Skip to main content
Log in

Fourier-based strength homogenization of porous media

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

Abstract

An efficient numerical method is proposed to upscale the strength properties of heterogeneous media with periodic boundary conditions. The method relies on a formal analogy between strength homogenization and non-linear elasticity homogenization. The non-linear problems are solved on a regular discretization grid using the Augmented Lagrangian version of Fast Fourier Transform based schemes initially introduced for elasticity upscaling. The method is implemented for microstructures with local strength properties governed either by a Green criterion or a Von Mises criterion, including pores or rigid inclusions. A thorough comparison with available analytical results or finite element elasto-plastic simulations is proposed to validate the method on simple microstructures. As an application, the strength of complex microstructures such as the random Boolean model of spheres is then studied, including a comparison to closed-form Gurson and Eshelby based strength estimates. The effects of the microstructure morphology and the third invariant of the macroscopic stress tensor on the homogenized strength are quantitatively discussed.

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

Similar content being viewed by others

Notes

  1. The specific case \(\omega _s=\omega _p=1\) corresponds to spherical inclusions for both phases and is given by (69).

References

  1. Barthélémy JF (2005) Approche micromécanique de la rupture et de la fissuration dans les géomatériaux. Ph.D. thesis, Ecole Nationale des Ponts et Chaussées

  2. Barthélémy JF, Dormieux L (2004) A micromechanical approach to the strength criterion of Drucker–Prager materials reinforced by rigid inclusions. Int J Numer Anal Methods Geomech 28(7–8):565–582

    Article  MATH  Google Scholar 

  3. Benallal A, Desmorat R, Fournage M (2014) An assessment of the role of the third stress invariant in the Gurson approach for ductile fracture. Eur J Mech A 47:400–414

    Article  MathSciNet  Google Scholar 

  4. Bilger N, Auslender F, Bornert M, Michel JC, Moulinec H, Suquet P, Zaoui A (2005) Effect of a nonuniform distribution of voids on the plastic response of voided materials: a computational and statistical analysis. Int J Solids Struct 42:517–538

    Article  MATH  Google Scholar 

  5. Bilger N, Auslender F, Bornert M, Moulinec H, Zaoui A (2007) Bounds and estimates for the effective yield surface of porous media with a uniform or a nonuniform distribution of voids. Eur J Mech A 26:810–836

    Article  MathSciNet  MATH  Google Scholar 

  6. Bleyer J, de Buhan P (2013) Yield surface approximation for lower and upper bound yield design of 3D composite frame structures. Comput Struct 129:86–98

    Article  Google Scholar 

  7. Brisard S, Dormieux L (2010) FFT-based methods for the mechanics of composites: a general variational framework. Comput Mater Sci 49:663–671

    Article  Google Scholar 

  8. Brisard S, Dormieux L (2012) Combining Galerkin approximation techniques and the principle of Hashin and Shtrikman to improve two FFT-based numerical methods for the homogenization of composites. Comput Methods Appl Mech Eng 217–220:197–212

    Article  MathSciNet  MATH  Google Scholar 

  9. Budiansky B (1965) On the elastic moduli of some heterogeneous materials. J Mech Phys Solids 13:223–227

    Article  Google Scholar 

  10. Danas K, Idiart M, Ponte Castañeda A (2008) A homogenization-based constitutive model for isotropic viscoplastic porous media. Int J Solids Struct 45:3392–3409

    Article  MATH  Google Scholar 

  11. de Buhan P (1986) Approche fondamentale du calcul à la rupture des ouvrages en sols renforcés. Ph.D. thesis, Université Pierre et Marie Curie, Paris

  12. de Buhan P, Mangiavacchi R, Nova R, Pellegrini G, con JS (1989) Yield design of reinforced earth walls by a homogenization method. Géotechnique 39(2):189–201

    Article  Google Scholar 

  13. Dormieux L, Jeannin L, Bemer E, Le TH, Sanahuja J (2010) Micromechanical models of the strength of a sandstone. Int J Numer Anal Methods Geomech 34:249–271

    MATH  Google Scholar 

  14. Dormieux L, Kondo D, Ulm FJ (2006) Microporomechanics. Wiley, Chichester

    Book  MATH  Google Scholar 

  15. Dormieux L, Sanahuja J, Maalej Y (2007) Résistance d’un polycristal avec interfaces intergranulaires imparfaites. C R Mec 335:25–31

    Article  MATH  Google Scholar 

  16. Frémond M, Friaà A (1978) Analyse limite. Comparaison des méthodes statique et cinématique. C R Acad Sci Paris 286:107–110

    MATH  Google Scholar 

  17. Fritzen F, Forest S, Bohlke T, Kondo D, Kanit T (2012) Computational homogenization of elasto-plastic porous metals. Int J Plast 29:102–119

    Article  Google Scholar 

  18. Fritzen F, Forest S, Kondo D, Bohlke T (2013) Computational homogenization of porous materials of Green type. Comput Mech 52:121–134

    Article  MathSciNet  MATH  Google Scholar 

  19. Gélébart L, Mondon-Cancel R (2013) Non-linear extension of FFT-based methods accelerated by conjugate gradients to evaluate the mechanical behavior of composite materials. Comput Mater Sci 77:430–439

    Article  Google Scholar 

  20. Gueguin M, Hassen G, Bleyer J, de Buhan P (2013) An optimization method for approximating the macroscopic strength criterion of stone column reinforced soils. In: Proceedings of the 3rd international symposium on computational geomechanics, pp 484–494. Pologne

  21. Gurson A (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 Trans ASME 99:2–15

    Article  Google Scholar 

  22. He Z, Dormieux L, Lemarchand E, Kondo D (2013) Cohesive Mohr–Coulomb interface effects on the strength criterion of materials with granular-based microstructure. Eur J Mech A 42:430–440

    Article  MathSciNet  Google Scholar 

  23. Hill R (1965) A self consistent mechanics of composite materials. J Mech Phys Solids 13:213–222

    Article  Google Scholar 

  24. Jeulin D, Moreaud M (2006) Percolation d’agrégats multi-échelles de sphêres et de fibres: application aux nanocomposites. In: Matériaux, pp 341–348. Dijon, France

  25. Kanit T, Forest S, Galliet I, Mounoury V, Jeulin D (2003) Determination of the size of the representative volume element for random composites: statistical and numerical approach. Int J Solids Struct 40:3647–3679

    Article  MATH  Google Scholar 

  26. Krabbenhoft K, Lyamin A, Sloan S (2007) Formulation and solution of some plasticity problems as conic programs. Int J Solids Struct 44:1533–1549

    Article  MATH  Google Scholar 

  27. Kushch V, Podoba Y, Shtern M (2008) Effect of micro-structure on yield strength of porous solid: a comparative study of two simple cell models. Comput Mater Sci 42:113–121

    Article  Google Scholar 

  28. Leblond JB, Perrin G, Suquet P (1994) Exact results and approximate models for porous viscoplastic solids. Int J Plast 10(3):213–235

    Article  MATH  Google Scholar 

  29. Maalej Y, Dormieux L, Sanahuja J (2009) Micromechanical approach to the failure criterion of granular media. Eur J Mech A 28:647–653

    Article  MATH  Google Scholar 

  30. Maghous S (1991) Détermination du critère de résistance macroscopique d’un matériau hétérogène à structure périodique. Ph.D. thesis, École Nationale des Ponts et Chaussées

  31. Maghous S, Dormieux L, Barthélémy JF (2009) Micromechanical approach to the strength properties of frictional geomaterials. Eur J Mech A 28:179–188

    Article  MathSciNet  MATH  Google Scholar 

  32. Mbiakop A, Constantinescu A, Danas K (2015) An analytical model for porous single crystals with ellipsoidal voids. J Mech Phys Solids 84:436–467

    Article  MathSciNet  Google Scholar 

  33. Mbiakop A, Danas K, Constantinescu A (2016) A homogenization based yield criterion for a porous Tresca material with ellipsoidal voids. Int J Fract. doi:10.1007/s10704-015-0071-9

  34. McElwain D, Roberts A, Wilkins A (2006) Yield criterion of porous materials subjected to complex stress states. Acta Materialia 54:1995–2002

    Article  Google Scholar 

  35. Michel JC, Moulinec H, Suquet P (2001) A computational scheme for linear and non-linear composites with arbitrary phase contrast. Int J Numer Methods Eng 52:139–160

    Article  Google Scholar 

  36. Mori T, Tanaka K (1973) Average stress in matrix and average elastic energy of materials with misfitting inclusions. Acta Metallurgica 21(5):1605–1609

    Article  Google Scholar 

  37. Moulinec H, Silva F (2014) Comparison of three accelerated FFT-based schemes for computing the mechanical response of composite materials. Int J Numer Methods Eng 97(13):960–985

    Article  MathSciNet  Google Scholar 

  38. Moulinec H, Suquet P (1994) A fast numerical method for computing the linear and non linear properties of composites. C R Acad Sci 2(318):1417–1423

    MATH  Google Scholar 

  39. 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 

  40. Norris A (1985) A differential scheme for the effective moduli of composites. Mech Mater 4:1–16

    Article  Google Scholar 

  41. Pastor F, Kondo D, Pastor J (2013) 3D-FEM formulations of limit analysis methods for porous pressure-sensitive materials. Int J Numer Methods Eng 95:847–870

    Article  MathSciNet  Google Scholar 

  42. Ponte-Castañeda P (1991) The effective mechanical properties of nonlinear isotropic composites. J Mech Phys Solids 39:45–71

    Article  MathSciNet  MATH  Google Scholar 

  43. Priour D Jr (2014) Percolation through voids around overlapping spheres: a dynamically based finite-size scaling analysis. Phys Rev E 89(1):1–5

    Article  Google Scholar 

  44. Revil-Baudard B, Cazacu O (2014) New three-dimensional strain-rate potentials for isotropic porous metals: role of the plastic flow of the matrix. Int J Plast 60:101–117

    Article  Google Scholar 

  45. Richelsen A, Tvergaard V (1994) Dilatant plasticity or upper bound estimates for porous ductile solids. Acta Metallurgica et Materialia 42(8):2561–2577

    Article  Google Scholar 

  46. Rintoul M, Torquato S (1997) Precise determination of the critical threshold and exponents in a three-dimensional continuum percolation model. J Phys 30(16):585–592

    MATH  Google Scholar 

  47. Salençon J (1990) An introduction to the yield design theory and its applications to soil mechanics. Eur J Mech 9:477–500

    MathSciNet  MATH  Google Scholar 

  48. Salençon J, Chatzigogos CT, Pecker A (2009) Seismic bearing capacity of circular footings: a yield deisgn approach. J Mech Mater Struct 4(2):427–440

  49. Sanahuja J, Dormieux L (2005) Résistance d’un milieu poreux à phase solide hétérogène. C R Mec 333:818–823

    Article  MATH  Google Scholar 

  50. Sanahuja J, Dormieux L, Chanvillard G (2007) Modelling elasticity of a hydrating cement paste. Cem Concr Res 37:1427–1439

    Article  Google Scholar 

  51. Suquet P (1983) Analyse limite et homogénéisation. CRAS 296:1355–1358

    MathSciNet  MATH  Google Scholar 

  52. Suquet P (1995) Overall properties of nonlinear composites: a modified secant moduli approach and its link with Ponte-Castaneda’s nonlinear variational procedure. C R Acad Sci Paris 320:563–571

    MATH  Google Scholar 

  53. Traxl R, Lackner R (2015) Multi-level homogenization of strength properties of hierarchical-organized matrix-inclusion materials. Mech Mater 89:98–118

    Article  Google Scholar 

  54. Tvergaard V (1981) Influence of voids on shear band instabilities under plane strain conditions. Int J Fract 17(4):389–407

    Article  Google Scholar 

  55. Tvergaard V (1982) On localization in ductile materials containing spherical voids. Int J Fract 18(4):237–252

    Google Scholar 

  56. van der Marck S (1996) Network approach to void percolation in a pack of unequal spheres. Phys Rev Lett 77(9):1785–1788

    Article  Google Scholar 

  57. Vincent PG, Suquet P, Monerie Y, Moulinec H (2014) Effective flow surface of porous materials with two populations of voids under internal pressure: II. full-field simulations. Int J Plast 56:74–98

    Article  Google Scholar 

  58. Wicklein M, Thoma K (2005) Numerical investigations of the elastic and plastic behaviour of an open-cell aluminium foam. Mater Sci Eng A 397:391–399

    Article  Google Scholar 

  59. Willam KJ, Warnke EP (1975) Constitutive models for the triaxial behavior of concrete. Proc Int Assoc Bridge Struct Eng 19:1–30

    Google Scholar 

  60. Willot F (2015) Fourier-based schemes for computing the mechanical response of composites with accurate local fields. C R Acad Sci Méc 340(3):232–245

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to François Bignonnet.

Appendices

Appendix 1: Green operator

1.1 (a) Continuous operator

Consider the following problem of prestressed linear elasticity:

$$\begin{aligned} \begin{aligned}&\text {find } {\varvec{\xi }} \in \mathcal {K}({\varvec{E}}), \, {\varvec{\sigma }} \in \mathcal {S} \text { such that } \forall {\varvec{z}} \in \varOmega , \\&\qquad {\varvec{\sigma }}({\varvec{z}}) = \mathbb {C}_0 : {\varvec{\epsilon }}({\varvec{z}}) + {\varvec{\tau }}({\varvec{z}}) \text { with } {\varvec{\epsilon }} = {\varvec{{{\mathrm{grad}}}}}^s {\varvec{\xi }} \end{aligned} \end{aligned}$$
(54)

In this problem, \({\varvec{\xi }}\) is the displacement field, \({\varvec{\epsilon }}\) the associated strain field and \({\varvec{\sigma }}\) the stress field. The set \(\mathcal {K}({\varvec{E}})\) (resp. \(\mathcal {S}\)) is defined by (8) (resp. (12)). The stiffness \(\mathbb {C}_0\) is homogeneous, \({\varvec{E}}\) is the imposed macroscopic deformation and \({\varvec{\tau }}\) is the so-called polarization field which is an imposed arbitrary loading parameter corresponding to a prestress field.

The displacement field \({\varvec{\xi }}\) solution to the prestressed problem (54) minimizes the potential energy among kinematically admissible displacement fields \({\varvec{\xi }}^\prime \in \mathcal {K}({\varvec{E}})\):

$$\begin{aligned} {\varvec{\xi }} = \text {arg } \min _{{\varvec{\xi }}^\prime \in \mathcal {K}({\varvec{E}})} \int _{\varOmega } \dfrac{1}{2} {\varvec{\epsilon }}^\prime : \mathbb {C}_0 : {\varvec{\epsilon }}^\prime + {\varvec{\epsilon }}^\prime : {\varvec{\tau }} {{\mathrm{d}}}V \end{aligned}$$
(55)

By definition, the Green operator \({\varvec{\varGamma }}_0\) is the operator such that the strain field solution to (54) is:

$$\begin{aligned} \begin{aligned}&{\varvec{\epsilon }}({\varvec{z}}) = {\varvec{E}} - \left( {\varvec{\varGamma }}_0 *{\varvec{\tau }}\right) ({\varvec{z}})\\&\text {where } \left( {\varvec{\varGamma }}_0 *{\varvec{\tau }}\right) ({\varvec{z}}) = \int _{\varOmega } {\varvec{\varGamma }}_0({\varvec{x}},{\varvec{y}}) : {\varvec{\tau }}({\varvec{y}}) {{\mathrm{d}}}V_y \end{aligned} \end{aligned}$$
(56)

In the Fourier space, let \({\varvec{k}}\) denote the wave vector, \(k =|{\varvec{k}}|\) its norm and \({\varvec{n}} = {\varvec{k}}/k\) its normalized direction. The Fourier transform of a \(\varOmega \)-periodic function f is denoted \(\hat{f}\) and is computed using the following convention:

$$\begin{aligned} \hat{f}({\varvec{k}}) = \dfrac{1}{|\varOmega |} \int _{\varOmega }{ f({\varvec{z}}) \exp \left( - \imath {\varvec{k}} \cdot {\varvec{z}} \right) {{\mathrm{d}}}V_z}. \end{aligned}$$
(57)

The FFT-based method benefits from the property on the Fourier transform of a convolution product:

$$\begin{aligned} \widehat{{\varvec{\varGamma }}_0 *{\varvec{\tau }}} = {\varvec{\hat{\varGamma }}}_{0} : {\varvec{\hat{\tau }}} \end{aligned}$$
(58)

When the stiffness tensor \(\mathbb {C}_0\) is isotropic with shear modulus \(\mu _0\) and Poisson’s ratio \(\nu _0\), the Fourier transform of the Green operator is:

$$\begin{aligned} {\varvec{\hat{\varGamma }}}_{0}({\varvec{k}}) = {\varvec{k}} \mathop {\otimes }\limits ^{s} {\varvec{\hat{G}}}_{0}({\varvec{k}}) \mathop {\otimes }\limits ^{s} {\varvec{k}} . \end{aligned}$$
(59)

where \({\varvec{\hat{G}}}_{0}\) is the Fourier transform of the Green function:

$$\begin{aligned} {\varvec{\hat{G}}}_{0}({\varvec{k}}) = {\left\{ \begin{array}{ll} \dfrac{1}{\mu _{0}k^{2}} \left( {\varvec{1}}- \dfrac{1}{2 (1- \nu _{0})} {\varvec{n}} \otimes {\varvec{n}} \right) &{} \text {if } {\varvec{k}} \ne {\varvec{0}},\\ {\varvec{0}} &{}\text {if } {\varvec{k}} = {\varvec{0}}. \end{array}\right. } \end{aligned}$$
(60)

The Green operator is null for \({\varvec{k}} = {\varvec{0}}\) consistently with the property \({\varvec{\varGamma }}_0 *{\varvec{\tau }} \in \mathcal {K}({\varvec{0}})\).

1.2 (b) Discretized operators

The discretization of the fictitious non-linear problem carried out in Sect. 3.2 leads to compute in (27) the average over any voxel of the strain rate \({\varvec{d}}\) solution to (54) for a voxel-wise constant polarization field. As shown in [7], the term involving the Green operator can be rigorously computed using the Discrete Fourier Transform as detailed in Sect. 3.3 provided that the consistent discrete Green operator \(\hat{{\varvec{\varGamma }}}_0^{\text {c}}\) is used, defined by:

$$\begin{aligned} \begin{aligned} \forall {\varvec{b}} \in \mathcal {I} , \, \hat{{\varvec{\varGamma }}}^{\text {c}}_{0,{\varvec{b}}} = \sum _{{\varvec{n}} \in \mathbb {Z}^{d}}&\left[ \prod _{i \in (1,..,d)} {{\mathrm{sinc}}}\left( \dfrac{\pi b_{i}}{N_{i}} \right) \right] ^{2} \\&\times \hat{{\varvec{\varGamma }}}_0({\varvec{k}}_{b_{1}+n_{1}N_{1}, ..., b_{d}+n_{d}N_{d}} ), \end{aligned} \end{aligned}$$
(61)

where for any multi-index \({\varvec{b}} = b_1,...,b_d\) of \(\mathbb {Z}^d\) the wave vector \({\varvec{k}}_{{\varvec{b}}}\) is defined by:

$$\begin{aligned} {\varvec{k}}_{{\varvec{b}}} = \dfrac{2\pi b_1}{L_1}{\varvec{e}}_1 + ... + \dfrac{2\pi b_d}{L_d}{\varvec{e}}_d. \end{aligned}$$
(62)

with the notations introduced in Sect. 3.2.

As the consistent discrete Green operator \(\hat{{\varvec{\varGamma }}}_0^{\text {c}}\) involves infinite sums, it is uneasy to implement. As an alternative, a so-called filtered, non consistent Green operator \(\hat{{\varvec{\varGamma }}}_0^{\text {fnc}}\) has been introduced in [8] as a good approximation to the consistent discrete Green operator:

$$\begin{aligned} \begin{aligned} \forall {\varvec{b}} \in \mathcal {I} , \, \hat{{\varvec{\varGamma }}}^{\text {fnc}}_{0,{\varvec{b}}} = \sum _{{\varvec{n}} \in \{0,1\}^{d}}&\left[ \prod _{i \in (1,..,d)} \cos \left( \dfrac{\pi b_{i}}{2 N_{i}} \right) \right] ^{2} \\&\times \hat{{\varvec{\varGamma }}}_0({\varvec{k}}_{b_{1}+n_{1}N_{1}, ..., b_{d}+n_{d}N_{d}} ), \end{aligned} \end{aligned}$$
(63)

This approximate operator can be readily computed at each iteration and does not require to be stored in memory.

Appendix 2: Estimates of the homogenized strength of porous material with Von Mises solid phase

1.1 (a) Criteria based on an exterior kinematic approach of yield design on a hollow sphere

Gurson criterion The Gurson criterion corresponds to an exterior kinematic approach of the yield design theory on a hollow sphere with a porosity \(\phi \) and a Von Mises solid matrix [21]. Using the notations (40) for the Von Mises criterion, the Gurson criterion reads:

$$\begin{aligned} \dfrac{\varSigma _d^2}{b(1+\phi ^2)} + \dfrac{2 \phi }{1+\phi ^2} \cosh \left( \sqrt{\dfrac{3}{2b}}\varSigma _m\right) - 1 \le 0 \end{aligned}$$
(64)

As this criterion does not take into account interaction between pore, [55] has proposed to replace \(\phi \) by the correcting term \(q_1 \phi \) with \(q_1\) around 1.5.

Criterion based on trial fields from linear elasticity While using the kinematic fields solution to the same hollow sphere with an incompressible linear elastic matrix which are richer than the trial fields used by Gurson, the derived criterion improves the Gurson criterion (see e.g. [3]). Although the expression are no longer known in closed-form, they only require a numerical volume integration and feature effects of the third invariant of the macroscopic stress tensor.

1.2 (b) Eshelby-based criteria

Non-linear techniques relying on Eshelby-based homogenization schemes have been used to estimate the solution to the fictitious problem (17).

In case of heterogeneous material comprising a Von Mises solid phase and pores, these techniques rely on the estimate of the homogenized stiffness \(\mathbb {C}^{\text {hom}}\) of a linear elastic composite with pores and incompressible solid phase with shear modulus \(\mu _s\). When \(\mathbb {C}^{\text {hom}}\) is isotropic,

$$\begin{aligned} \mathbb {C}^{\text {hom}} = 3k^{\text {hom}} \mathbb {J} + 2 \mu ^{\text {hom}} \mathbb {K} \quad \text {with} \quad {\left\{ \begin{array}{ll} k^{\text {hom}} = 2 \mu _s x&{} \\ 2\mu ^{\text {hom}} = 2\mu _s y&{} \\ \end{array}\right. } \end{aligned}$$
(65)

where x and y are functions depending on morphological parameters (porosity, aspect ratio of the constituents, ...).

Based on the modified secant method for non-linear homogenization techniques [52], the homogenized strength properties may then be estimated by a Green strength criterion [14]:

$$\begin{aligned} {\varvec{\varSigma }} \in G^{\text {hom}} \Leftrightarrow \dfrac{\varSigma _m^2}{A} + \dfrac{\varSigma _d^2}{B} - 1 \le 0 \quad \text {with} \quad {\left\{ \begin{array}{ll} A = (1-\phi ) x b&{} \\ B = (1-\phi ) y b&{} \\ \end{array}\right. } \end{aligned}$$
(66)
Fig. 14
figure 14

Schematic representation of the interpolation of octahedral cuts by the conic (70) for Lode angles \(\theta \in [0;\pi /3]\) based on the three values \(R_1=\varSigma _d(\theta =0)\), \(R_2=\varSigma _d(\theta =\pi /6)\) and \(R_3=\varSigma _d(\theta =\pi /3)\) and its symmetric and periodic reproductions. Yellow area admissible values of \(R_2\) to ensure convexity for given values of \(R_1\) and \(R_3\) [see (73)]. (Color figure online)

Values of x and y are recalled below for classical homogenization schemes (see [14] for details):

  • Mori–Tanaka scheme [36] with spherical pores

    $$\begin{aligned} x_{\text {mt}} = 2\dfrac{1-\phi }{3\phi }\quad ; \quad y_{\text {mt}} = \dfrac{1-\phi }{1+2\phi /3} \end{aligned}$$
    (67)
  • Differential scheme [40] with spherical pores

    $$\begin{aligned} x_{\text {ds}} = \dfrac{2y_{\text {ds}}}{3(1- y_{\text {ds}}^{3/5})} \quad ; \quad \dfrac{y_{\text {ds}}^3}{2-y_{\text {ds}}^{3/5}}=(1-\phi )^6 \end{aligned}$$
    (68)
  • Self-consistent scheme [9, 23] with spherical pores and solid grains

    $$\begin{aligned} x_{\text {sc}} = 2\dfrac{(1-2\phi )(1-\phi )}{\phi (3-\phi )}\quad ; \quad y_{\text {sc}} = 3\dfrac{1-2\phi }{3-\phi } \end{aligned}$$
    (69)

The expressions of x and y for the self-consistent scheme with spheroidal solid grains with aspect ratio \(\omega _s\) and spheroidal pores with aspect ratio \(\omega _p\) having both an isotropic distribution of orientation (see e.g. [50]) are implicit functions of \(\phi \), \(\omega _s\) and \(\omega _p\), too long to be reproduced here.

Appendix 3: Interpolation of octahedral projections of \(\partial G^{\text {hom}}\) by combinations of a conic

Projections of the boundary of the macroscopic strength criteria by octahedral planes \(\varSigma _m=\) constant can be excellently approximated by a combination of conics (ellipse or hyperbola). In such \((x=\varSigma _{d1},y=\varSigma _{d2})\) plane [refering to (50)], the equation of an arbitrary ellipse or hyperbola of center \((x_0,y_0)\) is:

$$\begin{aligned} \dfrac{(x-x_0)^2}{A}+\dfrac{(y-y_0)^2}{B}+\dfrac{2(x-x_0)(y-y_0)}{C} = 1 \end{aligned}$$
(70)

The conic (70) is an ellipse if \(AB-C^2<0\) and \(AB>0\). The parameters A, B, C, \(x_0\) and \(y_0\) are deduced from the values of the stress deviator \(\varSigma _d\) at three Lode angles \(\theta \): \(R_1=\varSigma _d(\theta =0)\), \(R_2=\varSigma _d(\theta =\pi /6)\) and \(R_3=\varSigma _d(\theta =\pi /3)\) (see Fig. 14). Imposing that the tangent to the conic in \(\theta =0\) and \(\pi /3\) is orthogonal to the radial vector in polar coordinates, geometrical considerations lead to the unique expression of the conic parameters given in equation (71).

$$\begin{aligned} \begin{aligned} \varDelta&=-27(R_1-R_3)^2R_2^2+2\sqrt{3}(8R_3^2-17R_1R_3+8R_1^2)(R_3+R_1)R_2 -R_1R_3(32R_1^2+32R_3^2-71R_1R_3)\\ B&=\dfrac{(R_1-2R_3)^2(-3R_1R_3+\sqrt{3}R_2R_1+R_2\sqrt{3}R_3)^2}{3 \varDelta }\\ A&=B\dfrac{(R_2\sqrt{3}-2R_1)(2R_3-R_2\sqrt{3})(2R_1-R_3)^2}{(4R_1R_3+2R_1^2-7R_3^2)R_2^2-2\sqrt{3}(R_3+R_1)(2R_1-3R_3)R_3R_2+6R_1R_3^2(R_1-2R_3)}\\ C&=B \dfrac{(-R_2\sqrt{3}+2R_1)(2R_3-R_2\sqrt{3})(2R_1-R_3)^2}{\sqrt{3}(-(R_1-2R_3)^2R_2^2-2\sqrt{3}(R_1-R_3)(R_3+R_1)R_3R_2+R_1R_3^2(-4R_3+5R_1))}\\ x_0&=\dfrac{1}{\varDelta } \times \left[ (18R_1^2R_3-6R_3^3-15R_1^3)R_2^2+2\sqrt{3}(-2R_3^2-R_1R_3+4R_1^2)(R_3+R_1)(R_1-R_3)R_2\right. \\&\qquad \qquad \left. -R_1R_3(16R_1^3-4R_1R_3^2-23R_1^2R_3+8R_3^3) \right] \\ y_0&= \dfrac{\sqrt{3}\left[ (R_1-2R_3)^2R_2^2+2\sqrt{3}(R_1-R_3)(R_3+R_1)R_3R_2-R_1R_3^2(-4R_3+5R_1)\right] (R_1-2R_3)}{\varDelta }\\ \end{aligned} \end{aligned}$$
(71)

For Lode angles \(\theta \in [0;\pi /3]\), the interpolation of the macroscopic stress deviator is then explicitly given by equation (72).

(72)

where

$$\begin{aligned} \epsilon = {\left\{ \begin{array}{ll} -1 &{} \text {if } \left( \text {the conic~(70) is an ellipse and } C<0 \right) \\ &{} \text {or } \left( \text {the conic~(70) is a hyperbola and } A<0 \right) \\ 1 &{} \text {otherwise}\\ \end{array}\right. } \end{aligned}$$

Due to the symmetry properties exposed in Sect. 4.1.4, the whole cut of the boundary of the macroscopic strength criteria by the octahedral plane \(\varSigma _m=\) constant is constructed first by symmetry with respect to the \(\varSigma _{d2}\) axis and then by periodic reproduction with a \(2\pi /3\) period.

The convexity of the strength domain imposes the conditions:

$$\begin{aligned} \dfrac{R_3}{2} \le R_1 \le 2R_3 \quad \text {and} \quad \dfrac{\sqrt{3}R_1R_3}{R_1+R_3} \le R_2 \le \dfrac{2}{\sqrt{3}} \min (R_1,R_3) \end{aligned}$$
(73)

In the case where \(R_2\) is chosen as a function of \(R_1\) and \(R_3\) such that \(y_0\) vanishes and \(C\rightarrow \infty \), the expression (72) corresponds to the Lode angle dependence of the Willam-Warnke criterion [59]. In this case, a circle is obtained if \(R_1=R_3\) or an equilateral triangle (as for the Rankine criterion) is obtained if \(R_3=2R_1\).

In the case \(R_1=R_3=2R_2/\sqrt{3}\), the expression (72) corresponds to a regular hexagon and hence to the Lode angle dependence of the Tresca and Mohr-Coulomb criteria. The case \(R_1=R_2=R_3\) corresponds to a circle.

Excepted in the limit cases \(R_1=R_3=2R_2/\sqrt{3}\) (hexagonal shape) and \(2R_1=\sqrt{3}R_2 = R_3\) (triangular shape), the obtained criterion is smooth: the tangent is continuously varying.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bignonnet, F., Hassen, G. & Dormieux, L. Fourier-based strength homogenization of porous media. Comput Mech 58, 833–859 (2016). https://doi.org/10.1007/s00466-016-1319-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00466-016-1319-6

Keywords

Navigation