Skip to main content
Log in

The Elastic Properties of Dilute Solid Suspensions with Imperfect Interfacial Bonding: Variational Approximations Versus Full-Field Simulations

  • Research
  • Published:
Journal of Elasticity Aims and scope Submit manuscript

Abstract

Approximations for the elastic properties of dilute solid suspensions with imperfect interfacial bonding are derived and assessed. A variational procedure is employed in such a way that the resulting approximations reproduce exact results for weakly anisotropic but otherwise arbitrarily large interfacial compliances. Two approximations are generated which display the exact same format but differ in the way the interfacial compliance is averaged over the interfaces: the first approximation depends on an ‘arithmetic’ mean while the second approximation depends on a ‘harmonic’ mean. Both approximations allow for arbitrary elastic anisotropy of the constitutive phases but are restricted to suspended inclusions of spherical shape. The approximations are applied to a class of isotropic suspensions and confronted to full-field numerical simulations for assessment. Simulations are performed by means of a Fast Fourier Transform algorithm suitably implemented to handle dilute suspensions with imperfect interfaces. Also included in the comparisons are available results for suspensions with extremely anisotropic bondings. Overall, the ‘harmonic’ approximation is found to be much more precise than the ‘arithmetic’ approximation. The finding is of practical relevance given the widespread use of ‘arithmetic’ approximations in existing descriptions based on modified Eshelby tensors.

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

Similar content being viewed by others

References

  1. McDanels, D.L.: Analysis of stress-strain, fracture, and ductility behaviour of aluminum matrix composites containing discontinuous silicon carbide reinforcement. Metall. Trans. 16A, 1105–1115 (1985)

    Article  Google Scholar 

  2. Chen, P., Huang, F., Ding, Y.: Microstructure, deformation and failure of polymer bonded explosives. J. Mater. Sci. 42, 5272–5280 (2007)

    Article  Google Scholar 

  3. Ghahremani, F.: Effect of grain boundary sliding on anelasticity of polycrystals. Int. J. Solids Struct. 16, 825–845 (1980)

    Article  MATH  Google Scholar 

  4. Lene, F., Leguillon, D.: Homogenized constitutive law for a partially cohesive composite material. Int. J. Solids Struct. 18, 443–458 (1982)

    Article  MathSciNet  MATH  Google Scholar 

  5. Suquet, P.: Plasticité et Homogénéisation. Ph.D. Thesis, Université Pierre et Marie Curie, Paris (1982)

  6. Firooz, S., Steinmann, P., Javili, A.: Homogenization of composites with extended general interfaces: comprehensive review and unified modeling. Appl. Mech. Rev. 73, 040802 (2022)

    Article  Google Scholar 

  7. Hashin, Z.: Extremum principles for elastic heterogeneous media with imperfect interfaces and their application to bounding of effective moduli. J. Mech. Phys. Solids 40, 767–781 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  8. Qu, J.: The effect of slightly weakened interfaces on the overall elastic properties of composite materiales. Mech. Mater. 14, 269–281 (1993)

    Article  Google Scholar 

  9. Willis, J.R.: Variational and related methods for the overall properties of composites. Adv. Appl. Mech. 21, 1–78 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  10. Willis, J.R.: Elasticity theory of composites. In: Hopkins, H.G., Sewell, M.J. (eds.) Mechanics of Solids, The Rodney Hill 60th Anniversary Volume, pp. 653–686. Pergamon, Elmsford (1982)

    Google Scholar 

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

    MATH  Google Scholar 

  12. Shibata, S., Jasiuk, I., Mori, T., Mura, T.: Successive iteration method applied to composites containing sliding inclusions: effective modulus and anelasticity. Mech. Mater. 9, 229–243 (1990)

    Article  Google Scholar 

  13. Jasiuk, I., Chen, J., Thorpe, M.F.: Elastic moduli of composites with rigid sliding inclusions. J. Mech. Phys. Solids 40, 373–391 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  14. Michel, J.-C., Suquet, P., Thébaud, F.: Une modélisation du rôle des interfaces dans le comportement des composites à matrice métallique. Rev. Europ. élém. Finis 3, 573–595 (2012)

    Article  MATH  Google Scholar 

  15. Lipton, R., Vernescu, B.: Variational methods, size effects and extremal microgeometries for elastic composites with imperfect interface. Math. Models Methods Appl. Sci. 5, 1139–1173 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  16. Ponte Castañeda, P., Suquet, P.: Nonlinear composites. Adv. Appl. Mech. 34, 171–302 (1998)

    Article  MATH  Google Scholar 

  17. Hashin, Z.: The spherical inclusion with imperfect interface. J. Appl. Mech. 58, 444–449 (1991)

    Article  Google Scholar 

  18. Eshelby, J.D.: The determination of the elastic field of an ellipsoidal inclusion, and related problems. Proc. R. Soc. Lond. A 241, 376–396 (1957)

    Article  MathSciNet  MATH  Google Scholar 

  19. Idiart, M.I., Ponte Castañeda, P.: Field statistics in nonlinear composites. I. Theory. Proc. R. Soc. Lond. A 463, 183–202 (2007)

    MathSciNet  MATH  Google Scholar 

  20. Hashin, Z.: Thermoelastic properties of particulate composites with imperfect interface. J. Mech. Phys. Solids 39, 745–762 (1991)

    Article  MathSciNet  Google Scholar 

  21. Hashin, Z.: Thin interphase/imperfect interface in elasticity with application to coated fiber composites. J. Mech. Phys. Solids 50, 2509–2537 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  22. Dinzart, F., Sabar, H.: New micromechanical modeling of the elastic behavior of composite materials with ellipsoidal reinforcements and imperfect interfaces. Int. J. Solids Struct. 108, 254–262 (2017)

    Article  Google Scholar 

  23. Duan, H.L., Yi, X., Huang, Z.P., Wang, J.: A unified scheme for prediction of effective moduli of multiphase composites with interface effects. Part I: theoretical framework. Mech. Mater. 39, 81–93 (2007)

    Article  Google Scholar 

  24. Zecevic, M., Bennett, K.C., Luscher, D.J., Lebensohn, R.A.: New self-consistent homogenization for thermo-elastic polycrystals with imperfect interfaces. Mech. Mater. 155, 103651 (2021)

    Article  Google Scholar 

  25. Nazarenko, N., Stolarski, H.: Energy-based definition of equivalent inhomogeneity for various interphase models and analysis of effective properties of particulate composites. Compos. B 94, 82–94 (2016)

    Article  Google Scholar 

  26. Zecevic, M., Lebensohn, R.A., Capolungo, L.: New large-strain FFT-based formulation and its application to model strain localization in nano-metallic laminates and other strongly anisotropic crystalline materials. Mech. Mater. 166, 104208 (2022)

    Article  Google Scholar 

  27. Bigoni, D., Movchan, A.B.: Statics and dynamics of structural interfaces in elasticity. Int. J. Solids Struct. 39, 4843–4865 (2002)

    Article  MATH  Google Scholar 

  28. Othmani, Y., Delannay, L., Doghri, I.: Equivalent inclusion solution adapted to particle debonding with a non-linear cohesive law. Int. J. Solids Struct. 48, 3326–3335 (2011)

    Article  Google Scholar 

  29. Eisenlohr, P., Diehl, M., Lebensohn, R.A., Roters, F.: A spectral method solution to crystal elasto-viscoplasticity at finite strains. Int. J. Plast. 46, 37–53 (2013)

    Article  Google Scholar 

  30. Lucarini, S., Upadhyay, M.V., Segurado, J.: FFT based approaches in micromechanics: fundamentals, methods and applications. Model. Simul. Mater. Sci. Eng. 30, 023002 (2021)

    Article  Google Scholar 

  31. Zecevic, M., Lebensohn, R.A., Capolungo, L.: Non-local large-strain FFT-based formulation and its application to interface-dominated plasticity of nano-metallic laminates. J. Mech. Phys. Solids 173, 105187 (2023)

    Article  MathSciNet  Google Scholar 

  32. Walpole, L.J.: Fourth-rank tensors of the thirty-two crystal classes: multiplication tables. Proc. R. Soc. Lond. A 391, 149–179 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  33. Kushch, V.I.: Elastic equilibrium of spherical particle composites with transversely isotropic interphase and incoherent material interface. Int. J. Solids Struct. 232, 111180 (2021)

    Article  Google Scholar 

  34. Zhong, Z., Meguid, S.A.: On the elastic field of a spherical inhomogeneity with an imperfectly bonded interface. J. Elast. 46, 91–113 (1997)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  Google Scholar 

  36. Lebensohn, R.A., Kanjarla, A.K., Eisenlohr, P.: An elasto-viscoplastic formulation based on fast Fourier transforms for the prediction of micromechanical fields in polycrystalline materials. Int. J. Plast. 32, 59–69 (2012)

    Article  Google Scholar 

  37. Idiart, M.I., Lahellec, N., Suquet, P.: Model reduction by mean-field homogenization in viscoelastic composites. I. Primal theory. Proc. R. Soc. A 476, 20200407 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  38. Lahellec, N., Idiart, M.I., Suquet, P.: Model reduction by mean-field homogenization in viscoelastic composites. III. Dual theory. Proc. R. Soc. A 477, 20200869 (2021)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The work of V.G. and M.I.I. was supported by the Air Force Office of Scientific Research (U.S.A.) under award number FA9550-19-1-0377 and by the Universidad Nacional de La Plata (Argentina) through grant I-2019-247. The work of M.Z. and R.A.L. was supported by Los Alamos National Laboratory’s Directed Research and Development (LDRD) program. V.G. is grateful to Prof. F. Bacchi of UNLP for assisting with the computer cluster.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Martín I. Idiart.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

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

Appendices

Appendix A: Auxiliary Problem

Consider the displacement field \({\mathbf{v}}({\mathbf{x}})\) that solves the set of equations

$$\begin{aligned} &\nabla \cdot \boldsymbol{\sigma}= {\mathbf{0}}, \quad \boldsymbol{\sigma}= \mathbb{C}^{(1)} \boldsymbol{\varepsilon}+ \chi ^{(2)}({\mathbf{x}})\ \boldsymbol{\tau}^{(2)}, \quad \boldsymbol{\varepsilon}= \overline{\boldsymbol{\varepsilon}}+ \nabla \otimes _{s} {\mathbf{v}}({ \mathbf{x}}) \quad \mbox{in} \quad \varOmega \backslash \varGamma ^{(1,2)}, \end{aligned}$$
(87)
$$\begin{aligned} &[\!\![\boldsymbol{\sigma}]\!\!]{\mathbf{n}}= {\mathbf{0}}, \quad [\!\![{\mathbf{v}}]\!\!]= {\mathbf{d}}({\mathbf{x}}), \quad {\mathbf{d}}({ \mathbf{x}}) = a{\mathbf{D}}^{(1,2)}{\mathbf{n}}({\mathbf{x}}) \quad \quad \mbox{on} \quad \varGamma ^{(1,2)}, \end{aligned}$$
(88)
$$\begin{aligned} &{\mathbf{v}}({\mathbf{x}}) = {\mathbf{0}} \quad \mbox{on} \quad \partial \varOmega , \end{aligned}$$
(89)

for prescribed second-order symmetric tensors \(\boldsymbol{\tau}^{(2)}\) and \({\mathbf{D}}^{(1,2)}\). We intend to determine the corresponding strain field within the inclusion phase. To that end, we begin by condensing these equations into

$$\begin{aligned} \nabla \cdot \left [ \mathbb{C}^{(1)} \nabla \otimes _{s} {\mathbf{v}}({ \mathbf{x}}) \right ] + \nabla \cdot \boldsymbol{\tau}({\mathbf{x}}) = { \mathbf{0}} \quad \mbox{in} \quad \varOmega \end{aligned}$$
(90)

with

$$\begin{aligned} \boldsymbol{\tau}({\mathbf{x}}) = \chi ^{(2)}({\mathbf{x}})\ \boldsymbol{\tau}^{(2)} - \delta ^{(1,2)}({\mathbf{x}})\ \mathbb{C}^{(1)} \mathbb{E}\left ({\mathbf{n}}({\mathbf{x}})\right ) {\mathbf{D}}^{(1,2)}, \end{aligned}$$
(91)

which should be understood in the sense of distributions. Here, \(\mathbb{E}({\mathbf{n}}) = a\ {\mathbf{I}}\boxtimes ({\mathbf{n}}\otimes {\mathbf{n}})\) and \(\delta ^{(1,2)}({\mathbf{x}})\) refers to the Dirac delta function with support on the interface \(\varGamma ^{(1,2)}\). Making use of the Radon transform of equation (90) and its inverse we can write the strain field as

$$\begin{aligned} \boldsymbol{\varepsilon}({\mathbf{x}}) = \overline{\boldsymbol{\varepsilon}}+ \frac{1}{8\pi ^{2}} \int _{| \boldsymbol{\xi}|=1} \mathbb{H}(\boldsymbol{\xi})\ \frac{\partial ^{2} \breve{\boldsymbol{\tau}}}{\partial p^{2}}( \boldsymbol{\xi},\boldsymbol{\xi}\cdot {\mathbf{x}})\ {\mathrm{d}} \boldsymbol{\xi} \end{aligned}$$
(92)

with

$$\begin{aligned} \mathbb{H}_{\mathit{ijkl}}(\boldsymbol{\xi}) = \left . K_{ik}^{-1}( \boldsymbol{\xi}) \xi _{j}\xi _{l} \right |_{(ij)(kl)} \quad \mbox{and} \quad K_{ik}(\boldsymbol{\xi}) = \mathbb{C}^{(1)}_{\mathit{ijkl}} \xi _{j} \xi _{l}, \end{aligned}$$
(93)

and the Radon transform of the polarization field \(\boldsymbol{\tau}({\mathbf{x}})\) given by

$$\begin{aligned} \breve{\boldsymbol{\tau}}(\boldsymbol{\xi},p) &= \int _{ \mathbb{R}^{3}} \boldsymbol{\tau}({\mathbf{x}})\ \delta ( \boldsymbol{\xi}\cdot {\mathbf{x}}- p)\ {\mathrm{d}}{\mathbf{x}}. \end{aligned}$$
(94)

Given that inclusions are well separated, the strain field within any particular inclusion can be computed by neglecting the presence of all the other inclusions. This amounts to identifying the functions \(\chi ^{(2)}({\mathbf{x}})\) and \(\delta ^{(1,2)}({\mathbf{x}})\) in (91) with those of a single inclusion centered at the origin of an infinite medium. In that case,

$$\begin{aligned} \breve{\boldsymbol{\tau}}(\boldsymbol{\xi},p) &= - \breve{\mathbb{T}}(\boldsymbol{\xi},p) \boldsymbol{\tau}^{(2)} - \mathbb{C}^{(1)} \breve{\mathbb{E}}(\boldsymbol{\xi},p) {\mathbf{D}}^{(1,2)}, \end{aligned}$$
(95)

where

$$\begin{aligned} \breve{\mathbb{T}}(\boldsymbol{\xi},p) &= \int _{\mathbb{R}^{3}} \chi ^{(2)}({\mathbf{x}})\ \delta (\boldsymbol{\xi}\cdot {\mathbf{x}}- p) \ \mathbb{I}\ {\mathrm{d}}{\mathbf{x}}= \int _{\mathcal{S}( \boldsymbol{\xi},p)} \mathbb{I}\ {\mathrm{d}}{\mathbf{x}}, \end{aligned}$$
(96)
$$\begin{aligned} \breve{\mathbb{E}}(\boldsymbol{\xi},p) &= \int _{\mathbb{R}^{3}} \delta ^{(1,2)}({\mathbf{x}})\ \mathbb{E}\left ({\mathbf{n}}({\mathbf{x}})\right )\ \delta (\boldsymbol{\xi}\cdot {\mathbf{x}}- p)\ {\mathrm{d}}{\mathbf{x}}=\int _{ \mathcal{C}(\boldsymbol{\xi},p)} \frac{\mathbb{E}\left ({\mathbf{n}}({\mathbf{x}})\right )}{\left [1 - (\boldsymbol{\xi}\cdot {\mathbf{n}}({\mathbf{x}}))^{2}\right ]^{1/2}} {\mathrm{d}}{\mathbf{x}}, \end{aligned}$$
(97)

and the sets \(\mathcal{S}(\boldsymbol{\xi},p) = \left \{{\mathbf{x}}:\ |{\mathbf{x}}| \leq a \ \mbox{and} \ \boldsymbol{\xi}\cdot {\mathbf{x}}=p \right \}\) and \(\mathcal{C}(\boldsymbol{\xi},p) = \left \{{\mathbf{x}}:\ |{\mathbf{x}}| = a \ \mbox{and} \ \boldsymbol{\xi}\cdot {\mathbf{x}}=p \right \}\) are the intersections of a Radon plane with the spherical inclusion and with its surface, respectively; computations give

$$\begin{aligned} \breve{\mathbb{T}}(\boldsymbol{\xi},p) &= \pi \left (a^{2} - p^{2} \right ) \mathbb{I}\ H\left (1 - \frac{|p|}{a}\right ) \quad \mbox{and} \end{aligned}$$
(98)
$$\begin{aligned} \breve{\mathbb{E}}(\boldsymbol{\xi},p) &= 2\pi \left [p^{2} \boldsymbol{\xi}\otimes \boldsymbol{\xi}+ \frac{a^{2} - p^{2}}{2} ({\mathbf{I}}- \boldsymbol{\xi}\otimes \boldsymbol{\xi}) \right ] \boxtimes {\mathbf{I}}\ H\left (1 - \frac{|p|}{a}\right ), \end{aligned}$$
(99)

where \(H(\cdot )\) is the Heaviside step function. Introducing these expressions into (92) then gives an expression for the strain field within the spherical inclusion (\(|{\mathbf{x}}|< a\)) of the form

$$\begin{aligned} \boldsymbol{\varepsilon}({\mathbf{x}}) = \overline{\boldsymbol{\varepsilon}}- \mathbb{P} \boldsymbol{\tau}^{(2)} - (\mathbb{I} - \mathbb{P} \mathbb{C}^{(1)}) {\mathbf{D}}^{(1,2)}, \end{aligned}$$
(100)

where ℙ is a microstructural tensor given by

$$\begin{aligned} \mathbb{P} = \frac{1}{4\pi} \int _{|\boldsymbol{\xi}|=1} \mathbb{H}(\boldsymbol{\xi}) \ {\mathrm{d}}\boldsymbol{\xi}. \end{aligned}$$
(101)

Thus, the strain field within every inclusion of the suspension is uniform and given by expression (100). An equivalent analysis in terms of Green’s functions was originally carried out by Qu [8] but contained an error later amended by Othmani et al. [28]. The above procedure avoids the problem by making use of Radon transforms instead of Green’s functions. The procedure was initially advocated by Willis [10] to study inclusion problems with perfect interfacial bonding.

Appendix B: A Numerical Method Based on the Fast Fourier Transform

2.1 B.1 Method

Following Michel et al. [14] and Hashin [21], the mechanical problem (3) for suspensions with imperfect interfaces is replaced by the mechanical problem

$$\begin{aligned} \widetilde{w}(\overline{\boldsymbol{\varepsilon}}) = \min _{ \boldsymbol{\varepsilon}\in \mathcal{K}( \overline{\boldsymbol{\varepsilon}})} \left [ \sum _{r=1}^{2} c^{(r)} \left \langle \frac{1}{2} \boldsymbol{\varepsilon}\cdot \mathbb{C}^{(r)}\boldsymbol{\varepsilon}\right \rangle ^{(r)} + c^{(1,2)} \Big\langle \frac{1}{2} \boldsymbol{\varepsilon}\cdot \mathbb{C}^{(1,2)}({\mathbf{n}}({\mathbf{x}})) \boldsymbol{\varepsilon} \Big\rangle ^{(1,2)} \right ] \end{aligned}$$
(102)

for suspensions with a thin compliant interphase \(\varOmega ^{(1,2)} = \varGamma ^{(1,2)}\times (0,h)\) of finite thickness \(h\) coating each inclusion and perfectly bonded to the matrix and inclusion phases. Here, \(c^{(1,2)}\) and \(\langle \cdot \rangle ^{(1,2)}\) denote —with slight abuse of notation— the volume fraction and the volume average over the interphase \(\varOmega ^{(1,2)}\), respectively, and \(\mathcal{K}(\overline{\boldsymbol{\varepsilon}})\) is now the set of strain fields of the form \(\boldsymbol{\varepsilon}({\mathbf{x}}) = \overline{\boldsymbol{\varepsilon}}+ \nabla \otimes _{s}{\mathbf{v}}({ \mathbf{x}})\) with \({\mathbf{v}}:\varOmega \rightarrow \mathbb{R}^{3}\) square-integrable with compact support. Note that the interphasial elasticity tensor \(\mathbb{C}^{(1,2)}({\mathbf{n}})\) is allowed to depend on the unit vector \({\mathbf{n}}({\mathbf{x}}) = {\mathbf{x}}/|{\mathbf{x}}|\) defined by the position vector \({\mathbf{x}}\) relative to the center of each spherical inclusion. By choosing positive-definite tensors \(\mathbb{C}_{\mathit{ijkl}}^{(1,2)}({\mathbf{n}})\) such that

$$\begin{aligned} \eta _{ik}^{-1}({\mathbf{n}}) = h^{-1} \mathbb{C}_{\mathit{ijkl}}^{(1,2)}({\mathbf{n}}) n_{j} n_{l} \quad \mbox{and} \quad ||\mathbb{C}^{(1,2)}({\mathbf{n}})|| < < || \mathbb{C}^{(r)}||, \end{aligned}$$
(103)

the mechanical problems (3) and (102) coincide in the limit of vanishing thickness \(h \rightarrow 0\).

The mechanical fields in (102) are numerically determined by considering a periodic cubic array of sufficiently small spherical inclusions and solving the Euler-Lagrange equations via an FFT-based algorithm. This FFT-based method has been extensively employed to determine mechanical fields in periodic media with a wide variety of local constitutive responses. When the number of elements/voxels is large —as in this work—, FFT-based methods are faster and require less computer memory than finite element methods (FEM) [29, 30]. In addition, periodic boundary conditions are natural in FFT-based methods while cumbersome and complex in FEM methods. In any event, the presence of a thin interphase in FFT simulations requires special treatment. Standard FFT-based implementations would discretize the spherical inclusion and the interphase into a regular grid of cuboidal voxels with two deleterious consequences. First, the smoothly curved boundaries of the thin interphase are represented in that case by a set of mutually orthogonal planes leading to spurious local stress concentrations and numerical contingencies; this is especially pronounced in the case of soft interphase behavior like the one required in this work. Second, small interphase thickness-to-inclusion radius ratios like the ones required in this work could only be achieved with unreasonably fine grid resolutions in view of the fact that the interphase thickness should contain at least one voxel. These contigencies are circumvented here following the strategy of Zecevic et al. [26], whose central idea is to embed the mechanical problem (102) into a large-deformation formulation with a suitably chosen reference configuration that differs from the initial and deformed configurations in such a way that FFT computations are facilitated. Specifically, a reference configuration is identified here with a square array of aligned cuboidal inclusions. The initial configuration is then generated by first applying the conformal mapping

$$\begin{aligned} x_{1}=X_{1}\sqrt{1-\frac{1}{2}\left (\frac{X_{2}}{a}\right )^{2}- \frac{1}{2}\left (\frac{X_{3}}{a}\right )^{2}+\frac{1}{3}\left ( \frac{X_{2}X_{3}}{a^{2}}\right )^{2}} \\ x_{2}=X_{2}\sqrt{1-\frac{1}{2}\left (\frac{X_{1}}{a}\right )^{2}- \frac{1}{2}\left (\frac{X_{3}}{a}\right )^{2}+\frac{1}{3}\left ( \frac{X_{1}X_{3}}{a^{2}}\right )^{2}} \\ x_{3}=X_{3}\sqrt{1-\frac{1}{2}\left (\frac{X_{1}}{a}\right )^{2}- \frac{1}{2}\left (\frac{X_{2}}{a}\right )^{2}+\frac{1}{3}\left ( \frac{X_{1}X_{2}}{a^{2}}\right )^{2}} \end{aligned}$$
(104)

between reference Cartesian coordinates \(X_{i}\) of points belonging to a cuboidal inclusion of edge \(2a\) and initial Cartesian coordinates \(x_{i}\) of points belonging to a spherical inclusion of radius \(a\). Second, the coordinates of two layers of grid points belonging to the interphase in the initial configuration are obtained using an analogous mapping from cubes to spheres with radii \((1+f)a\) and \((1+2f)a\), where \(f\) defines the distance between interphase grid points in the initial configuration and the thickness of the interphase. The layer of grid points belonging to the matrix adjacent to the interphase is mapped to a sphere with radius \((1+3f)a\). Finally, the coordinates of the remaining grid points of the matrix domain in the initial configuration are obtained by interpolation with the sides of the unit cell. In doing so, the spacing between matrix coordinates is increased with increasing distance from the inclusion. Corresponding conformal displacements are calculated as \({\mathbf{u}}^{c} = {\mathbf{x}}- {\mathbf{X}}\), and the conformal deformation gradient is given by \({\mathbf{F}}^{c} = {\mathbf{I}}+ \partial {\mathbf{u}}^{c}/\partial {\mathbf{X}}\). A cross section of the resulting initial configuration is shown in Fig. 3b for \(f = 0.001\).

Fig. 3
figure 3

Zoom-in of the y-z cross section, showing upper right quadrant of the inhomogeneity in the (a) reference and (b) initial configurations

Finally, the deformed configuration is generated by composing this stress-free deformation with a small-strain deformation dictated by the equilibrium equations associated with the mechanical problem (102). The latter amount to solving for the Cauchy stress and Eulerian strain rate fields in the current/initial configuration satisfying the equilibrium, compatibility and constitutive equations using periodic Green’s function. The original unit cell, discretized on an irregular grid, is replaced by a reference linear heterogeneous medium in the initial configuration, to which a stress polarization field is applied. The heterogeneity of the reference medium accounts for the distortion of the grid, while the polarization field accounts for the heterogeneity of the unit cell. The stress in the heterogeneous linear reference medium is given by

$$\begin{aligned} \boldsymbol{\sigma}({\mathbf{x}})=\mathbb{L({\mathbf{x}})} \frac{\partial \dot{{\mathbf{u}}}}{\partial {\mathbf{x}}}({\mathbf{x}})+ \boldsymbol{\phi}({\mathbf{x}}), \end{aligned}$$
(105)

where \(\mathbb{L}_{\mathit{ijkl}}({\mathbf{x}}) = \mathbb{L}^{\mathit{ref}}_{\mathit{iokp}}F_{\mathit{jo}}({\mathbf{X}}) F_{lp}({ \mathbf{X}})/J({\mathbf{X}})\) is the spatially varying linear viscous stiffness in the current configuration, \(\mathbb{L}^{\mathit{ref}}\) is a constant linear viscous stiffness in the reference configuration, \(J\) is determinant of the deformation gradient, and \(\boldsymbol{\phi}\) is the Cauchy stress polarization field in the current configuration. For a given polarization field, the velocity gradient in the linear reference medium is given by

$$\begin{aligned} \frac{\partial \dot{{\mathbf{u}}}}{\partial {\mathbf{x}}}({\mathbf{x}})&=-\left [ \int _{\varOmega ^{\mathit{ref}}}\mathbb{G}^{\mathit{ref}}({\mathbf{X}}-{\mathbf{X}}')[J({\mathbf{X}}')( \boldsymbol{\phi}({\mathbf{x}}')-\left < \boldsymbol{\phi} \right >){\mathbf{F}}^{-T}({\mathbf{X}}')] d{\mathbf{X}}'\right ]{\mathbf{F}}^{-1}({\mathbf{X}}) \\ &{}+ \frac{\partial \dot{\mathbf{{U}}}}{\partial {\mathbf{X}}}{\mathbf{F}}^{-1}({\mathbf{X}})+ \left < \frac{\partial \dot{{\mathbf{u}}}}{\partial {\mathbf{x}}}\right >, \end{aligned}$$
(106)

where \(\mathbb{G}^{\mathit{ref}}\) is the periodic Green’s operator in reference configuration. The above convolution integral is performed over the unit cell in the reference configuration, \(\varOmega ^{\mathit{ref}}\), where grid is regular. Therefore, the convolution theorem can be used and the convolution integral is efficiently calculated as a point-wise product in Fourier space. The system of equations given by Eq. (106) is solved using an augmented Lagrangian method [35, 36] modified for the case of large grid distortions [31].

2.2 B.2 Algorithm

In the augmented Lagrangian (AL) method, auxiliary stress and velocity gradient fields, \(\boldsymbol{\lambda}\) and \(\partial \dot{{\mathbf{u}}}/\partial {\mathbf{x}}\), are introduced and iteratively adjusted until the auxiliary stress and the symmetric part of the auxiliary velocity gradient fields converge to the constitutively related stress and strain rate, respectively. The AL algorithm reads:

  1. 1.

    Calculate new guess for the auxiliary velocity gradient field:

    $$\begin{aligned} \frac{\partial \dot{{\mathbf{u}}}}{\partial {\mathbf{x}}}^{i+1}= \frac{\partial \dot{{\mathbf{u}}}}{\partial {\mathbf{x}}}^{i}-\left [ \mathbb{G}^{\mathit{ref}}*(J(\boldsymbol{\lambda}^{i}-\left < \boldsymbol{\lambda}^{i}\right >){\mathbf{F}}^{-T})\right ]{\mathbf{F}}^{-1}+ \frac{\partial \dot{\mathbf{{U}}}}{\partial {\mathbf{X}}}{\mathbf{F}}^{-1}, \end{aligned}$$

    where “∗” represents convolution.

  2. 2.

    Calculate new guess for the stress field by solving the residual at each grid point:

    $$\begin{aligned} {\mathbf{R}}(\boldsymbol{\sigma}^{i+1})=\boldsymbol{\sigma}^{i+1}+ \mathbb{L}\dot{\boldsymbol{\varepsilon}}( \boldsymbol{\sigma}^{i+1})-\boldsymbol{\lambda}^{i}- \mathbb{L}\frac{\partial \dot{{\mathbf{u}}}}{\partial {\mathbf{x}}}^{i+1}+ \mathbb{L}[\mathbb{G}^{\mathit{ref}}(\mathbf{0})(J(\boldsymbol{\sigma}^{i+1}- \boldsymbol{\sigma}^{i}){\mathbf{F}}^{-T}){\mathbf{F}}^{-1}]=\mathbf{0}. \end{aligned}$$
  3. 3.

    The new guess for the auxiliary stress field is:

    $$\begin{aligned} \boldsymbol{\lambda}^{i+1}=\boldsymbol{\lambda}^{i}+ \mathbb{L}\left (\frac{\partial \dot{{\mathbf{u}}}}{\partial {\mathbf{x}}}^{i+1}- \dot{\boldsymbol{\varepsilon}}^{i+1}\right )= \boldsymbol{\sigma}^{i+1}. \end{aligned}$$
  4. 4.

    Convergence test.

In comparison to the original augmented Lagrangian method, the algorithm given above differs in the expression for residual in step 2. The last term in the expression of the residual is introduced, which accounts for the effect of the new guess for stress at each grid point, \(\boldsymbol{\sigma}^{i+1}\), on the velocity gradient at the same grid point, \((\partial \dot{{\mathbf{u}}}/\partial {\mathbf{x}})^{i+1}\), through the convolution Eq. (106) [31]. Note that effect of the new guess for stress in the surrounding grid points on the velocity gradient at the current grid point is not considered. Also note that once the fields are converged, \(\boldsymbol{\sigma}^{i}\approx \boldsymbol{\sigma}^{i+1}\), and the last term vanishes.

2.3 B.3 Calibration and Verification

Full-field simulations are carried out to determine the effective properties of the isotropic dilute suspensions introduced in Sect. 6. To that end, interphasial elasticity tensors of the form

$$\begin{aligned} \mathbb{C}^{(1,2)}({\mathbf{n}}) = f_{1}\mathbb{E}^{[1]}({\mathbf{n}}) + f_{2} \mathbb{E}^{[2]}({\mathbf{n}}) + f_{3}\mathbb{E}^{[3]}({\mathbf{n}}) + f_{4} \mathbb{E}^{[4]}({\mathbf{n}}) + f_{5}\mathbb{E}^{[5]}({\mathbf{n}}) + f_{6} \mathbb{E}^{[6]}({\mathbf{n}}) \end{aligned}$$
(107)

are employed, where the set \(\mathbb{E}^{[b]}({\mathbf{n}})\) (\(b=1,\ldots,6\)) constitutes Walpole’s basis for transversely isotropic tensors with symmetry axis \({\mathbf{n}}\) [32], and the moduli \(f_{b}\) must satisfy the conditions

$$\begin{aligned} f_{1} + f_{2} > 0, \quad f_{1} f_{2} - f_{3}^{2} > 0, \quad f_{5} > 0, \quad f_{6} > 0, \quad f_{3} = f_{4}, \end{aligned}$$
(108)

so that the tensor \(\mathbb{C}^{(1,2)}({\mathbf{n}})\) is symmetric and positive-definite. For interfacial compliances of the form (64), condition (103)1 requires that

$$\begin{aligned} f_{1} = \frac{h}{\eta _{\parallel }} \quad \mbox{and} \quad f_{6} = \frac{2h}{\eta _{\perp}}. \end{aligned}$$
(109)

The remaining moduli can be chosen arbitrarily provided conditions (103)2 and (108) are satisfied; a convenient choice is

$$\begin{aligned} f_{2} = f_{1}, \quad f_{3} = f_{4} = 0, \quad f_{5} = f_{6}. \end{aligned}$$
(110)

It is recalled that, while mathematically simpler, interphasial elastiticy tensors with isotropic —rather than transversely isotropic— symmetry do not give access to the entire range of possible values for \(\eta _{\parallel }\) and \(\eta _{\perp}\), as already noted by Hashin [17]. The connection between thin interphases with transversely isotropic elasticity and sharp imperfect interfaces has been recently investigated by Kushch [33]. It should be noted, however, that even though the above choice gives access to a much wider range of interfacial compliances, the range remains constrained by the additional inequality (103)2. Indeed, noting that for the above choice of interphasial elasticity tensor

$$\begin{aligned} ||\mathbb{C}^{(1,2)}|| \leq \frac{h}{\eta _{\parallel }} ||\mathbb{E}^{[1]} + \mathbb{E}^{[2]}|| + \frac{2h}{\eta _{\perp}} ||\mathbb{E}^{[5]} + \mathbb{E}^{[6]}|| = \frac{\sqrt{2} h}{\eta _{\parallel }} + \frac{4h}{\eta _{\perp}} \end{aligned}$$
(111)

in view of the triangle inequality, and that

$$\begin{aligned} ||\mathbb{C}^{(r)}|| > 2\mu ^{(r)} ||\mathbb{K}|| = 2\sqrt{5} \mu _{*} \end{aligned}$$
(112)

with \(\mu _{*} = {\mathrm{min}}\{\mu ^{(1)}, \mu ^{(2)} \}\), it is easy to verify that a sufficient condition for the inequality (103)2 to hold is

$$\begin{aligned} \eta _{s} >> \frac{h/a}{\sqrt{5}} \left (\frac{\sqrt{2}/2}{1+2\alpha} + \frac{2}{1-\alpha} \right ) a\mu _{*}^{-1}, \end{aligned}$$
(113)

where the interfacial parameters \(\eta _{s}\) and \(\alpha \) are defined in terms of \(\eta _{\parallel }\) and \(\eta _{\perp}\) by (79) to provide the alternative representation (78) for the interfacial compliance tensor. Given the smallness of the ratio \(h/a\) for the microgeometries of interest in this work, we conclude that the inequality

$$\begin{aligned} \eta _{s} \geq \frac{1}{\sqrt{5}} \left (\frac{\sqrt{2}/2}{1+2\alpha} + \frac{2}{1-\alpha} \right ) a\mu _{*}^{-1} \end{aligned}$$
(114)

warrants admissibility of interfacial parameters. Given that the right-hand side of this inequality is strictly positive, the parameter \(\eta _{s}\) cannot vanish. Consequently, the thin interphase model cannot mimic a perfect interface. According to this criterion, when mimicking isotropic interfaces (\(\alpha =0\)) the model is constrained to \(\eta _{s} \gtrsim 1.21\ a \mu _{*}^{-1}\), whereas when mimicking strongly anisotropic interfaces (\(\alpha \rightarrow -1/2\) or \(\alpha \rightarrow 1\)) the model is constrained to imperfect interfaces with large compliances \(\eta _{s}\).

In any event, for this choice of local constitutive tensors, the effective bulk modulus of the isotropic suspension can be determined exactly in terms of the interphase thickness \(h\) following the strategy of Hashin [20] for concentric shells; the result is

$$\begin{aligned} \widetilde{\kappa} = \kappa ^{(1)} + c \frac{(1+h/a)^{3}}{\displaystyle \frac{1}{\kappa ^{(1,2)} - \kappa ^{(1)}} + \frac{1}{\kappa ^{(1)} + 4\mu ^{(1)}/3}}, \end{aligned}$$
(115)

where

$$\begin{aligned} \kappa ^{(1,2)} = \frac{h}{3\eta _{\parallel }} + \frac{1}{\displaystyle \frac{(1+h/a)^{3}}{\kappa ^{(2)} - h/(3\eta _{\parallel })} + \frac{(1 + h/a )^{3} - 1}{h/\eta _{\parallel }}} \end{aligned}$$
(116)

Central to the derivation of this result are the choices (110)1 and (110)2; this class of transversely isotropic elasticity tensors produces the same form of radial displacement fields as their isotropic counterparts. In the limit of sharp imperfect interfaces (\(h \rightarrow 0\)), expression (115) reduces to (66). This exact result can be exploited for calibration and verification of the numerical simulations. Figure 4a compares the dilute correction factors \(\beta _{\kappa }= (\widetilde{\kappa}/\kappa ^{(1)}-1)/c\) predicted by the exact result (115) and the FFT simulations, as a function of the ratio \(h/a\), for interfaces with a spherical compliance \(\eta_{s} = a/\mu^{(1)}\) and various interfacial anisotropies (\(\alpha = -0.25,0,0.75\)). The FFT simulations discretize the unit cell into \(256^{3}\) voxels such that \(92^{3}\) grid points lie within the inclusion, and make use of 2 voxels across the interphase. These settings generate inclusion contents of approximately \(c \approx 2.4\times 10^{-3}\), which were found small enough to mimic dilute suspensions. In all simulations a tolerance of \(10^{-6}\) for the change in stress/strain-rate fields was required for convergence. Also included in the figure is the limiting value of (66) for sharp imperfect interfaces (\(h \rightarrow 0\)). Two observations are in order. First, the FFT results are seen to reproduce the exact result for the entire range of interphasial thicknesses considered, thus confirming the validity of the various parameters employed in the simulations. Second, the FFT simulations with the smallest thickness considered (\(h/a = 10^{-4}\)) are seen to accurately reproduce the exact result for suspensions with sharp imperfect interfaces. This grid size and interphase thickness were therefore chosen to produce reference results for suspensions with sharp imperfect interfaces more generally. As the thickness of the interphase reduces,the interaction between different points within the interphase as defined by the integral equation (106) also reduces due to the decrease in relative volume of the interphase, leading to imperfect interface conditions. The numerical model is further verified by performing a more stringent comparison with the exact local fields within dilute suspensions with imperfect interfaces subject to uniaxial tension derived by Zhong and Meguid [34]. Figure 4b shows the variation of the local tensile stress colinear with the applied tension (\(x_{3}\) direction) along its perpendicular direction (\(x_{1}\) direction) starting from the center of the inclusion. The Poisson’s ratios and shear moduli of matrix and inclusion phases are \(\nu ^{(1)} = 0.35\), \(\nu ^{(2)}=0.2\), and \(\mu ^{(2)} = 20\mu ^{(1)}\). Three different combinations of interfacial compliances are considered. For simplicity, this particular comparison makes use of isotropic interphasial elasticity tensors, \(128^{3}\) voxels such that \(48^{3}\) grid points lie within the inclusion, an inclusion content of \(c \approx 2.2 \times 10^{-3}\), and a thickness-to-radius ratio of \(h/a = 2 \times 10^{-3}\). In all cases, excellent agreement is observed for the entire domain considered, thus confirming the validity of the various parameters employed in the simulations for different loading conditions.

Fig. 4
figure 4

Comparisons between full-field simulations (FFT) and exact results: a) dilute correction for the bulk modulus of a suspension with thin interfacial coating of thickness \(h\), spherical compliance \(\eta_{s} = a/\mu^{(1)}\) and various interfacial anisotropies (\(\alpha = -0.25,0,0.75\)); b) tensile stress distribution in a specimen subject to uniaxial tension for various interfacial compliances [34]

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gallican, V., Zecevic, M., Lebensohn, R.A. et al. The Elastic Properties of Dilute Solid Suspensions with Imperfect Interfacial Bonding: Variational Approximations Versus Full-Field Simulations. J Elast 153, 373–398 (2023). https://doi.org/10.1007/s10659-023-10001-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10659-023-10001-4

Keywords

Mathematics Subject Classification

Navigation