17
International Journal of Solids and Structures 118–119 (2017) 24–40 Contents lists available at ScienceDirect International Journal of Solids and Structures journal homepage: www.elsevier.com/locate/ijsolstr 3D hierarchical multiscale analysis of heterogeneous SMA based materials Peyman Fatemi Dehaghani, Saeed Hatefi Ardakani, Hamid Bayesteh, Soheil Mohammadi High Performance Computing Laboratory, School of Civil Engineering, University of Tehran, Tehran, Iran a r t i c l e i n f o Article history: Received 14 December 2016 Revised 12 March 2017 Available online 19 April 2017 Keywords: Multiscale analysis Homogenization Shape memory alloy SMA composite Porous SMA a b s t r a c t A 3D hierarchical multiscale computational homogenization framework with periodic Representative Vol- ume Element (RVE) is proposed in the present study for analysis of shape memory alloy (SMA) fiber reinforced composite and porous SMA specimen. The shape memory alloy, which exhibits forward and reverse transformations, is wired through a matrix, while a dense SMA matrix encircles voids as inclu- sions in the porous model. The asymptotic mathematical homogenization approach is used to express the microscopic and macroscopic governing equations. An iterative solution is used at each point in or- der to derive the tangential constitutive response of the corresponding periodic RVE. Several numerical problems are solved and the results are compared with those obtained from a fully meshed direct numer- ical model. Detailed discussions are presented regarding the distribution of field variables, in particular the martensitic volume fraction and the stress state inside RVEs located at critical points of the macro models, subjected to cyclic stress loads in the form of tension, flexure and torsion. Moreover, a microme- chanical analysis is conducted to investigate the effect of the shape of ellipsoid porosities on the overall response of porous SMA. © 2017 Elsevier Ltd. All rights reserved. 1. Introduction Shape memory alloys (SMA) are one of the most appealing smart materials, with growing engineering applications in struc- tures, biomedical industry and aerospace specimen (Lecce and Concilio, 2015). SMAs feature two very distinguished capabilities, namely the pseudo-elastic response and the shape memory effect, that allow engineers to design devices capable of withstanding large strains and internal forces when subjected to cyclic stress and temperature loads. Once SMA is shaped into a wire, tube or any other forms of inclusion and embedded into other high strength and durable matrices such as metallic or reinforced materials, it is possible to make high performance composite devices that can provide active control of the response of the whole structure. On the other hand, gains from the pseudo-elastic response and the shape memory effect can be combined with those from porous structures such as high surface area and permeability and low den- sity to create porous SMA composites with inclusions in the form of porosities of different sizes, shapes and arrangements. Porous SMAs that are generally made of sintered NiTi powder enjoy rela- tively good corrosion resistance and biocompatibility and thus are Corresponding author. E-mail address: [email protected] (S. Mohammadi). widely used in biomedical applications such as human body im- plants (Lecce and Concilio, 2015). In recent years, advances in SMA-integrated smart material manufacturing technology have been accompanied by efforts to de- velop appropriate simulation techniques to study these heteroge- neous materials. The well-established continuum approaches ex- press homogeneous material response through adoption of appro- priate constitutive laws. As a result, they are incapable of repre- senting complicated response of such materials and providing in- sight on microscopic behavior of constituents and their interac- tion. To address these complex issues, modern multiscale simula- tion methods have been developed. Generally, multiscale methods are divided into two main cate- gories of sequential and concurrent methods (Tadmor and Miller, 2011). In a sequential method, the two scales are solved inde- pendently and the results from micro model are used as the in- put variables for macro simulation, without any information being passed during the analysis (Eftekhari et al., 2014). The main ad- vantage of this method is the relatively low computational cost. In a concurrent method, appropriate scale transition techniques are invoked to simultaneously solve micro and macro scale problems. The concurrent methods are also subdivided into the hierarchical and partitioned-domain methods (Tadmor and Miller, 2011). The hierarchical method, which is alternatively referred to as “com- putational homogenization” or “multiscale computational homog- http://dx.doi.org/10.1016/j.ijsolstr.2017.04.025 0020-7683/© 2017 Elsevier Ltd. All rights reserved.

International Journal of Solids and Structures hierarchical... · 2019. 10. 12. · 26 P. Fatemi Dehaghani et al. / International Journal of Solids and Structures 118–119 (2017)

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

  • International Journal of Solids and Structures 118–119 (2017) 24–40

    Contents lists available at ScienceDirect

    International Journal of Solids and Structures

    journal homepage: www.elsevier.com/locate/ijsolstr

    3D hierarchical multiscale analysis of heterogeneous SMA based

    materials

    Peyman Fatemi Dehaghani, Saeed Hatefi Ardakani, Hamid Bayesteh, Soheil Mohammadi ∗

    High Performance Computing Laboratory, School of Civil Engineering, University of Tehran, Tehran, Iran

    a r t i c l e i n f o

    Article history:

    Received 14 December 2016

    Revised 12 March 2017

    Available online 19 April 2017

    Keywords:

    Multiscale analysis

    Homogenization

    Shape memory alloy

    SMA composite

    Porous SMA

    a b s t r a c t

    A 3D hierarchical multiscale computational homogenization framework with periodic Representative Vol-

    ume Element (RVE) is proposed in the present study for analysis of shape memory alloy (SMA) fiber

    reinforced composite and porous SMA specimen. The shape memory alloy, which exhibits forward and

    reverse transformations, is wired through a matrix, while a dense SMA matrix encircles voids as inclu-

    sions in the porous model. The asymptotic mathematical homogenization approach is used to express

    the microscopic and macroscopic governing equations. An iterative solution is used at each point in or-

    der to derive the tangential constitutive response of the corresponding periodic RVE. Several numerical

    problems are solved and the results are compared with those obtained from a fully meshed direct numer-

    ical model. Detailed discussions are presented regarding the distribution of field variables, in particular

    the martensitic volume fraction and the stress state inside RVEs located at critical points of the macro

    models, subjected to cyclic stress loads in the form of tension, flexure and torsion. Moreover, a microme-

    chanical analysis is conducted to investigate the effect of the shape of ellipsoid porosities on the overall

    response of porous SMA.

    © 2017 Elsevier Ltd. All rights reserved.

    1. Introduction

    Shape memory alloys (SMA) are one of the most appealing

    smart materials, with growing engineering applications in struc-

    tures, biomedical industry and aerospace specimen ( Lecce and

    Concilio, 2015 ). SMAs feature two very distinguished capabilities,

    namely the pseudo-elastic response and the shape memory effect,

    that allow engineers to design devices capable of withstanding

    large strains and internal forces when subjected to cyclic stress and

    temperature loads. Once SMA is shaped into a wire, tube or any

    other forms of inclusion and embedded into other high strength

    and durable matrices such as metallic or reinforced materials, it

    is possible to make high performance composite devices that can

    provide active control of the response of the whole structure. On

    the other hand, gains from the pseudo-elastic response and the

    shape memory effect can be combined with those from porous

    structures such as high surface area and permeability and low den-

    sity to create porous SMA composites with inclusions in the form

    of porosities of different sizes, shapes and arrangements. Porous

    SMAs that are generally made of sintered NiTi powder enjoy rela-

    tively good corrosion resistance and biocompatibility and thus are

    ∗ Corresponding author. E-mail address: [email protected] (S. Mohammadi).

    widely used in biomedical applications such as human body im-

    plants ( Lecce and Concilio, 2015 ).

    In recent years, advances in SMA-integrated smart material

    manufacturing technology have been accompanied by effort s to de-

    velop appropriate simulation techniques to study these heteroge-

    neous materials. The well-established continuum approaches ex-

    press homogeneous material response through adoption of appro-

    priate constitutive laws. As a result, they are incapable of repre-

    senting complicated response of such materials and providing in-

    sight on microscopic behavior of constituents and their interac-

    tion. To address these complex issues, modern multiscale simula-

    tion methods have been developed.

    Generally, multiscale methods are divided into two main cate-

    gories of sequential and concurrent methods ( Tadmor and Miller,

    2011 ). In a sequential method, the two scales are solved inde-

    pendently and the results from micro model are used as the in-

    put variables for macro simulation, without any information being

    passed during the analysis ( Eftekhari et al., 2014 ). The main ad-

    vantage of this method is the relatively low computational cost. In

    a concurrent method, appropriate scale transition techniques are

    invoked to simultaneously solve micro and macro scale problems.

    The concurrent methods are also subdivided into the hierarchical

    and partitioned-domain methods ( Tadmor and Miller, 2011 ). The

    hierarchical method, which is alternatively referred to as “com-

    putational homogenization” or “multiscale computational homog-

    http://dx.doi.org/10.1016/j.ijsolstr.2017.04.025

    0020-7683/© 2017 Elsevier Ltd. All rights reserved.

    http://dx.doi.org/10.1016/j.ijsolstr.2017.04.025http://www.ScienceDirect.comhttp://www.elsevier.com/locate/ijsolstrhttp://crossmark.crossref.org/dialog/?doi=10.1016/j.ijsolstr.2017.04.025&domain=pdfmailto:[email protected]://dx.doi.org/10.1016/j.ijsolstr.2017.04.025

  • P. Fatemi Dehaghani et al. / International Journal of Solids and Structures 118–119 (2017) 24–40 25

    Fig. 1. A macroscopic heterogeneous material and definition of RVE in the mathe-

    matical homogenization theory.

    (a)

    (b)

    Fig. 2. Global periodicity (a) versus local periodicity (b).

    enization”, provides the response of the micro constituents while

    no assumption on the macroscopic constitutive law is needed. This

    emphasizes the preference of this method for modeling of compos-

    ite and heterogeneous materials and designing their micro struc-

    ture. Classical homogenization methods such as self-consistent

    ( Mori and Tanaka, 1973 ) and transformation field analysis ( Dvorak,

    1992 ) make some simplifying assumptions in order to accelerate

    the computation. In contrast to the classical methods, recently de-

    veloped computational methods have fewer restrictions and are

    applicable to diverse problems. Among different methods, the

    first and second order computational homogenization techniques

    Fig. 3. A homogeneous macro material �, its heterogeneous microstructure �ζ and

    a typical RVE �Y .

    Fig. 4. Schematic procedure of hierarchical multiscale homogenization.

    ( Geers et al., 2010; Kouznetsova, 2002 ), multiresolution methods

    ( McVeigh et al., 2006; Tian et al., 2010 ), the computational con-

    tinua ( Fish and Kuznetsov, 2010 ), enrichment based methods ( Fish

    and Yuan, 2007 ), the eigenstrain method ( Oskay and Fish, 2007 )

    and the asymptotic mathematical homogenization approach ( Fish

    and Shek, 1999; Fish et al., 1997; Hassani and Hinton, 1998b ) can

    be mentioned.

    Multiscale and homogenization methods have been developed

    and applied to problems involving SMA composites and porous

    SMAs. The self-consistent method was used by Lagoudas et al.

    (1994), Cherkaoui et al. (20 0 0) and Marfia (2005) , among oth-

    ers, in order to model the micromechanical response of compos-

    ites with SMA constituents. Zhao and Taya (2007) and Zhu and

    Dui (2011) adopted the same methodology for modeling the re-

    sponse of porous SMAs. Many researchers conducted finite ele-

    ment based micro and micro-macro mechanical modeling of SMA

  • 26 P. Fatemi Dehaghani et al. / International Journal of Solids and Structures 118–119 (2017) 24–40

    Fig. 5. Algorithm of nonlinear multiscale homogenization.

    Fig. 6. Super-elastic response of the SMA fiber and the elastic response of the ma-

    trix under uniaxial tensile loading/unloading.

    fiber reinforced composites and porous SMAs. Marfia (2005) de-

    veloped micro-macro approaches for analysis of SMA fiber rein-

    forced composites based on the periodic homogenization method

    to analyze 2D examples and then extended his work to shell struc-

    tures made of SMA composite laminates ( Marfia and Sacco, 2007 ).

    Chatzigeorgiou et al. (2015) proposed a hierarchical multiscale

    framework for analysis of periodic composite laminates with SMA

    inclusions under quasi-static thermomechanical conditions and

    concluded that these laminates might show a non-convex trans-

    formation surface. Thermo-micromechanical analysis of SMA com-

    posites under multi-axial proportional/non-proportional loadings

    for generalized plane strain state was presented by Damanpack et

    al. (2015a, 2015b, 2015c ) and micromechanical modeling of SMA

    fiber reinforced composites and porous SMAs using non-uniform

    transformation field analysis (NUTFA) was performed by Sepe et

    al. (2013, 2016) . Moreover, the effects of size and shape of porosity

    on the response of porous SMAs were investigated by Sepe et al.

    (2014, 2015) through the micromechanical analysis of RVEs.

    To the best knowledge of the authors, available 3D multiscale

    modeling frameworks are restricted to the generalized plane strain

    case, which is not generally applicable to porous SMA problems

    due to non-prismatic microstructure. Also, most of the few avail-

    able works in this context are restricted to micromechanics of

    SMA based materials. Accordingly, in the present work, a 3D fi-

    nite element based hierarchical multiscale analysis framework is

    presented to solve the macro and micro scale problems simultane-

    ously. The proposed approach is applied to a number of problems

    involving SMA fiber reinforced composites and porous SMAs to as-

    sess its accuracy and effectiveness. The pseudo-elastic response of

    SMA constituents in micro scale is modeled based on the work by

    Boyd and Lagoudas (1996) and Qidwai and Lagoudas (20 0 0) . Macro

    problems are solved under various stress loading/unloading condi-

    tions including tension, flexure, and torsion. Detailed illustrations

    are presented regarding the distribution and concentration of dif-

    ferent field variables such as the martensitic volume fraction and

    the stress state in the micro scale along the austenite to marten-

    site transformation and the reverse path. Due to lack of plane sym-

    metry, 2D general plane strain models cannot accurately predict

    the microscopic behavior of porous materials. Hence, the effect

    of shape of 3D ellipsoid pores on the mechanical response of the

    porous SMA is investigated through a detailed parametric study.

    Organization of the paper is as follows: Section 2 presents the

    constitutive model adopted for the SMA inclusion and matrix and

    the asymptotic mathematical homogenization formulation is dis-

    cussed in Section 3 . Finally, hierarchical multiscale analyses of 3D

    SMA-fiber reinforced composite beams and porous SMA specimens

    subjected to various stress load cycles are presented and various

    results are discussed in detail.

  • P. Fatemi Dehaghani et al. / International Journal of Solids and Structures 118–119 (2017) 24–40 27

    Fig. 7. Geometry of the model (a) homogenization analysis and (b) direct finite element analysis.

    Fig. 8. Comparison of the stress-strain responses based on the homogenization and

    direct solutions.

    2. SMA constitutive model

    A number of constitutive models are available for SMAs, which

    include the phenomenological models ( Auricchio et al., 1997;

    Bekker and Brinson, 1997; Govindjee and Kasper, 1999; Lubliner

    and Auricchio, 1996 ) and the continuum formulations based on

    the Helmholtz free energy ( Abeyaratne and Knowles, 1993; Anand

    and Gurtin, 2003; Helm and Haupt, 2003; Paiva et al., 2005; Reese

    and Christ, 2008 ) or the Gibbs free energy ( Ahmadian et al., 2015;

    Bo and Lagoudas, 1999; Boyd and Lagoudas, 1996; Lagoudas and

    Entchev, 2004; Popov and Lagoudas, 2007; Raniecki and Lexcel-

    lent, 1998 ). In this study, the constitutive model for poly-crystalline

    shape memory alloys, proposed by Boyd and Lagoudas (1996) , has

    been adopted within the framework of the infinitesimal strains.

    The martensitic volume fraction ξ and the phase transformation induced strain ε t

    i j are chosen as the internal state variables since

    they can best characterize the phase transformation with a single

    variant martensitic state. Additive decomposition of the total in-

    finitesimal strain εij can be written as,

    ε i j = ε th −e i j + ε t i j (1)

    where ε th −e i j

    is the thermo-elastic strain part and ε t i j

    is the trans-

    formation part. Using the Cauchy stress tensor σ ij and the current temperature T as the independent state variables, the Gibbs free

  • 28 P. Fatemi Dehaghani et al. / International Journal of Solids and Structures 118–119 (2017) 24–40

    Fig. 9. Distribution of the martensitic volume fraction in the RVE at various load levels.

    Fig. 10. Comparison of the martensitic volume fraction in direct (left) and homogenization (right) solutions at two load levels (a) load level B and (b) load level D.

  • P. Fatemi Dehaghani et al. / International Journal of Solids and Structures 118–119 (2017) 24–40 29

    Fig. 11. Model of the flexural beam.

    Fig. 12. Load-displacement diagram from the homogenization and two direct anal-

    yses.

    energy function G s is defined in the explicit form by

    G s ( σi j , T , ε t i j , ξ ) = −

    1

    2 ρS i jkl σi j σkl −

    1

    ρσi j

    (αi j (T − T 0 ) + ε t i j

    )+ c

    (( T − T 0 ) − T ln

    (T

    T 0

    ))−s 0 T + u 0 + 1

    ρη( ξ ) (2)

    where ρ , T 0 , S ijkl , αij , c, s 0 and u 0 are the phase independent mate- rial density, the reference temperature, the effective material com-

    pliance tensor, the effective thermal expansion tensor, the effective

    specific heat, the effective specific entropy at the reference state

    and the effective specific internal energy at the reference state,

    respectively. f ( ξ ) is the hardening function ( Boyd and Lagoudas, 1996 ), defined as a quadratic form by

    f ( ξ ) = {

    1 2 ρb M ξ

    2 + ( μ1 + μ2 ) ξ ; ˙ ξ > 0 1 2 ρb A ξ

    2 + ( μ1 − μ2 ) ξ ; ˙ ξ < 0 (3)

    where b A , b M , μ1 and μ2 are material parameters to be determined later. The effective material parameters are assumed to be constant

    at fully austenitic and fully martensitic states and change linearly

    with respect to the martensitic volume fraction during the trans-

    formation

    ( . ) (ξ ) = ( . ) A + ξ(( . )

    M − ( . ) A )

    (4)

    where (.) A and (.) M are the effective material parameters at fully

    austenitic and fully martensitic states, respectively.

    Under the assumptions of:

    • The change in the transformation strain is a function of the change in the martensitic volume fraction.

    • Material training and fatigue effects are neglected.

    evolution of the internal state variables can be defined by:

    ˙ ε t i j = i j ˙ ξ (5) where the transformation tensor ij is given by

    i j = {

    3 2 H

    σ ′ i j σ̄ ′ ; ˙ ξ > 0

    H ε t−re v erse i j

    ε̄ t−re v erse ; ˙ ξ < 0 (6)

    where H is an experimentally determined material parameter

    linked to the maximum transformation strain, σ ′ ij is the devia- toric stress tensor and ε t−re v erse

    i j is the transformation strain at the

  • 30 P. Fatemi Dehaghani et al. / International Journal of Solids and Structures 118–119 (2017) 24–40

    Fig. 13. Distribution of the martensitic volume fraction (left) and von-Misses stress (right) for two RVEs (a) above and (b) under the neutral axis.

    Fig. 14. Stress-strain diagram of the dense SMA specimen.

    turning point of the transformation. The following norms are also

    adopted

    σ̄ ′ = √

    3

    2 σ ′ i j σ ′ i j (7)

    ε̄ t−re v erse = √

    2

    3 ε t−re v erse i j

    ε t−re v erse i j

    (8)

    The Clausius–Planck inequality can now be stated as (−ρ ∂ G s

    ∂ε t i j

    )︸ ︷︷ ︸

    σi j

    ˙ ε t i j + (

    −ρ ∂ G s ∂ξ

    )˙ ξ ≥ 0

    ⇒ σi j ˙ ε t i j ︸︷︷︸ = i j ̇ ξ

    + (

    −ρ ∂ G s ∂ξ

    )˙ ξ ≥ 0

    ⇒ (

    σi j i j − ρ∂ G s ∂ξ

    )˙ ξ = π ˙ ξ ≥ 0 (9)

    where π is regarded as the thermodynamic force and can be writ- ten as

    π = σi j i j + 1

    2

    S i jkl σi j σkl + σi j αi j (T − T 0 )

    −ρc [ (T − T 0 ) − T ln

    (T

    T 0

    )] + ρs 0 T − ρu 0 − ∂ f (ξ )

    ∂ξ(10)

  • P. Fatemi Dehaghani et al. / International Journal of Solids and Structures 118–119 (2017) 24–40 31

    Fig. 15. Finite element mesh of RVE with a spherical void ( f = 13%).

    Definition of the constitutive model is completed by setting

    a material parameter Y as the threshold value for π at which the transformation initiates. The transformation function � can be

    stated accordingly as

    � = {

    π − Y ; ˙ ξ > 0 −π − Y ; ˙ ξ < 0

    (11)

    and the transformation surface is characterized by �= 0. Finally, the material parameters μ1 , μ2 , b A , b M and Y are given by

    μ1 = 1 2 ρs 0

    (T M s + T A f

    )− ρu 0 (12)

    μ2 = 1 4 ρs 0

    (T A s − T A f − T M f + T M s

    )(13)

    b A = −s 0 (T A f − T A s

    )(14)

    b M = −s 0 (T M s − T M f

    )(15)

    Y = 1 4 ρs 0

    (T M s + T M f − T A f − T A s

    )(16)

    where T is the temperature and subscripts s and f refer to start and

    finish points, respectively, and superscripts A and M represent the

    austenite and martensite sates, respectively.

    3. 3D periodic hierarchical multiscale homogenization method

    The conventional theory of continuum mechanics has widely

    been used for analysis of homogeneous materials. However, the

    nature of all materials is not necessarily homogeneous and may

    be heterogeneous in micro and meso scales and discrete in nano

    scale. The key concept in replacing microscopically heterogeneous

    materials with macroscopically homogeneous materials is to facili-

    tate employment of mathematical expression of constitutive laws

    that consider different phenomena in micro level using internal

    variables. Despite several advantages of conventional homogeneous

    Fig. 16. Stress-strain diagram of the porous SMA specimen.

    continuum formulation, they suffer severe shortcomings in a num-

    ber of new applications, particularly in design of new advanced

    materials in the forms of composites and porous materials. One of

    the most important drawbacks is that no additional information is

    provided for micro and/or meso scales by the conventional theo-

    ries, while it is necessary for material design purposes. Advanced

    numerical methods of “multiscale”, “computational homogeniza-

    tion” or “multiscale computational homogenization” are designed

    to capture more sophisticated material response of the macro-

    scopic homogeneous point and to provide additional information

    about structural and material response in the micro level.

    The homogenization method has been presented from differ-

    ent points of view in the literature. In the present study, the

    asymptotic mathematical homogenization perspective is adopted

    ( Hassani and Hinton, 1998a,b; Yuan and Fish, 2008 ). The basic idea

    is to express the displacement field of a heterogeneous medium,

    u ζ( i, j )

    , with the double scale asymptotic expansion

    u ζi ( x , y ) = u 0 i ( x , y ) + ζu 1 i ( x , y ) + ζ 2 u 2 i ( x , y ) + ... (17)

    where the superscript ζ denotes the dependency of the field vari- able on the existing inhomogeneity of the RVE, and x = ( x 1 , x 2 , x 3 ) and y = ( y 1 , y 2 , y 3 ) are the spatial variables in the macro material and the RVE domain is y = x /ζ , as depicted in Fig. 1 . Eq. (17) as- sumes that u �

    i ( x , y ) ( � = 0, 1, 2, ...) are Y -periodic where Y =

    〈 Y 1 Y 2 Y 3 〉 T ( Hassani and Hinton, 1998b ). While the periodicity is a basic assumption in conventional mathematical homogeniza-

    tion, the local periodicity is also acceptable for many applications.

    In comparison with the global periodicity, which is the repetition

    of RVE in the whole domain, local periodicity assumes periodicity

    of RVEs only in the vicinity of a macroscopic point, as depicted in

    Fig. 2 . These assumptions are held particularly in man-made com-

    posites and materials.

    Three subsets of R 3 denoted by �, �ζ and �Y , are distinguished

    for the macro homogeneous material, the heterogeneous domain

    and the RVE, respectively, as depicted in Fig. 3 . Furthermore, the

    indicial notations �i, x j and �i, y j are used for derivatives with re-

    spect to x i and y i , ∂ �i ∂ x j

    and ∂ �i ∂ y j

    , respectively.

    It should be noted that, based on the present formulation, a

    typical RVE could possess both inclusion and porosity with differ-

    ent arrangements for various problems, as schematically shown in

    Fig. 3 . RVEs used for the numerical examples of this study are all

    special cases of the general RVE depicted in Fig. 3 and will be dis-

  • 32 P. Fatemi Dehaghani et al. / International Journal of Solids and Structures 118–119 (2017) 24–40

    (a)

    (b)

    Fig. 17. (a) Approximating porous SMA stress-strain pseudo-elastic response with two trilinear curves and (b) distribution of the martensitic volume fraction in any arbitrarily

    chosen RVE at various load levels.

    cussed in Section 4 . While the present formulation is applicable

    to random heterogeneous materials, the main focus of the current

    study is devoted to the man-made composites with periodic RVEs.

    The stress–strain and compatibility relations are expressed by

    ε ζi j

    = u ζ( i, x j )

    (18)

    σ ζi j

    = f (ε ζi j

    )= f

    (u 0 ( i, x j )

    ( x , y ) + ζu 1 ( i, x j )

    ( x , y ) + ... )

    (19)

    where �ζ states that � is defined in �ζ . Moreover, u ζ( i, x j )

    is the

    symmetric part of the deformation gradient u ζi, x j

    . The weak form

    of the equation can be written as,

    ∫ �ζ

    σ ζi j w i, j d� =

    ∫ �ζ

    w i f ζi d� +

    ∫ �t

    w i t i d�

    + ∫ s ζw i p

    ζi ds ∀ w i ∈ V ζ (20)

    where the prescribed traction t i is defined on �t , and S ζ is the

    boundary of pores in RVEs and p ζi is the traction on pores. Addi-

    tionally, w i is the weight function of the Galerkin method and V ζ

    is the space of all admissible functions w i .

    Using the double-scale asymptotic expansion ( 17 ) and the

    chain rule, the governing equations in macro and micro scales

    for the first order multiscale homogenization can be obtained, as

    described by Yuan and Fish (2008) . The microscopic governing

  • P. Fatemi Dehaghani et al. / International Journal of Solids and Structures 118–119 (2017) 24–40 33

    Fig. 18. Finite element meshes for RVEs with different shape factors n .

    equation can be written as, ∫ �Y

    σ ζi j ( x , y ) w i, y j ( y ) d� −

    ∫ S

    p i ( y ) w i ( y ) dS = 0 ∀ w i ∈ V �Y (21) where σ ζ

    i j ( x , y ) is a function of u 0

    ( i, x j ) (x ) and u 1

    ( i, x j ) ( x , y ) from Eq.

    (19) . It is proved that u 0 k, x l

    is constant on RVE ( Yuan and Fish, 2008 )

    and is just a function of x and not y . V �Y is the space of all admis-

    sible periodic functions of w i ( y ) in RVE

    V �Y = {w i (y ) in �

    Y | w i ( y ) is Y − periodic }

    (22)

    Then, the macro scale governing equation can be stated as, ∫ �

    σ 0 i j ( x ) w i, x j ( x ) d� −∫ �

    f̄ i w i (x ) d�

    −∫ �t

    t i w i ( x ) d�= 0 ∀ w i ∈ V � (23) where

    σ 0 i j ( x ) = 1

    | �Y | ∫ �Y

    σ ζi j ( x , y ) d� (24)

    f̄ i = 1

    | �Y | ∫ �Y

    f i ( y ) d� (25)

    and the space V � in the homogenized macro scale is defined as

    V � = {w i (x ) in �

    ∣∣w i smooth enoughand w ∣∣�d = 0 } (26) σ 0 i j (x ) is calculated by solving the micro scale Eq. (21) and em-

    ploying Eq. (24) . Functions u 0 i (x ) and u 1

    i (y ) in macro and micro

    scale equations are unknown and should be solved simultaneously

    and reciprocally. The final solutions of u 0 i (x ) and u 1

    i (y ) are ob-

    tained when both Eqs. (21) and ( 23 ) are simultaneously satisfied.

    First, Eq. (21) is solved for u 1 i (y ) using the known values of u 0

    k, x l (x )

    for the RVE located at the macroscopic point x , and is followed by

    calculation of the stress σ ζi j ( x , y ) at each Gauss point of RVE. Then,

    the macroscopic stress σ 0 i j (x ) is calculated from Eq. (24) and the

    residual of Eq. (23) is obtained accordingly. u 0 k, x l

    (x ) is then cor-

    rected, and Eq. (21) is solved again. This procedure continues un-

  • 34 P. Fatemi Dehaghani et al. / International Journal of Solids and Structures 118–119 (2017) 24–40

    Fig. 19. Influence of the shape of porosity on the macroscopic behavior of porous

    SMA specimen.

    til the residuals of both Eqs. (21) and ( 23 ) become infinitesimally

    small. The iterative solution scheme is illustrated in Fig. 4 .

    In the linear case where σ ζi j

    = c i jkl ε ζkl , u 1 i (y ) in Eq. (21) can be written as

    u 1 i (y ) = χ pq i ( y ) ε 0 pq + ψ i ( y ) (27) such that

    ε 0 pq = u 0 (p, x q ) (28)

    and ψ i ( y ) and χpq i

    (y ) are obtained from Eqs. (29) and ( 30 ), respec-

    tively ( Hassani and Hinton, 1998b ). ∫ �Y

    c i jkl ψ k, y l ( y ) w i, x j ( y ) d� = −∫ S

    p i w i ( y ) dS ∀ w i ∈ V �Y (29)

    ∫ �Y

    c i jkl χkl k, y l

    ( y ) w i, y j ( y ) d� = −∫ �Y

    c i jkl w i, y j ( y ) d� ∀ w i ∈ V �Y (30)

    Substituting Eq. (27) into Eq. (23) leads to the material consti-

    tutive matrix as

    c̄ i jkl = 1

    | �Y | ∫ �Y

    (c i jkl + c i jpq χ kl p, y q ( y )

    )d� (31)

    and the macro scale governing equation can then be written as ∫ �c̄ i jkl u

    0 k, x l

    (x ) w i, x j (x ) d� = ∫ �p̄ i j w i, x j ( x ) d� +

    ∫ �

    f̄ i w i ( x ) d�

    + ∫ �t

    t i w i ( x ) d�∀ w i ∈ V � (32)

    where

    p̄ i j = 1

    | �Y | ∫ �Y

    c i jkl ψ k, y l d� (33)

    and f̄ was defined in Eq. (25) . The tangential constitutive tensor

    c̄ i jkl from Eqs. (29) to ( 31 ) is used to calculate the initial stiffness

    matrix in the beginning of each macroscopic iteration. c ijkl ( y ) in

    Eqs. (29) to ( 31 ) is the tangential tensor at each Gauss point of

    RVE. The iterative solution algorithm is summarized in Fig. 5 .

    Table 1

    Material properties of SMA.

    Material parameters Values Material parameters Values

    E A (GPa) 70 T M f

    (K) 275

    E M (GPa) 30 T M s (K) 290

    υA 0.33 T A s (K) 294

    υM 0.33 T A f (K) 309

    H 0.05 C A (MPa K −1 ) 10

    T 0 (K) 310 C M (MPaK -1 ) 10

    Table 2

    Material properties of matrix.

    Material parameters Values

    E (GPa) 20

    υ 0.3

    4. Numerical results

    In this section, the developed homogenization method is

    adopted to analyze the macroscopic behavior of 3D SMA-fiber re-

    inforced composite beams and porous SMA specimens subjected

    to various loads. The macroscopic behavior of a 3D SMA-fiber re-

    inforced composite beam under the tensile and flexural loadings

    is analyzed using the direct numerical method and the homoge-

    nization method and the results are compared to verify the devel-

    oped approach. Then, a 3D porous SMA model is analyzed under

    the tensile loading and the results are compared with the available

    numerical data. Two porous SMA models are also analyzed under

    the bending and torsional loadings. Moreover, the influence of the

    shape of a constant- volume porosity on the macroscopic response

    of the porous SMA is investigated.

    4.1. 3D SMA-fiber reinforced composite

    The macroscopic behavior of a 3D SMA-fiber reinforced com-

    posite subjected to two loading/unloading cycles (tension and

    bending) is analyzed using the direct finite element solution and

    the present homogenization method. The specimen is composed of

    SMA fibers, wired through an elastic matrix, with the fiber vol-

    ume fraction of 0.267. Material properties of the elastic matrix and

    the SMA fiber are given in Tables 1 and 2 . The super-elastic re-

    sponse of the SMA fiber and the elastic response of the matrix

    under uniaxial tensile loading/unloading are shown in Fig. 6 . The

    inclusion and the matrix are assumed fully bonded. It is possible

    to introduce a debonding softening law between the matrix and

    inclusions, which requires major changes in the homogenization

    algorithm ( Nguyen et al., 2012, 2011 ), and is out of scope of the

    present work. It should be noted that all the tests are carried out

    in an isothermal condition throughout the paper.

    The RVE and the macro model used for the homogenization

    analysis are meshed by 224 and 32 8-noded linear hexahedral el-

    ements, respectively (see Fig. 7 (a)). The model used for the direct

    finite element analysis is meshed by 161280 8-noded linear hexa-

    hedral elements for the arrangement of fibers shown in Fig. 7 (b).

    In the homogenization solution procedure, the external forces

    are applied incrementally on the macro model. Then, the ob-

    tained macroscopic strain tensor at each Gauss point of the macro-

    element is transferred to a periodically constrained RVE. The RVE

    is solved using an iterative FE scheme and the resultant effec-

    tive stress and the algorithmic tangent tensor are conveyed to the

    macro solution procedure. The macro solution continues iteratively

    until the residual falls under the designated tolerance.

    First, the composite beam is restrained in a way that at X = 0, u X = 0; at Y = 0, u Y = 0; and at Z = 0, u Z = 0. Then, the beam is pulled in the Z direction by a surface load σ =1240MPa applied

  • P. Fatemi Dehaghani et al. / International Journal of Solids and Structures 118–119 (2017) 24–40 35

    (a)

    (b)

    Fig. 20. Distribution of the martensitic volume fraction at (a) load level A and (b) load level B for RVEs with different shape factors n.

    at Z = 100 mm. The stress-strain diagrams from the homogenization and direct solutions, shown in Fig. 8 , illustrate a very good agree-

    ment with a maximum error of less than 0.1%.

    Fig. 9 illustrates distribution of the martensitic volume frac-

    tion in a specific RVE, shown in Fig. 7 (a), at various load levels.

    The SMA fiber is in fully austenite state at level A. With the in-

    crease of load level, fully martensitic state is developed at level D.

    Reverse transformation can be tracked with the model being un-

    loaded down to level H. Fig. 10 depicts distribution of the marten-

    sitic volume fraction in the direct solution at load levels B and D,

    which closely coincide with the homogenization solutions.

    The same composite is then subjected to a flexural load-

    ing/unloading while is fully restrained at Z = 0. Due to inaccu- racy of linear elements for bending problems with coarse meshes,

    the macro model in the homogenization analysis is meshed by

    32 20-noded hexahedral elements. The surface load σ =33.33MPa is transversely applied at Z = 100mm (see Fig. 11 ). The tip load- displacement diagram from the homogenization and direct anal-

    yses are compared in Fig. 12 that expresses a maximum of 0.4%

    difference while the run time has decreased from 19,003s for the

    direct analysis to 3757s for the homogenization analysis (Both sim-

    ulations have been performed on a PC with an Intel(R) Xeon(R)

  • 36 P. Fatemi Dehaghani et al. / International Journal of Solids and Structures 118–119 (2017) 24–40

    Fig. 21. Distribution of the normal stress at the end of loading for RVEs with different shape factors n.

    E5-2690 CPU (2.9 GHz, 32 cores) and 64GB of RAM). This empha-

    sizes how the homogenization analysis decreases the calculation

    time and preserves the accuracy of the nonlinear analysis. The dif-

    ference in elapsed computational time would be more notable in

    real engineering problems with sophisticated micro structures.

    Distribution of the martensitic volume fraction and von-Misses

    stress for two RVEs, located at a pair of transversely iso-planar

    macro-element Gauss points, at farthest positions with respect to

    the neutral axis, are shown on the deformed RVE (magnification

    scale factor = 15) in Fig. 13 . It is observed that, the RVE above the neutral axis is stretched due to interaction of the shear stress and

    positive normal stress, whereas the RVE below the neutral axis is

    compressed due to the interaction of the shear stress and negative

    normal stress. Locations of the Gauss points are illustrated in Fig.

    11 .

    4.2. 3D porous SMA

    In this example, the macroscopic behavior of 3D porous SMA

    specimens is discussed. The problem, originally presented by Sepe

    et al. (2015) , is considered in order to verify the constitutive be-

    havior of the porous SMA. The SMA material parameters based

    on the Boyd and Lagoudas (1996) model are calibrated in order

    to represent the SMA constitutive behavior proposed by Sepe et

    al. (2015) (see Table 3 ). The stress-strain diagram for a dense SMA

    Table 3

    Material properties of SMA.

    Material parameters Values Material parameters Values

    E A (GPa) 75 T M f

    (K) 296.24

    E M (GPa) 66 T M s (K) 316

    υA 0.33 T A s (K) 310.4

    υM 0.33 T A f (K) 329

    H 0.0276 C A (MPa K −1 ) 30

    T 0 (K) 331 C M (MPaK -1 ) 30

    specimen under the tensile loading/unloading is shown in Fig. 14 .

    Clearly, the material parameter calibration has resulted in a good

    agreement between the two models.

    In order to investigate the macroscopic tensile behavior of the

    porous SMA specimen, a single 8-noded linear hexahedral macro

    element subjected to unidirectional tensile loading/unloading is

    analyzed by the present homogenization method. At each macro

    Gauss point, an RVE with a spherical void and porosity of f = 13% is solved concurrently with the finite element mesh of Fig. 15 . The

    macroscopic stress-strain response is presented in Fig. 16 , which is

    in a very good agreement with the response reported by Sepe et

    al. (2015) .

    Fig. 17 (a) postulates that the porous SMA constitutive behavior

    resembles the dense SMA to a great extent and the stress-strain

  • P. Fatemi Dehaghani et al. / International Journal of Solids and Structures 118–119 (2017) 24–40 37

    (a)

    (b)

    Fig. 22. Geometry of the models, macro mesh and RVE mesh (a) flexural specimen and (b) torsional specimen.

    Fig. 23. Load-displacement diagrams from the homogenization and the direct finite

    element analysis.

    response can be approximated by a trilinear curve for the loading

    path. Regions I & III represent the austenite and martensite state,

    respectively, while the martensitic transformation ( A → M ) occurs in region II. The same is true for the unloading path. In region

    IV, the material is in the martensite state while in region V, the

    reverse transformation ( M → A ) occurs, followed by the austenite state in region VI. Distribution of the martensitic volume fraction

    for an arbitrarily chosen RVE is shown in Fig. 17 (b). Note that

    RVEs at all Gauss points are in the same mechanical state due to

    the nature of the tensile test. It can be seen that the transforma-

    tion process initiates around the void and then propagates through

    the most of RVE by increasing the load level. Complete reverse

    transformation to the austenite state is observed along the unload-

    ing path. At the load level A, the whole RVE is in the austenite

    state, which can also be predicted from Fig. 17 (a). The martensitic

    transformation initiates gradually by increasing the load up to the

    level B, situated in region II of Fig. 17 (b), at which the maximum

    martensitic volume fraction in RVE is 0.75. The load is then in-

    creased to the level C which is also located in the same region II of

    Fig. 17 (a). At this load level, the fully martensite state is developed

    in the vicinity of the ellipsoid poles, while most of the load bearing

    region still remains in the transformation phase. By increasing the

    load to level D, most of the load bearing region transforms to the

    fully martensite state and thus the load level D is located in region

    III of Fig. 17 (a). As expected, unloading down to the level E, which

    is located in region IV of Fig. 17 (a), does not make a noticeable

    change in distribution of the martensitic volume fraction in RVE.

    Further decrease of the load level to F and then G, which are lo-

    cated in region V of Fig. 17 (a), will trigger and extend the reverse

    transformation, respectively. Finally, at the load level H the fully

    austenite state is reversed in RVE and the macroscopic response is

    positioned in region VI of Fig. 17 (a).

    In order to investigate the influence of the shape of porosity on

    the macroscopic behavior of porous SMA specimen, series of anal-

  • 38 P. Fatemi Dehaghani et al. / International Journal of Solids and Structures 118–119 (2017) 24–40

    Fig. 24. Distribution of the martensitic volume fraction for two RVEs located at (a) Gauss point A and (b) Gauss point B.

    Fig. 25. Load-displacement response of the plate at node A.

    yses have been conducted on the same macro model of the pre-

    vious example using several three-dimensional RVEs with different

    shape factors, n, as defined in Fig. 18 . It should be noted that the

    porosity is taken constant in all RVEs ( f = 13%). The macroscopic stress-strain responses are compared in Fig. 19 , which shows that

    with decreasing the shape factor n from 3 to 0.6, the effective load

    carrying surface normal to the load direction is increased and thus

    a noticeable growth in load bearing capacity is observed.

    Fig. 20 shows distribution of the martensitic volume fraction at

    load levels A and B for different types of RVEs. According to Fig.s

    19 and 20(a), at the load level A, the RVE with shape factor n = 0.6 is at the threshold of the transformation with negligible marten-

    sitic region, while the RVE with shape factor n = 3.0 has already started the martensitic transformation and even completed trans-

    formation can be observed in the vicinity of the ellipsoid poles. In

    fact, the higher stress in the latter RVE has resulted in faster initi-

    ation of the forward transformation. At the load level B (see Figs.

    19 and 20 (b)), almost all of the load bearing region of the RVE

    with shape factor n = 3.0 has completed the martensitic transfor- mation and thus the stress-strain response has marginally entered

    the third part of the trilinear curve (which was phenomenologi-

    cally explained in Fig. 17 (a)), while most of the load bearing region

    of the RVE with shape factor n = 0.6 is still in transformation. Fig. 21 presents distribution of the normal stress in the direc-

    tion of applied load at the end of loading. By increasing the poros-

    ity shape factor from n = 0.6–3, most of the load bearing surface at mid-height of the RVE decays and the geometry of the void tends

    to represent a surface crack perpendicular to the load direction, re-

    sulting in a noticeable 220% increase in the stress field in the vicin-

    ity of the ellipsoid poles, which underlines the important effect of

    shape of porosity on microscopic behavior of the porous SMA.

    Now, two new models are considered. The first model is an-

    alyzed under the flexural loading/unloading, while the second one

    is studied subjected to the torsional loading/unloading. The macro-

    scopic models are meshed using 48 and 96 20-noded hexahedral

    elements, respectively and the RVE is meshed by 2163 4-noded

    linear tetrahedral elements. Geometry of the specimens, applied

    loads, macro mesh and RVE mesh are shown in Fig. 22 .

    The flexural loading/unloading case is investigated first. The tip

    load-displacement diagrams from the homogenization and the di-

    rect finite element analysis are compared in Fig. 23 , which are

    in good agreement. The direct model is composed of 192 patches

    with a total of 415,296 elements. It should be noted that all the

    required history variables must be saved in each step of analysis

    in both direct and multiscale solutions. Due to the less number of

    Gauss points in the multiscale model, fewer variables should be

    saved compared with the direct model. In this example, 830,592

    martensitic volume fraction variables, ξ , should be saved for the multiscale model in contrast with 3,322,368 variables for the di-

    rect model.

  • P. Fatemi Dehaghani et al. / International Journal of Solids and Structures 118–119 (2017) 24–40 39

    Fig. 26. Distribution of the homogenized von-Misses stress on the deformed macro mesh and distribution of the martensitic volume fraction on the specified RVE at different

    load levels.

    Distribution of the martensitic volume fraction is shown in Fig.

    24 for the two RVEs located at Gauss points A and B at the end

    of loading. Shrinkage and elongation of voids in bending are ob-

    served at A and B, respectively, due to their relative positions with

    respect to the neutral axis. It is also observed that since the SMA

    has a similar tensile and compressive behavior, distributions of the

    martensitic volume fraction are identical in RVEs located at A and

    B.

    Finally, the plate of Fig. 22 (b) is analyzed subjected to a tor-

    sional loading/unloading using the homogenization method. The

    plate is distorted due to the applied Neumann and Dirichlet

    boundary conditions. The load-displacement response of the plate

    at point A is shown in Fig. 25 .

    Distribution of the homogenized von-Misses stress on the de-

    formed macro mesh (magnification scale factor = 15) and the

    martensitic volume fraction distribution on a specific RVE (lo-

    cated at the specified Gauss point shown on the Fig. 26 ) at dif-

    ferent load levels are shown in Fig. 26 . Note that the results are

    displayed under the micro fluctuation field (magnification scale

    factor = 50). Higher deformation gradients at the restrained cor- ners of the macroscopic model have caused stress concentration

    in those regions.

    5. Conclusion

    The proposed finite element based 3D hierarchical (micro-

    macro) multiscale analysis framework using periodic homogeniza-

    tion method has been successfully applied to the problems of SMA

    reinforced composite and porous SMA. The effectiveness of the

    method for analysis of composite materials, where the macroscopic

    constitutive law is not available and exact minute meshing is a te-

  • 40 P. Fatemi Dehaghani et al. / International Journal of Solids and Structures 118–119 (2017) 24–40

    dious task, has been emphasized. A simple bending test was used

    to underscore how the homogenization analysis decreased the cal-

    culation time by more than 5 times while the accuracy of the non-

    linear analysis was preserved. The difference in calculation time

    would be more notable in real engineering problems with sophis-

    ticated micro structures. Analyses have been performed on spec-

    imens subjected to various stress loadings cycles and the results

    have been compared with the solutions obtained by fully meshed

    direct finite element, for verification purposes. The multiscale tech-

    nique allows in-depth investigation of distribution of the field vari-

    ables and matrix-inclusion interaction in the micro scale. The in-

    fluence of the shape of a constant-volume porosity on the macro-

    scopic response of the porous SMA has also been investigated

    through a parametric study on the role of the shape of porosity on

    the concentration of the field variables such as the martensitic vol-

    ume fraction and the von Misses and normal stresses. It has been

    shown that the shape and orientation of the porosity can consid-

    erably make a shift in pre-transformation load carrying capacity of

    porous SMAs.

    Acknowledgment

    The authors wish to gratefully acknowledge the technical sup-

    port of the High Performance Computing Lab, School of Civil Engi-

    neering, University of Tehran. The financial support of Iran National

    Science Foundation (INSF) is gratefully acknowledged.

    References

    Abeyaratne, R. , Knowles, J.K. , 1993. A continuum model of a thermoelastic solid ca-

    pable of undergoing phase transitions. J. Mech. Phys. Solids 41, 541–571 . Ahmadian, H. , Ardakani, S.H. , Mohammadi, S. , 2015. Strain-rate sensitivity of unsta-

    ble localized phase transformation phenomenon in shape memory alloys using a non-local model. Int. J. Solids Struct. 63, 167–183 .

    Anand, L. , Gurtin, M.E. , 2003. Thermal effects in the superelasticity of crystalline shape-memory materials. J. Mech. Phys. Solids 51, 1015–1058 .

    Auricchio, F. , Taylor, R.L. , Lubliner, J. , 1997. Shape-memory alloys: macromodelling

    and numerical simulations of the superelastic behavior. Comput. Meth. Appl. Mech. Eng. 146, 281–312 .

    Bekker, A. , Brinson, L. , 1997. Temperature-induced phase transformation in a shape memory alloy: phase diagram based kinetics approach. J. Mech. Phys. Solids 45,

    949–988 . Bo, Z. , Lagoudas, D.C. , 1999. Thermomechanical modeling of polycrystalline SMAs

    under cyclic loading, Part I: theoretical derivations. Int. J. Eng. Sci. 37, 1089–1140 .

    Boyd, J.G. , Lagoudas, D.C. , 1996. A thermodynamical constitutive model for shape

    memory materials. Part I. The monolithic shape memory alloy. Int. J. Plasticity 12, 805–842 .

    Chatzigeorgiou, G. , Chemisky, Y. , Meraghni, F. , 2015. Computational micro to macro transitions for shape memory alloy composites using periodic homogenization.

    Smart Mater. Struct. 24, 035009 . Cherkaoui, M. , Sun, Q. , Song, G. , 20 0 0. Micromechanics modeling of composite with

    ductile matrix and shape memory alloy reinforcement. Int. J. Solids Struct. 37,

    1577–1594 . Damanpack, A. , Aghdam, M. , Shakeri, M. , 2015a. Micro-mechanics of composite with

    SMA fibers embedded in metallic/polymeric matrix under off-axial loadings. Eur. J. Mech.-A/Solids 49, 467–480 .

    Damanpack, A. , Aghdam, M. , Shakeri, M. , 2015b. Micromechanics of shape mem- ory alloy fiber–reinforced composites subjected to multi-axial non-proportional

    loadings. J. Intell. Mater. Syst. Struct. 26, 2431–2445 .

    Damanpack, A. , Liao, W. , Aghdam, M. , Shakeri, M. , Bodaghi, M. , 2015c. Micro–macro thermo-mechanical analysis of axisymmetric shape memory alloy composite

    cylinders. Compos. Struct. 131, 1001–1016 . Dvorak, G.J. , 1992. Transformation field analysis of inelastic composite materials. In:

    Proceedings of the Royal Society of London A: Mathematical, Physical and Engi- neering Sciences. The Royal Society, pp. 311–327 .

    Eftekhari, M. , Ardakani, S.H. , Mohammadi, S. , 2014. An XFEM multiscale approach

    for fracture analysis of carbon nanotube reinforced concrete. Theor. Appl. Fract. Mech. 72, 64–75 .

    Fish, J. , Kuznetsov, S. , 2010. Computational continua. Int. J. Numer. Methods Eng. 84, 774–802 .

    Fish, J. , Shek, K. , 1999. Finite deformation plasticity for composite structures: com- putational models and adaptive strategies. Comput. Meth. Appl. Mech. Eng. 172,

    145–174 .

    Fish, J. , Shek, K. , Pandheeradi, M. , Shephard, M.S. , 1997. Computational plasticity for composite structures based on mathematical homogenization: theory and prac-

    tice. Comput. Meth. Appl. Mech. Eng. 148, 53–73 . Fish, J. , Yuan, Z. , 2007. Multiscale enrichment based on partition of unity for non-

    periodic fields and nonlinear problems. Comput. Mech. 40, 249–259 . Geers, M.G. , Kouznetsova, V.G. , Brekelmans, W. , 2010. Multi-scale computational ho-

    mogenization: trends and challenges. J. Comput. Appl. Math. 234, 2175–2182 . Govindjee, S. , Kasper, E.P. , 1999. Computational aspects of one-dimensional shape

    memory alloy modeling with phase diagrams. Comput. Meth. Appl. Mech. Eng.

    171, 309–326 . Hassani, B. , Hinton, E. , 1998a. A review of homogenization and topology opimiza-

    tion II—analytical and numerical solution of homogenization equations. Comput. Struct. 69, 719–738 .

    Hassani, B. , Hinton, E. , 1998b. A review of homogenization and topology opti- mization I—homogenization theory for media with periodic structure. Comput.

    Struct. 69, 707–717 .

    Helm, D. , Haupt, P. , 2003. Shape memory behaviour: modelling within continuum thermomechanics. Int. J. Solids Struct. 40, 827–849 .

    Kouznetsova, V.G. , 2002. Computational Homogenization for the Multi-scale Analy- sis of Multi-phase Materials. Eindhoven University of Technology .

    Lagoudas, D. , Boyd, J. , Bo, Z. , 1994. Micromechanics of active composites with SMA fibers. J. Eng. Mater. Technol. 116, 337–347 .

    Lagoudas, D.C. , Entchev, P.B. , 2004. Modeling of transformation-induced plasticity

    and its effect on the behavior of porous shape memory alloys. Part I: constitu- tive model for fully dense SMAs. Mech. Mater. 36, 865–892 .

    Lecce, L. , Concilio, A. , 2015. Shape Memory Alloy Engineering: For Aerospace, Struc- tural and Biomedical Applications. Elsevier .

    Lubliner, J. , Auricchio, F. , 1996. Generalized plasticity and shape memory alloys. Int. J. Solids Struct. 33, 991–1003 .

    Marfia, S. , 2005. Micro–macro analysis of shape memory alloy composites. Int. J.

    Solids Struct. 42, 3677–3699 . Marfia, S. , Sacco, E. , 2007. Analysis of SMA composite laminates using a multiscale

    modelling technique. Int. J. Numer. Methods Eng. 70, 1182–1208 . McVeigh, C. , Vernerey, F. , Liu, W.K. , Brinson, L.C. , 2006. Multiresolution analysis for

    material design. Comput. Meth. Appl. Mech. Eng. 195, 5053–5076 . Mori, T. , Tanaka, K. , 1973. Average stress in matrix and average elastic energy of

    materials with misfitting inclusions. Acta Metall. 21, 571–574 .

    Nguyen, V.P. , Lloberas-Valls, O. , Stroeven, M. , Sluys, L.J. , 2012. Computational homog- enization for multiscale crack modeling. Implementational and computational

    aspects. Int. J. Numer. Methods Eng. 89, 192–226 . Nguyen, V.P. , Stroeven, M. , Sluys, L.J. , 2011. Multiscale continuous and discontinu-

    ous modeling of heterogeneous materials: a review on recent developments. J. Multiscale Modell. 3, 229–270 .

    Oskay, C. , Fish, J. , 2007. Eigendeformation-based reduced order homogenization for

    failure analysis of heterogeneous materials. Comput. Meth. Appl. Mech. Eng. 196, 1216–1243 .

    Paiva, A. , Savi, M.A. , Braga, A.M.B. , Pacheco, P.M.C.L. , 2005. A constitutive model for shape memory alloys considering tensile–compressive asymmetry and plastic-

    ity. Int. J. Solids Struct. 42, 3439–3457 . Popov, P. , Lagoudas, D.C. , 2007. A 3-D constitutive model for shape memory alloys

    incorporating pseudoelasticity and detwinning of self-accommodated marten- site. Int. J. Plast. 23, 1679–1720 .

    Qidwai, M. , Lagoudas, D. , 20 0 0. On thermomechanics and transformation surfaces

    of polycrystalline NiTi shape memory alloy material. Int. J. Plast. 16, 1309–1343 . Raniecki, B. , Lexcellent, C. , 1998. Thermodynamics of isotropic pseudoelasticity in

    shape memory alloys. Eur. J. Mech.-A/Solids 17, 185–205 . Reese, S. , Christ, D. , 2008. Finite deformation pseudo-elasticity of shape memory

    alloys–constitutive modelling and finite element implementation. Int. J. Plast. 24, 455–482 .

    Sepe, V. , Auricchio, F. , Marfia, S. , Sacco, E. , 2015. Micromechanical analysis of porous

    SMA. Smart Mater. Struct. 24, 085035 . Sepe, V. , Auricchio, F. , Marfia, S. , Sacco, E. , 2016. Homogenization techniques for the

    analysis of porous SMA. Comput. Mech. 57, 755–772 . Sepe, V., Marfia, S., Auricchio, F., 2014. Response of porous SMA: a micromechanical

    study. Frattura ed Integrità Strutturale, 85. Sepe, V. , Marfia, S. , Sacco, E. , 2013. A nonuniform TFA homogenization technique

    based on piecewise interpolation functions of the inelastic field. Int. J. Solids

    Struct. 50, 725–742 . Tadmor, E.B. , Miller, R.E. , 2011. Modeling materials: continuum, Atomistic and Mul-

    tiscale Techniques. Cambridge University Press . Tian, R. , Chan, S. , Tang, S. , Kopacz, A.M. , Wang, J.-S. , Jou, H.-J. , Siad, L. , Lindgren, L.-E. ,

    Olson, G.B. , Liu, W.K. , 2010. A multiresolution continuum simulation of the duc- tile fracture process. J. Mech. Phys. Solids 58, 1681–1700 .

    Yuan, Z. , Fish, J. , 2008. Toward realization of computational homogenization in prac-

    tice. Int. J. Numer. Methods Eng. 73, 361–380 . Zhao, Y. , Taya, M. , 2007. Analytical modeling for stress-strain curve of a porous NiTi.

    J. Appl. Mech. 74, 291–297 . Zhu, Y. , Dui, G. , 2011. A model considering hydrostatic stress of porous NiTi shape

    memory alloy. Acta Mech. Solida Sin. 24, 289–298 .

    http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0001http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0001http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0001http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0002http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0002http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0002http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0002http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0003http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0003http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0003http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0004http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0004http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0004http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0004http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0005http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0005http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0005http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0006http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0006http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0006http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0007http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0007http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0007http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0008http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0008http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0008http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0008http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0009http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0009http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0009http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0009http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0010http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0010http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0010http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0010http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0011http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0011http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0011http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0011http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0012http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0012http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0012http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0012http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0012http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0012http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0013http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0013http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0014http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0014http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0014http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0014http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0015http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0015http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0015http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0016http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0016http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0016http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0017http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0017http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0017http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0017http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0017http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0018http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0018http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0018http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0019http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0019http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0019http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0019http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0020http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0020http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0020http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0021http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0021http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0021http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0022http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0022http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0022http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0023http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0023http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0023http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0024http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0024http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0025http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0025http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0025http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0025http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0026http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0026http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0026http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0027http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0027http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0027http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0028http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0028http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0028http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0029http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0029http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0030http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0030http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0030http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0031http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0031http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0031http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0031http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0031http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0032http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0032http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0032http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0033http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0033http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0033http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0033http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0033http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0034http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0034http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0034http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0034http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0035http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0035http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0035http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0036http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0036http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0036http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0036http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0036http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0037http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0037http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0037http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0038http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0038http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0038http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0039http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0039http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0039http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0040http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0040http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0040http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0041http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0041http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0041http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0041http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0041http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0042http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0042http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0042http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0042http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0042http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0043http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0043http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0043http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0043http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0044http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0044http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0044http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0045http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0045http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0045http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0045http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0045http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0045http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0045http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0045http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0045http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0045http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0045http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0046http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0046http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0046http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0047http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0047http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0047http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0048http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0048http://refhub.elsevier.com/S0020-7683(17)30176-2/sbref0048

    3D hierarchical multiscale analysis of heterogeneous SMA based materials1 Introduction2 SMA constitutive model3 3D periodic hierarchical multiscale homogenization method4 Numerical results4.1 3D SMA-fiber reinforced composite4.2 3D porous SMA

    5 Conclusion Acknowledgment References