Skip to main content
Log in

Shape and packing effects in particulate composites: micromechanical modelling and numerical verification

  • Original Article
  • Published:
Archives of Civil and Mechanical Engineering Aims and scope Submit manuscript

Abstract

The aim of this study is to analyse the joint effect of reinforcement shape and packing on the effective behaviour of particulate composites. The proposed semi-analytical modelling method combines the Replacement Mori–Tanaka scheme, by means of which the concentration tensors for non-ellipsoidal inhomogeneities are found numerically, and the analytical morphologically representative pattern approach to account for particle packing. Five shapes of inhomogeneities are selected for the analysis: a sphere, a prolate ellipsoid, a sphere with cavities, an oblate spheroid with a cavity as well as an inhomogeneity created by three prolate spheroids crossing at right angles. Semi-analytical estimates are compared with the results of numerical simulations performed using the finite element method and with the outcomes of classical mean-field models based on the Eshelby solution, e.g. the Mori–Tanaka model or the self-consistent scheme.

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

Similar content being viewed by others

Data Availability Statement

The authors can confirm that all relevant data are included in the article.

References

  1. Olson G. Computational design of hierarchically structured materials. Science. 1997;277:1237–42. https://doi.org/10.1126/science.277.5330.1237.

    Article  CAS  Google Scholar 

  2. Matouš K, Geers MGD, Kouznetsova VG, Gillman A. A review of predictive nonlinear theories for multiscale modeling of heterogeneous materials. J Comput Phys. 2017;330:192–220. https://doi.org/10.1016/j.jcp.2016.10.070.

    Article  ADS  MathSciNet  Google Scholar 

  3. Clyne TW, Hull D. An introduction to composite materials. 3rd ed. Cambridge: Cambridge University Press; 2019. https://doi.org/10.1017/9781139050586.

    Book  Google Scholar 

  4. Rajak DK, Pagar DD, Kumar R, Pruncu CI. Recent progress of reinforcement materials: a comprehensive overview of composite materials. J Mater Res Technol. 2019;8:6354–74. https://doi.org/10.1016/j.jmrt.2019.09.068.

    Article  CAS  Google Scholar 

  5. Christensen RM. A critical evaluation for a class of micro-mechanics models. J Mech Phys Solids. 1990;38:379–404. https://doi.org/10.1016/0022-5096(90)90005-O.

    Article  ADS  Google Scholar 

  6. Eshelby JD. The determination of the elastic field of an ellipsoidal inclusion, and related problems. Proc R Soc A-Math Phy. 1957;241:376–96. https://doi.org/10.1098/rspa.1957.0133.

    Article  ADS  MathSciNet  Google Scholar 

  7. Kanaun SK, Levin VM. Self-consistent methods for composites, vol 1: static problems. Dordrecht: Springer; 2008. p. 172–86. https://doi.org/10.1007/978-1-4020-6664-1.

    Book  Google Scholar 

  8. Castañeda PP, Willis JR. The effect of spatial distribution on the effective behavior of composite materials and cracked media. J Mech Phys Solids. 1995;43:1919–51. https://doi.org/10.1016/0022-5096(95)00058-Q.

    Article  ADS  MathSciNet  Google Scholar 

  9. Torquato S. Effective stiffness tensor of composite media: II. Applications to isotropic dispersions. J Mech Phys Solids. 1998;46:1411–40. https://doi.org/10.1016/S0022-5096(97)00083-5.

    Article  ADS  MathSciNet  CAS  Google Scholar 

  10. Marcadon V, Herve E, Zaoui A. Micromechanical modeling of packing and size effects in particulate composites. Int J Solids Struct. 2007;44:8213–28. https://doi.org/10.1016/j.ijsolstr.2007.06.008.

    Article  Google Scholar 

  11. Majewski M, Kursa M, Holobut P, Kowalczyk-Gajewska K. Micromechanical and numerical analysis of packing and size effects in elastic particulate composites. Compos B Eng. 2017;124:158–74. https://doi.org/10.1016/j.compositesb.2017.05.004.

    Article  Google Scholar 

  12. Diani J, Gilormini P, Merckel Y, Vion-Loisel F. Micromechanical modeling of the linear viscoelasticity of carbon-black filled styrene butadiene rubbers: the role of the filler-rubber interphase. Mech Mater. 2013;59:65–72. https://doi.org/10.1016/j.mechmat.2012.12.007.

    Article  Google Scholar 

  13. Majewski M, Holobut P, Kursa M, Kowalczyk-Gajewska K. Packing and size effects in elastic-plastic particulate composites: micromechanical modelling and numerical verification. Int J Eng Sci. 2020;151:103271. https://doi.org/10.1016/j.ijengsci.2020.103271.

    Article  MathSciNet  Google Scholar 

  14. Chaboche JL, Kanouté P, Roos A. On the capabilities of mean-field approaches for the description of plasticity in metal matrix composites. Int J Plast. 2005;21:1409–34. https://doi.org/10.1016/j.ijplas.2004.07.001.

    Article  CAS  Google Scholar 

  15. Sevostianov I, Mogilevskaya SG, Kushch VI. Maxwell’s methodology of estimating effective properties: alive and well. Int J Eng Sci. 2019;140:35–88. https://doi.org/10.1016/j.ijengsci.2019.05.001.

    Article  MathSciNet  CAS  PubMed  PubMed Central  Google Scholar 

  16. Kowalczyk-Gajewska K, Majewski M, Mercier S, Molinari A. Mean field interaction model accounting for the spatial distribution of inclusions in elasticviscoplastic composites. Int J Solids Struct. 2021;224:111040. https://doi.org/10.1016/j.ijsolstr.2021.111040.

    Article  Google Scholar 

  17. Segurado J, Llorca J. Computational micromechanics of composites: the effect of particle spatial distribution. Mech Mater. 2006;38:873–83. https://doi.org/10.1016/j.mechmat.2005.06.026.

    Article  Google Scholar 

  18. Dastgerdi JN, Anbarlooie B, Miettinen A, Hosseini-Toudeshky H, Remes H. Effects of particle clustering on the plastic deformation and damage initiation of particulate reinforced composite utilizing X-ray CT data and finite element modeling. Compos B Eng. 2018;153:57–69. https://doi.org/10.1016/j.compositesb.2018.07.027.

    Article  CAS  Google Scholar 

  19. Escoda J, Willot F, Jeulin D, Sanahuja J, Toulemonde C. Influence of the multiscale distribution of particles on elastic properties of concrete. Int J Eng Sci. 2016;98:60–71. https://doi.org/10.1016/j.ijengsci.2015.07.010.

    Article  CAS  Google Scholar 

  20. de Francqueville F, Gilormini P, Diani J, Vandenbroucke A. Comparison of the finite strain macroscopic behavior and local damage of a soft matrix highly reinforced by spherical or polyhedral particles. Eur J Mech A Solids. 2020;84:104070. https://doi.org/10.1016/j.euromechsol.2020.104070.

    Article  Google Scholar 

  21. Liu LP. Solutions to the Eshelby conjectures. Proc R Soc A-Math Phys. 2008;464:573–94. https://doi.org/10.1098/rspa.2007.0219.

    Article  ADS  MathSciNet  Google Scholar 

  22. Klusemann B, Böhm HJ, Svendsen B. Homogenization methods for multi-phase elastic composites with non-elliptical reinforcements: comparisons and benchmarks. Eur J Mech A Solids. 2012;34:21–37. https://doi.org/10.1016/j.euromechsol.2011.12.002.

    Article  MathSciNet  Google Scholar 

  23. Zheng QS, Du DX. An explicit and universally applicable estimate for the effective properties of multiphase composites which accounts for inclusion distribution. J Mech Phys Solids. 2001;49:2765–88. https://doi.org/10.1016/S0022-5096(01)00078-3.

    Article  ADS  Google Scholar 

  24. Du DX, Zheng QS. A further exploration of the interaction direct derivative (IDD) estimate for the effective properties of multiphase composites taking into account inclusion distribution. Acta Mech. 2002;157:61–80. https://doi.org/10.1007/BF01182155.

    Article  Google Scholar 

  25. Nogales S, Böhm HJ. Modeling of the thermal conductivity and thermomechanical behavior of diamond reinforced composites. Int J Eng Sci. 2008;46:606–19. https://doi.org/10.1016/j.ijengsci.2008.01.011.

    Article  CAS  Google Scholar 

  26. Kachanov M, Tsukrov I, Shafiro B. Effective moduli of solids with cavities of various shapes. Appl Mech Rev. 1994;47:151–74. https://doi.org/10.1115/1.3122810.

    Article  ADS  Google Scholar 

  27. Segurado J, Llorca J. A numerical approximation to the elastic properties of sphere-reinforced composites. J Mech Phys Solids. 2002;50:2107–21. https://doi.org/10.1016/S0022-5096(02)00021-2.

    Article  ADS  CAS  Google Scholar 

  28. Hassani B, Hinton E. A review of homogenization and topology optimization I—homogenization theory for media with periodic structure. Comput Struct. 1998;69:707–17. https://doi.org/10.1016/S0045-7949(98)00131-X.

    Article  Google Scholar 

  29. Ranganathan SI, Ostoja-Starzewski M. Scaling function, anisotropy and the size of RVE in elastic random polycrystals. J Mech Phys Solids. 2008;56:2773–91. https://doi.org/10.1016/j.jmps.2008.05.001.

    Article  ADS  MathSciNet  Google Scholar 

  30. Hashin Z. The elastic moduli of heterogeneous materials. J Appl Mech. 1962;29:143–50. https://doi.org/10.1115/1.3636446.

    Article  MathSciNet  CAS  Google Scholar 

  31. Herve E, Zaoui A. n-Layered inclusion-based micromechanical modelling. Int J Eng Sci. 1993;31:1–10. https://doi.org/10.1016/0020-7225(93)90059-4.

    Article  Google Scholar 

  32. Hori M, Nemat-Nasser S. Double-inclusion model and overall moduli of multi-phase composites. Mech Mater. 1993;14:189–206. https://doi.org/10.1016/0167-6636(93)90066-Z.

    Article  Google Scholar 

  33. Suquet P. Effective properties of nonlinear composites. In: Suquet P, editor. Continuum micromechanics. Vienna: Springer; 1997. p. 197–264. https://doi.org/10.1007/978-3-7091-2662-2.

    Chapter  Google Scholar 

  34. Mori T, Tanaka K. Average stress in matrix and average elastic energy of materials with misfitting inclusions. Acta Metall. 1973;21:571–4. https://doi.org/10.1016/0001-6160(73)90064-3.

    Article  Google Scholar 

  35. Schöberl J. NETGEN—an advancing front 2D/3D-mesh generator based on abstract rules. Comput Vis Sci. 1997;1:41–52. https://doi.org/10.1007/s007910050004.

    Article  Google Scholar 

  36. Korelc J. Multi-language and multi-environment generation of nonlinear finite element codes. Eng Comput. 2002;18:312–27. https://doi.org/10.1007/s003660200028.

    Article  Google Scholar 

  37. Markov A, Trofimov A, Sevostianov I. A unified methodology for calculation of compliance and stiffness contribution tensors of inhomogeneities of arbitrary 2D and 3D shapes embedded in isotropic matrix - open access software. J Mech Phys Solids. 2020;157:103390. https://doi.org/10.1016/j.ijengsci.2020.103390.

    Article  MathSciNet  Google Scholar 

  38. Kowalczyk-Gajewska K, Ostrowska-Maciejewska J. Review on spectral decomposition of Hooke’s tensor for all symmetry groups of linear elastic material. Eng Trans. 2009; 57:145–183. https://et.ippt.gov.pl/index.php/et/article/view/172.

  39. Šmilauer V et al. Yade Documentation. 3rd ed. The Yade Project. Zenodo; 2021. https://doi.org/10.5281/zenodo.5705394.

  40. Mazzucco G, Pomaro B, Salomoni VA, Majorana CE. Numerical modelling of ellipsoidal inclusions. Constr Build Mater. 2018;167:317–24. https://doi.org/10.1016/j.conbuildmat.2018.01.160.

    Article  Google Scholar 

  41. Kursa M, Kowalczyk-Gajewska K, Petryk H. Multi-objective optimization of thermo-mechanical properties of metal-ceramic composites. Compos B Eng. 2014;60:586–96. https://doi.org/10.1016/j.compositesb.2014.01.009.

    Article  CAS  Google Scholar 

  42. Hill R. Continuum micro-mechanics of elastoplastic polycrystals. J Mech Phys Solids. 1965;13:89–101. https://doi.org/10.1016/0022-5096(65)90023-2.

    Article  ADS  CAS  Google Scholar 

  43. Kowalczyk-Gajewska K, Maździarz M. Elastic properties of nanocrystalline materials of hexagonal symmetry: the core-shell model and atomistic estimates. Int J Eng Sci. 2020;157:103393. https://doi.org/10.1016/j.ijengsci.2020.103393.

    Article  MathSciNet  CAS  Google Scholar 

  44. Celentano DJ, Chaboche JL. Experimental and numerical characterization of damage evolution in steels. Int J Plast. 2007;23:1739–62. https://doi.org/10.1016/j.ijplas.2007.03.008.

    Article  CAS  Google Scholar 

  45. Steinmann P. Formulation and computation of geometrically non-linear gradient damage. Int J Numer Methods Eng. 1999;46:757–79. https://doi.org/10.1002/(SICI)1097-0207(19991020)46:5757::AID-NME7313.0.CO;2-NN.

    Article  MathSciNet  Google Scholar 

  46. Rutecka A, Kursa M, Pietrzak K, Kowalczyk-Gajewska K, Makowska K, Wyszkowski M. Damage evolution in AA2124/SiC metal matrix composites under tension with consecutive unloadings. Arch Civ Mech. 2020;20:135. https://doi.org/10.1007/s43452-020-00134-x.

    Article  Google Scholar 

Download references

Funding

The research was partially supported by the project No. 2017/25/N/ST8/01968 of the National Science Center (NCN), Poland. The work of the second author was supported by the National Science Centre (NCN) in Poland under Grant No. 2015/19/N/ST8/03924.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to M. Majewski.

Ethics declarations

Conflict of interest

The authors have no conflicts of interest to declare that are relevant to the content of this article.

Ethical approval

This article does not contain any studies with human participants or animals performed by any of the authors.

Informed consent

Not applicable.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Stiffness tensor decompositions: transverse isotropy and cubic symmetry

In the paper, five different shapes are studied (see Fig. 3a–e). Let us assume that the spheroid is prolate (Fig. 3c) or oblate and drilled (Fig. 3e) in the \(\mathbf {m}_1\) direction. Then the stiffness tensor \(\overline{\mathbb {L}}^{dil }_{i }\) of a tiny spheroid embedded in a large cube (Fig. 4) and the effective stiffness tensor \(\overline{\mathbb {L}}_{NDil }\) in Eq. 14 exhibit transverse isotropy and can be written in the following form [43]:

$$\begin{aligned} \overline{\mathbb {L}}^{Trans }= 3 \overline{K} \mathbb {I}^{P }+ 2 \overline{G}_{1} \mathbb {P}_{1}+ 2 \overline{G}_{2} \mathbb {P}_{2}+ 2 \overline{G}_{3} \mathbb {P}_{3}+ 2 \overline{G}_{12} \frac{1}{\sqrt{6}}\left( \mathbf {d}\otimes \mathbf {I} + \mathbf {I}\otimes \mathbf {d} \right) \,. \end{aligned}$$
(A.1)

The orthogonal projectors \(\mathbb {P}_K\) (\(K=1,2,3\)) define:

  • 1D space of isochoric tension/compression along \(\mathbf {m}_1\)

    $$\begin{aligned} \mathbb {P}_{1}= \mathbf {d}\otimes \mathbf {d}\,,\quad \mathbf {d}=\frac{1}{\sqrt{6}}\left( 3\mathbf {m}_1\otimes \mathbf {m}_1-\mathbf {I}\right) \,, \end{aligned}$$
    (A.2)
  • 2D space of in-plane shears

    $$\begin{aligned} \mathbb {P}_{2}= \frac{1}{2} \left[ \left( \mathbf {m}_2\otimes \mathbf {m}_3+\mathbf {m}_3\otimes \mathbf {m}_2\right) \otimes \left( \mathbf {m}_2\otimes \mathbf {m}_3+\mathbf {m}_3\otimes \mathbf {m}_2\right) + \left( \mathbf {m}_2\otimes \mathbf {m}_2-\mathbf {m}_3\otimes \mathbf {m}_3\right) \otimes \left( \mathbf {m}_3\otimes \mathbf {m}_3-\mathbf {m}_2\otimes \mathbf {m}_2\right) \right] \,, \end{aligned}$$
    (A.3)
  • 2D space of out-of-plane shears where

    $$\begin{aligned} \mathbb {P}_{3}= \frac{1}{2} \sum _{k=2,3} \left( \mathbf {m}_1\otimes \mathbf {m}_k+\mathbf {m}_k\otimes \mathbf {m}_1\right) \otimes \left( \mathbf {m}_1\otimes \mathbf {m}_k+\mathbf {m}_k\otimes \mathbf {m}_1\right) \,. \end{aligned}$$
    (A.4)

As it was mentioned, to calculate the five independent components: \(3\overline{K}\), \(2\overline{G}_1\), \(2 \overline{G}_{12}\), \(2 \overline{G}_2\), \(2 \overline{G}_3\), a set of four analyses with micro-periodic displacement boundary conditions (Eq. 16)   were performed for a unit volume element. The four strain tensors \(\mathbf {E}^{(n)}\) (\(n=0,1,2,3\)) imposed in these analyses are given by Eq. (19).

The strains \(\mathbf {E}^{\left( n\right) }\) result in the overall stresses, calculated as the local stresses averaged over the RVE’s volume, \(\mathbf {\Sigma }^{(n)}=1/V\int _V\varvec{\sigma }dV\), which allow one to derive each independent component of \(\overline{\mathbb {L}}^{Trans }\) according to the relations:

$$\begin{aligned}&3 \overline{K}=\frac{{\Sigma }^{\left( 0\right) }_{11}+2{\Sigma }^{\left( 0\right) }_{22}}{3d}\,,\quad 2\overline{G}_{12}=\frac{{\Sigma }^{\left( 0\right) }_{11}-{\Sigma }^{\left( 0\right) }_{22}+{\Sigma }^{\left( 1\right) }_{11}+2{\Sigma }^{\left( 1\right) }_{22}}{3d}\,,\quad 2 \overline{G}_1=\frac{2\left( {\Sigma }^{\left( 1\right) }_{11}-{\Sigma }^{\left( 1\right) }_{22}\right) }{3d}\,,\quad \nonumber \\ 2\overline{G}_2=\frac{{\Sigma }^{\left( 2\right) }_{23}}{d}\,,\quad 2 \overline{G}_3=\frac{{\Sigma }^{\left( 3\right) }_{12}}{d}\,,\quad \end{aligned}$$
(A.5)

where \({\Sigma }^{\left( n\right) }_{i\,j}\) denotes the ij component of the averaged stress tensor \(\mathbf {\Sigma }^{\left( n\right) }\) obtained as a material response to the strain \(\mathbf {E}^{\left( n\right) }\).

For the other shapes, i.e. the drilled sphere and crossed spheroids (Fig. 3c and d), the effective elastic stiffness tensor \(\overline{\mathbb {L}}^{Cub }\) of the unit cells is the anisotropic fourth-order tensor relevant for cubic symmetry with three Kelvin moduli \(3\overline{K}\), \(2\overline{G}_1\), and \(2\overline{G}_2\), namely [38]:

$$\begin{aligned} \overline{\mathbb {L}}^{Cub }= 3\overline{K}\mathbb {I}^{\mathrm{P}}+ 2\overline{G}_1(\mathbb {K}-\mathbb {I}^{P })+ 2\overline{G}_2(\mathbb {I}-\mathbb {K})\,, \end{aligned}$$
(A.6)

where

$$\begin{aligned} \mathbb {K}=\sum _{k=1}^3\mathbf {m}_k\otimes \mathbf {m}_k\otimes \mathbf {m}_k\otimes \mathbf {m}_k\,,\quad \mathbb {I}^{P }=\frac{1}{3}\mathbf {I}\otimes \mathbf {I}\,, \end{aligned}$$
(A.7)

and \(\mathbf {m}_k\) are the main symmetry axes of the unit cell. The eigensubspaces corresponding to \(3\overline{K}\), \(2\overline{G}_1\), and \(2\overline{G}_2\) are one dimensional (the space of hydrostatic states), two dimensional, and three dimensional (two orthogonal subspaces of deviatoric states), respectively.

Only three FE analyses with strains \(\mathbf {E}^{\left( 0\right) }\), \(\mathbf {E}^{\left( 1\right) }\), and \(\mathbf {E}^{\left( 2\right) }\) imposed by periodic boundary conditions (Eq. 19) are needed to derive each Kelvin modulus independently, according to the relations:

$$\begin{aligned} 3 \overline{K}=\frac{\mathbf {\Sigma }^{\left( 0\right) }_{11}}{d}\,,\quad 2 \overline{G}_1=\frac{\mathbf {\Sigma }^{\left( 1\right) }_{11}}{d}\,,\quad 2 \overline{G}_2=\frac{\mathbf {\Sigma }^{\left( 2\right) }_{23}}{d}\,.\quad \end{aligned}$$
(A.8)

It can be seen that the above three strains \(\mathbf {E}^{\left( 0\right) }\), \(\mathbf {E}^{\left( 1\right) }\), and \(\mathbf {E}^{\left( 2\right) }\) belong to three distinct eigensubspaces of the tensor \(\overline{\mathbb {L}}^{Cub }\) and result in the overall stresses \(\mathbf {\Sigma }^{\left( 0\right) }\), \(\mathbf {\Sigma }^{\left( 1\right) }\), and \(\mathbf {\Sigma }^{\left( 2\right) }\) of the same structure as strains. Finally, the response of a unit volume element reinforced with a single tiny spherical inhomogeneity (Fig. 4) is isotropic.

The situation changes in the case of unit cells with the FCC placement of inhomogeneities (Fig. 9), which for spherical inclusions are of cubic symmetry. If the composite is reinforced with prolate (Fig. 9c) or oblate (Fig. 9e) spheroids, the effective stiffness tensor \(\overline{\mathbb {L}}\) has tetragonal symmetry and can be written in the following form:

$$\begin{aligned} \overline{\mathbb {L}}^{Tetra }= 3 \overline{K} \mathbb {I}^{P }+ 2 \overline{G}_{1} \mathbb {P}_{1}+ 2 \overline{G}_{2} \hat{\mathbb {P}}_{2}+ 2 \overline{G}_{3} \mathbb {P}_{3}+ 2 \overline{G}_{4} \hat{\mathbb {P}}_{4}+ 2 \overline{G}_{12} \frac{1}{\sqrt{6}}\left( \mathbf {d}\otimes \mathbf {I} + \mathbf {I}\otimes \mathbf {d} \right) \,, \end{aligned}$$
(A.9)

where \(3\overline{K}\), \(2\overline{G}_1\), \(2 \overline{G}_{12}\), \(2 \overline{G}_2\), \(2 \overline{G}_3\), \(2 \overline{G}_4\) are six independent components. Note that it is assumed that the main axis of a prolate or oblate spheroid, or a drilled sphere, is coaxial with the \(\mathbf {m}_1\) direction, which is at the same time coaxial with one of the unit cell’s edges \(\mathbf {m}_k\). Under such conditions, the projectors \(\mathbb {P}_K\) (K=1,3) for tetragonal symmetry are the same as for the transversal one. The projectors \(\hat{\mathbb {P}}_2\) and \(\hat{\mathbb {P}}_4\) sum up to \(\mathbb {P}_2\) for transverse isotropy. They define two 1D orthogonal subspaces of shearing in the 2–3 plane, namely:

  • pure shear along the \(\mathbf {m}_2\) and \(\mathbf {m}_3\) directions,

    $$\begin{aligned} \hat{\mathbb {P}}_{2}= \frac{1}{2} \left[ \left( \mathbf {m}_2\otimes \mathbf {m}_3+\mathbf {m}_3\otimes \mathbf {m}_2\right) \otimes \left( \mathbf {m}_2\otimes \mathbf {m}_3+\mathbf {m}_3\otimes \mathbf {m}_2\right) \right] \,, \end{aligned}$$
    (A.10)
  • pure shear along directions inclined by \(45^o\) with respect to the \(\mathbf {m}_2\) and \(\mathbf {m}_3\) directions,

    $$\begin{aligned} \hat{\mathbb {P}}_{4}= \frac{1}{2} \left[ \left( \mathbf {m}_2\otimes \mathbf {m}_2-\mathbf {m}_3\otimes \mathbf {m}_3\right) \otimes \left( \mathbf {m}_3\otimes \mathbf {m}_3-\mathbf {m}_2\otimes \mathbf {m}_2\right) \right] \,. \end{aligned}$$
    (A.11)

To find the six independent components of \(\bar{\mathbb {L}}^{Tetra }\), the four analyses with the overall strains \(\mathbf {E}^{(n)}\) (n = 0, 1, 2, 3) given by Eqs. (19), performed in the case of the overall transverse isotropy, are completed with a fifth analysis with an imposed strain of the following representation in the basis \(\{\mathbf {m}_i\}\):

$$\begin{aligned} E^{\left( 4\right) }_{ij}= \left( \begin{array}{ccc} 0 &{} 0&{} 0 \\ 0 &{} d &{} 0 \\ 0 &{} 0 &{} -d \end{array}\right) \,, \end{aligned}$$
(A.12)

where d specifies the strain magnitude.

The imposed strains \(\mathbf {E}^{\left( n\right) }\) result in the overall stresses \(\mathbf {\Sigma }^{\left( n\right) }\), which allow one to derive each independent component of \(\overline{\mathbb {L}}^{Tetra }\) according to Eq. A.5 with the addition of \(\overline{G}_4\):

$$\begin{aligned} 2 \overline{G}_4=\frac{\mathbf {\Sigma }^{\left( 4\right) }_{22}}{d}\,. \end{aligned}$$
(A.13)

The effective stiffness of a composite with the FCC arrangement of inhomogeneities shown in Fig. 9a, b and d is of cubic symmetry as long as the main axes of ingomogeneities coincide with the axes of the unit cell.

Appendix B: The form of concentration tensors: transverse isotropy and cubic symmetry

The numerical strain concentration tensor \(\mathbb {A}_{i }^{NDil }\) of an inhomogeneity has the same symmetry group as the effective stiffness tensor of a large unit cell (Fig. 4) with a tiny inhomogeneity. Thus, for inhomogeneities having the shape of the prolate spheroid or drilled oblate spheroid (Fig. 3b and e) its symmetry group is that for transverse isotropy:

$$\begin{aligned} \mathbb {A}^{Trans }= A^{Trans }_0 \mathbb {I}^{P }+ A^{Trans }_1 \mathbb {P}_{1}+ A^{Trans }_2 \mathbb {P}_{2}+ A^{Trans }_3 \mathbb {P}_{3}+ \frac{1}{\sqrt{6}}\left( A^{Trans }_{12} \mathbf {d}\otimes \mathbf {I} +A^{Trans }_{21} \mathbf {I}\otimes \mathbf {d} \right) \,, \end{aligned}$$
(B.1)

and for the other shapes: drilled sphere and crossed spheroids (Fig. 3c and d), it has cubic symmetry, viz.

$$\begin{aligned} \mathbb {A}^{Cub }= A^{Cub }_0\mathbb {I}^{\mathrm{P}}+ A^{Cub }_1(\mathbb {K}-\mathbb {I}^{P })+ A^{Cub }_2(\mathbb {I}-\mathbb {K})\,, \end{aligned}$$
(B.2)

where the projectors \(\mathbb {P}_K\) and the remaining tensorial quantities, e.g. \(\mathbb {K}\), are listed above; see Eqs. A.2, A.3, and A.7. Finally, if a particle is spherical then the numerical strain concentration tensor is isotropic.

For the set of micro-periodic displacement boundary conditions specified by the strains \(\mathbf {E}^{(n)}\) from Eq. (19), the components \(A^{Cub }\) and \(A^{Trans }\), established from Eq. (15), are:

$$\begin{aligned}&A^{Trans }_0=\frac{\langle \varvec{\varepsilon }\rangle ^{\left( 0\right) }_{11}+2\langle \varvec{\varepsilon }\rangle ^{\left( 0\right) }_{22}}{3d}\,,\quad A^{Trans }_{12}\approx A^{Trans }_{21}=\frac{\langle \varvec{\varepsilon }\rangle ^{\left( 0\right) }_{11}-\langle \varvec{\varepsilon }\rangle ^{\left( 0\right) }_{22}+\langle \varvec{\varepsilon }\rangle ^{\left( 1\right) }_{11}+2\langle \varvec{\varepsilon }\rangle ^{\left( 1\right) }_{22}}{3d}\,,\quad A^{Trans }_1=\frac{2\left( \langle \varvec{\varepsilon }\rangle ^{\left( 1\right) }_{11}-\langle \varvec{\varepsilon }\rangle ^{\left( 1\right) }_{22}\right) }{3d}\,,\quad \nonumber \\&A^{Trans }_2=\frac{\langle \varvec{\varepsilon }\rangle ^{\left( 2\right) }_{23}}{d}\,,\quad A^{Trans }_3=\frac{\langle \varvec{\varepsilon }\rangle ^{\left( 3\right) }_{12}}{d}\,,\quad \nonumber \\&A^{Cub }_0=\frac{\langle \varvec{\varepsilon }\rangle ^{\left( 0\right) }_{11}}{d}\,,\quad A^{Cub }_1=\frac{\langle \varvec{\varepsilon }\rangle ^{\left( 1\right) }_{11}}{d}\,,\quad A^{Cub }_2=\frac{\langle \varvec{\varepsilon }\rangle ^{\left( 2\right) }_{23}}{d}\,,\quad \end{aligned}$$
(B.3)

where \(\langle \cdot \rangle\) is the volume averaging operation defined as \(1/V \int _V (\cdot )dV\), \(\varvec{\varepsilon }\) is the local strain tensor in the inhomogeneity domain, \(\langle \varvec{\varepsilon }\rangle ^{\left( n\right) }_{ij}\) is the component ij of the inhomogeneity’s average strain in response to the displacement BC given by \(\mathbf {E}^{\left( n\right) }\). The simplification \(A^{Trans }_{12}\approx A^{Trans }_{21}\) was assumed taking into account very small values of this component compared to the remaining ones (see Table 2).

Appendix C: The impact of inclusion shape on damage evolution

As it has been shown, the shape of inclusions may substantially influence the local strain and stress fields in the composite phases. These local fields have a prominent effect on the initiation of damage in the material [20]. Below, we analyze damage evolution in composites with the considered shapes of heterogeneities by means of numerical homogenization.

Material degradation is simulated by FEM using the concept of the damage parameter d and the framework of continuous damage mechanics [44]. Within the local constitutive model of the phases, the free energy density \(\Pi\) is postulated in the form

$$\begin{aligned} \Pi = (1-d) \Pi _e(\varvec{\varepsilon })\,, \end{aligned}$$
(C.1)

where \(\Pi _e\) is the elastic energy density specified as

$$\begin{aligned} \Pi _e = \frac{1}{2}\varvec{\varepsilon }\cdot \mathbb {L}_{i/m }\cdot \varvec{\varepsilon }\,. \end{aligned}$$
(C.2)

The damage parameter d depends on the history parameter \(\omega\) which is related to the elastic energy density according to the exponential law [45]:

$$\begin{aligned} d=d_{max } \left( 1 - Exp [- H \omega ]\right) , \end{aligned}$$
(C.3)

where \(d_{max }=0.8\) is taken as the maximum damage, \(H=0.1\) is the damage ductility, and \(\omega\) is equal to the maximum value of the elastic energy density \(\Pi _e\) achieved during the deformation process up to the considered time step.

Figure 16a presents the average value of the von Mises stress, calculated from the effective stress \(\bar{\sigma }\), versus the overall strain component \(\varepsilon _{23}\) of the MMC (Table 1) subjected to periodic boundary conditions (16) with \(\mathbf {E}=\mathbf {E}^{(2)}\). Similarly to [46], damage evolution is enabled only in the matrix phase. Figure 16b shows the average damage evolution \(d_{m }=1/V_{m } \int d \, d V_{m }\,\) in the matrix phase for five different inclusion shapes in the FCC spatial arrangement. The studied microstructures are presented in Fig. 9.

Fig. 16
figure 16

MMC with the FCC particle distribution with a volume fraction \(f_{i }=0.3\), subjected to the shear specified by \(\mathbf {E}^{\left( 2\right) }\) (Eq. 19). Damage development is enabled in the matrix phase only. a The overall equivalent von Mises stress \(\overline{\sigma }_{Mises }\), and b the averaged matrix damage d vs the overall strain component \(\overline{\varepsilon }_{23}\)

The results indicate that the shape of inhomogeneities plays a critical role in the non-linear response of the particulate composite. As can be seen in Fig. 16a, the difference in the overall von Mises stress between the crossed spheroids and the prolate spheroids is almost 500[MPa]. In all cases we observe first an increase in the stresses accompanied by a smooth reduction of stiffness. Next, some stabilization of the stress level is seen, so the strain increases under an approximately constant stress. Finally, the overall stresses start to increase again. The reason for this behaviour is the damage evolution law (Eq. C.1) and the assumption of the maximum damage \(d_{max }=0.8\) which is less than 1.0, so the elastic stiffness of the matrix never degrades to zero. When d reaches its maximum value \(d_{max }\), the locally damaged matrix becomes elastic but with a smaller stiffness: \((1-d_{max })\mathbb {L}_{m }\). The contour maps shown in Fig. 17 present the distribution of the damage parameter d at \(\overline{\varepsilon }_{23}=0.02\). Details of the distribution vary with the inclusion shape. In the MMC reinforced with prolate spheroids, wide and almost straight damage bands develop (see Fig. 17c), while for crossed spheroids damage bands are more curved (see Fig. 17d), and for drilled spheres damage develops in the matrix inside the inclusions (see Fig. 17b). In general, higher damage localization is observed in specimens with a smaller distance between the particles’ external surfaces and with more spherical shapes of inclusions.

Fig. 17
figure 17

Distribution of the damage parameter d (contour maps) at the strain \(\overline{\varepsilon }_{23}=0.02\) in the shear test specified by \(\mathbf {E}^{(2)}\) (Eq. 19). The MMC is reinforced with ceramic particles of five different shapes: a spheres, b drilled spheres, c prolate spheroids, d crossed spheroids, and e drilled oblate spheroids (see Fig. 3) placed in the FCC arrangement (Fig. 9). The damage evolution law of the matrix phase is specified by Eq. C.1 (damage evolution is disabled in the inclusions)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Majewski, M., Wichrowski, M., Hołobut, P. et al. Shape and packing effects in particulate composites: micromechanical modelling and numerical verification. Archiv.Civ.Mech.Eng 22, 86 (2022). https://doi.org/10.1007/s43452-022-00405-9

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s43452-022-00405-9

Keywords

Navigation