Skip to main content
Log in

A model-reduction approach to the micromechanical analysis of polycrystalline materials

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

Abstract

The present study is devoted to the extension to polycrystals of a model-reduction technique introduced by the authors, called the nonuniform transformation field analysis (NTFA). This new reduced model is obtained in two steps. First the local fields of internal variables are decomposed on a reduced basis of modes as in the NTFA. Second the dissipation potential of the phases is replaced by its tangent second-order (TSO) expansion. The reduced evolution equations of the model can be entirely expressed in terms of quantities which can be pre-computed once for all. Roughly speaking, these pre-computed quantities depend only on the average and fluctuations per phase of the modes and of the associated stress fields. The accuracy of the new NTFA-TSO model is assessed by comparison with full-field simulations on two specific applications, creep of polycrystalline ice and response of polycrystalline copper to a cyclic tension-compression test. The new reduced evolution equations is faster than the full-field computations by two orders of magnitude in the two examples.

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

Similar content being viewed by others

Notes

  1. Another possible and frequent choice is to choose the plastic slips \(\gamma _s\) as internal variables, at the expense of a larger number of variables.

  2. Other boundary conditions can be considered provided the Hill-Mandel condition is satisfied (Suquet [41]).

  3. Alternatively one could prescribe the path of macroscopic stress.

References

  1. Armstrong P, Frederick C (1966) A mathematical representation of the multiaxial Bauschinger effect. Central Electricity Generating Board and Berkeley Nuclear Laboratories, Research & Development Department Report RD/B/N731, reprinted in. Mat High Temp 24(2007):11–26

    Google Scholar 

  2. Asaro R (1983) Micromechanics of crystals and polycrystals. In: Hutchinson J, Wu T (eds) Advances in applied mechanics. Academic Press, New-York, pp 1–114

    Google Scholar 

  3. Ashby M, Duval P (1985) The creep of polycrystalline ice. Cold Reg Sci Technol 11:285–300

    Article  Google Scholar 

  4. Castelnau O, Canova G, Lebensohn R, Duval P (1997) Modelling viscoplastic behavior of anisotropic polycrystalline ice with a self-consistent approach. Acta Mater 45:4823–4834

    Article  Google Scholar 

  5. Castelnau O, Duval P, Montagnat M, Brenner R (2008) Elastoviscoplastic micromechanical modeling of the transient creep of ice. J Geophys Res 113:B11203

    Article  Google Scholar 

  6. Chinesta F, Cueto E (2014) PGD-based modeling of materials, structures and processes. Springer, Heidelberg

    Book  Google Scholar 

  7. Duval P, Ashby M, Anderman I (1983) Rate controlling processes in the creep of polycrystalline ice. J Phys Chem 87:4066–4074

    Article  Google Scholar 

  8. Dvorak G (1992) Transformation field analysis of inelastic composite materials. Proc R Soc Lond A 437:311–327

    Article  MathSciNet  MATH  Google Scholar 

  9. Dvorak G, Bahei-El-Din Y, Wafa A (1994) The modeling of inelastic composite materials with the transformation field analysis. Model Simul Mater Sci Eng 2:571–586

    Article  MATH  Google Scholar 

  10. Feyel F, Chaboche JL (2000) FE2 multiscale approach for modelling the elastoviscoplastic behaviour of long fibre SiC/Ti composite materials. Comput Methods Appl Mech Eng 183:309–330

    Article  MATH  Google Scholar 

  11. Fish J, Shek K, Pandheeradi M, Shepard M (1997) Computational plasticity for composite structures based on mathematical homogenization: Theory and practice. Comput Methods Appl Mech Eng 148:53–73

    Article  MathSciNet  MATH  Google Scholar 

  12. Fish J, Yu Q (2002) Computational mechanics of fatigue and life predictions for composite materials and structures. Comput Methods Appl Mech Eng 191:4827–4849

    Article  MATH  Google Scholar 

  13. Fritzen F, Böhlke T (2010) Three-dimensional finite element implementation of the nonuniform transformation field analysis. Int J Numer Methods Eng 84:803–829

    Article  MathSciNet  MATH  Google Scholar 

  14. Fritzen F, Leuschner M (2013) Reduced basis hybrid computational homogenization based on a mixed incremental formulation. Comput Methods Appl Mech Eng 260:143–154

    Article  MathSciNet  MATH  Google Scholar 

  15. Gérard C (2008) Field measurements and identification of crystal plasticity models. Ph.D. thesis, Université Paris-Nord—Paris XIII. https://tel.archives-ouvertes.fr/tel-00315792

  16. Gérard C, N’Guyen F, Osipov N, Cailletaud G, Bornert M, Caldemaison D (2009) Comparison of experimental results and finite element simulation of strain localization scheme under cyclic loading. Comput Mater Sci 46(3):755–760

    Article  Google Scholar 

  17. Germain P, Nguyen QS, Suquet P (1983) Continuum thermodynamics. J Appl Mech 50:1010–1020

    Article  MATH  Google Scholar 

  18. Halphen B, Nguyen QS (1975) Sur les matériaux standard généralisés. J Mécanique 14:39–63

    MATH  Google Scholar 

  19. Herrera-Solaz V, Llorca J, Dogan E, Karaman I, Segurado J (2014) An inverse optimization strategy to determine single crystal mechanical behavior from polycrystal tests: Application to AZ31 Mg alloy. Int J Plasticity 57:1–15

    Article  Google Scholar 

  20. Holmes P, Lumley J, Berkooz G (1996) Structures, dynamical systems and symmetry. Cambridge University Press, Cambridge

    MATH  Google Scholar 

  21. Homescu C, Petzold L, Serban R (2007) Error estimation for reduced-order models of dynamical systems. SIAM Rev 49:277–299

    Article  MathSciNet  MATH  Google Scholar 

  22. Kanouté P, Boso D, Chaboche JL, Schrefler B (2009) Multiscale methods for composites: a review. Arch Comput Methods Eng 16:31–75

    Article  MATH  Google Scholar 

  23. Kattan P, Voyiadjis G (1993) Overall damage and elastoplastic deformation in fibrous metal matrix composites. Int J Plasticity 9:931–949

    Article  MATH  Google Scholar 

  24. Largenton R, Michel JC, Suquet P (2014) Extension of the nonuniform transformation field analysis to linear viscoelastic composites in the presence of aging and swelling. Mech. Mater. 73:76–100. doi:10.1016/j.mechmat.2014.02.004

    Article  Google Scholar 

  25. Lebensohn R (2001) N-site modelling of a 3d viscoplastic polycrystal using fast Fourier transforms. Acta Mater. 49:2723–2737

    Article  Google Scholar 

  26. Lucia D, Beran P, Silva W (2004) Reduced-order modeling: new approaches for computational physics. Progr Aerospace Sci 40:51–117

    Article  Google Scholar 

  27. Méric L, Cailletaud G (1991) Single crystal modeling for structural calculations. part 2: finite element implementation. J Eng Mat Technol 113:171–182

    Article  Google Scholar 

  28. Michel JC, Galvanetto U, Suquet P (2000) Constitutive relations involving internal variables based on a micromechanical analysis. In: Maugin G, Drouot R, Sidoroff F (eds) Continuum Thermomechanics: the art and science of modelling material behaviour. Klüwer Academic Publishers, Boston, pp 301–312

    Google Scholar 

  29. Michel JC, Moulinec H, Suquet P (1999) Effective properties of composite materials with periodic microstructure: a computational approach. Comput Methods Appl Mech Eng 172:109–143

    Article  MathSciNet  MATH  Google Scholar 

  30. Michel JC, Suquet P (2003) Nonuniform transformation field analysis. Int J Solids Struct 40:6937–6955

    Article  MathSciNet  MATH  Google Scholar 

  31. Michel JC, Suquet P (2004) Computational analysis of nonlinear composite structures using the nonuniform transformation field analysis. Comput Methods Appl Mech Eng 193:5477–5502

    Article  MathSciNet  MATH  Google Scholar 

  32. Michel JC, Suquet P (2009) Nonuniform transformation field analysis: a reduced model for multiscale nonlinear problems in Solid Mechanics. In: Aliabadi F, Galvanetto U (Eds.) Multiscale modelling in solid mechanics—computational approaches. Imperial College Press, pp. 159–206, chapter 4

  33. Michel JC, Suquet P (2015) A model-reduction approach in micromechanics of materials preserving the variational structure of constitutive relations. J. Mech. Phys. Solids Submitted

  34. Moulinec H, Suquet P (1998) A numerical method for computing the overall response of nonlinear composites with complex microstructure. Comput Methods Appl Mech Eng 157:69–94

    Article  MathSciNet  MATH  Google Scholar 

  35. Pellegrino C, Galvanetto U, Schrefler B (1999) Numerical homogenization of periodic composite materials with non-linear material components. Int J Numer Methods Eng 46:1609–1637

    Article  MATH  Google Scholar 

  36. Rice J (1970) On the structure of stress-strain relations for time-dependent plastic deformation in metals. J Appl Mech 37:728–737

    Article  Google Scholar 

  37. Rice J (1971) Inelastic constitutive relations for solids: an internal-variable theory and its application to metal plasticity. J Mech Phys Solids 19:433–455

    Article  MATH  Google Scholar 

  38. Roters F, Eisenlohr P, Bieler T, Raabe D (2010) Crystal plasticity finite element methods. materials science and engineering. Wiley-VCH, Weinheim

    Book  Google Scholar 

  39. Ryckelynck D, Benziane D (2010) Multi-level a priori Hyper-reduction of mechanical models involving internal variables. Comput Methods Appl Mech Eng 199:1134–1142

    Article  MathSciNet  MATH  Google Scholar 

  40. Sirovich L (1987) Turbulence and the dynamics of coherent structures. Q Appl Math 45:561–590

    MathSciNet  MATH  Google Scholar 

  41. Suquet P (1987) Elements of homogenization for inelastic solid mechanics. In: Sanchez-Palencia E, Zaoui A (eds) Homogenization techniques for composite media. Vol. 272 of Lecture Notes in Physics. Springer Verlag, New York, pp 193–278

    Chapter  Google Scholar 

  42. Suquet P (1997) Effective properties of nonlinear composites. In: Suquet P (ed) Continuum micromechanics. Vol. 377 of CISM Lecture Notes. Springer Verlag, New York, pp 197–264

    Chapter  Google Scholar 

  43. Suquet P, Moulinec H, Castelnau O, Lahellec N, Montagnat M, Grennerat F, Duval P, Brenner R (2012) Multi-scale modeling of the mechanical behavior of polycrystalline ice under transient creep. Proc IUTAM 3:76–90. doi:10.1016/j.piutam.2012.03.006

    Article  Google Scholar 

  44. Taylor G (1934) The mechanics of plastic deformation of crystals. Part I: theoretical. Proc R Soc A 145:362–387

  45. Terada K, Kikuchi N (2001) A class of general algorithms for multi-scale analyses of heterogeneous media. Comp Methods Appl Mech Eng 190:5427–5464

    Article  MathSciNet  MATH  Google Scholar 

  46. Weertman J (1973) Creep of ice. In: Whalley E, Jones S, Gold L (eds) Physics and chemistry of ice. Royal Society, Ottawa, pp 320–337

    Google Scholar 

Download references

Acknowledgments

The authors acknowledge the support of the Labex MEC and of A*Midex through Grants ANR-11-LABX-0092 and ANR-11-IDEX-0001-02. The authors want to thank the referees for their comments which led to the discussion in Sect. 6.5

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Pierre Suquet.

Appendices

Appendix 1: Alternative writing of the constitutive relations for single crystals

The aim of this section is to show that the two crystal plasticity models presented in Sects. 2.2.1 and 2.2.2 can written in the form

$$\begin{aligned} \dot{{\varvec{\alpha }}} = {\varvec{\mathcal {F}}}({\varvec{\mathcal {A}}}), \end{aligned}$$
(58)

where the thermodynamic forces associated with the state variables are \({\varvec{\mathcal {A}}}=({\varvec{\mathcal {A}}}_{{\textsf {v}}}, {\varvec{\mathcal {A}}}_\beta ,{\varvec{\mathcal {A}}}_{{\textsf {p}}})\) given in (20). In both models, the evolution of the viscoplastic strain-rate is governed by a potential \(\psi \)

$$\begin{aligned} \dot{{\varvec{\varepsilon }}}_{{\textsf {v}}}= \frac{\partial \psi }{\partial {\varvec{\mathcal {A}}}_{{\textsf {v}}}}({\varvec{\mathcal {A}}}), \end{aligned}$$
(59)

where the potential \(\psi \) is the combination of individual potentials \(\psi _s\) on the different slip systems

$$\begin{aligned} \psi ({\varvec{\mathcal {A}}})= \sum _{s=1}^S \psi _s({\varvec{\mathcal {A}}}_{{\textsf {v}}},{\varvec{\mathcal {A}}}_\beta ,{\varvec{\mathcal {A}}}_{{\textsf {p}}}). \end{aligned}$$
(60)

The \(\psi _s\) take the form (61) and (62) for the model of Méric-Cailletaud and for ice, respectively,

$$\begin{aligned}&\psi _s({\varvec{\mathcal {A}}}_{{\textsf {v}}},{\varvec{\mathcal {A}}}_\beta ,{\varvec{\mathcal {A}}}_{{\textsf {p}}})\nonumber \\&\quad = \frac{K_s \dot{\gamma }_{0,s}}{n_s+1} \left( \frac{\left( \left| {{\varvec{\mathcal {A}}}_{{\textsf {v}}}:{\varvec{m}}_s + \mathcal {A}_{\beta _s}}\right| +\mathcal {A}_{{p}_s}\right) ^+}{K_s} \right) ^{n_s+1}, \end{aligned}$$
(61)
$$\begin{aligned}&\psi _s({\varvec{\mathcal {A}}}_{{\textsf {v}}},{\varvec{\mathcal {A}}}_\beta ,{\varvec{\mathcal {A}}}_{{\textsf {p}}}) = \frac{\dot{\gamma }_{0,s}}{n_s+1} \frac{\left| {{\varvec{\mathcal {A}}}_{{\textsf {v}}}:{\varvec{m}}_s + \mathcal {A}_{\beta _s}}\right| ^{n_s+1}}{\left| {\mathcal {A}_{{p}_s}}\right| ^{n_s}}. \end{aligned}$$
(62)

The other evolution equations for \(\beta _s\) and \(p_s\) have to be considered separately for the two models.

Méric-Cailletaud crystal plasticity model Note that (10) can be re-written as (recall that \(x_{\beta _s}= c_s \beta _s=- \mathcal {A}_{\beta _s}\) and divide the two sides of (10) by \(c_s\))

$$\begin{aligned} \dot{\beta }_s = \dot{\gamma }_s + \frac{d_s}{c_s} \mathcal {A}_{\beta _s}\left| { \dot{\gamma }_s}\right| . \end{aligned}$$
(63)

Furthermore

$$\begin{aligned} \dot{\gamma }_s = \frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _s}}({\varvec{\mathcal {A}}}), \quad \left| {\dot{\gamma }}\right| _s = \left| {\frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _s}}({\varvec{\mathcal {A}}})}\right| . \end{aligned}$$

Eq. (63) can be re-written as

$$\begin{aligned} \dot{\beta }_s = \frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _s}}({\varvec{\mathcal {A}}}) + \frac{d_s}{c_s} \left| {\frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _s}}({\varvec{\mathcal {A}}})}\right| \mathcal {A}_{\beta _s}. \end{aligned}$$

The first relation in (11) can be inverted as

$$\begin{aligned} p_s = \frac{1}{Qb} \sum _{s'=1}^S (h^{-1})_{s,s'}(r_{s'}-r_{ini,s'}), \end{aligned}$$

and the differential equation for \(p_s\) reads as

$$\begin{aligned} \dot{p}_s = \left[ 1+ \frac{1}{Q} \sum _{s'=1}^S (h^{-1})_{s,s'} (\mathcal {A}_{{p}_{s'}}- \mathcal {A}_{ini,{p}_{s'}})\right] \left| { \frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _s}}({\varvec{\mathcal {A}}})}\right| , \end{aligned}$$

where \( \mathcal {A}_{ini,{p}_{s'}}=-r_{ini,s'}\). Finally the constitutive relations (8), (9), (10) and (11) can be written as \(\dot{{\varvec{\alpha }}}= {\varvec{\mathcal {F}}}({\varvec{\mathcal {A}}})\) where \({\varvec{\alpha }}\) and \({\varvec{\mathcal {A}}}\) are given by (16) and (20) respectively and

$$\begin{aligned} {\varvec{\mathcal {F}}}({\varvec{\mathcal {A}}})&= \left( \begin{array}{c} {\varvec{\mathcal {F}}}_{{\textsf {v}}}({\varvec{\mathcal {A}}}) \\ \\ {\varvec{\mathcal {F}}}_{\beta _s}({\varvec{\mathcal {A}}}) \\ \\ {\varvec{\mathcal {F}}}_{{p}_s}({\varvec{\mathcal {A}}}) \end{array} \right) \nonumber \\&= \left( \begin{array}{c} \displaystyle \frac{\partial \psi }{\partial {\varvec{\mathcal {A}}}_{{\textsf {v}}}}({\varvec{\mathcal {A}}}) \\ \displaystyle \frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _s}}({\varvec{\mathcal {A}}}) + \frac{d_s}{c_s} \left| {\frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _s}}({\varvec{\mathcal {A}}})}\right| \mathcal {A}_{\beta _s}\\ \displaystyle \left[ 1+ \frac{1}{Q} \sum _{s'=1}^S (h^{-1})_{s,s'} (\mathcal {A}_{{p}_{s'}}- \mathcal {A}_{ini,{p}_{s'}}) \right] \left| { \frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _s}}({\varvec{\mathcal {A}}})}\right| \end{array} \right) . \end{aligned}$$
(64)

\(I_h\) ice. Note that (13) can be re-written as (recall that \(x_{\beta _s}= c_s \beta _s=- \mathcal {A}_{\beta _s}\) and divide the two sides of (13) by \(c_s\))

$$\begin{aligned} \dot{\beta }_s = \dot{\gamma }_s + \frac{d_s}{c_s} \mathcal {A}_{\beta _s}\left| {\dot{\gamma }_s}\right| + \frac{e_s}{c_s} \mathcal {A}_{\beta _s}. \end{aligned}$$
(65)

Furthermore

$$\begin{aligned} \dot{\gamma }_s = \frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _s}}({\varvec{\mathcal {A}}}),\quad \left| {\dot{\gamma }_s}\right| = \left| {\frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _s}}({\varvec{\mathcal {A}}})}\right| , \end{aligned}$$

where \(\psi \) is the potential (62) for ice. Eq. (65) can be re-written as

$$\begin{aligned} \dot{\beta }_s = \frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _s}}({\varvec{\mathcal {A}}}) + \frac{d_s}{c_s} \left| {\frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _s}}({\varvec{\mathcal {A}}})}\right| \mathcal {A}_{\beta _s}+ \frac{e_s}{c_s} \mathcal {A}_{\beta _s}. \end{aligned}$$

Eq. (14) is simply

$$\begin{aligned} \dot{p}_s = \sum _{s'=1}^S h_{s,s'} \left| { \frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _{s'}}}({\varvec{\mathcal {A}}})}\right| . \end{aligned}$$

Finally the constitutive relations (8), (12), (13) and (14) can be written in the form (3) with \({\varvec{\alpha }}\) and \({\varvec{\mathcal {A}}}\) given by (16) and (20) respectively and

$$\begin{aligned}&{\varvec{\mathcal {F}}}({\varvec{\mathcal {A}}}) = \left( \begin{array}{c} {\varvec{\mathcal {F}}}_{{\textsf {v}}}({\varvec{\mathcal {A}}}) \\ \\ {\varvec{\mathcal {F}}}_{\beta _s}({\varvec{\mathcal {A}}}) \\ \\ {\varvec{\mathcal {F}}}_{{p}_s}({\varvec{\mathcal {A}}}) \end{array} \right) \nonumber \\&\quad = \left( \begin{array}{c} \displaystyle \frac{\partial \psi }{\partial {\varvec{\mathcal {A}}}_{{\textsf {v}}}}({\varvec{\mathcal {A}}}) \\ \displaystyle \frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _s}}({\varvec{\mathcal {A}}}) + \frac{d_s}{c_s} \left| {\frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _s}}({\varvec{\mathcal {A}}})}\right| \mathcal {A}_{\beta _s}+ \frac{e_s}{c_s} \mathcal {A}_{\beta _s}\\ \displaystyle \sum _{s'=1}^S h_{s,s'} \left| { \frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _{s'}}}({\varvec{\mathcal {A}}})}\right| . \end{array} \right) . \end{aligned}$$
(66)

Appendix 2: Effective potentials

In order to check that (25) and (26) define the effective response of the polycrystal (the correct average stress history), note first that the solution \({\varvec{\varepsilon }}\) of the variational problem in (26) is precisely the solution of the local problem (23). It remains to prove that the force associated with \(\overline{{\varvec{\varepsilon }}}\) is precisely the average of the stress solution of (23) and that the differential equation satisfied by \(\widetilde{{\varvec{\alpha }}}\) is consistent with (24). To this end, one has to compute the thermodynamic forces associated with \(\overline{{\varvec{\varepsilon }}}\) and \(\widetilde{{\varvec{\alpha }}}\),

$$\begin{aligned}&\frac{\partial \widetilde{w}}{\partial \overline{{\varvec{\varepsilon }}}} \left( \overline{{\varvec{\varepsilon }}},\{{\varvec{\alpha }}({\varvec{x}}) \}_{{\varvec{x}} \in V} \right) , \quad \widetilde{{\varvec{{\mathcal {A}}}}} = \{ \widetilde{{\varvec{{\mathcal {A}}}}}_{{\varvec{x}}} \}_{{\varvec{x}} \in V},\quad \\&\widetilde{{\varvec{{\mathcal {A}}}}}_{{\varvec{x}}}= - \frac{\partial \widetilde{w}}{\partial {\varvec{\alpha }}({\varvec{x}})} \left( \overline{{\varvec{\varepsilon }}},\{{\varvec{\alpha }}({\varvec{x}}) \}_{{\varvec{x}} \in V} \right) . \end{aligned}$$

The first force can be computed using Hill’s lemma

$$\begin{aligned} \frac{\partial \widetilde{w}}{\partial \overline{{\varvec{\varepsilon }}}} \left( \overline{{\varvec{\varepsilon }}},\widetilde{{\varvec{\alpha }}} \right)&=\left\langle \frac{\partial w}{\partial {\varvec{\varepsilon }}}({\varvec{\varepsilon }},{\varvec{\alpha }}): \frac{\partial {\varvec{\varepsilon }}}{\partial \overline{{\varvec{\varepsilon }}}} \right\rangle =\left\langle {\varvec{\sigma }} : \frac{\partial {\varvec{\varepsilon }}}{\partial \overline{{\varvec{\varepsilon }}}} \right\rangle \\&= \left\langle {\varvec{\sigma }} \right\rangle :\left\langle \frac{\partial {\varvec{\varepsilon }}}{\partial \overline{{\varvec{\varepsilon }}}} \right\rangle = \overline{{\varvec{\sigma }}}, \end{aligned}$$

since \( \left\langle \frac{\partial {\varvec{\varepsilon }}}{\partial \overline{{\varvec{\varepsilon }}}} \right\rangle = {\varvec{I}}\). The first force is therefore the average stress associated with the solution of (23). The second force is computed by taking the Frechet derivative of \(\widetilde{w}\)

$$\begin{aligned} - \widetilde{{\varvec{\mathcal {A}}}} \cdot \widetilde{{\varvec{\alpha }}}^*&= \lim _{t \rightarrow 0} \frac{\widetilde{w} \left( \overline{{\varvec{\varepsilon }}},\widetilde{{\varvec{\alpha }}}+ t \widetilde{{\varvec{\alpha }}}^* \right) - \widetilde{w} \left( \overline{{\varvec{\varepsilon }}},\widetilde{{\varvec{\alpha }}} \right) }{t} \nonumber \\&= \lim _{t \rightarrow 0} \left\langle \frac{ {w} \left( {{\varvec{\varepsilon }}},{{\varvec{\alpha }}}+ t \widetilde{{\varvec{\alpha }}}^* \right) - {w} \left( {{\varvec{\varepsilon }}},{{\varvec{\alpha }}}\right) }{t} \right\rangle \nonumber \\&= \left\langle \frac{\partial w}{\partial {\varvec{\alpha }}}({\varvec{x}}). {{\varvec{\alpha }}}^*({\varvec{x}}) \right\rangle , \end{aligned}$$
(67)

which implies that \(\widetilde{{\varvec{{\mathcal {A}}}}}_{{\varvec{x}}}\) coincides with the local force \({\varvec{\mathcal {A}}}({\varvec{x}})\) at point \({\varvec{x}}\). Following an identical procedure it can be shown that

$$\begin{aligned} \frac{\partial \widetilde{\varphi }}{\partial \dot{\widetilde{{\varvec{\alpha }}}}}(\dot{\widetilde{{\varvec{\alpha }}}})\cdot \widetilde{{\varvec{\alpha }}}^* = \left\langle \frac{\partial {\varphi }}{\partial \dot{{{\varvec{\alpha }}}}}(\dot{{{\varvec{\alpha }}}}). {{\varvec{\alpha }}}^*({\varvec{x}}) \right\rangle . \end{aligned}$$
(68)

The evolution equations for the standard model corresponding to the potentials (26) read as

$$\begin{aligned} \frac{\partial \widetilde{w}}{\partial \widetilde{{\varvec{\alpha }}}}(\overline{{\varvec{\varepsilon }}},\widetilde{{\varvec{\alpha }}})+ \frac{\partial \widetilde{\varphi }}{\partial \dot{\widetilde{{\varvec{\alpha }}}}}(\dot{\widetilde{{\varvec{\alpha }}}}) = 0. \end{aligned}$$
(69)

which, by virtue of (67) and (68) shows that the evolution equations for the field of internal variables \({\varvec{\alpha }}({\varvec{x}})\) are precisely (24).

Appendix 3: TSO approximation for the two crystal plasticity models

1.1 C.1 Evolution equation for \({\varvec{\mathfrak {a}}}_\beta \)

For the non-standard variable \(\beta _s\) (and similarly for \(p_s\)) the TSO approximation of the evolution equations for the generalized forces consists in making use of a second-order expansion of \({\mathcal {F}}_{\beta _s}\) in (53), where

$$\begin{aligned} {\varvec{\mathcal {F}}} _{\beta _s}({{{\varvec{\mathcal {A}}}}}) = \frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _s}}({\varvec{\mathcal {A}}}) + \frac{d_s}{c_s} \left| {\frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _s}}({\varvec{\mathcal {A}}})}\right| {\mathcal {A}}_{\beta _s}+ \frac{e_s}{c_s} {\mathcal {A}}_{\beta _s}, \end{aligned}$$
(70)

with \( {\mathcal {A}}_{\beta _s}=-x_s\) and \(e_s=0\) in the model of Méric and Cailletaud. The first term in the expression of \(\displaystyle \left\langle {\varvec{\mathcal {F}}}_{\beta _s}({\varvec{\mathcal {A}}})\frac{\partial \mathcal {A}_{\beta _s}}{\partial \xi _{\beta _s}^{(\ell )}} \right\rangle \) is standard and the TSO approximation of this is

$$\begin{aligned}&\displaystyle \frac{\partial \widetilde{\psi }_{TSO}}{\partial \xi ^{(\ell )}_{\beta _s}} ({\varvec{\mathcal {A}}}) = \displaystyle \sum _{g=1}^N c^{(g)} \left[ - \frac{\partial f_s}{\partial \tau }\left( \overline{\tau }^{(g)}_s-\overline{x}_s^{(g)},r_s^{(g)}\right) \frac{\partial \overline{x}_s^{(g)} }{\partial \xi ^{(\ell )}_{\beta _s}} \right. \nonumber \\&\qquad + \displaystyle \frac{1}{2} \frac{\partial ^2 f_s}{\partial \tau ^2}\left( \overline{\tau }^{(g)}_s-\overline{x}_s^{(g)},r_s^{(g)}\right) \frac{\partial C^{(g)}(\tau _s-x_s)}{\partial \xi ^{(\ell )}_{\beta _s}} \nonumber \\&\qquad - \displaystyle \left. \frac{1}{2} \frac{\partial ^3 f_s}{\partial \tau ^3} \left( \overline{\tau }^{(g)}_s-\overline{x}_s^{(g)},r_s^{(g)}\right) C^{(g)}(\tau _s-x_s) \frac{\partial \overline{x}_s^{(g)} }{\partial \xi ^{(\ell )}_{\beta _s}} \right] . \end{aligned}$$
(71)

The TSO approximation of the second term in \(\left\langle {\varvec{\mathcal {F}}}_{\beta _s}({\varvec{\mathcal {A}}})\frac{\partial \mathcal {A}_{\beta _s}}{\partial \xi _{\beta _s}^{(\ell )}} \right\rangle \) is

$$\begin{aligned}&\displaystyle \sum _{g=1}^N c^{(g)} \left\langle \frac{d_s}{c_s} \left| {\frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _s}}({\varvec{\mathcal {A}}})}\right| x_s \frac{\partial x_s }{\partial \xi ^{(\ell )}_{\beta _s}} \right\rangle ^{(g)}\nonumber \\&\quad = \displaystyle \sum _{g=1}^N c^{(g)} \left\{ \frac{d_s}{c_s} \left[ \frac{\partial f_s}{\partial \tau }\left( \overline{\tau }^{(g)}_s-\overline{x}_s^{(g)},r_s^{(g)}\right) \right. \right. \nonumber \\&\qquad \times \left( \overline{x}_s^{(g)} \frac{\partial \overline{x}_s^{(g)} }{\partial \xi ^{(\ell )}_{\beta _s}} + \frac{1}{2} \frac{\partial C^{(g)}(x_s)}{\partial \xi ^{(\ell )}_{\beta _s}} \right) \nonumber \\&\qquad + \displaystyle \frac{\partial ^2 f_s}{\partial \tau ^2}\left( \overline{\tau }^{(g)}_s-\overline{x}_s^{(g)},r_s^{(g)}\right) \left( C^{(g)}(\tau _s-x_s,x_s) \frac{\partial \overline{x}_s^{(g)} }{\partial \xi ^{(\ell )}_{\beta _s}} \right. \nonumber \\&\qquad \left. - \displaystyle \frac{1}{2} \frac{\partial C^{(g)}(\tau _s-x_s)}{\partial \xi ^{(\ell )}_{\beta _s}} \overline{x}_s^{(g)} \right) \nonumber \\&\qquad \left. + \displaystyle \frac{1}{2} \frac{\partial ^3 f_s}{\partial \tau ^3} \left( \overline{\tau }^{(g)}_s-\overline{x}_s^{(g)},r_s^{(g)}\right) C^{(g)}(\tau _s-x_s) \overline{x}_s^{(g)} \frac{\partial \overline{x}_s^{(g)} }{\partial \xi ^{(\ell )}_{\beta _s}} \right] \nonumber \\&\qquad \times \left. \text {sign}\left( {\frac{\partial f_s}{\partial \tau }\left( \overline{\tau }^{(g)}_s-\overline{x}_s^{(g)},r_s^{(g)}\right) }\right) \right\} , \end{aligned}$$
(72)

where

$$\begin{aligned} C^{(g)}(\tau _s-x_s,x_s) = \left\langle (\tau _s-\overline{\tau }^{(g)}_s-x_s+\overline{x}_s^{(g)} ) (x_s-\overline{x}_s^{(g)}) \right\rangle ^{(g)}. \end{aligned}$$

The third term in \(\displaystyle \left\langle {\varvec{\mathcal {F}}}_{\beta _s}({\varvec{\mathcal {A}}})\frac{\partial \mathcal {A}_{\beta _s}}{\partial \xi _{\beta _s}^{(\ell )}} \right\rangle \) does not require any approximation and is simply

$$\begin{aligned}&\sum _{g=1}^N c^{(g)} \frac{e_s}{c_s} \left\langle x_s \frac{\partial x_s }{\partial \xi ^{(\ell )}_{\beta _s}} \right\rangle ^{(g)} \nonumber \\&\quad = \sum _{g=1}^N c^{(g)} \frac{e_s}{c_s} \left( \overline{x}_s^{(g)} \frac{\partial \overline{x}_s^{(g)} }{\partial \xi ^{(\ell )}_{\beta _s}} + \frac{1}{2} \frac{\partial C^{(g)}(x_s)}{\partial \xi ^{(\ell )}_{\beta _s}} \right) . \end{aligned}$$
(73)

Finally the evolution equation for \(\mathfrak {a}^{(\ell )}_{\beta _s}\) is obtained by equalling \(\dot{\mathfrak {a}}_{\beta _s}^{(\ell )}\) to the sum of the right-hand sides of (71), (72) and (73).

1.2 C.2 Evolution equation for the isotropic hardening variables

According to (46), the evolution equation for the reduced variable \( p_s^{(g)}\) is

$$\begin{aligned} \dot{p}_s^{(g)} = \left\langle \left( \mathcal {F}({\varvec{\mathcal {A}}}) \frac{\partial {\varvec{\mathcal {A}}}}{\partial \mathfrak {a}_{{p}_s}^{(g)}}\right) _{TSO} \right\rangle . \end{aligned}$$

According to (34) only \(\mathcal {A}_{{p}_s}=-r_s\) depends on \(\mathfrak {a}_{p_s}^{(g)}\) with

$$\begin{aligned} \mathcal {A}_{{p}_s}({\varvec{x}}) = \sum _{g=1}^N \frac{1}{c^{(g)}} \mathfrak {a}_{p_s}^{(g)} \chi ^{(g)}({\varvec{x}}), \quad \frac{\partial \mathcal {A}_{{p}_s}}{\partial \mathfrak {a}_{p_s}^{(g)}} = \frac{1}{c^{(g)}} \chi ^{(g)}({\varvec{x}}) . \end{aligned}$$

Finally the differential equation for \( p_s^{(g)}\) reduces to

$$\begin{aligned} \dot{p}_s^{(g)} = \left\langle \mathcal {F}_{{p}_s}({\varvec{\mathcal {A}}}) \right\rangle ^{(g)}. \end{aligned}$$
(74)

It remains to expand \( \left\langle \mathcal {F}_{{\textsf {p}}_s}({\varvec{\mathcal {A}}}) \right\rangle ^{(g)}\) to second-order in the fluctuations of \({\varvec{\mathcal {A}}}\). This expansion depends on the constitutive relations under consideration.

Méric-Cailletaud The differential equation takes the form

$$\begin{aligned} \dot{p}_s^{(g)}&= \left\langle \left( 1 -bp_s^{(g)} \right) \left| {\frac{\partial \psi }{\partial \mathcal {A}_{\beta _s}}({\varvec{\mathcal {A}}})}\right| \right\rangle ^{(g)} \nonumber \\&= \left( 1 -bp_s^{(g)} \right) \left\langle \left| {\frac{\partial \psi }{\partial \mathcal {A}_{\beta _s}}({\varvec{\mathcal {A}}})}\right| \right\rangle ^{(g)}. \end{aligned}$$
(75)
Table 3 Ice single crystal. a Slip systems. b Hardening matrix

It results from (48) that up to second-order in the fluctuations of \(\tau _s\) and \(x_s\) (\(p_s\) has no fluctuations),

$$\begin{aligned}&\displaystyle \left| {\frac{\partial \psi }{\partial {\mathcal {A}}_{\beta _s}}({\varvec{\mathcal {A}}})}\right| \simeq \displaystyle \Biggl [ {\frac{\partial f_s}{\partial \tau }\left( \overline{\tau }^{(g)}_s-\overline{x}_s^{(g)},r_s^{(g)}\right) } \nonumber \\&\qquad + \displaystyle \frac{\partial ^2 f_s}{\partial \tau ^2}\left( \overline{\tau }^{(g)}_s-\overline{x}_s^{(g)},r_s^{(g)}\right) (\tau _s-x_s- \overline{\tau }_s^{(g)} + \overline{x}_s^{(g)}) \nonumber \\&\qquad + \displaystyle \frac{1}{2} \frac{\partial ^3 f_s}{\partial \tau ^3} \left( \overline{\tau }^{(g)}_s-\overline{x}_s^{(g)},r_s^{(g)}\right) \left( \tau _s-x_s- \overline{\tau }_s^{(g)} + \overline{x}_s^{(g)}\right) ^2 \Biggr ] \nonumber \\&\qquad \times \text {sign}\left( {\frac{\partial f_s}{\partial \tau }\left( \overline{\tau }^{(g)}_s-\overline{x}_s^{(g)},r_s^{(g)}\right) }\right) , \end{aligned}$$
(76)

Therefore

$$\begin{aligned} \dot{p}_s^{(g)}= & {} \left( 1 -bp_s^{(g)} \right) \epsilon \Biggl [ {\frac{\partial f_s}{\partial \tau }\left( \overline{\tau }^{(g)}_s-\overline{x}_s^{(g)},r_s^{(g)}\right) } \nonumber \\&+ \frac{1}{2} \frac{\partial ^3 f_s}{\partial \tau ^3} \left( \overline{\tau }^{(g)}_s-\overline{x}_s^{(g)},r_s^{(g)}\right) C^{(g)}(\tau _s-x_s) \Biggr ] , \end{aligned}$$
(77)

where

$$\begin{aligned} \epsilon =\text {sign}\left( {\frac{\partial f_s}{\partial \tau }\left( \overline{\tau }^{(g)}_s-\overline{x}_s^{(g)},r_s^{(g)}\right) }\right) . \end{aligned}$$

Ice. For ice, the differential equation (74) takes the form

$$\begin{aligned} \dot{p}_s^{(g)} = \left\langle \sum _{s'=1}^S h_{s,s'} \left| {\frac{\partial \psi }{\partial \mathcal {A}_{{\beta }_{s'}}}({\varvec{\mathcal {A}}})}\right| \right\rangle ^{(g)}. \end{aligned}$$
(78)

Using (76) the following differential equation is obtained

$$\begin{aligned} \dot{p}_s^{(g)}= & {} \sum _{s'=1}^S h_{s,s'} \epsilon ' \Biggl [ {\frac{\partial f_{s'}}{\partial \tau }\left( \overline{\tau }^{(g)}_{s'}-\overline{x}_{s'}^{(g)},r_{s'}\right) } \nonumber \\&+ \frac{1}{2} \frac{\partial ^3 f_s}{\partial \tau ^3} \left( \overline{\tau }^{(g)}_{s'}-\overline{x}_{s'}^{(g)},r_{s'}^{(g)}\right) C^{(g)}(\tau _{s'}-x_{s'}) \Biggr ], \end{aligned}$$
(79)

where

$$\begin{aligned} \epsilon '=\text {sign}\left( {\frac{\partial f_{s'}}{\partial \tau }\left( \overline{\tau }^{(g)}_{s'}-\overline{x}_{s'}^{(g)},r_{s'}^{(g)}\right) }\right) . \end{aligned}$$

Alternatively a differential equation for \(r_s^{(g)}\) can be obtained. The evolution equation (53) for the reduced force \(\mathfrak {a}_{{p}_s}^{(g)}\) reads

$$\begin{aligned} \dot{\mathfrak {a}}_s^{(g)}&= \sum _{s'=1}^S \sum _{g'=1}^N c^{(g')} \left\langle \mathcal {F}_{p_{s'}}({\varvec{\mathcal {A}}}) \frac{\partial \mathcal {A}_{p_{s'}}}{\partial p_s^{(g)}} \right\rangle ^{(g')}\\&= c^{(g)} \left\langle \mathcal {F}_{{p}_s}({\varvec{\mathcal {A}}}) \right\rangle ^{(g)} \frac{\partial \mathcal {A}_{{p}_s}^{(g)}}{\partial p_s^{(g)}}, \end{aligned}$$

where the last equality has been obtained by noting that only \(\mathcal {A}_{{p}_s}\) in grain g depends on \(p_s^{(g)}\). Using the relations

$$\begin{aligned}&\mathfrak {a}_{{p}_s}^{(g)} = c^{(g)} \mathcal {A}_{{p}_s}^{(g)}, \quad r_s^{(g)} = - \mathcal {A}_{{p}_s}^{(g)}, \\&r_{s}(p) = \tau _{sta,s}+ \left( \tau _{ini,s}- \tau _{sta,s}\right) e^{-p}, \end{aligned}$$

one finds a differential equation for \(r_{s}^{(g)}\)

$$\begin{aligned} \dot{r}_{s}^{(g)}= & {} \displaystyle \left( r_{sta,s} -{r}_{s}^{(g)}\right) \sum _{s'=1}^S h_{s,s'} \Biggl [ \left| {\frac{\partial f_{s'}}{\partial \tau }\left( \overline{\tau }^{(g)}_{s'}-\overline{x}^{(g)}_{\beta _{s'}},{r}_{s'}^{(g)}\right) }\right| \\&+ \displaystyle \frac{1}{2} \frac{\partial ^3 f_s}{\partial \tau ^3} \left( \overline{\tau }^{(g)}_{s'}-\overline{x}^{(g)}_{{s'}} ,{r}_{s'}^{(g)}\right) \,\\&\times \, \text {sign}\left( {\frac{\partial f_{s'}}{\partial \tau }\left( \overline{\tau }^{(g)}_{s'}-\overline{x}_{s'}^{(g)},{r}_{s'}^{(g)}\right) }\right) C^{(g)}(\tau _{s'}-x_{{s'}}) \Biggr ] . \end{aligned}$$
Table 4 Material parameters used in the full-field simulations for single crystals of ice at \(-{1}0\,^{\circ }\mathrm{C}\). Units are MPa and s\(^{-{1}}\). After Suquet et al. [43]
Table 5 Copper single cristal. a Slip systems. b Hardening matrix
Table 6 Méric-Cailletaud model. Material parameters for copper. After [16]

1.3 C.3 Useful relations

The following definitions and relations can be useful to the reader:

$$\begin{aligned}&\displaystyle {\varvec{m}}_s({\varvec{x}}) = \sum _{g=1}^N {\varvec{m}}^{(g)}_s \chi ^{(g)}({\varvec{x}}), \\&\displaystyle {\tau }_s ({\varvec{x}}) = {\varvec{L}}({\varvec{x}}):{\varvec{A}}({\varvec{x}})::\overline{{\varvec{\varepsilon }}}\otimes {\varvec{m}}_s({\varvec{x}}) + \sum _{k=1}^{M_{{\textsf {v}}}} \xi _{{\textsf {v}}}^{(k)} {\rho }_s^{(k)}({\varvec{x}}),\\&{\rho }^{(k)}_s({\varvec{x}}) = {\varvec{\rho }}^{(k)}({\varvec{x}}):{\varvec{m}}_s({\varvec{x}}),\quad \displaystyle x_s({\varvec{x}}) = c_s \sum _{\ell =1}^{M_{\beta _s}} \xi _{\beta _s}^{(\ell )} \mu _{\beta _s}^{(\ell )}({\varvec{x}}), \\&r_s({\varvec{x}}) = \sum _{g=1}^{N} r_{s}^{(g)} \chi ^{(g)}({\varvec{x}}), \quad r_s^{(g)}=r_s({\varvec{p}}^{(g)}), \\&\displaystyle \overline{\tau }^{(g)}_s = \left\langle \tau _s \right\rangle ^{(g)} = \left\langle {\varvec{L}}:{\varvec{A}} \right\rangle ^{(g)}::\overline{{\varvec{\varepsilon }}}\otimes {\varvec{m}}_s^{(g)} + \sum _{k=1}^{M_{{\textsf {v}}}} \xi _{{\textsf {v}}}^{(k)} \overline{\rho }_s^{(k,g)},\\&\overline{\rho }^{(k,g)}_s = \left\langle {\rho }^{(k)}_s \right\rangle ^{(g)},\quad \displaystyle \overline{x}_s^{(g)} = c_s \sum _{\ell =1}^{M_{\beta _s}} \xi _{\beta _s}^{(\ell )} \overline{\mu }_{\beta _s}^{(\ell ,g)}, \quad \\&\displaystyle \overline{\mu }_{\beta _s}^{(\ell ,g)} = \left\langle \mu _{\beta _s}^{(\ell )} \right\rangle ^{(g)},\quad \frac{\partial \overline{\tau }^{(g)}_s }{\partial \xi ^{(k)}_{{\textsf {v}}}} = \overline{\rho }^{(k,g)}_s, \quad \frac{\partial \overline{x}_s^{(g)} }{\partial \xi ^{(\ell )}_{\beta _s}} = c_s \overline{\mu }^{(\ell ,g)}_{\beta _s},\\&\displaystyle \frac{\partial C^{(g)}(\tau _s-x_s)}{\partial \xi ^{(k)}_{{\textsf {v}}}} = 2 C^{(g)}\left( \tau _s-x_s, {\rho }_s^{(k)}\right) \\&\qquad = 2 \left\langle \left( \tau _s - \overline{\tau }^{(g)}_s -x_s+\overline{x}^{(g)}_s\right) \left( {\rho }_s^{(k)} - \overline{\rho }_s^{(k,g)}\right) \right\rangle ^{(g)},\\&\displaystyle \frac{\partial C^{(g)}(\tau _s-x_s)}{\partial \xi ^{(\ell )}_{\beta _s}} = -2c_s C^{(g)}\left( \tau _s-x_s, \mu _{\beta _s}^{(\ell )}\right) \\&\qquad = \displaystyle - 2 c_s \left\langle \left( \tau _s - \overline{\tau }^{(g)}_s -x_s+\overline{x}^{(g)}_s \right) \left( {\mu }_{\beta _s}^{(\ell )}-\overline{\mu }_{\beta _s}^{(\ell ,g)}\right) \right\rangle ^{(g)}. \end{aligned}$$

In the above, the symbol  :  :  is to be understood as the full contraction of two fourth-order tensors over their four indices.

Appendix 4: Material data for ice

Ice \(I_h\) is an HCP material with 12 slip systems. The list of the slip sytems can be found in Table 3. The hardening matrix is taken in the simple form shown in Table 3. It depends on 4 independent parameters. The hardening coefficients identified in Suquet et al. [43] are

$$\begin{aligned} h_1=70 \ \text {MPa},\quad h_2=110 \ \text {MPa},\quad h_3=125 \ \text {MPa},\quad h_4=0. \end{aligned}$$
(80)

The elastic moduli are (in MPa),

$$\begin{aligned}&L_{1111}=L_{2222}=13 930,\ L_{1122}=7 082,\ L_{1133}=5 765,\\&L_{2323}=L_{1313}=3 014,\ L_{1212}=3 424. \end{aligned}$$

The remaining material parameters are identical within the same family of slip systems, but differ from one family to the other (see Table 4).

Appendix 5: Material data for copper

Copper is a FCC material with twelve slip systems. The list of the slip sytems can be found in Table 5.

The cubic symmetries can be used to reduce the hardening matrix to 6 parameters \(h_i\): self-hardening is expressed through \(h_0\), whereas latent hardening is expressed by the other 5 \(h_i\)’s.

As suggested by [15] (chapter 6), the elastic moduli are taken to be isotropic (although this is probably a crude approximation) with Young’s modulus \(E=120{,}000\) MPa and Poisson ratio \(\nu =1/3\). Following [16] the four coefficients \(c_s\), \(d_s\), \(r_{0,s}\), \(b_s\) were taken to be the same for all systems (no dependence on s) (see Table 6).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Michel, JC., Suquet, P. A model-reduction approach to the micromechanical analysis of polycrystalline materials. Comput Mech 57, 483–508 (2016). https://doi.org/10.1007/s00466-015-1248-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00466-015-1248-9

Keywords

Navigation