Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina Barello; Anastasia H. Muliana; Martin Lé -- Coupled heat conduction and thermal

Embed Size (px)

Citation preview

  • 8/10/2019 Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina

    1/18

    Coupled heat conduction and thermal stress analysesin particulate composites

    Kamran A. Khan a, Romina Barello b, Anastasia H. Muliana a,, Martin Lvesque b

    a Department of Mechanical Engineering, Texas A&M University, USAb CREPEC, Department of Mechanical Engineering, Ecole Polytechnique de Montreal, Canada

    a r t i c l e i n f o

    Article history:

    Received 2 May 2010

    Received in revised form 15 April 2011

    Available online 13 July 2011

    Keywords:

    Heat conduction

    Thermal stresses

    Particulate composites

    Finite element

    Micromechanical model

    a b s t r a c t

    This study introduces two micromechanical modeling approaches to analyze spatial vari-

    ations of temperatures, stresses and displacements in particulate composites during tran-

    sient heat conduction. In the first approach, a simple micromechanical model based on a

    first order homogenization scheme is adopted to obtain effective mechanical and thermal

    properties, i.e., coefficient of linear thermal expansion, thermal conductivity, and elastic

    constants, of a particulate composite. These effective properties are evaluated at each

    material (integration) point in three dimensional (3D) finite element (FE) models that rep-

    resent homogenized composite media. The second approach treats a heterogeneous com-

    posite explicitly. Heterogeneous composites that consist of solid spherical particles

    randomly distributed in homogeneous matrix are generated using 3D continuum elements

    in an FE framework. For each volume fraction (VF) of particles, the FE models of heteroge-

    neous composites with different particle sizes and arrangements are generated such that

    these models represent realistic volume elements cut out from a particulate composite.

    An extended definition of a RVE for heterogeneous composite is introduced, i.e., the num-

    ber of heterogeneities in a fixed volume that yield the same expected effective response for

    the quantity of interest when subjected to similar loading and boundary conditions. Ther-

    mal and mechanical properties of both particle and matrix constituents are temperature

    dependent. The effects of particle distributions and sizes on the variations of temperature,

    stress and displacement fields are examined. The predictions of field variables from the

    homogenized micromechanical model are compared with those of the heterogeneous

    composites. Both displacement and temperature fields are found to be in good agreement.

    The micromechanical model that provides homogenized responses gives average values of

    the field variables. Thus, it cannot capture the discontinuities of the thermal stresses at the

    particlematrix interface regions and local variations of the field variables within particle

    and matrix regions.

    2011 Elsevier Ltd. All rights reserved.

    1. Introduction

    The existence of thermal stresses has always been a

    subject of discussion when a body is subjected to coupled

    heat conduction and mechanical loadings. In composites,

    significant thermal stresses can arise due to the mismatch

    in the coefficient of thermal expansions (CTEs) of the con-

    stituents, affecting their overall performance. The use of

    composite materials in structural components requires

    understanding the variations in the field variables, such

    as stress, strain, temperature and displacement, both at

    micro- and macro scales. At the macro scale, composite

    structures are often analyzed as homogeneous structures

    through the use of effective properties, which allow per-

    forming large-scale structural analyses. Composites are

    heterogeneous materials that can exhibit large variations

    and even discontinuities in the field variables. These

    0167-6636/$ - see front matter 2011 Elsevier Ltd. All rights reserved.doi:10.1016/j.mechmat.2011.06.013

    Corresponding author.

    E-mail address: [email protected](A.H. Muliana).

    Mechanics of Materials 43 (2011) 608625

    Contents lists available at ScienceDirect

    Mechanics of Materials

    j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c at e / m e c h m a t

    http://dx.doi.org/10.1016/j.mechmat.2011.06.013mailto:[email protected]://dx.doi.org/10.1016/j.mechmat.2011.06.013http://www.sciencedirect.com/science/journal/01676636http://www.elsevier.com/locate/mechmathttp://www.elsevier.com/locate/mechmathttp://www.sciencedirect.com/science/journal/01676636http://dx.doi.org/10.1016/j.mechmat.2011.06.013mailto:[email protected]://dx.doi.org/10.1016/j.mechmat.2011.06.013
  • 8/10/2019 Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina

    2/18

    variations and discontinuities at the micro-scale cannot be

    captured if one treats composites as homogenous (homog-

    enized) materials.

    Different types of micromechanical models with simpli-

    fied microstructures have been developed to obtain effec-

    tive thermal and mechanical properties. The composite

    spheres model (Hashin, 1962), the self consistent approach

    (Budiansky, 1965 and Hill, 1965); the generalized self

    consistent scheme (Christensen and Lo, 1979, 1986), the

    Mori-Tanaka model (Mori and Tanaka, 1973), the probabi-

    listic approach ofChen and Acrivos (1978), the differential

    method (McLaughlin, 1977 and Norris, 1985) are some

    examples. Detailed discussion of various micromechanical

    models and bounds on the effective mechanical properties

    can be found inAboudi (1991), Mura (1987),Nemat-Nas-

    ser and Hori (1999). Khan and Muliana (2010) presented

    Fig. 1. Schematic diagram of integration of micro-macro scale approach for particulate composites.

    Fig. 2. Homogenization of the sub-regions of a composite component.

    K.A. Khan et al. / Mechanics of Materials 43 (2011) 608625 609

  • 8/10/2019 Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina

    3/18

    a short review of the capabilities and shortcoming of few

    analytical and numerical models for effective CTE and

    effective thermal conductivity (ETC) of composites. Analyt-

    ical expressions for ETC for two-phase composites made of

    randomly distributed and dilute concentrations of spheres

    in a homogeneous medium are given, for example, by Max-

    well (1954), Verma et al. (1991) and Hasselman and John-

    son (1987). Expressions for non-dilute concentrations of

    spheres are provided by the models ofJeffrey (1973), Davis

    (1986) and Sangani and Yao (1988). These models lead to

    accurate ETC predictions, when compared to experimental

    results, when the volume fraction (VF) is relatively small or

    when the conductivities of the constituents are compara-

    ble (Matt and Cruz, 2002). Bounds on the effective thermal

    conductivity have been derived using variational principle

    (Hashin and Shtrikman, 1962; Beran, 1965) and three point

    probability function for the distribution of microstructures.

    Several numerical modeling approaches have been

    developed to estimate effective thermal and mechanical

    behavior of a composite. To obtain effective heat conduc-

    tion response and ETC of a composite the composites are

    often described with random microstructures (Ostoja-Star-

    zewski and Schulte, 1996; Nogales and Bohm, 2008) as

    well as with periodic microstructures (Auriault, 1983; Ver-

    ma et al., 1991; Jiang et al., 2002). Real composite micro-

    structures are generally non periodic; however, the use

    of periodic microstructures can give approximate effective

    properties for a smaller computational cost when com-

    pared to that of random microstructures. To obtain the

    effective thermo-mechanical response, among the numer-

    ical methods, the Finite Element (FE) technique has been

    used for determining the micro- and macro-structural per-

    formance of composites by meshing their detailed micro-

    structures; see for example the works of Llorca and

    Segurado (2002, 2003 and 2006), Lvesque et al. (2004,

    2008) and Zohdi and Wriggers (2001). The composite

    behavior was obtained from simulations of a 3D cubic Rep-

    resentative Volume Element (RVE) containing several ran-

    domly distributed non-overlapping identical spheres.

    Periodic boundary conditions were imposed on the meshes

    in order to obtain the effective properties. Linearly elastic,

    elasticplastic and linearly (Lvesque and Barello, 2009)

    and nonlinearly viscoelastic (Lvesque et al. 2004) re-

    sponses were analyzed. Kwon and Kim (1998) developed

    a 3D micromechanical model to compute thermal stresses

    of particulate composites. Using a similar type of microme-

    chanical model, Meijer et al. (2000) studied the influence of

    inclusion geometry and thermal residual stresses- and

    strains on the mechanical behavior of an aluminum alloy

    reinforced with aluminum oxide particles. A unit-cell con-

    sisting of eight subcells was considered. One cell repre-

    sents the particle and the rest of the cells surrounding

    the particle were considered as matrix. Both constituents

    were assumed to be linearly elastic and the thermal prop-

    erties were considered to be temperature independent. The

    effects of the size of the subcells on the stresses and strains

    Fig. 3. 3D FE models of (a) homogenized and (b) heterogeneous composites. (c) Schematic of thermal and mechanical loading directions and profiles along

    which field variables are evaluated, i.e., AB, CD, EF and GH.

    Table 1

    Temperature dependent mechanical and physical properties of materials of Ti6Al4V and ZrO 2used in 3D FE analyses.

    Property Ti6Al4V Zirconia (Zr02)

    Young modulus (E) (Pa) 1.23 101156.46 106T 2.44 1011334.28 106T+ 295.24 103T2 89.79T3

    Poisson ratio (t) 0.3 0.3Coefficient of thermal expansion (a)106,

    1/K

    7.58 106 + 4.93 109

    T+ 2.39 1012 T21.28 105 19.07 109 T+ 1.28 1011

    T2 8.67 1017 T3

    Thermal conductivity, (k), W/m/K 1.21 + 0.0169T 1.7+ 2.17 104 T+ 1.13 105 T2

    Specific heat (c), J/kg K 625.297 0.264T+ 4.49 104 T2 487.343 + 0.149T 2.94 105 T2

    Density (q), kg/m3 4429 5700

    T is temperature in Kelvin (K).

    610 K.A. Khan et al./ Mechanics of Materials 43 (2011) 608625

  • 8/10/2019 Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina

    4/18

    were analyzed. It was found that the sharp corner and

    edges of the cube resulted in localized stresses and strains.

    Recently, Aboudi (2008)presented a micromodel for fully

    coupled thermo-mechanical analysis for multiphase com-

    posites where heat generation from the interconvertibility

    of the mechanical and thermal energy was accounted for.

    This study focuses on understanding the effects of con-

    stituents properties and microstructural details on the

    variation of stress, displacement and temperature in com-

    posites. The results are used to justify the capability of

    micromechanical models to analyze the overall response

    of composites subjected to simultaneous mechanical and

    thermal stimuli, within a certain degree of accuracy. A

    simplified micromechanical model for predicting the effec-

    tive thermo-mechanical behavior of particulate composites

    subjected to both time and space varying temperature field

    is introduced. The micromechanical model is called at each

    integration point in a FE mesh to calculate thelocal effective

    properties of the composite. The responses thus obtained

    are then compared to FE simulations of composites using

    the meshes ofBarello and Lvesque (2008). A sequentially

    thermo-mechanical coupled problem, i.e., the temperature

    field influences the deformation field, is considered. The

    effects of particle volume contents and temperature depen-

    dent constituent properties on the overall thermo-elastic

    behavior of the composites are examined.

    The paper is organized as follows. A brief outline of the

    simplified micromechanicalmodel is presented in Section2.

    The effective thermoelastic stress-strain relations, effective

    heat flux equation and uncoupled energy equation for an

    isotropic homogenized composite are also discussed.

    Section 3 presents a micro-mechanical framework used

    for computing the response of a real-size composite. FE

    modeling of composites with microstructural details is

    given in Section4. Coupled heat conduction and thermal

    stress analysis of particulate composites are discussed in

    Section5. A summary of the research findings is given in

    Section6.

    2. A simplified micromechanical model for particle

    reinforced composites

    Muliana and Kim (2007), Muliana (2008) and Khan and

    Muliana (2010) developed a micromechanical model for

    determining the effective viscoelastic responses and

    300

    400

    500

    600

    700

    Temperature(K)

    Temperature(K)

    a

    t =12s

    t=26s

    t=2s

    Steady State Time = 150 seconds

    Detail Microstructural Model

    Micromechanical Model(CPU Time = 277s)

    Model-1 (CPU Time = 12785s)

    Model-2 (CPU Time = 13386s)

    Model-3 (CPU Time = 12481s)

    Model-4 (CPU Time = 12519s)

    Model-5 (CPU Time = 12152s)

    Model-6 (CPU Time = 11319s)

    (Volume fraction = 20%, 20 particles)

    b

    0 2 4 6 8 10

    Distance (mm) Distance (mm)

    300

    400

    500

    600

    700

    300

    400

    500

    600

    700

    Temperature(K)

    Temperature(K)

    300

    400

    500

    600

    700

    t=12s

    t =26s

    t=2s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    (Volume fraction = 20%, 40 particles)

    t=12s

    t=26s

    t =2s

    Detail Microstructural Model (Volume fraction = 20%, 20 particles)

    Micromechanical Model

    c

    0 2 4 6 8 10

    0 2 4 6 8 10

    Distance (mm) Distance (mm)0 2 4 6 8 10

    t=12s

    t=26s

    t=2s

    Detail Microstructural Model

    Micromechanical Model

    (Volume fraction = 20%, 40 particles)

    d

    Fig. 4. Comparison of temperature profiles for FE models with the unit cell (micromechanical model) at each integration point (solid line) and the FEmodels with 3D microstructural detail (symbols) for volume fraction of 20%. (a) and (b) are actual values of temperature at top (corner) edge {( X1, 10, 10);

    06X1 6 10}, (c) and (d), mean value of temperatures of different FE models measured at extreme top and bottom (corner) edges of the cubes along the

    temperature gradient direction..

    K.A. Khan et al. / Mechanics of Materials 43 (2011) 608625 611

    http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-
  • 8/10/2019 Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina

    5/18

    thermal properties (CTE, and thermal conductivity) of a par-

    ticle reinforced polymer composite. The model is modified

    for simulating sequentially coupled heat conduction and

    deformation in particulate composites. The model idealizes

    particles in the microstructure as cubes. The cubic particles

    are arranged uniformly in a homogeneous matrix. The RVE

    is defined as a single particle embedded in a cubic matrix.

    Periodic boundary conditions are imposed to the RVE. Due

    to the three-plane symmetry, a unit-cell model thatconsists

    of four particle and matrix subcells is considered. Fig. 1

    t =12s

    t=26s

    t =2s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    (Volume fraction = 30%, 15 particles)

    t =12s

    t=26s

    t=2s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    (Volume fraction = 30%, 30 particles)

    t =12s

    t =26s

    t =2s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    (Volume fraction = 30%, 45 particles)

    t=12s

    t =26s

    t= 2s

    Detail Microstructural Model

    Micromechanical Model

    (Volume fraction = 30%, 15 particles)

    t=12s

    t=26s

    t =2s

    Detail Microstructural Model

    Micromechanical Model

    (Volume fraction = 30%, 30 particles)

    t=12s

    t=26s

    t=2s

    Detail Microstructural Model

    Micromechanical Model

    (Volume fraction = 30%, 45 particles)

    300

    400

    500

    600

    700

    Temperature(K)

    a

    300

    400

    500

    600

    700

    Temperature(K)

    d

    0 2 4 6 8 10

    Distance (mm)0 2 4 6 8 10

    Distance (mm)

    300

    400

    500

    600

    700

    T

    emperature(K)

    b

    300

    400

    500

    600

    700

    T

    emperature(K)

    e

    0 2 4 6 8 10

    Distance (mm)0 2 4 6 8 10

    Distance (mm)

    300

    400

    500

    600

    700

    Temperature(

    K)

    c

    300

    400

    500

    600

    700

    Temperature(K)

    f

    0 2 4 6 8 10

    Distance (mm)0 2 4 6 8 10

    Distance (mm)

    Fig. 5. Comparison of temperature profiles for FE models with the unit cell (micromechanical model) at each integration point (solid line) and the FE

    models with 3D microstructural detail (symbols) for volume fraction of 30%. (a), (b) and (c) are actual values of temperature at top (corner) edge{(X1, 10,

    10); 06X1 6 10}, (d), (e) and (f), mean value of temperatures of different FE models measured at extreme top and bottom(corner) edges of the cubes alongthe temperature gradient direction.

    612 K.A. Khan et al./ Mechanics of Materials 43 (2011) 608625

  • 8/10/2019 Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina

    6/18

    shows the RVE idealization, unit-cell and its integration

    with the FE framework. The choice of the four-cell model

    was primarily done to reduce computational costs; how-

    ever, this turned out to cause extra effort in incorporating

    material symmetry. It is noted that the chosen unit-cell

    model, which is only one-eighth of the full model, was

    due to the threeplanes of symmetrythat allows interchang-

    ing the principal axes in formulating the stress, strain, tem-

    perature gradient and heat flux quantities. As a result, the

    unit-cell response is independent of the loading direction.

    When both particle and matrix are isotropic, the outcome

    of the homogenized micromechanical model does not

    always fulfill the isotropic condition with regard to the

    mechanical response. The micromechanical formulation

    has been discussed in detail elsewhere (Muliana and Kim,

    2007). Perfect bonds are assumed along the subcells inter-

    faces. Thermo-elastic constitutive models with tempera-

    ture dependent material parameters are used for both

    isotropic constituents. Linearized micromechanical rela-

    tions are formulated in terms of incremental average field

    variables, i.e., stress, strain, heat flux and temperature

    gradient, of the subcells. The effective CTE is derived by sat-

    isfying total displacement compatibility and traction conti-

    nuity at the interfaces during thermo-elastic deformations.

    This formulation leads to an effective temperature-depen-

    dent CTE. The effective thermal conductivity is formulated

    by imposing heat flux and temperature continuities at the

    subcells interfaces. This micromechanical model is

    compatible with general displacement based FE software,

    which can be used to perform thermo-mechanical analyses

    of composites structures.

    For linearly thermo-elastic problems, the effective

    stress (rij) and strain (eij) are related through:

    rij Cijklekl aklT T0 1

    or

    eij Sijklrkl aijT T0 2

    whereCijkl and Sijkl are the components of the effective elas-

    tic stiffness and compliance tensors, respectively. The aklare the components of the effective CTE tensor. The param-

    eters TandT0 are the effective current and reference tem-

    peratures, respectively.

    The average heat flux equation for a homogeneous com-

    posite medium is expressed by Fouriers law of heat con-

    duction as:

    qi Kij uj; where uj @T

    @xj3

    where qi and uj are the components of the average heatflux and temperature gradient vectors, respectively. In or-

    der to obtain the temperature profiles during heat conduc-

    tion in the composite, the energy equation needs to be

    solved. For the thermo-elastic case, in the absence of inter-

    nal heat generation and thermo-mechanical coupling ef-

    fect, the energy equation can be written as:

    qcxk_T qi;i i; k 1; 2; 3 4

    whereq

    cx

    kis the effective heat capacity that depends on

    the composition, density and specific heat of the two

    constituents in the composite body. The effective heat

    capacity is obtained using a volume average method.

    The linearized thermo-elastic constitutive equations are

    expressed in an incremental form. A macroscopic strain

    and temperature gradient are known and used as input

    variables to the micromechanical formulation. The current

    microscopic strain and temperature gradient are expressed

    as etij

    etDtij

    detij

    anduti

    utDti

    duti

    , respectively. For

    simplicity, the superscript t indicating current time, will

    be dropped. The macroscopic incremental strain (dekl) islinked to the average incremental strain of each sub-cell

    (deaij ) using the strain concentration tensor (Baijkl), written

    as

    dea

    ij Ba

    ijkldekl 5

    Similarly, the average temperature gradient in each

    subcell duai ) is related through to overall temperaturegradient (d uj) by the concentration tensor (M

    aij ), written

    as

    dua

    i M

    a

    ij duj 6

    where superscript (a) denotes the subcell number, i.e.,a = 1, 2, 3, 4. The strain concentration tensor (Baijkl) is

    t=8s

    t =26s

    t=2s

    Detail Microstructural Model (Volume fraction = 30%)

    15 particles

    30 particles

    t=12st=18s

    45 particles

    t=8s

    t=26s

    t=2s

    Detail Microstructural Model (Volume fraction = 20%)

    20 particles

    40 particles

    t =12s

    t=18s

    300

    400

    500

    600

    700

    Tem

    perature(K)

    a

    0 2 4 6 8 10

    Distance (mm)

    300

    400

    500

    600

    700

    Temperature(K

    )

    b

    0 2 4 6 8 10

    Distance (mm)

    Fig. 6. Mean temperature profiles for FE models with the unit cell

    (micromechanicalmodel) at each integration point (solid line) and the FE

    models with 3D microstructural detail (symbols) for volume fraction of(a) 20% and (b) 30%.

    K.A. Khan et al. / Mechanics of Materials 43 (2011) 608625 613

  • 8/10/2019 Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina

    7/18

    determined by imposing the constitutive relation of each

    subcell and micromechanical relations such that the dis-

    placement compatibility and traction continuity condi-

    tions are satisfied. Two sets of equations are formed. The

    first set of the equations are determined from the strain

    compatibility equations which are given as:

    fReg121

    AM11224

    e1

    e2

    e3

    e4

    8>>>>>:

    9>>>=>>>;

    241

    DM1126

    feg61

    7

    where {Re} is the strain residual vector. The second set of

    the equations satisfies the traction continuity relations

    within subcells:

    fRrg121

    AM21224

    e1

    e2

    e3

    e4

    8>>>>>:

    9>>>=

    >>>;241

    O126

    feg61

    8

    where {Rr} is the stress residual vector. The matrixO is the

    zero matrix and the components of matrix AM1 ;A

    M2 andD

    M1

    can be found elsewhere (Muliana and Kim, 2007). For lin-

    earized elasticity problems, the components of the residual

    vectors are zero and thus the micromechanical relations

    are exactly satisfied. When any of the constituents exhibit

    nonlinear response, imposing linearized micromechanical

    relations leads to non-zero residual vectors {Re} and {Rr}.

    The Newton Raphson iterative method is used to minimize

    these residual vectors. Upon minimizing the residual, the

    strain concentration matrixB(a) is obtained from:

    Ba;t246

    A

    M1

    AM2

    " #12424

    DM1

    O

    " #246

    9

    To formulate the concentration tensor, Maij ;the micro-

    mechanical relations and the constitutive equations for

    heat flux are imposed. This requires forming twelve (12)

    equations based on the temperature and heat flux continu-

    ities at the interface of each subcell as:

    -0.005

    0

    0.005

    0.01

    0.015

    Di

    splacement(mm)

    t=26s

    t=12s

    t=2s

    Steady State Time = 150 seconds

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    Model-5

    Model-6

    a

    -0.005

    0

    0.005

    0.01

    0.015

    Displacement(mm)

    t=26s

    t=12s

    t=2s

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    b

    -0.005

    0

    0.005

    0.01

    0.015

    Displacement(mm)

    t=12s

    t=26s

    t=2s

    Detail Microstructural Model

    Micromechanical Model

    (Volume fraction = 20%, 40 particles)

    d

    -0.005

    0

    0.005

    0.01

    0.015

    Displacement(mm)

    t =12s

    t =26s

    t=2s

    Detail Microstructural Model

    Micromechanical Model

    (Volume fraction = 20%, 20 particles)

    cDetail Microstructural Model(Volume fraction = 20%, 20 particles)

    Detail Microstructural Model (Volume fraction = 20%, 40 particles)

    0 2 4 6 8 10

    Distance (mm)0 2 4 6 8 10

    Distance (mm)

    0 2 4 6 8 10

    Distance (mm)0 2 4 6 8 10

    Distance (mm)

    Fig. 7. Comparison of axial displacements for FE models with the unit cell (micromechanical model) at each integration point (solid line) andthe FE models

    with 3D microstructural detail (symbols) for volume fraction of 20%.(a) and (b) are actual values of displacements at top (corner) edge {(X1, 10, 10);

    06X1 6 10}, (c) and (d), mean value of displacements of different FE models measured at extreme top and bottom (corner) edges of the cubes along thetemperature gradient direction.

    614 K.A. Khan et al./ Mechanics of Materials 43 (2011) 608625

  • 8/10/2019 Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina

    8/18

    fRug91

    AT1912

    du1

    du2

    du3

    du4

    8>>>>>:

    9>>>=>>>;

    121

    DT193

    fd ug31

    10 fRqg31

    AT2312

    du1

    du2

    du3

    du4

    8>>>>>:

    9>>>=>>>;

    121

    O33

    fdug31

    11

    -0.005

    0

    0.005

    0.01

    0.015

    0.02

    Displacement(mm)

    t =26s

    t =12s

    t =2s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    -0.005

    0

    0.005

    0.01

    0.015

    0.02

    Displacement(mm)

    t =26s

    t=12s

    t=2s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    -0.005

    0

    0.005

    0.01

    0.015

    0.02

    Displacement(mm)

    t=26s

    t =12s

    t=2s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    -0.005

    0

    0.005

    0.01

    0.015

    0.02

    Displacement(mm)

    t=12s

    t=26s

    t=2s

    Detail Microstructural Model

    Micromechanical Model

    (Volume fraction = 30%, 15 particles)(Volume fraction = 30%, 15 particles)

    -0.005

    0

    0.005

    0.01

    0.015

    0.02

    Displacement(mm)

    t=12s

    t=26s

    t=2s

    Detail Microstructural Model

    Micromechanical Model

    (Volume fraction = 30%, 30 particles)

    (Volume fraction = 30%, 30 particles)

    -0.005

    0

    0.005

    0.01

    0.015

    0.02

    Displacement(

    mm)

    t=12s

    t=26s

    t=2s

    Detail Microstructural Model

    Micromechanical Model

    (Volume fraction = 30%, 45 particles)

    (Volume fraction = 30%, 45 particles)

    0 2 4 6 8 10

    Distance (mm)0 2 4 6 8 10

    Distance (mm)

    0 2 4 6 8 10

    Distance (mm)0 2 4 6 8 10

    Distance (mm)

    0 2 4 6 8 10

    Distance (mm)0 2 4 6 8 10

    Distance (mm)

    a d

    b e

    c f

    Fig. 8. Comparison of axial displacements for FE models with the unit cell (micromechanical model) at each integration point (solid line) and the FE models

    with 3D microstructural detail (symbols) for volume fraction of 30%. (a), (b) and (c) are actual values of displacements at top (corner) edge {( X1, 10, 10);

    06X1 6 10}, (d), (e) and (f), mean value of displacements of different FE models measured at extreme top andbottom(corner) edges of the cubes along thetemperature gradient direction.

    K.A. Khan et al. / Mechanics of Materials 43 (2011) 608625 615

  • 8/10/2019 Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina

    9/18

    where {R/} and {Rq} are the temperature gradient and heat

    flux residual vectors, respectively. By substituting Eq. (6)

    into Eqs.(10) and (11), theM(a) matrix can be determined

    as:

    Ma121

    A

    T1

    AT2

    " #11212

    DT1

    O

    " #121

    12

    where AT1, A

    T2 andD

    T1 can be found elsewhere (Khan and

    Muliana, 2009). The effective (homogenized) stress and

    effective tangent stiffness matrix can be obtained from

    the following equations:

    rij1

    V

    X4a1

    VaCa

    ijklBa

    klrsers a

    a

    kl DT Cijklekl aklDT 13

    Cijkl1

    V

    X4a1

    VaCa

    ijklB

    a

    klrs 14

    whereVandV(a) are the total volume of the unit-cell mod-

    el and sub-cell volume, respectively. From Eq. (13), theeffective CTE, (aij) can be obtained and written as:

    aijC1ijklV

    X4a1

    VaCa

    klmnaamn 15

    Using the heat conduction equation for each sub-cell

    and the effective heat flux relation (Eq. (3)), the tangent

    effective thermal conductivity matrix of the composite

    can be expressed as:

    Kik 1

    V

    X4a1

    VaKa

    ij Ma

    jk 16

    3. A multi-scale model for computing the response of a

    homogenized composite

    The micro-mechanical model of Section 2 has been inte-

    grated into a multi-scale FE framework in order to com-

    pute the field variables of a real-size composite.

    Boundary conditions are imposed on the real-size compos-

    ite model and initial values of the field variables are as-

    sumed. Local effective properties (thermal, mechanical)

    are computed at each integration point using the micro-

    mechanical model. Computations of the effective proper-

    ties are based on the assumption that the each integration

    point is associated with a much smaller volume than that

    of the whole composite. As a result, we assumed that each

    volume associated with each integration point was at a

    uniform temperature. Therefore, at the micro-scale, both

    the matrix and reinforcement were assumed to be at the

    same temperature. This allows determining temperature-

    dependent thermal conductivities for the particle and ma-

    trix constituents and calculating the effective thermal con-

    ductivities of the composite at that instant of time in each

    material point. Therefore, in this context, the periodic

    boundary conditions are fully justified for obtaining the

    effective properties. At macroscopic scale during the tran-

    sient heat conduction, the temperature-dependent proper-

    ties of the constituents lead to spatially dependent ETC.

    To simulate the effective thermo-elastic response of the

    particulate composite, the micromechanical model is inte-

    grated with the ABAQUS/standard FE code. At each integra-

    tion point in the FE mesh, the user subroutine UMATH is

    first called to evaluate the effective thermal conductivity,

    heat fluxes, and temperature gradient for solving the equa-

    tion that governs the conduction of heat in the composite

    body. The temperature distributions obtained from heat

    transfer analyses at various instants of time in the compos-

    ites are used as input transient thermal load to determine

    the thermo-elastic deformation in the composite body.

    UMAT and UEXPAN subroutines were used to evaluate

    the effective mechanical response and CTE, respectively.

    Fig. 1illustrates the whole multi-scale framework.

    4. Three dimensional FE models of particulate

    composites

    4.1. General methodology for evaluating the performance of

    the multi-scale framework

    The multi-scale model of Section 3 can be considered as

    an approximation to a complicated thermo-mechanical

    -0.005

    0

    0.005

    0.01

    0.015

    Displacement(m

    m)

    t =12s

    t=26s

    t=2s

    Detail Microstructural Model (Volume fraction = 30%)

    15 particles

    b

    30 particles

    45 particles

    -0.005

    0

    0.005

    0.01

    0.015

    Disp

    lacement(mm)

    t=12s

    t=26s

    t=2s

    Detail Microstructural Model (Volume fraction = 20%)

    20 particles

    a

    40 particles

    0 2 4 6 8 10

    Distance (mm)

    0 2 4 6 8 10

    Distance (mm)

    Fig. 9. Mean displacement profiles for FE models with the unit cell

    (micromechanical model) at each integration point (solid line) and the FE

    models with 3D microstructural detail (symbols) for volume fraction of(a) 20% and (b) 30%.

    616 K.A. Khan et al./ Mechanics of Materials 43 (2011) 608625

    http://-/?-http://-/?-http://-/?-http://-/?-
  • 8/10/2019 Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina

    10/18

    problem: the problem of computing the temperature and

    stresses inside a heterogeneous material subjected to both

    thermal and mechanical loads. In order to evaluate the

    reliability of the micro-mechanical model, it is therefore

    required to compare its predictions against the numeri-

    cally exact solution. The objective of this section is to

    generate this so called numerically exact solution.

    When dealing with effective properties, one of the key

    issues is the definition of the RVE. A RVE can be seen as a

    volume of material having the same behavior as any larger

    volume of the same material. The size of the RVE is mea-

    sured in terms of inhomogeneities it contains (e.g. the

    number of particles meshed, like 5, 10, 15, etc., particles).

    One of the techniques used for obtaining the RVE size is

    to use numerical homogenization based on FE. The method

    consists in generating many FE meshes of the composite

    microstructure with a fixed number of reinforcements

    (10, 15, 50, etc.). Since the particles are distributed accord-

    ing to a statistical distribution, each mesh, or realization,

    will be different. Therefore, each realization should lead

    -400

    -200

    0

    200

    400

    AxialStresses(

    11

    )MPa

    t=2s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    Model-5

    Model-6

    (Volume fraction = 20%, 20 particles)

    -400

    -200

    0

    200

    400

    AxialStresses(

    11

    )MPa

    t=12s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    Model-5

    Model-6

    (Volume fraction = 20%, 20 particles)

    -400

    -200

    0

    200

    400

    AxialStresses(

    11

    )MPa

    t=26s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    Model-5

    Model-6

    (Volume fraction = 20%, 20 particles)

    -400

    -200

    0

    200

    400

    AxialStresses(

    11

    )MPa

    t

    =2s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    (Volume fraction = 20%, 40 particles)

    -400

    -200

    0

    200

    400

    AxialStresses(

    11

    )MPa

    t=12s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    (Volume fraction = 20%, 40 particles)

    -400

    -200

    0

    200

    400

    AxialStresses(

    11

    )MP

    a

    t=26s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3Model-4

    (Volume fraction = 20%, 40 particles)

    0 2 4 6 8 10

    Distance (mm)0 2 4 6 8 10

    Distance (mm)

    0 2 4 6 8 10

    Distance (mm)

    0 2 4 6 8 10

    Distance (mm)

    0 2 4 6 8 10

    Distance (mm)

    0 2 4 6 8 10

    Distance (mm)

    a d

    b e

    c f

    Fig. 10. Axial thermal stresses for FE models with the unit cell (micromechanical model) at each integration point (solid line) and the FE models with 3Dmicrostructural detail (symbols) for volume fraction of 20% at different times.

    K.A. Khan et al. / Mechanics of Materials 43 (2011) 608625 617

  • 8/10/2019 Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina

    11/18

    to different (within certain accuracy) effective properties.

    For the same number of reinforcements and load history,

    the effective responses are computed for each realization

    and then averaged. Computing a confidence interval (for

    example a two-tail 95% confidence interval) on this data

    could give an estimation of the composites effective prop-

    erties and its precision for a given number of reinforce-

    ments. The relative precision (for example the Youngs

    modulus is estimated to bexy%) can be adjusted by vary-

    ing the number of realizations. If many realizations are

    performed, the confidence interval can be adjusted to the

    desired width. Kanit et al. (2003) mentioned that for

    microstructures containing numerous reinforcements,

    smaller numbers of realizations are required to estimate

    0 2 4 6 8 10

    Distance (mm)

    -400

    -300

    -200

    -100

    0

    100

    200

    300

    400

    AxialStresses(

    11

    )MPa

    t=2s

    Detail Microstructural Model (Volume fraction = 20%, 20 particles)

    Micromechanical Model

    -400

    -300

    -200

    -100

    0

    100

    200

    300

    400

    Axial

    Stresses(

    11

    )MPa

    t =12s

    Detail Microstructural Model (Volume fraction = 20%, 20 particles)

    Micromechanical Model

    -400

    -300

    -200

    -100

    0

    100

    200

    300

    400

    AxialStresses(

    11

    )MPa

    t=26s

    Detail Microstructural Model (Volume fraction = 20%, 20 particles)

    Micromechanical Model

    0 2 4 6 8 10

    Distance (mm)

    -400

    -300

    -200

    -100

    0

    100

    200

    300

    400

    AxialStresses(

    11

    )MPa

    t=2s

    Detail Microstructural Model (Volume fraction = 20%, 40 particles)

    Micromechanical Model

    -400

    -300

    -200

    -100

    0

    100

    200

    300

    400

    AxialStresses(

    11

    )MPa

    t=12s

    Detail Microstructural Model (Volume fraction = 20%, 40 particles)

    Micromechanical Model

    -400

    -300

    -200

    -100

    0

    100

    200

    300

    400

    AxialStresses(

    11

    )MPa

    t=26s

    Detail Microstructural Model (Volume fraction = 20%, 40 particles)

    Micromechanical Model

    0 2 4 6 8 10

    Distance (mm)0 2 4 6 8 10

    Distance (mm)

    0 2 4 6 8 10

    Distance (mm)0 2 4 6 8 10

    Distance (mm)

    a d

    b e

    c f

    Fig. 11. Axial thermal stresses for FE models with the unit cell (micromechanical model) at each integration point (solid line) and the FE models with 3D

    microstructural detail (symbols) for volume fraction of 20% at different times with C.I. of 95%. (a), (b) and (c) are mean value of stresses of different FE

    models with 20 particles and (d), (e) and (f) with 40 particles, measured at extreme top and bottom (corner) edges of the cubes along the temperaturegradient direction.

    618 K.A. Khan et al./ Mechanics of Materials 43 (2011) 608625

  • 8/10/2019 Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina

    12/18

    the wanted overall property within desired precision. For

    relatively small numbers of particles, the homogenized

    properties vary statistically until a certain number of par-

    ticles are meshed. The number of particles after which

    the effective response does not change anymore is called

    the representative volume element.

    In this work, emphasis was put on field variables distri-

    butions rather than on effective properties. For illustration

    purposes, consider a macroscopic component made of a

    nonlinear composite constituted of many reinforcements,

    as shown in Fig. 2(a). Consider that the component is di-

    vided into many sub-regions of length l(Fig. 2(b)). For each

    (Volume fraction = 30%,, 15 particles)

    -400

    -200

    0

    200

    400

    AxialStresses(

    11

    )MPa

    t=2s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    -400

    -200

    0

    200

    400

    AxialStresses(

    11

    )MPa

    t=2s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    (Volume fraction = 30%,, 30 particles)

    -400

    -200

    0

    200

    400

    AxialStresses(

    11

    )MPa

    t=2s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3Model-4

    (Volume fraction = 30%, 45 particles)

    -400

    -200

    0

    200

    400

    AxialStresses(

    11

    )MPa

    (Volume fraction = 30%,, 15 particles)

    t=12s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    -400

    -200

    0

    200

    400

    AxialStresses(

    11

    )MPa

    t =12s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    (Volume fraction = 30%,, 30 particles)

    -400

    -200

    0

    200

    400

    AxialStresses(

    11

    )MP

    a

    t=12s

    Detail Microstructural Model

    Micromechanical Model

    Model-1

    Model-2

    Model-3Model-4

    (Volume fraction = 30%, 45 particles)

    0 2 4 6 8 10

    Distance (mm)0 2 4 6 8 10

    Distance (mm)

    0 2 4 6 8 10

    Distance (mm)0 2 4 6 8 10

    Distance (mm)

    0 2 4 6 8 10

    Distance (mm)0 2 4 6 8 10

    Distance (mm)

    a b

    c d

    e f

    Fig. 12. Axial thermal stresses for FE models with the unit cell (micromechanical model) at each integration point (solid line) and the FE models with 3Dmicrostructural detail (symbols) for volume fraction of 30% at different times.

    K.A. Khan et al. / Mechanics of Materials 43 (2011) 608625 619

  • 8/10/2019 Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina

    13/18

    sub-region, assume that the RVE size has been reached so

    that effective properties could be calculated (Fig. 2(c)).

    Therefore, the composite component can be divided into a

    number of homogeneous sub-regions, as was done in Sec-

    tion 3. If the component is subjected to an external loading

    (thermal, mechanical, etc.), the effective properties of each

    sub-region can be iteratively computed in order to estimate

    the field variable distribution into the homogenized com-

    ponent. It is recalled that each effective property is known

    within a given accuracy and different realization of the

    same composite component should lead to different field

    variables distributions. As l decreases and the number of

    sub-regions increases, the difference in the field variables

    distributions between each realization should decrease un-

    til an acceptable scatter is reached. Since each sub-region is

    constituted of a finite number of reinforcements, there is a

    finite number of reinforcements inside the composite com-

    ponent for which the homogenized response from one real-

    ization to the other stays within a given tolerance. In this

    work, the meshed composites are considered as the macro-

    scopic components mentioned above. The simulations

    performed in Section5 aim at determining the number of

    reinforcements required for obtaining field variables distri-

    butions corresponding to the homogenized component, as

    well as the distribution themselves.

    4.2. Generation of the FE meshes used in this study

    The FE meshes used in this study are those ofBarello

    and Lvesque (2008). Their generation is recalled here.

    The detailed composite consists of randomly distributed

    identical spherical particles reinforced matrix. The micro-

    structures were generated using the Random Sequential

    Adsorption Algorithm (Segurado and LLorca, 2002). Thealgorithm consists in generating the position of a first

    spherical particle center into a cubic volume using a uni-

    form random number generator. Then, the center position

    of a second sphere is generated. If the distance between the

    centers of the first two particles and the distance from the

    particle center from the cubes faces is smaller than a pre-

    set value, then the second particle is rejected and a new

    center position is generated until the minimum distance

    criterion mentioned above is met. The other particles are

    sequentially added, following the same process where

    the distance criteria are checked with all the existing par-

    ticles. The particles are added until the desired volume

    fraction is reached.

    The particles were allowed to cut the edges and the

    faces of the cube. When this happened, the particles were

    completed periodically on the corresponding faces and

    edges. The realizations thus obtained were therefore peri-

    odic and always had an integer number of complete

    spheres. The minimum distance between two particles

    centers was set to 2.07r, where ris the particle radius while

    the minimum distance from a particle center to a cubes

    face was set to 0.1r. These distance criteria were obtained

    through trial and error with the meshing software until

    elements of acceptable aspect ratios were obtained.

    A Matlab program was used for generating the particle

    centers. This program wrote an ANSYS command file for

    generating the FE mesh of the microstructure. Finally, a

    Matlab program was used for converting the ANSYS

    model to ABAQUS. The mesh consisted of 10-noded

    tetrahedra.

    5. Simulation execution and results

    5.1. Simulations performed in this study

    For both the multi-scale framework and the detailed

    models of Section 4, cubic models of dimensions10 10 10 mm were used.Fig. 3shows these models as

    well as the axes used for defining the boundary conditions

    below. The studied composite is a ZrO2 matrix reinforced by

    randomly distributed Ti-6Al-4V spherical particles. The

    heterogeneous composites directly incorporate nonlinear

    thermo-elastic behaviors for the particle and matrix re-

    gions. The thermal as well as the mechanical properties

    used in the simulations can be found in Khan and Muliana

    (2010)and are given inTable 1. Two volume fractions of

    reinforcements were studied, namely 20% and 30%. For

    the detailed FE meshes, cubes containing 15, 20, 30, 40

    and 45 spheres were generated. The transient thermal anal-

    ysis consisted in a problem where a composite was initiallyat 300 K, except for one face that was at 600 K. This tran-

    sient heat transfer problem was solved until a steady state

    was reached. A uniformstress of 10 MPa was applied on the

    face that was at 600 K in order to simulate effective tran-

    sient thermal stresses. The models were subjected to the

    following mixed uniform boundary conditions:

    u10;x2;x3; t 0:0; 0 6x2 6 10; 0 6x3 6 10; tP 0

    u2x1;0;x3; t 0:0; 0 6x1 6 10; 0 6x3 6 10; tP 0

    u3x1;x2;0; t 0:0; 0 6x1 6 10; 0 6x2 6 10; tP 0

    p110;x2;x3; t 10:0 MPa; 0 6x2 6 10; 0 6x3 6 10; tP 0

    p2x1;10;x3; t 0:0 MPa; 0 6x1 6 10; 0 6x3 6 10; tP 0

    p3x1;x2;10; t 0:0 MPa; 0 6x1 6 10; 0 6x2 6 10; tP 0

    18

    Tx1;x2;x3; 0 300 K; 0 6x1 6 10; 0 6 x2 6 10; 0 6 x3 6 10

    T10;x2;x3; t 600 K; 0 6x2 6 10; 0 6 x3 6 10; tP 0

    @Tx1; 0;x3; t

    @x2

    @Tx1; 10;x3; t

    @x2 0:0; 0 6x1 6 10; 0 6 x3 6 10; tP 0

    @Tx1;x2; 0; t

    @x3

    @Tx1;x2; 10; t

    @x30:0; 0 6x1 6 10; 0 6 x2 6 10; tP 0

    17

    620 K.A. Khan et al./ Mechanics of Materials 43 (2011) 608625

    http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-
  • 8/10/2019 Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina

    14/18

    whereui and pi (i= 1, 2, 3) are the components of the dis-

    placements and the surface tractions, respectively. It is re-

    called that the set of boundary conditions affects the size of

    the representative volume element, and hence, that of the

    representative component. It is therefore expected that the

    same component subjected to different boundary condi-

    tions requires a different number of heterogeneities in or-

    der to lead to converged field variables distributions.

    -400

    -300

    -200

    -100

    0

    100

    200

    300

    400

    AxialStresses(

    11

    )MPa

    t=2s

    Detail Microstructural Model(Volume fraction = 30%, 45 particles)

    Micromechanical Model

    -400

    -300

    -200

    -100

    0

    100

    200

    300

    400

    AxialStresses(

    11

    )MPa

    t=12s

    Detail Microstructural Model(Volume fraction = 30%, 45 particles)

    Micromechanical Model

    -400

    -300

    -200

    -100

    0

    100

    200

    300

    AxialStresses(

    11

    )MPa

    t=2s

    Detail Microstructural Model(Volume fraction = 30%, 15 particles)

    Micromechanical Model

    -400

    -300

    -200

    -100

    0

    100

    200

    300

    400

    400

    AxialStresses(

    11

    )MPa

    t=2s

    Detail Microstructural Model(Volume fraction = 30%, 30 particles)

    Micromechanical Model

    -400

    -300

    -200

    -100

    0

    100

    200

    300

    AxialStresses(

    11

    )MPa

    t =12s

    Detail Microstructural Model(Volume fraction = 30%, 15 particles)

    Micromechanical Model

    -400

    -300

    -200

    -100

    0

    100

    200

    300

    400

    400

    AxialStresses(

    11

    )MPa

    t=12s

    Detail Microstructural Model(Volume fraction = 30%, 30 particles)

    Micromechanical Model

    0 2 4 6 8 10Distance (mm) 0 2 4 6 8 10Distance (mm)

    0 2 4 6 8 10

    Distance (mm)0 2 4 6 8 10

    Distance (mm)

    0 2 4 6 8 10

    Distance (mm)0 2 4 6 8 10

    Distance (mm)

    a b

    c d

    e f

    Fig. 13. Axial thermal stresses for FE models with the unit cell (micromechanical model) at each integration point (solid line) and the FE models with 3D

    microstructural detail (symbols) for volume fraction of 30% at different times with C.I. of 95%. (a), (b) and (c) are mean value of stresses of different FE

    models with 15 particles and (d), (e) and (f) with 30 particles, (g), (h) and (i) with 45 particles, measured at extreme top and bottom (corner) edges of thecubes along the temperature gradient direction.

    K.A. Khan et al. / Mechanics of Materials 43 (2011) 608625 621

  • 8/10/2019 Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina

    15/18

    Due to the boundary conditions, the field variables

    distribution on the four cube segments oriented along x1(see Fig. 3) should be identical for a large number of

    spheres. For the detailed models, the field variables were

    extracted at identical x1 coordinates and then averaged.

    At each x1 coordinate, 95% confidence intervals on the

    mean value were computed. Finally, the averaged distribu-

    tions of the detailed models were compared to the

    distributions of the multi-scale model.

    In the following subsections, distributions of the field

    variables predicted from the multi-scale framework are

    compared to those of the detailed FE meshes of Section4.

    5.2. Temperature distribution

    Fig. 4(a) and (b) show the temperature distributions ob-

    tained from the homogenized model as well as from the

    heterogeneous composite reinforced by 20% of Ti6Al4V

    particles for model sizes of 20 and 40 particles, respec-

    tively, for different times.Fig. 4(c) and (d) show the mean

    responses of the various realizations, along with 95% con-

    fidence intervals for models of 20 and 40 particles, respec-

    tively. For the 20 particle model, the largest width of the

    confidence interval is 1.74% of the mean value while it is

    of 3.42% for the 40 particles model. The width of the confi-

    dence interval decreases as the time increases. Fig. 5(a)(f)

    show temperature profiles from similar type of analyses

    but for a composite reinforced by 30% of Ti6Al4V parti-

    cles and for models containing 15, 30 and 45 particles.

    The largest widths of the confidence intervals are of

    3.25%, 1.88% and 3.7% for the 15, 30 and 45 particle models,

    respectively. It is interesting to note that the confidence

    interval widths did not decrease as the number of particles

    increased, as would have been expected. The reasons for

    this phenomenon are unclear at this time. Conducting

    more realizations and/or simulating larger number of par-

    ticles might lead to the expected tendency and this phe-

    nomenon might be of statistical nature.

    Fig. 6(a) shows the mean temperature curves for a com-

    posite with 20% of Ti6Al4V particle volume content and

    for models containing 20 and 40 particles. Considering

    their relatively narrow confidence intervals, it can be seen

    that the RVE has been reached for these microstructures

    since both the 20 and 40 particle models lead to very sim-

    ilar results. Fig. 6(b) shows the mean temperature curves

    for the 15, 30 and 45 particle models for composites with

    30% particle volume content. It can also be concluded that

    the RVE size has been reached and overcome. Figs. 4 and 5

    show that the micromechanical model predicts fairly well

    the temperature profiles for the range of material proper-

    ties simulated.

    5.3. Displacement distribution

    Fig. 7(a) and (b) show the displacement distributions

    obtained from the homogeneous and heterogeneous mod-

    els for a composite containing 20% of Ti6Al4V particles

    for models having 20 and 40 particles, respectively.

    Fig. 7(c) and (d) show the average response of the various

    realizations, along with 95% confidence intervals on the

    mean value for models containing 20 and 40 particles,

    respectively. For the 20 particles model, the largest width

    of the confidence interval is 96% of the mean value while

    it is of 52% for the 40 particles model.

    Fig. 8(a)(f) show displacements from similar analyses

    butfor a composite reinforced by 30% of Ti6Al4V particles

    for models containing 15, 30 and 45 particles. The largest

    -400

    -300

    -200

    -100

    0

    100

    200

    300

    400

    AxialStresses(

    11

    )MPa

    t =2s

    Detail Microstructural Model (Volume fraction = 20%)

    40 particles

    20 particles

    -400

    -300

    -200

    -100

    0

    100

    200

    300

    400

    Axial

    Stresses(

    11

    )MPa

    t =12s

    Detail Microstructural Model (Volume fraction = 20%)

    40 particles

    20 particles

    -400

    -300

    -200

    -100

    0

    100

    200

    300

    400

    AxialStresses(

    11

    )MPa

    t=26s

    Detail Microstructural Model (Volume fraction = 20%)

    40 particles

    20 particles

    0 2 4 6 8 10

    Distance (mm)

    0 2 4 6 8 10

    Distance (mm)

    0 2 4 6 8 10

    Distance (mm)

    a

    b

    c

    Fig. 14. Mean stress profiles for FE models with the unit cell (microme-

    chanical model) at each integration point (solid line) and the FE models

    with 3D microstructural detail (symbols) for volume fraction of (a) 20% atdifferent times.

    622 K.A. Khan et al./ Mechanics of Materials 43 (2011) 608625

    http://-/?-http://-/?-
  • 8/10/2019 Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina

    16/18

    widths of the confidence intervals were of 30%, 65% and

    143% for the 15, 30 and 45 particle models, respectively.

    Fig. 9(a) shows the superimposed curves for 20 and 40

    particles model for a sphere volume fraction of 20%. Consid-

    ering thewidth of theconfidence intervals ofFig. 7, itcan be

    seen that the average responses are reasonably close and

    hence that the RVE has been reached. These observations

    allow to conclude that the micro-mechanical model pre-

    dicts relatively well the macroscopic response of the

    composite for this specific microstructure. Fig. 9(b) shows

    the superimposed curves for 15, 30 and 45 particles model

    for a sphere volume fraction of 30%. For times t= 12 s and

    t= 26 s (seconds), the huge widths of the confidence inter-

    vals (see Fig. 8(d)(f)) do not allow to conclude whether the

    size of the RVE has been reached or not within a reasonable

    precision and hence render these RVE analyses meaning-

    less. However, fort= 2 s, the confidence intervals are rela-

    tively narrow and it is possible to conclude fromFig. 9(b)

    that for this time, the RVE size has been reached. For

    t= 2 s, it seems that the micro-mechanical model predicts

    relatively well the homogenized displacement distribution,

    although it is less accurate than the microstructure having

    20% of reinforcements. Moreover, it seems that performing

    simulations with more than 45 reinforcements might lead

    to narrower confidence intervals for a better assessment

    of the RVE size. Finally, it can be observed that the micro-

    mechanical model predicts with more accuracy the tem-

    perature distribution than the displacement field, for the

    cases studied here.

    5.4. Thermal stresses distributions

    The contrast in the CTEs values of the constituents and

    high temperature gradient are the main cause for the gen-

    eration of high thermal stresses.Figs. 1013show the var-

    iation of thermal stresses for spheres volume fractions of

    20% and 30% at different times for the homogenized and

    heterogeneous composites, respectively. For all the figures,

    except for t= 2 s over a certain distance, the width of the

    confidence intervals cannot be used to determine if the

    RVE size has been reached with a high degree of confi-

    dence. For t= 2 s, it seems that the micro-mechanical

    TOP ELEMENTS

    BOTTOM ELEMENTS

    -400

    -200

    0

    200

    400

    AxialStresses(

    11

    )MPa

    t =26s

    a

    TOP ELEMENTS

    BOTTOM ELEMENTS

    -400

    -200

    0

    200

    400

    AxialStresses(

    11

    )MPa

    t=26s

    Detail Microstructural Model

    Model-1

    b

    Detail Microstructural Model

    Model-2

    0 2 4 6 8 10

    Distance (mm)0 2 4 6 8 10

    Distance (mm)

    Fig. 15. Axial thermal stresses for FE models with 3D microstructural detail for volume fraction of 20% att= 26 s. (a) and (b) are actual values of stresses at

    top (corner) edge {(X1, 10, 10); 06X1 6 10} for model-2 and model-1, respectively.

    K.A. Khan et al. / Mechanics of Materials 43 (2011) 608625 623

  • 8/10/2019 Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina

    17/18

    model can predict reasonably well the thermal stresses

    distribution. However, for all the other cases, the results

    suggest that the micromechanical model is not capable of

    capturing the thermal stresses with good accuracy. To cor-

    roborate the above-mentioned hypothesis, the mean val-

    ues of axial thermal stress at different times are shown

    in Fig. 14(a)(c) for the microstructures having 20% of rein-

    forcements. More realizations, and possibly with models

    having more reinforcements, are required for confirming

    this hypothesis with more confidence.

    The localized stresses are found in some models which

    are generally due to the specific micro-geometrical fea-

    tures and the high fluctuation about the mean stress pro-

    file is due to the presence or absence of the particle

    along the profile where the stresses are computed. These

    high compressive stresses are found in those matrix ele-

    ments which surround the particle region. In this study

    the thermal expansion of the particle is higher than the

    surrounding matrix at all temperatures. Therefore, during

    transient heat conduction the free expansion of the particle

    is constrained by the surrounding matrix elements. The

    larger CTE mismatch of particle/matrix elements results

    in such high values of compressive stresses in the neigh-

    boring elements of particle.

    For example, consider model-2 shown in Fig. 15(a) for

    which the high compressive stresses are found in the ma-

    trix region that restraints the free expansion of two parti-

    cle regions. Similar behavior is found for the elements

    neighboring the particle region approximately at 2.5 mm

    and 8.3 mm, respectively. For the same temperature differ-

    ence the particle expands more than the matrix but the

    surrounding restraints provided by the matrix elements

    are the main cause for the generation of such high values

    of compressive stresses. The same description is applicable

    to other models where such micro-geometrical features

    are found; for example, see belowFig. 15(b) of model-1.

    5.5. Effective displacement

    The effective displacement, (d1), is defined asd1 e11 L,wheree11 is the volume average of the strains in x1 direc-tion and L is the length of the cube. For both the multi-scale

    and the detailed models, (d1) was computed at the face of

    loading (BDHF inFig. 3(c)) for composites having a sphere

    volume fraction of 20% and 30%, respectively. The

    d1 as afunction of time is plotted inFig. 16(a) and (b). The mean

    values of effective displacements (along with 95%

    confidence intervals) for heterogeneous composite models

    0

    0.005

    0.01

    0.015

    0.02

    Displacements(mm

    )

    Displacements(mm)

    Detail Microstructural Model(Volume fraction = 30%, 30 particles)

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    0

    0.005

    0.01

    0.015

    0.02

    Steady State Time = 150 seconds

    Detail Microstructural Model(Volume fraction = 20%, 20 particles)

    Micromechanical Model

    Model-1

    Model-2

    Model-3

    Model-4

    Model-5

    Model-6

    300K

    600KCenter line

    0

    0.005

    0.01

    0.015

    0.02

    Displacements(mm)

    Detail Microstructural Model(Volume fraction = 30%, 30 particles)

    Micromechanical Model

    0 5 10 15 20 25 30

    Time (seconds)

    0 5 10 15 20 25 30

    Time (seconds)

    0 5 10 15 20 25 30

    Time (seconds)

    0 5 10 15 20 25 30

    Time (seconds)

    0

    0.005

    0.01

    0.015

    0.02

    Displa

    cements(mm)

    Detail Microstructural Model(Volume fraction = 20%, 20 particles)

    Micromechanical Model

    a c

    b d

    Fig. 16. Effective Axial displacements for FE models with the unit cell (micromechanical model) at each integration point (solid line) and the FE models

    with 3D microstructural detail (symbols) for volume fraction of (a) 20% and (b) 30%. Mean values of effective displacements for (c) 20% and (d) 30% with C.Iof 95%.

    624 K.A. Khan et al./ Mechanics of Materials 43 (2011) 608625

  • 8/10/2019 Mechanics of Materials Volume 43 issue 10 2011 [doi 10.1016_j.mechmat.2011.06.013] Kamran A. Khan; Romina

    18/18

    having 20% and 30% reinforcement particles are shown in

    Fig. 16(c) and (d). Agreement of these results corroborate

    that the present micromechanical formulation is suitable

    for the prediction of effective responses of composites

    through the incorporation of a nonlinear thermo-elastic

    constitutive material model.

    6. Summary

    The transient responses of the homogenized and heter-

    ogeneous composites due to coupled heat conduction and

    mechanical loading have been studied. For the tempera-

    ture response, the RVE size was reached for both models

    having 20% and 30% reinforcements. It was found that

    the temperature distribution is relatively well predicted

    with the multi-scale model. The width of the confidence

    intervals for the displacements were larger than those for

    the temperature but allowed nevertheless to conclude that

    the multi-scale framework can also predict with a reason-

    able accuracy the displacement field inside the composite.

    The RVE size was not reached for the thermal stresses andit is not possible to conclude that the multi-scale frame-

    work is suitable for representing accurately these stresses.

    Larger RVEs or many more simulations for the same RVE

    sizes would be required in order to narrow the confidence

    intervals. However, the mean results obtained are encour-

    aging and running more simulations might reveal that the

    multi-scale framework is also suitable for evaluating the

    thermal stresses. Finally, the multi-scale model reasonably

    predicts the effective displacement. Therefore, the main

    contribution of this work was the development and the

    partial validation of a multi-scale framework that allows

    predicting the field variables of a temperature dependent

    thermo-mechanical problem.

    Acknowledgements

    This research is sponsored by the Air Force Office of Sci-

    entific Research (AFOSR) under Grant No. FA 9550-10-1-

    0002. We also thank the Texas A&M Supercomputing Facil-

    ity (http://sc.tamu.edu/) for providing computing resources

    useful in conducting the research reported in this paper.

    References

    Aboudi, J., 1991. Mechanics of Composite Materials. A Unified

    Micromechanical Approach. Elsevier, New York.

    Aboudi, J., 2008. Thermomechanically coupled micromechanical analysisof multiphase composites. J. Eng. Math. 61 (24), 111132.

    Auriault, J.L., 1983. Effective macroscopic description for heat conduction

    in periodic composites. Int. J. Heat Mass Transfer 26, 861869.

    Barello, R.B., Lvesque, M., 2008. Comparison between relaxation spectra

    obtained from homogenization model and FE simulation. Int. J. Solids

    Struct. 45, 850867.

    Beran, M., 1965. Nuovo Cimento 38, 771.

    Budiansky, B., 1965. On the elastic moduli of some heterogeneous

    materials. J. Mech. Phys. Solids 13, 223227.

    Chen, H.S., Acrivos, A., 1978. The effective elastic moduli of composite

    materials containing spherical inclusions at non-dilute

    concentrations. Int. J. Solids Struct. 14 (5), 349364.

    Christensen, R.M., Lo, K.H., 1979. Solutions for effective properties of

    composite materials. J. Mech. Phys. Solids 27, 315.

    Christensen, R.M., Lo, K.H., 1986. J. Mech. Phys. Solids 34, 639.

    Davis, R.H., 1986. The effective thermal conductivity of a composite

    material with spherical inclusions. Int. J. Thermophys. 7 (3), 609620.

    Hashin, Z., 1962. The elastic moduli of heterogeneous materials. In: Trans.

    ASME 84 (E) Ser. J. Appl. Mech., vol. 29. pp. 143150.

    Hashin, Z., Shtrikman, S., 1962. J. Mech. Phys. Solids 10, 343352.

    Hasselman, D.P.H., Johnson, L.F., 1987. Effective thermal conductivity of

    composites with interfacial thermal barrier resistance. J. Compos.

    Mater. 21, 508515.

    Hill, R., 1965. A self consistent mechanics of composite materials. J. Mech.

    Phys. Solids 13, 213222.

    Jeffrey, D.J., 1973. Conduction through a random suspension of spheres.

    Proceedings of the Royal Society of London Series A. Mathematical

    and Physical Sciences 335 (1602), 355367.Jiang, M., Jasiuk, I., Ostoja-Starzewski, M., 2002. Apparent thermal

    conductivity of periodic two-dimensional composites. Comput.

    Mater. Sci. 25, 329338.

    Kanit, T., Forest, S., Galliet, I., Mounoury, V., Jeulin, D., 2003.

    Determination of the size of the representative volume element for

    random composites: statistical and numerical approach. Int. J. Solids

    Struct. 40, 36473679.

    Khan, K.A., Muliana, A.H., 2010. Effective thermal properties of

    viscoelastic composites having field-dependent constituent

    properties. Acta Mech. 209 (12), 153178.

    Khan, K.A., Muliana, A.H., 2009. A multi-scale model for coupled heat

    conduction and deformations of viscoelastic functionally graded

    materials. Compos Part B: Eng 40 (6), 511521.

    Kwon, Y.W., Kim, C., 1998. Micromechanical model for thermal analysis of

    particulate and fibrous composites. J. Therm. Stress. 21, 2139.

    Lvesque, M., Barello, R.B., 2009. On the homogenization of nonlinearly

    viscoelastic composite materials. In: ASME International MechanicalEngineering Congress and Exposition, Proceedings, vol. 12. pp. 303304.

    Lvesque, M., Derrien, K., Mishnaevski, L., Baptiste, D., Gilchrist, M.D.,

    2004. A micromechanical model for nonlinear viscoelastic particle

    reinforced polymeric composite materials: undamaged state.

    Composites A 35, 905913.

    Lvesque, M., Derrien, K., Baptiste, D., Gilchrist, M.D., 2008. On the

    development and parameter identification of Schapery-type

    constitutive theories. Mech. Time-Depend. Mat. 12 (2), 95127.

    Maxwell, J.C., 1954. A Treatise on Elec. and Magnetism, third ed. Dover,

    New York.

    Matt, C.F., Cruz, M.A.E., 2002. Application of a multiscale finite-element

    approach to calculate the effective thermal conductivity of particulate

    media. Comput. Appl. Math. 21, 429460.

    McLaughlin, R., 1977. A study of the differential scheme for composite

    materials. Int. J. Eng. Sci. 15, 237.

    Meijer, G., Ellyin, F., Xia, Z., 2000. Aspects of residual thermal stress/strain in

    particlereinforcedmetalmatrixcomposites. Composites,Part B 31,2937.Mori, T., Tanaka, K., 1973. Average stress in matrixand average elasticenergy

    of materials with misfitting inclusions. Acta Metall. 231, 571574.

    Muliana, A.H., Kim, J.S., 2007. A concurrent micromechanical model for

    predicting nonlinear viscoelastic responses of composites reinforced

    with solid spherical particles. Int. J. Solids Struct. 44 (21), 68916913.

    Muliana, A.H., 2008. Multi-scale framework for the thermo-viscoelastic

    analyses of polymer composites. Mech. Res. Commun. 35 (12), 8995.

    Mura, T., 1987. Micromechanics of Defects in Solids. Martinus Nijhoff, The

    Hague.

    Nemat-Nasser, S., Hori, M., 1999. Micromechanics: Overall Properties of

    Heterogeneous Materials, second ed. Elsevier, Amsterdam.

    Nogales, S., Bohm, H.J., 2008. Modeling of the thermal conductivity and

    thermomechanical behavior of diamond reinforced composites. Int. J.

    Eng. Sci. 46 (6), 606619.

    Norris, A.N., 1985. A differential scheme for the effective moduli of

    composites. Mech. Mater. 4, 116.

    OstojaStarzewski, M., Schulte, J., 1996. Bounding of effective thermalconductivities of multiscale materials by essential and natural

    boundary conditions. Phys. Rev. B 54 (1), 278285.

    Sangani, A.S., Yao, C., 1988. Bulk thermal conductivity of composites with

    spherical inclusions. J. Appl. Phys. 63, 1334.

    Segurado, J., Gonzalez, C., Llorca, J., 2003. A numerical investigation of the

    effect of particle clustering on the mechanical properties of

    composites. Acta Mater. 51, 23552369.

    Segurado, J., Llorca, 2002. A numericalapproximationto the elastic properties

    of sphere-reinforced composites. J. Mech. Phys. Solids 50, 21072121.

    Segurado, J., Llorca, 2006. Computational micromechanics of composites:

    The effect of particle spatial distribution. Mech. Mater. 38, 873883.

    Verma, L.S., Shrotriya, A.K., Singh, R., Chaudhary, D.R., 1991. Thermal

    conduction in two phase materials with spherical and non spherical

    inclusions. J. Phys. D 24, 17291737.

    Zohdi, T.I., Wriggers, P., 2001. Aspects of the computational testing of the

    mechanical properties of microheterogeneous material samples. Int. J.

    Numer. Methods Eng. 50, 25732599.

    K.A. Khan et al. / Mechanics of Materials 43 (2011) 608625 625

    http://sc.tamu.edu/http://sc.tamu.edu/