling of interfacial debonding in simply supported plated beams. Int J Solids Struct 50:2477–2494

    Article  Google Scholar 

  24. Roychowdhury YDA, Jr Dodds RH (2002) Ductile tearing in thin aluminum panels: experiments and analyses using large-displacement, 3D surface cohesive elements. Eng Fract Mech 69:983–1002

    Article  Google Scholar 

  25. Cornec A, Scheider I, Schwalbe KH (2003) On the practical application of the cohesive model. Eng Fract Mech 70(14):1963– 1987

    Article  Google Scholar 

  26. Dimitri R, De Lorenzis L, Wriggers P, Zavarise G (2014) NURBS- and T-spline-based isogeometric cohesive zone modeling of interface debonding. Comput Mech 54:369–388

    Article  MathSciNet  Google Scholar 

  27. Scheider I, Brocks W (2003) Simulation of cup–cone fracture using the cohesive model. Eng Fract Mech 70(14):1943–1961

    Article  Google Scholar 

  28. Moes N, Belytschko T (2002) Extended finite element method for cohesive crack growth. Eng Fract Mech 69(7):813–833

    Article  Google Scholar 

  29. Seabra MRR, Sustarić P, Cesar de Sa JMA, Rodić T (2013) Damage driven crack initiation and propagation in ductile metals using XFEM. Comput Mech 52:161–179

    Article  MathSciNet  Google Scholar 

  30. Broumand P, Khoei AR (2013) The extended finite element method for large deformation ductile fracture problems with a non-local damage-plasticity model. Eng Fract Mech 112–113:97–125

    Article  Google Scholar 

  31. Crété JP, Longère P, Cadou JM (2014) Numerical modelling of crack propagation in ductile materials combining the GTN model and X-FEM. Comput Methods Appl Mech Eng 275:204–233

    Article  Google Scholar 

  32. Bourdin B, Francfort GA, Marigo JJ (2000) Numerical experiments in revisited brittle fracture. J Mech Phys Solids 48:797–826

    Article  MathSciNet  Google Scholar 

  33. Kuhn C, Müller R (2008) A phase field model for fracture. Proc Appl Math Mech 8:10223–10224

    Article  Google Scholar 

  34. Kuhn C, Müller R (2010) A continuum phase field model for fracture. Eng Fract Mech 77:3625–3634

    Article  Google Scholar 

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

    Article  Google Scholar 

  36. Miehe C, Welschinger F, Hofacker M (2010) Thermodynamically consistent phase-field models of fracture: variational principles and multi-field FE implementations. Int J Numer Methods Eng 83:1273–1311

    Article  MathSciNet  Google Scholar 

  37. Miehe C, Hofacker M, Welschinger F (2010) A phase field model for rate-independent crack propagation: robust algorithmic implementation based on operator splits. Comput Methods Appl Mech Eng 199:2765–2778

    Article  MathSciNet  Google Scholar 

  38. Borden MJ, Hughes TJR, Landis CM, Verhoosel CV (2014) A higher-order phase-field model for brittle fracture: formulation and analysis within the isogeometric analysis framework. Comput Methods Appl Mech Eng 273:100–118

    Article  MathSciNet  Google Scholar 

  39. Ambati M, Gerasimov T, De Lorenzis L (2014) A review on phase-field models of brittle fracture and a new fast hybrid formulation. Comput Mech. doi:10.1007/s00466-014-1109-y

  40. Hofacker M, Miehe C (2012) A phase field model for ductile to brittle failure mode transition. Proc Appl Math Mech 12:173–174

    Article  Google Scholar 

  41. Ulmer H, Hofacker M, Miehe C (2013) Phase field modeling of brittle and ductile fracture. Proc Appl Math Mech 13:533–536

    Article  Google Scholar 

  42. Duda FP, Ciarbonetti A, Sanchez PJ, Huespe AE (2014) A phase-field/gradient damage model for brittle fracture in elastic–plastic solids. Int J Plast. doi:10.1016/j.ijplas.2014.09.005

  43. Francfort GA, Marigo JJ (1998) Revisiting brittle fractures as an energy minimization problem. J Mech Phys Solids 46:1319–1342

    Article  MathSciNet  Google Scholar 

  44. Bourdin B, Francfort GA, Marigo JJ (2008) The variational approach to fracture. J Elast 91:5–148

    Article  MathSciNet  Google Scholar 

  45. Larsen CJ, Ortner C, Süli E (2010) Existence of solutions to a regularized model of dynamic fracture. Math Models Methods Appl Sci 20:1021–1048

    Article  MathSciNet  Google Scholar 

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

    Article  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

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

    Article  Google Scholar 

  49. 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 93:276–301

    Article  MathSciNet  Google Scholar 

  50. Schlüter A, Willenbücher A, Kuhn C, Müller R (2014) Phase field approximation of dynamic brittle fracture. Comput Mech 54:1141–1161

    Article  MathSciNet  Google Scholar 

  51. Simo JC, Hughes TJR (1998) Computational Inelasticity. Springer, New York

    Google Scholar 

  52. Neto EdS, Perić D, Owen DRJ (2008) Computational methods for plasticity. Wiley, New York

    Book  Google Scholar 

  53. Mediavilla J, Peerlings RHJ, Geers MGD (2006) A robust and consistent remeshing-transfer operator for ductile fracture simulations. Comput Struct 84:604–623

    Article  Google Scholar 

  54. Xue L (2007) Ductile fracture modeling: theory, experimental investigation and numerical verification. Massachusetts Institute of Technology, Cambridge

    Google Scholar 

  55. Xue L (2008) Ductile fracture initiation and propagation modeling using damage plasticity theory. Eng Fract Mech 75:3276–3293

    Article  Google Scholar 

  56. Alves M, Jones N (1999) Influence of hydrostatic stress on failure of axisymmetric notched specimens. J Mech Phys Solids 47:643–667

    Article  Google Scholar 

  57. Amstutz BE, Sutton MA, Dawicke DS, Boone ML (1997) Effects of mixed mode I/II loading and grain orientation on crack initiation and stable tearing in 2024–T3 aluminum. Fatigue Fract Mech 27:24–217

    Google Scholar 

  58. Arcan M, Hashin Z, Voloshin A (1978) Method to produce uniform plane stress states with applications to fiber-reinforced materials. Exp Mech 18:6–141

    Article  Google Scholar 

  59. Nayfeh AH (2011) Introduction to perturbation techniques. Wiley-VCH, Weinheim

    Google Scholar 

  60. Gurtin ME (1996) Generalized Ginzburg–Landau and Cahn–Hilliard equations based on a microforce balance. Phys D 92(3–4):178–192

    Article  MathSciNet  Google Scholar 

  61. Gurtin M, Fried E, Anand L (2010) The mechanics and thermodynamics of continua. Cambridge University Press, New York

    Book  Google Scholar 

Download references

Acknowledgments

This research was funded by the European Research Council, ERC Starting Researcher Grant INTERFACES, Grant Agreement N. 279439.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to T. Gerasimov.

Appendix

Appendix

1.1 Perturbation analysis for the \(s\) and \(p\) mutual interaction

One of the most important aspects in our phase-field model of elasto-plastic ductile fracture is the coupling function \(g(s,p):=s^{2p}\). Introduction of the exponent \(p\), which accounts for the accumulation and localization of plastic strains, into the classic brittle-case function \(g(s)=s^{2}\) allows us to let \(s\) explicitly depend also on the plastic energy density \(\Psi _\mathrm{p}\), and not only on the elastic one, \(\Psi _\mathrm{e}\). This can be formally seen from the corresponding evolution equation for \(s\) (25). More importantly, we show below that the evolution of \(s\) in this case will mainly be governed by \(p\sim \Psi _\mathrm{p}\) rather than \(\Psi _\mathrm{e}\), thus delaying the brittle-like fracture development and enabling the desired plastic-driven fracture mechanism.

Without loss of generality, we restrict ourselves to the pure tensile loading situation so that no split of the elastic energy density function \(\Psi _\mathrm{e}\) is required and the evolution equation for \(s\) is given by

$$\begin{aligned} -4\ell ^2\Delta s+s+\frac{4\ell }{G_c}\Psi _\mathrm{e}(\varvec{\varepsilon })p s^{2p-1}=1. \end{aligned}$$
(31)

Equation (31) is a non-linear partial differential equation whose explicit analytical solution for \(s\) is out of reach: one can only conclude that with \(p\equiv 0\) (zero plastic strains), a trivial solution is \(s=1\) (no phase-field development occurs). With a suitable rescaling of variables (i.e. non-dimensionalization), one can obtain a dimensionless version of (31) containing a small (perturbation) parameter. An approximate analytical solution of such perturbed non-linear equation can then be constructed explicitly, establishing a relation between \(s\) and \(p\).

Let \(L\) be the maximal macroscopic dimension of the considered body. We introduce the following non-dimensional quantities:

$$\begin{aligned} (\bar{x},\bar{y}):=\left( \frac{x}{L},\frac{y}{L}\right) , \quad \bar{\varvec{u}}:=\frac{\varvec{u}}{A}, \quad \bar{\mathbb C}:=\frac{\mathbb C}{E}, \quad \bar{\ell }:=\frac{\ell }{\frac{1}{2}L}, \end{aligned}$$
(32)

where \(A>0\) is a constant whose particular choice will be motivated below. The scaled strain tensor \(\bar{\varvec{\varepsilon }}=\frac{1}{2}(\bar{\nabla }\bar{\varvec{u}}+\bar{\nabla }\bar{\varvec{u}}^T)\), with \(\bar{\nabla }\) denoting the gradient w.r.t. the non-dimensional variables \((\bar{x},\bar{y})\), relates to the actual strain tensor \(\varvec{\varepsilon }\) as

$$\begin{aligned} \varvec{\varepsilon }(\varvec{u})=\frac{A}{L}\bar{\varvec{\varepsilon }}(\bar{\varvec{u}}), \end{aligned}$$
(33)

yielding for \(\Psi _\mathrm{e}\) the relation

$$\begin{aligned} \Psi _\mathrm{e}(\varvec{\varepsilon }) =\frac{1}{2}\mathbb {C}\varvec{\varepsilon }^2 =\frac{1}{2} E \bar{\mathbb {C}} \left( \frac{A}{L} \right) ^2 \bar{\varvec{\varepsilon }}^2 =E\left( \frac{A}{L}\right) ^2 \bar{\Psi }_\mathrm{e}(\bar{\varvec{\varepsilon }}). \end{aligned}$$
(34)

Also, we set \(\bar{s}:=s\) so that \(\Delta s=\frac{1}{L^2}\bar{\Delta }\bar{s}\), with \(\bar{\Delta }\) standing for the Laplace operator w.r.t. \((\bar{x},\bar{y})\). Finally, a non-dimensional plastic strain variable is defined as \(\overline{\varvec{\varepsilon }^\mathrm{p}}:=\varvec{\varepsilon }^\mathrm{p}\) leading to the corresponding definition of \(\bar{p}\):

$$\begin{aligned} \bar{p}:=\frac{ \overline{\varepsilon ^\mathrm{p}_\mathrm{eq}} }{\varepsilon ^\mathrm{p}_\mathrm{eq,crit}}, \quad \overline{\varepsilon ^\mathrm{p}_\mathrm{eq}}(t):= \sqrt{\frac{2}{3}}\int _{0}^t \sqrt{ \dot{\overline{\varvec{\varepsilon }^\mathrm{p}}}:\dot{\overline{\varvec{\varepsilon }^\mathrm{p}}} } \;d\tau . \end{aligned}$$
(35)

Note that the denominator remains the same as in the definition of \(p\).

Using (32)–(35) we turn (31) into

$$\begin{aligned} -\bar{\ell }^2\bar{\Delta }\bar{s}+\bar{s} +\frac{2\bar{\ell } L}{G_c} E \left( \frac{A}{L} \right) ^2 \bar{\Psi }_\mathrm{e}(\bar{\varvec{\varepsilon }}) \bar{p} \bar{s}^{2\bar{p}-1}=1. \end{aligned}$$

In the above, the scaling constant \(A\) can be chosen to provide \(\frac{2L}{G_c} E \left( \frac{A}{L} \right) ^2=1\), that is, \(A:=\sqrt{\frac{G_c L}{2E}}\), and the dimensionless counterpart of (31) we end up with reads as

$$\begin{aligned} -\bar{\ell }^2\bar{\Delta }\bar{s}+\bar{s} +\bar{\ell } \bar{\Psi }_\mathrm{e}(\bar{\varvec{\varepsilon }}) \bar{p} \bar{s}^{2\bar{p}-1}=1. \end{aligned}$$
(36)

In what follows, we set \(\gamma :=\bar{\ell }\) and also drop the bars over all remaining variables in (36). This yields

$$\begin{aligned} -\gamma ^2\Delta s+s+\gamma \Psi _\mathrm{e}(\varvec{\varepsilon }) p s^{2p-1}=1, \end{aligned}$$
(37)

where, by definition, \(\gamma \) is a small (perturbation) parameter, that is, \(0<\gamma \ll 1\).

According to the perturbation analysis framework [59], an asymptotic expansion of \(s\) is taken in a form of a power series in \(\gamma \):

$$\begin{aligned} s=s_0+\gamma s_1+\gamma ^2 s_2+\cdots , \end{aligned}$$
(38)

what yields for \(\Delta s\) and \(s^{2p-1}\) the expansions

$$\begin{aligned} \Delta s=\Delta s_0+\gamma \Delta s_1+\gamma ^2 \Delta s_2+\cdots , \end{aligned}$$
(39)

and

$$\begin{aligned} s^{2p-1}&= s_0^{2p-1}+\gamma (2p-1)s_0^{2p-2}s_1 \nonumber \\&\quad \ + \gamma ^2(2p-1)s_0^{2p-3}\left( s_0s_2+s_1^2(p-1)\right) +\cdots , \end{aligned}$$
(40)

respectively. In (39) and (40), only the terms up to order \(O(\gamma ^2)\) have been retained, consistently with the assumed expansion (38). We now substitute (38)–(40) into the governing equation (37) and collect the coefficients of like powers of \(\gamma \) to obtain:

$$\begin{aligned}&s_0+\gamma \left( s_1+p\Psi _\mathrm{e}s_0^{2p-1} \right) \nonumber \\&\quad + \gamma ^2 \left( -\Delta s_0+s_2+p(2p-1)\Psi _\mathrm{e}s_0^{2p-2}s_1 \right) +\cdots =1. \end{aligned}$$
(41)

Equating the coefficients of each power of \(\gamma \) to zero we ’extract’ the equations for \(s_0,s_1\) and \(s_2\). The corresponding solutions are

$$\begin{aligned} s_0=1, \quad s_1=-p\Psi _\mathrm{e}, \quad s_2=p^2(2p-1)\Psi _\mathrm{e}^2, \end{aligned}$$
(42)

and, due to (38), we finally obtain

$$\begin{aligned} s = 1-\gamma p\Psi _\mathrm{e} + \gamma ^2 p^2(2p-1)\Psi _\mathrm{e}^2+\cdots , \end{aligned}$$
(43)

to be treated as an approximate analytical solution of equation (37). Note that with \(\gamma \rightarrow 0\), equation (37) reduces to \(s=1\). Sending \(\gamma \) to zero also in (43), we recover the same result, i.e. \(s=1\), as expected. Recall also that setting \(p=0\) in (37), the trivial solution of the reduced equation is \(s=1\). Similar result is obtained while substituting \(p=0\) in (43).

From (43) the influence of \(p\) on the phase-field \(s\) can be grasped. In the linear elastic regime (no plastic strains, \(p=0\)) and at the beginning of the plastic regime (plastic strains are still negligibly small, \(0<p\ll 1\)), the evolution of \(s\) is minor even with a (possibly large) non-zero contribution of \(\Psi _\mathrm{e}\). In other words, the presence of \(p\) in \(g(s,p)=s^{2p}\) delays the brittle-like fracture formation. This is in a contrast to the brittle phase-field formulation, when the use of \(g(s)=s^2\) results in \(s\) depending solely on \(\Psi _\mathrm{e}\) (expansion (43) with \(p\equiv 1\)) meaning that the crack development occurs independently of the evolution of the plastic strains.

1.2 Proof of thermodynamic consistency

In this Appendix, the governing equations of the proposed model are rewritten in the framework of the principle of virtual power as developed by Gurtin, see [60, 61], and their thermodynamic consistency is investigated.

1.2.1 Kinematics

The kinematics is based on the decomposition of the displacement gradient into elastic and plastic components \(\nabla \mathbf {u}=\mathbf {H}^\mathrm{e}+\mathbf {H}^\mathrm{p}\). Correspondingly, we define the elastic and plastic strain tensors as

$$\begin{aligned} {\varvec{\varepsilon }}^\mathrm{e}=\frac{1}{2}\left( \mathbf {H}^\mathrm{e}+\mathbf {H}^{\mathrm{e}T}\right) \quad {\varvec{\varepsilon }}^\mathrm{p}=\frac{1}{2}\left( \mathbf {H}^\mathrm{p}+\mathbf {H}^{\mathrm{p}T}\right) \end{aligned}$$
(44)

so that the total strain tensor is given by \({\varvec{\varepsilon }}={\varvec{\varepsilon }}^\mathrm{e}+{\varvec{\varepsilon }}^\mathrm{p}\). Finally, we assume that \(\mathbf {H}^\mathrm{p}\) is purely deviatoric, i.e. \(\mathrm{tr}\,\mathbf {H}^\mathrm{p}=\mathrm{tr}\,{\varvec{\varepsilon }}^\mathrm{p}=0\).

1.2.2 Balance equations

Let us introduce the internal power in the form

$$\begin{aligned} \mathcal {I}\left( P\right)= & {} \int _{P}{{\varvec{\sigma }}}^\mathrm{e}:\dot{\mathbf {H}}^\mathrm{e}\;dv+\int _{P}{\varvec{\sigma }}^\mathrm{p}:\dot{\mathbf {H}}^\mathrm{p}\;dv+\int _{P}{\varvec{\xi }}\cdot \nabla \dot{s}\;dv\nonumber \\&+ \int _{P}\zeta \cdot \dot{s}\;dv \end{aligned}$$
(45)

where \({{\varvec{\sigma }}}^\mathrm{e}\) and \({{\varvec{\sigma }}}^\mathrm{p}\) are respectively an elastic stress, power-conjugate to \(\dot{\mathbf {H}}^\mathrm{e}\), and a plastic stress, power-conjugate to \(\dot{\mathbf {H}}^\mathrm{p}\). Since \(\mathbf {H}^\mathrm{p}\) is deviatoric, we may assume without loss of generality that \({{\varvec{\sigma }}}^\mathrm{p}\) is deviatoric, i.e. that \(\mathrm{tr}{\varvec{\sigma }}^\mathrm{p}=0\). Moreover, \({\varvec{\xi }}\) is the microscopic stress power-conjugate to \(\nabla \dot{s}\), and \(\zeta \) is the microscopic internal body force power-conjugate to \(\dot{s}\), where \(s\) is the crack field parameter, \(0\le s\le 1\). The integration domain is any subregion \(P\) of the body under consideration. The external power is given by

$$\begin{aligned} \mathcal {W}\left( P\right)= & {} \int _{\partial P}\varvec{t}\left( \varvec{n}\right) \cdot \dot{\varvec{u}}\;da+\int _{P}\varvec{b}\cdot \dot{\varvec{u}}\;dv\nonumber \\&+\int _{\partial P}\chi \left( \varvec{n}\right) \dot{s}\;da+\int _{P}\gamma \dot{s}\;dv \end{aligned}$$
(46)

where \(\varvec{t}\) is the traction vector on the elementary area \(da\) of the surface of \(P,\, \partial P\), with outward unit normal \(\varvec{n},\, \varvec{b}\) is the body force, \(\chi \) and \(\gamma \) are respectively the microscopic external traction and the microscopic external body force, both power-conjugate to \(\dot{s}\). Considering a generalized virtual velocity

$$\begin{aligned} \mathcal {V}=\left( \tilde{\varvec{u}},\tilde{\mathbf {H}}^\mathrm{e},\tilde{\mathbf {H}}^\mathrm{p},\tilde{s}\right) \end{aligned}$$
(47)

satisfying the kinematical constraints above, the principle of virtual power reads

$$\begin{aligned}&\int _{P}{{\varvec{\sigma }}}^\mathrm{e}:\tilde{\mathbf {H}}^\mathrm{e}\;dv\!+\!\int _{P}{{\varvec{\sigma }}}^\mathrm{p}:\tilde{\mathbf {H}}^\mathrm{p}\;dv\!+\!\int _{P}{\varvec{\xi }}\cdot \nabla \tilde{s}\;dv\!+\!\int _{P}\zeta \cdot \tilde{s}\;dv\nonumber \\&=\int _{\partial P}\varvec{t}\left( \varvec{n}\right) \cdot \varvec{\tilde{u}}\;da\!+\!\int _{P}\varvec{b}\cdot \varvec{\tilde{u}}\;dv\!+\!\int _{\partial P}\chi \left( \varvec{n}\right) \tilde{s}\;da\!+\!\int _{P}\gamma \tilde{s}\;dv\nonumber \\ \end{aligned}$$
(48)

for any subregion \(P\) of the body and any \(\mathcal {V}\). Frame invariance implies that \({\varvec{\sigma }}^\mathrm{e}\) is symmetric, therefore \({\varvec{\sigma }}^\mathrm{e}:\tilde{\mathbf {H}}^\mathrm{e}={\varvec{\sigma }}^\mathrm{e}:\tilde{{{\varvec{\varepsilon }}}}^\mathrm{e}\). Applying Eq. (48) with \(\mathcal {V}=\left( \tilde{\varvec{u}},\nabla \tilde{\varvec{u}}, \mathbf {0},0\right) \) leads to the macroscopic force balance \(\mathrm{div}\varvec{\sigma }+\varvec{b}=\mathbf {0}\) and to the expression of the macroscopic traction \(\varvec{t}\left( \varvec{n}\right) ={\varvec{\sigma }}\cdot \varvec{n}\), where we have introduced \({\varvec{\sigma }}:={\varvec{\sigma }}^\mathrm{e}\) as the (standard) Cauchy stress. Considering next a virtual velocity \(\mathcal {V}=\left( \mathbf {0},-\tilde{\mathbf {H}}^\mathrm{p},{\tilde{\mathbf {H}}}^\mathrm{p},0\right) \) delivers the plastic microscopic force balance [61], \({\varvec{\sigma }}_{dev}={\varvec{\sigma }}^\mathrm{p}\), where \({\varvec{\sigma }}_{dev}\) is the deviatoric component of the stress tensor. Note that thus also \({\varvec{\sigma }}^\mathrm{p}\) is symmetric, therefore \({\varvec{\sigma }}^\mathrm{p}:\tilde{\mathbf {H}}^\mathrm{p}={\varvec{\sigma }}^\mathrm{p}:\tilde{{{\varvec{\varepsilon }}}}^\mathrm{p}\). Finally, insertion in Eq. (48) of a virtual velocity \(\mathcal {V}=\left( \mathbf {0},\mathbf {0},\mathbf {0},\tilde{s}\right) \) yields

$$\begin{aligned} \int _{P}{\varvec{\xi }}\cdot \nabla \tilde{s}\;dv+\int _{P}\zeta \cdot \tilde{s}\;dv=\int _{\partial P}\chi \left( \varvec{n}\right) \tilde{s}\;da+\int _{P}\gamma \tilde{s}\;dv. \end{aligned}$$
(49)

Through the divergence theorem, Eq. (49) leads to the phase-field microscopic force balance

$$\begin{aligned} \mathrm{div}{\varvec{\xi }}-\zeta +\gamma =0 \end{aligned}$$
(50)

and to the expression of the phase-field microscopic traction \(\chi \left( \varvec{n}\right) ={{\varvec{\xi }}}\cdot \varvec{n}\).

As a consequence of the above, and if we further introduce the codirectionality constraint valid for \(J_{2}\) plasticity, \(\frac{\dot{{\varvec{\varepsilon }}}^\mathrm{p}}{\left\| \dot{{\varvec{\varepsilon }}}^\mathrm{p}\right\| }=\varvec{n}^\mathrm{p}\) with \(\varvec{n}^\mathrm{p}=\frac{{\varvec{\sigma }}_{dev}}{\left\| {\varvec{\sigma }}_{dev}\right\| }\), the power balance takes the final form

$$\begin{aligned}&\int _{P}{{\varvec{\sigma }}}:\dot{{\varvec{\varepsilon }}}^\mathrm{e}\;dv+\int _{P}\tau ^\mathrm{p}\dot{e}^\mathrm{p}\;dv+\int _{P}{\varvec{\xi }}\cdot \nabla \dot{s}\;dv+\int _{P}\zeta \cdot \dot{s}\;dv\\&= \int _{\partial P}\varvec{t}\left( \varvec{n}\right) \cdot \dot{\varvec{u}}\;da+\int _{P}\varvec{b}\cdot \dot{\varvec{u}}\;dv\!+\!\int _{\partial P}\chi \left( \varvec{n}\right) \dot{s}\;da\!+\!\int _{P}\!\gamma \dot{s}\;dv, \end{aligned}$$

where

$$\begin{aligned} \tau ^\mathrm{p}:=\left\| {\varvec{\sigma }}_{dev}\right\| , \quad \dot{e}^\mathrm{p}:=\left\| \dot{{\varvec{\varepsilon }}}^\mathrm{p}\right\| \ge 0, \quad e^\mathrm{p}\left( t\right) =\int _{0}^{t}\left\| \dot{{\varvec{\varepsilon }}}^\mathrm{p}\right\| \;d\tau \end{aligned}$$
(51)

1.2.3 Dissipation inequality and constitutive laws

Based on the formulation in Sect. 1, the dissipation inequality can be written (in local form) as

$$\begin{aligned} \dot{E}_\ell -{\varvec{\sigma }}:\dot{{\varvec{\varepsilon }}}^\mathrm{e}-\tau ^\mathrm{p}\dot{e}^\mathrm{p}-{\varvec{\xi }}\cdot \nabla \dot{s}-\zeta \cdot \dot{s}=-D\le 0 \end{aligned}$$
(52)

with \(E_\ell \) as the free energy and \(D\) as the dissipation rate, both per unit volume. Let us consider the following form of the free energy

$$\begin{aligned} E_\ell =E_\ell \left( {\varvec{\varepsilon }}^\mathrm{e},e^\mathrm{p},\alpha ,s,\nabla s\right) . \end{aligned}$$
(53)

This postulated form contains the “standard” dependencies on the elastic strain \({\varvec{\varepsilon }}^\mathrm{e}\) and on the hardening variable \(\alpha \), plus further dependencies on the phase field and on its gradient, as well as on \(e^\mathrm{p}\). The latter is a scalar measure of plastic strain accumulation defined in Eq. (51). The dependency of the free energy on \(e^\mathrm{p}\) is needed to realize a coupling between the evolution of the phase-field parameter and the accumulation of plastic strains, and is an essential ingredient of the proposed model. Substitution of Eq. (53) in the dissipation inequality (52) leads to

$$\begin{aligned}&\left( \frac{\partial E_\ell }{\partial {\varvec{\varepsilon }}^\mathrm{e}}-{\varvec{\sigma }}\right) :\dot{{\varvec{\varepsilon }}}^\mathrm{e}+\left( \frac{\partial E_\ell }{\partial e^\mathrm{p}}-\tau ^\mathrm{p}\right) \dot{e}^\mathrm{p}+\frac{\partial E_\ell }{\partial \alpha }\dot{\alpha }\nonumber \\&+\left( \frac{\partial E_\ell }{\partial s}-\zeta \right) \dot{s}+\left( \frac{\partial E_\ell }{\partial \nabla s}-{\varvec{\xi }}\right) \cdot \nabla \dot{s}=-D\le 0. \end{aligned}$$
(54)

Well-known arguments lead immediately to the elastic stress-strain relationship

$$\begin{aligned} {\varvec{\sigma }}=\frac{\partial E_\ell }{\partial {\varvec{\varepsilon }}^{e}}. \end{aligned}$$
(55)

Additional consequences can be driven by applying the inequality individually to the group of terms related to the phase field

$$\begin{aligned} \left( \frac{\partial E_\ell }{\partial s}-\zeta \right) \dot{s}+\left( \frac{\partial E_\ell }{\partial \nabla s}-{\varvec{\xi }}\right) \cdot \nabla \dot{s}\le 0. \end{aligned}$$
(56)

From (56) follow the phase-field microscopic constitutive equations

$$\begin{aligned} {\varvec{\xi }}=\frac{\partial E_\ell }{\partial \nabla s}, \qquad \zeta =\frac{\partial E_\ell }{\partial s}. \end{aligned}$$
(57)

Substituting eqs. (57) into (50) leads to the phase-field evolution equation

$$\begin{aligned} \mathrm{div}\left( \frac{\partial E_\ell }{\partial \nabla s}\right) -\frac{\partial E_\ell }{\partial s}=0, \end{aligned}$$
(58)

where we have further assumed \(\gamma =0\).

As a result of Eq. (55) and inequality (56), and introducing the thermodynamic force power-conjugate to \(\dot{\alpha }\),

$$\begin{aligned} t_{\alpha }:=-\frac{\partial E_\ell }{\partial \alpha }, \end{aligned}$$
(59)

the following reduced dissipation inequality is obtained from (54)

$$\begin{aligned} D:=\tau ^\mathrm{p}\dot{e}^\mathrm{p}+t_{\alpha }\dot{\alpha }-\frac{\partial E_\ell }{\partial e^\mathrm{p}}\dot{e}^\mathrm{p}\ge 0. \end{aligned}$$
(60)

Note that, being \(\dot{e}^\mathrm{p}\ge 0\), inequality (60) can be further reduced to its “classical” version

$$\begin{aligned} \tau ^\mathrm{p}\dot{e}^\mathrm{p}+t_{\alpha }\dot{\alpha }\ge 0, \end{aligned}$$
(61)

(whose satisfaction is ensured by the “classical” flow rule and hardening evolution equation of \(J_{2}\)-plasticity used in this work) provided that the following inequality holds

$$\begin{aligned} \frac{\partial E_\ell }{\partial e^\mathrm{p}}\le 0. \end{aligned}$$
(62)

From the specific choice of the free energy in Eq. (16) follows \(\frac{\partial E_\ell }{\partial e^\mathrm{p}}=\frac{\partial g}{\partial e^\mathrm{p}}\Psi _\mathrm{e}^{+}({\varvec{\varepsilon }}^\mathrm{e})\) and therefore inequality (62) holds [and the reduced dissipation inequality takes the form (61)] provided that

$$\begin{aligned} \frac{\partial g}{\partial e^\mathrm{p}}\le 0. \end{aligned}$$
(63)

A possible specific choice of degradation function \(g\) complying with (63) is the following

$$\begin{aligned} g\left( s,e^\mathrm{p}\right) =s^{2e^\mathrm{p}/e_\mathrm{crit}^\mathrm{p}}+\eta \end{aligned}$$
(64)

with coincides with the one in Eq. (17) by taking

$$\begin{aligned} e_\mathrm{crit}^\mathrm{p}=\sqrt{\frac{3}{2}}\varepsilon _\mathrm{eq,crit}^\mathrm{p} \end{aligned}$$
(65)

This proves the thermodynamic consistency of the proposed model.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ambati, M., Gerasimov, T. & De Lorenzis, L. Phase-field modeling of ductile fracture. Comput Mech 55, 1017–1040 (2015). https://doi.org/10.1007/s00466-015-1151-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00466-015-1151-4

Keywords

Navigation