Skip to main content
Log in

A method of two-scale analysis with micro-macro decoupling scheme: application to hyperelastic composite materials

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

Abstract

The aim of this study is to propose a strategy for performing nonlinear two-scale analysis for composite materials with periodic microstructures (unit cells), based on the assumption that a functional form of the macroscopic constitutive equation is available. In order to solve the two-scale boundary value problems (BVP) derived within the framework of the homogenization theory, we employ a class of the micro-macro decoupling scheme, in which a series of numerical material tests (NMTs) is conducted on the unit cell model to obtain the data used for the identification of the material parameters in the assumed constitutive model. For the NMTs with arbitrary patterns of macro-scale loading, we propose an extended system of the governing equations for the micro-scale BVP, which is equipped with the external material points or, in the FEM, control nodes. Taking an anisotropic hyperelastic constitutive model for fiber-reinforced composites as an example of the assumed macroscopic material behavior, we introduce a tensor-based method of parameter identification with the ‘measured’ data in the NMTs. Once the macro-scale material behavior is successfully fitted with the identified parameters, the macro-scale analysis can be performed, and, as may be necessary, the macro-scale deformation history at any point in the macro-structure can be applied to the unit cell to evaluate the actual micro-scale response.

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

Similar content being viewed by others

References

  1. Benssousan A, Lions JL, Papanicoulau G (1978) Asymptotic analysis for periodic structures. North-Holland, Amsterdam

    Google Scholar 

  2. Sanchez-Palencia E (1980) Non-homogeneous media and vibration theory. Springer, Berlin

    MATH  Google Scholar 

  3. Lions JL (1981) Some methods in mathematical analysis of systems and their control. Kexue Chubanshe Science Press and Gordon & Breach Science Pub, Beijing

    MATH  Google Scholar 

  4. Suquet PM (1987) Elements of homogenization theory for inelastic solid mechanics. In: Sanchez-Palencia E, Zaoui A (eds) Homogenization techniques for composite media. Springer, Berlin, pp 193–278

    Chapter  Google Scholar 

  5. Nemat-Nasser S, Hori M (1993) Micromechanics: overall properties of heterogeneous materials. North-Holland, Amsterdam

    MATH  Google Scholar 

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

    Article  MATH  Google Scholar 

  7. Devries F, Dumontet H, Duvaut G, Léné F (1989) Homogenization and damage for composite structures. Int J Numer Methods Eng 27:285–298

    Article  MATH  Google Scholar 

  8. Guedes JM, Kikuchi N (1990) Preprocessing and postprocessing for materials based on the homogenization method with adaptive finite element methods. Comput Methods Appl Mech Eng 83:143–198

    Article  MathSciNet  MATH  Google Scholar 

  9. Swan CC, Cakmak AS (1994) A hardening orthotropic plasticity model for non-frictional composites: rate formulation and integration algorithm. Int J Numer Methods Eng 27:839–860

    Article  Google Scholar 

  10. Ghosh S, Lee K, Moorthy S (1995) Multiple scale analysis of heterogeneous elastic structures using homogenization theory and voronoi cell finite element method. Int J Solids Struct 32:27–62

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  12. van Rens BJE, Brekelmans WAM, Baaijens FPT (1998) Homogenization of the elastoplastic behavior of perforated plates. Comput Struct 69:537–545

    Article  MATH  Google Scholar 

  13. Konke C (1995) Damage evolution in ductile materials: from micro- to macro-damage. Comput Mech 15:497–510

    Article  Google Scholar 

  14. Lee K, Moorthy S, Ghosh S (1999) Multiple scale computational model for damage in composite materials. Comput Methods Appl Mech Eng 172:175–201

    Article  MATH  Google Scholar 

  15. Fish J, Yu O, Shek K (1999) Computational damage mechanics for composite materials based on mathematical homogenization. Int J Numer Methods Eng 45:1657–1679

    Article  MATH  Google Scholar 

  16. Terada K, Kikuchi N (1995) Nonlinear homogenization method for practical applications. In: Ghosh S, Ostoja-Starzewski M (eds) Computational methods in micromechanics, vol AMD-212/MD-62. AMSE, New York, pp 1–16

    Google Scholar 

  17. Smit RJM, Brekelmans WAM, Meijer HEH (1998) Prediction of the mechanical behavior of nonlinear heterogeneous systems by multi-level finite element modeling. Comput Methods Appl Mech Eng 155:181–192

    Article  MATH  Google Scholar 

  18. Wieckowski Z (2000) Dual finite element methods in homogenization for elastic-plastic fibrous composite material. Int J Plast 16:199–221

    Article  MATH  Google Scholar 

  19. Zheng SF, Ding K, Denda M, Weng GJ (2000) A dual homogenization and finite-element study on the in-plane local and global behavior of a nonlinear coated fiber composite. Comput Methods Appl Mech Eng 183:141–155

    Article  MATH  Google Scholar 

  20. Feyel F, Chaboche JL (2000) FE\(^2\) multiscale approach for modelling the elastoviscoplastic behaviour of long fibre SiC/Ti composite materials. Comput Methods Appl Mech Eng 183:309–330

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

    Article  MathSciNet  MATH  Google Scholar 

  22. Kaneko K, Terada K, Kyoya T, Kishino Y (2003) Global-local analysis of granular media in quasi-static equilibrium. Int J Solids Struct 40:4043–4069

    Article  MATH  Google Scholar 

  23. Terada K, Kurumatani M (2010) Two-scale diffusion-deformation coupling model for material deterioration involving micro-crack propagation. Int J Numer Meth Eng 83:426–451

    Google Scholar 

  24. Yamada T (2006) Iterative algorithms for computing the averaged response of nonlinear composites under stress-controlled loadings. Int J Multiscale Comput Eng 4:475–486

    Article  Google Scholar 

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

    Google Scholar 

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

    Google Scholar 

  27. Yvonnet J, He Q-C (2007) The reduced model multiscale method (R3M) for the non-linear homogenization of hyperelastic media at finite strains. J Comput. Physics 223:341–368

    Google Scholar 

  28. Oskay C, Fish J (2007) Eigendeformation-based reduced order homogenization for failure analysis of heterogeneous materials. Comput Methods Appl Mech Eng 196:1216–1243

    Article  MathSciNet  MATH  Google Scholar 

  29. Matsui K, Terada K, Yuge K (2004) Two-scale finite element analysis of heterogeneous solids with periodic microstructures. Comput Struct 82:593–606

    Article  Google Scholar 

  30. Takano N, Zako M, Ohnishi Y (1996) Macro-micro uncoupled homogenization procedure for microscopic nonlinear behavior analysis of composites. Mater Sci Res Int 2:81–86

    Article  Google Scholar 

  31. Temizer İ, Wriggers P (2007) An adaptive method for homogenization in orthotropic nonliner elasticity. Comput Methods Appl Mech Eng 196:3409–3423

    Article  MathSciNet  MATH  Google Scholar 

  32. Watanabe I, Terada K (2005) Decoupled micro-macro analysis method for two-scale BVPs in nonlinear homogenization theory. JSCE J Appl Mech 8:277–285 (in Japanese)

    Google Scholar 

  33. Ren X, Chen JS, Li J, Slawson TR, Roth MJ (2011) Micro-cracks informed damage models for brittle solids. Int J Solids Struct 48:1560–1571

    Article  MATH  Google Scholar 

  34. Terada K, Saiki I, Matsui K, Yamakawa Y (2003) Two-scale kinematics and linearization for simultaneous two-scale analysis of periodic heterogeneous solids at finite strain. Comput Methods Appl Mech Eng 192:3531–3563

    Article  MathSciNet  MATH  Google Scholar 

  35. Ogden RW, Saccomandi G, Sgura I (2004) Fitting hyperelastic models to experimental deta. Comput Mech 34:484–502

    Article  MATH  Google Scholar 

  36. Ogden RW (1984) Non-linear elastic deformations. Dover, New York

    Google Scholar 

  37. ANSYS Inc. (2008) Release 11.0 Documentation for ANSYS

  38. Kaliske M (2000) A formulation of elasticity and viscoelastcity for fibre reinforced material at small and finite strains. Comput Methods Appl Mech Eng 185:225–243

    Article  MATH  Google Scholar 

  39. Kaliske M, Schmidt J, (2004) Nonlinear anisotropic elasticity at finite strains. In: 22nd CAD-FEM Users’ Meeting 2004, International Congress on FEM Technology with ANSYS CFX & ICEM CFD Conference. Int Congress Center Dresden, Germany

  40. Holzapfel GA (2000) Nonlinear solid mechanics: a continuum approach for engineering. Wiley, New York

    Google Scholar 

Download references

Acknowledgments

The authors would like to acknowledge the Nonlinear Homogenization Research Group consisting of Tohoku University, Nitto Boseki Co., Ltd. and Quint Corporation, in particular Dr. Keizo Ishii and Mr. Makoto Tsukino (Quint Corporation ), Mr. Koji Yamamoto, Mr. Tomohiro Ishida, Mr. Akio Miyori, Mr. Fukagawa (Cybernet Systems, Co. Ltd.), and Mr. Naohiro Miyanaga (Nitto Boseki Co., Ltd.) for their contributions and discussions in developing a linear and nonlinear mutli-scale analysis code “Multiscale.Sim”, which is embedded in ANSYS\(^\mathrm{{\textregistered }}\) software.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to J. Kato.

Appendices

Relationship between the macro- and micro-scale traction vectors

In this appendix, we derive relationship (11), which is critical to the numerical material testing with a general-purpose FEM software. For the sake of simplicity, a rectangular parallelepipe-shaped unit cell is considered, with its boundary surfaces assumed to be perpendicular to one of the axes of the micro-scale coordinate system so that only the \(J\)-th component of the side vector \(\varvec{L}^{[J]}\) is a non-zero value \(L_{J}~(J=1, 2, 3)\). In other words, the \(J\)-th basis vector \(\varvec{E}^{[J]}\) coincides with the outward unit normal vector of the side vector \(\varvec{L}^{[J]}\) of the unit cell of \(L_{1} \times L_{2} \times L_{3}\). It is also assumed that the basis vectors \(\varvec{E}^{[1]}, \varvec{E}^{[2]}, \varvec{E}^{[3]}\) are common to the micro- and macro-scale coordinate systems, \(O\text{- }Y_1 Y_2 Y_3\) and \(O\text{- }X_1 X_2 X_3\).

As indicated in Eq. (6), the macroscopic 1st Piola–Kirchhoff (PK) stress \(\tilde{\varvec{P}}\) is the volumetric average of the corresponding microscopic stress \({\varvec{P}}\) over the domain of the initial configuration \(\mathcal{Y _0}\) of a unit cell. This homogenization formula for the stress can be transformed to

$$\begin{aligned} \tilde{\varvec{P}}&=\frac{1}{|\mathcal Y _0|} \int _\mathcal{Y _0}{ \left[ \nabla _Y \cdot \left( {{\varvec{P}} \otimes {\varvec{Y}}} \right) - \left( {\nabla _Y \cdot {\varvec{P}}} \right) \otimes {\varvec{Y}} \right] } dY \nonumber \\&=\frac{1}{|\mathcal Y _0|}\int _\mathcal{Y _0} {\nabla _Y \cdot \left( {{\varvec{P}} \otimes {\varvec{Y}}} \right) } dY \end{aligned}$$
(37)

where we have used the micro-scale self-equilibrium equation (3) along with \( \nabla _Y \cdot \left( {{\varvec{P}} \otimes {\varvec{Y}}} \right) = \left( {\nabla _Y \cdot {\varvec{P}}} \right) \otimes {\varvec{Y}} + {\varvec{P}}. \) The application of the Gauss divergence theorem yields the following relationship:

$$\begin{aligned} \tilde{\varvec{P}} =\frac{1}{|\mathcal Y _0|}\int _{\partial \mathcal Y _0} {\left( \varvec{P} \otimes \varvec{Y} \right) \cdot \varvec{N} } ds =\frac{1}{|\mathcal Y _0|}\int _{\partial \mathcal Y _0} {{\varvec{T}}^{(\varvec{N})} \otimes {\varvec{Y}}} ds \end{aligned}$$
(38)

where \({\varvec{T}}^{(\varvec{N})}:= \varvec{P} \cdot \varvec{N}\) is the microscopic Piola traction vector, wtth \(\varvec{N}\) being an arbitrary outward unit normal vector at the boundary surface \(\partial \mathcal Y _0\). Note here that, due to the assumption on the geometry of the unit cell, the outward unit normal vector \(\varvec{N}\) at the unit cell boundairs \(\partial \mathcal Y _0^{[J]}\) coincides with \(\varvec{E}^{[J]}\).

Since the same basis vectors are used for the micro- and macro-scale coordinate systems, the macro-scale Piola traction vector on the macroscopic surface, which is parallel to the boundary surface \(\partial \mathcal Y _0^{[J]}\) with the outward unit normal vector \(\varvec{E}^{[J]}\) can be can be written as the area average of the corresponding micro-scale Piola traction vector on \(\partial \mathcal Y _0^{[J]}\). For example, the macro-scale Piola traction vector \(\tilde{\varvec{T}}^{({\varvec{E}}^{[1]} )} \) associated with \(\partial \mathcal Y _0^{[1]}\) and \(\varvec{E}^{[1]}\), whose compnents are \(\{1, ~0,~0\}^\mathrm{T}\), can be expressed as follows:

$$\begin{aligned} \tilde{\varvec{T}}^{({\varvec{E}}^{[1]} )}&= \tilde{\varvec{P}}\cdot \varvec{E}^{[1]} \nonumber \\&= \left( {\frac{1}{{\left| \mathcal Y _0 \right| }}\int _{\partial \mathcal Y _0} {{\varvec{T}}^{({\varvec{N}})} \otimes {\varvec{Y}}dS} } \right) \cdot {\varvec{E}}^{[1]} = \frac{1}{{\left| \mathcal Y _0 \right| }}\int _{\partial \mathcal Y _0} {{\varvec{T}}^{({\varvec{N}})} Y_1 dS} \nonumber \\&= \frac{1}{{\left| \mathcal Y _0 \right| }}\left( {\int _{\partial \mathcal Y _0^{[1]} } {{\varvec{T}}^{({\varvec{E}}^{[1]} )} Y_1 dS} + \int _{\partial \mathcal Y _0^{[{ - 1}]} } {{\varvec{T}}^{({\varvec{E}}^{[-1]} )} Y_1 S} } \right. \nonumber \\&\quad \quad + \int _{\partial \mathcal Y _0^{[2]} } {{\varvec{T}}^{({\varvec{E}}^{[2]} )} Y_1 dS} + \int _{\partial \mathcal Y _0^{[- 2]} } {{\varvec{T}}^{({\varvec{E}}^{[ - 2]} )} Y_1 dS} \nonumber \\&\quad \quad \left. + \int _{\partial \mathcal Y _0^{[3]} } {{\varvec{T}}^{({\varvec{E}}^{[3]} )} Y_1 dS} + \int _{\partial \mathcal Y _0^{[ - 3]} } {{\varvec{T}}^{({\varvec{E}}^{[ - 3]} )} Y_1 dS} \right) \end{aligned}$$
(39)

where Eq. (38) has been utilized. Then, due to the anti-periodicity of the Piola traction vector (4), we apply \({\varvec{T}}^{({\varvec{E}}^{[-J]} )}=-{\varvec{T}}^{({\varvec{E}}^{[J]} )}\) along with \(\left. Y_1 \right| _{\partial \mathcal Y _0^{[J]}} = \left. Y_1 \right| _{\partial \mathcal Y _0^{[-J]} } ~ (J \ne 1)\) to (39) to have

$$\begin{aligned} \tilde{\varvec{T}}^{({\varvec{E}}^{[1]} )} = \frac{\left. Y_1 \right| _{\partial \mathcal Y _0^{[1]} } - \left. Y_1 \right| _{\partial \mathcal Y _0^{[-1]} }}{{{\left| \mathcal Y _0 \right| }}} \int _{\partial \mathcal Y _0^{[1]} } {{\varvec{T}}^{({\varvec{E}}^{[1]} )} dS}. \end{aligned}$$
(40)

Using the equivalent expressions \(L_1\!=\!{\left. Y_1 \right| _{\partial \mathcal Y _0^{[1]} } \!-\! \left. Y_1 \right| _{\partial \mathcal Y _0^{[-1]} }}, \left| \mathcal Y _0 \right| \!=\! L_1 L_2 L_3 \) and \(\left| {\partial \mathcal Y _0^{[1]} } \right| \!=\! L_2 L_3 \), we arrive at the following relationship:

$$\begin{aligned} \tilde{\varvec{T}}^{({\varvec{E}}^{[1]} )} =\tilde{\varvec{P}} \cdot \varvec{E}^{[1]} = \frac{1}{|\partial \mathcal Y _0^{[1]}|} \int _{\partial \mathcal Y _0^{[1]}} { \varvec{T}^{({\varvec{E}}^{[1]} )} } ds \end{aligned}$$
(41)

Since the same expression can be obtained for the traction vectors on the other two boundary surfaces, Eq. (11) has been proven.

Anisotropic hyperelastic model with fabric vectors

We consider one class of anisotropic hyperelastic constitutive models, whose functional form is expressed by means of the invariants of the right Cauchy-Green (CG) deformation tensor \(\varvec{C}\) along with the so-called “fabric” vector indicating the direction of reinforcements such as fibers. Although the application for the macroscopic BVP is assumed, we do not distinguish the micro- and macro-scale variables below.

The elastic energy functional of the employed anisotropic hyperelastic model is given as

$$\begin{aligned} W = W_{\mathrm{vol}} \left( {J} \right) +W_{\mathrm{iso}} \left( \bar{\varvec{C}}; {\varvec{A}}, {\varvec{B}} \right) \end{aligned}$$
(42)

where \(\varvec{A}\) and \(\varvec{B}\) are the two distinct directions of fiber alignment in the reference configuration and can be referred to as the fibric vectors. Here, \(W_{\mathrm{vol}}(J)\) is the energy function of the Jacobian \(J:=\det \varvec{F}\) associated with the volumetric deformation. \(W_{\mathrm{iso}} \left( \bar{\varvec{C}}; {\varvec{A}}, {\varvec{B}} \right) \) is the isochoric component of the energy functional by means of the deviatoric part of the right CG deformation tensor, which is defined as

$$\begin{aligned} \bar{\varvec{C}} = \bar{\varvec{F}}^\mathrm{{T}} \bar{\varvec{F}} = J^{ - 2/3} {\varvec{F}}^\mathrm{{T}} {\varvec{F}} = I_3^{ - 1/3} {\varvec{C}} \end{aligned}$$
(43)

with \(\bar{\varvec{F}} = J^{ - 1/3} {\varvec{F}}\) and \(I_3 =\det \varvec{C}= J^2 \).

The 2nd Piola–Kirchhoff (PK) stress can be obtained by differentiating the energy function (42) with respect to the right CG deformation tensor \(\varvec{C}\) as follows:

$$\begin{aligned} {\varvec{S}} = 2\frac{{\partial W}}{{\partial {\varvec{C}}}} = 2\frac{{\partial W_{\mathrm{{vol}}} }}{{\partial {\varvec{C}}}} + 2\frac{{\partial W_{\mathrm{{iso}}} }}{{\partial {\varvec{C}}}} = {\varvec{S}}_{\mathrm{{vol}}} + {\varvec{S}}_{\mathrm{{iso}}} \end{aligned}$$
(44)

Here, we have defined the volumetric and isochoric components of \({\varvec{S}}, {\varvec{S}}_{\mathrm{{vol}}}\) and \({\varvec{S}}_{\mathrm{{iso}}}\), are respectively expressed as

$$\begin{aligned} {\varvec{S}}_{\mathrm{{vol}}}&= 2\frac{{\partial W_{\mathrm{{vol}}} }}{{\partial {\varvec{C}}}} = J\frac{{\partial W_{\mathrm{{vol}}} }}{{\partial J}}{\varvec{C}}^{ - 1}\end{aligned}$$
(45)
$$\begin{aligned} {\varvec{S}}_{\mathrm{{iso}}}&= I_3^{ - 1/3} \mathbb Q :{\varvec{\bar{S}}} \end{aligned}$$
(46)

where

$$\begin{aligned} \mathbb Q&= {\varvec{I}} - \frac{1}{3}{\varvec{C}}^{ - 1} \otimes {\varvec{C}}\end{aligned}$$
(47)
$$\begin{aligned} \bar{\varvec{S}}&= 2\frac{{\partial W_{\mathrm{{iso}}} }}{{\partial \bar{\varvec{C}}}} \nonumber \\&= \bar{\gamma }_1 {\varvec{1}} + \bar{\gamma }_2 \bar{\varvec{C}} + \bar{\gamma }_4 \left( {{\varvec{A}} \otimes {\varvec{A}}} \right) + \bar{\gamma }_5 \left( {{\varvec{A}} \otimes \bar{\varvec{C}} \varvec{A} + \bar{\varvec{CA}} \otimes {\varvec{A}}} \right) \nonumber \\&\quad + \bar{\gamma }_6 \left( {{\varvec{B}} \otimes {\varvec{B}}} \right) + \bar{\gamma }_7 \left( {{\varvec{B}} \otimes \bar{\varvec{CB}} + \bar{\varvec{C}}\bar{\varvec{B}} \otimes {\varvec{B}}} \right) \nonumber \\&\quad + \bar{\gamma }_8 \left( {{\varvec{A}} \cdot {\varvec{B}}} \right) \left( {{\varvec{A}} \otimes {\varvec{B}}} \right) \end{aligned}$$
(48)

along with

$$\begin{aligned} \left. \begin{array}{l} \bar{\gamma }_1 = 2\left( {{\displaystyle \frac{{\partial W_{\mathrm{{iso}}} }}{{\partial \bar{I}_1 }}} + \bar{I}_1 {\displaystyle \frac{{\partial W_{\mathrm{{iso}}} }}{{\partial \bar{I}_2 }}}} \right) , \\ \bar{\gamma }_2 = - 2{\displaystyle \frac{{\partial W_{\mathrm{{iso}}} }}{{\partial \bar{I}_2 }}}, \quad \bar{\gamma }_4 = 2{\displaystyle \frac{{\partial W_{\mathrm{{iso}}} }}{{\partial \bar{I}_4 }}}, \quad \bar{\gamma }_5 = 2{\displaystyle \frac{{\partial W_{\mathrm{{iso}}} }}{{\partial \bar{I}_5 }}}, \\ \bar{\gamma }_6 = 2{\displaystyle \frac{{\partial W_{\mathrm{{iso}}} }}{{\partial \bar{I}_6 }}}, \quad \ \ \ \bar{\gamma }_7 = 2{\displaystyle \frac{{\partial W_{\mathrm{{iso}}} }}{{\partial \bar{I}_7 }}}, \quad \bar{\gamma }_8 = 2{\displaystyle \frac{{\partial W_{\mathrm{{iso}}} }}{{\partial \bar{I}_8 }}} \\ \end{array} \right\} \end{aligned}$$
(49)

Here, \(\mathbf{1}\) and \(\varvec{I}\) are the 2nd-order identity tensor and the 4th-order symmetric identify tensor, respectively. Information about the derivation of these formulae is found in [40].

One of the examples of this class is typified by Kaliske et al. [38, 39], which seems to be reasonable within the present framework of two-scale coupling analysis with the micro-macro decoupling scheme. We employ their model in this study and provide its concrete functional form below. The volumetric and isochoric parts of the energy functional in [39] by are respectively given as

$$\begin{aligned} W_{\mathrm{{vol}}}&= \frac{1}{D}\left( {J - 1} \right) ^2 \end{aligned}$$
(50)
$$\begin{aligned} W_{\mathrm{iso}}&= W_{\mathrm{{iso}}} (\bar{I}_1 ,\bar{I}_2 ,\bar{I}_4 ,\bar{I}_5 ,\bar{I}_6 ,\bar{I}_7 ,\bar{I}_8; a_i, b_j, c_k ,d_l, e_m, f_n, g_o; \varvec{A},~\varvec{B}) \nonumber \\&= \sum \limits _{i = 1}^3 {a_i \left( {\bar{I}_1 - 3} \right) ^i } + \sum \limits _{j = 1}^3 {b_j \left( {\bar{I}_2 - 3} \right) ^j } \nonumber \\&\quad + \sum \limits _{k = 2}^6 {c_k \left( {\bar{I}_4 - 1} \right) ^k } + \sum \limits _{l = 2}^6 {d_l \left( {\bar{I}_5 - 1} \right) ^l } + \sum \limits _{m = 2}^6 {e_m \left( {\bar{I}_6 - 1} \right) ^m } \nonumber \\&\quad + \sum \limits _{n = 2}^6 {f_n \left( {\bar{I}_7 - 1} \right) ^n } + \sum \limits _{o = 2}^6 {g_o \left( {\bar{I}_8 - \varsigma } \right) ^o } \end{aligned}$$
(51)

Here, \(D, a_i, b_j, c_k, d_l, e_m, f_n, g_o\) are scalar-valued material parameters, and \(\bar{I}_1, \bar{I}_2, \bar{I}_4, \bar{I}_5, \bar{I}_6, \bar{I}_7, \bar{I}_8\) are the invariant of \(\bar{\varvec{C}}\) defined as follows:

$$\begin{aligned} \left. \begin{array}{ll} \bar{I}_1 = \mathrm{tr }\bar{\varvec{C}}, &{} \bar{I}_2 = {\displaystyle \frac{1}{2}} \left( {\mathrm{{tr}}^2 \bar{\varvec{C}} - \mathrm{tr }\bar{\varvec{C}}^2 } \right) , \\ \bar{I}_4 = {\varvec{A}} \cdot \bar{\varvec{C}}\varvec{A}, &{} \bar{I}_5 = {\varvec{A}} \cdot \bar{\varvec{C}}^2 {\varvec{A}}, \quad \bar{I}_6 = {\varvec{B}} \cdot \bar{\varvec{C}}\varvec{B}, \\ \bar{I}_7 = {\varvec{B}} \cdot \bar{\varvec{C}}^2 {\varvec{B}}, &{} \bar{I}_8 = \left( {{\varvec{A}} \cdot {\varvec{B}}} \right) {\varvec{A}} \cdot \bar{\varvec{C}}\varvec{B} \\ \end{array} \right\} \end{aligned}$$
(52)

where \(\varsigma = \left( {{\varvec{A}} \cdot {\varvec{B}}} \right) ^2\).

This model is capable of representing a certain class of orthotropic hyperelastic behavior with the fibric vectors \(\varvec{A}\) and \(\varvec{B}\) as input data. If, for example, we assume \(\varvec{B}=\mathbf{0}\), then the resulting energy functional can be used for a class of transversely isotropic materials introduced in [38].

Rights and permissions

Reprints and permissions

About this article

Cite this article

Terada, K., Kato, J., Hirayama, N. et al. A method of two-scale analysis with micro-macro decoupling scheme: application to hyperelastic composite materials. Comput Mech 52, 1199–1219 (2013). https://doi.org/10.1007/s00466-013-0872-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00466-013-0872-5

Keywords

Navigation