28
OVERALL VISCOELASTIC RESPONSE OF RANDOM FIBROUS COMPOSITES WITH STATISTICALLY UNIFORM DISTRIBUTION OF REINFORCEMENTS M. ˇ Sejnoha a,1 and J. Zeman a a Czech Technical University, Faculty of Civil Engineering Department of Structural Mechanics Th´ akurova 7, 166 29 Prague 6 Abstract An accurate representation of time dependent response of polymeric composite systems with disordered microstructure is developed within the framework of clas- sical homogenization methods. A graphite fiber tow impregnated by an epoxy resin, Fig. 1(a), is just an example of such systems. The investigation is focused on mod- eling issues pertinent to random, non-periodic, material systems, while the loading conditions are left to those promoting the linear viscoelastic deformation only. Two different approaches are examined. The first approach assumes a well defined ge- ometry of the fiber arrangement and specific boundary conditions. In the modeling framework, the complicated real microstructure is replaced by a material represen- tative volume element consisting of a small number of particles, which statistically resembles the real microstructure. Periodic distribution of such unit cells is consid- ered and the finite element method is called to carry out the numerical computation. The theoretical basis for the second approach are the Hashin-Shtrikman variational principles. The random character of the fiber distribution is incorporated directly into the variational formulation employing certain statistical descriptors. At the Preprint submitted to Elsevier Preprint 6 August 2001

OVERALL VISCOELASTIC RESPONSE OF RANDOM FIBROUS …ksm.fsv.cvut.cz/~sejnom/download/viscos_hsmkp.pdf · 2001. 12. 14. · OVERALL VISCOELASTIC RESPONSE OF RANDOM FIBROUS COMPOSITES

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

  • OVERALL VISCOELASTIC RESPONSE OF

    RANDOM FIBROUS COMPOSITES WITH

    STATISTICALLY UNIFORM

    DISTRIBUTION OF REINFORCEMENTS

    M. Šejnoha a,1 and J. Zeman a

    aCzech Technical University, Faculty of Civil EngineeringDepartment of Structural Mechanics

    Thákurova 7, 166 29 Prague 6

    Abstract

    An accurate representation of time dependent response of polymeric composite

    systems with disordered microstructure is developed within the framework of clas-

    sical homogenization methods. A graphite fiber tow impregnated by an epoxy resin,

    Fig. 1(a), is just an example of such systems. The investigation is focused on mod-

    eling issues pertinent to random, non-periodic, material systems, while the loading

    conditions are left to those promoting the linear viscoelastic deformation only. Two

    different approaches are examined. The first approach assumes a well defined ge-

    ometry of the fiber arrangement and specific boundary conditions. In the modeling

    framework, the complicated real microstructure is replaced by a material represen-

    tative volume element consisting of a small number of particles, which statistically

    resembles the real microstructure. Periodic distribution of such unit cells is consid-

    ered and the finite element method is called to carry out the numerical computation.

    The theoretical basis for the second approach are the Hashin-Shtrikman variational

    principles. The random character of the fiber distribution is incorporated directly

    into the variational formulation employing certain statistical descriptors. At the

    Preprint submitted to Elsevier Preprint 6 August 2001

  • present time the applications are limited to microstructures, which are sufficiently

    described by the two-point probability function. The presented results support ap-

    plicability of both methods to the description of viscoelastic behavior of the selected

    material system.

    Key words: Unit cell, variational principle, Maxwell chain model, eigenstrain

    (a) (b)

    Fig. 1. Geometry of fiber bundle and distribution of fibers within transverse plane section.

    1 Introduction

    Reliability, ease of processibility and relatively low cost offered by the polymer matrix com-

    posites contribute to their continuous rise in both mechanical and civil engineering industry.

    An adequate description of the deformation behavior of such systems is therefore in order.

    The time dependent response of polymers when subject to sustained loading is well known

    and has been under thorough study since their introduction. Inventory of contributions to

    analysis of polymer systems includes [?,?, among others]. The deformation behavior of poly-

    mers is generally quite complex ranging from viscoelastic regime at very low stresses, over the

    non-linear viscoelastic deformation at moderate stresses up to the yield behavior at hight

    1 Corresponding author.

    2

  • stresses. Such deformation mechanisms are also found in the polymer matrix composites.

    Examples include polymer matrix systems reinforced by aligned fibers, whiskers or fabrics

    typically supplied in form of laminated plates or woven tubes. Constitutive modeling of such

    systems usually assumes linearly elastic response of reinforcements while the matrix phase

    may undergo an inelastic or time dependent deformation. e.g., [?,?].

    Apart from proper selection of the constitutive model, the overall response of composite

    structures is highly influenced by their microstructure configuration. Estimates of the overall

    response of composite materials are often derived on the basis of periodic microstructures,

    but such are difficult to realize in practice. Instead, a highly disordered microstructures arise

    in most practical applications, see Fig. 1(b). Therefore, recognizing a random nature of such

    systems is a key step to a successful modeling of composite materials and structures. This

    viewpoint is also adopted in the present contribution. In particular, we focus on an accurate

    description of microstructure configuration and its effect on the overall response of polymer

    composites under sustain loading, while admitting only such loading conditions which result

    in a liner viscoelastic response of the matrix phase. A detailed description of more complex

    material behavior of polymer systems is in progress and will be presented in a forthcoming

    paper.

    Two classical computational techniques in the analysis of composites are reviewed and ap-

    plied to modeling of random composites. The first approach hinge on the existence of a

    periodic unit cell having the same statistics up to two-point correlations as the real mi-

    crostructure. A thorough study on this subject can be found in [?]. Note that regardless of

    their origin, the present formulation is applicable to any composite systems which can be

    represented by periodic unit cells. The second approach builds upon the well known Hashin-

    Shtrikman (H-S) variational principles [?]. The interested reader may also consult the work

    of [?,?], for generalization of these principles to treat random composites. Evaluation of ef-

    fective thermoelastic properties of fiber tow assuming statistically uniform distribution of

    3

  • fibers is presented in [?]. A strong attention is paid to an efficient evaluation of certain

    microstructure dependent matrices, given in terms of the two-point probability functions,

    which arise in the solution procedure based on the H-S principles. The most important as-

    pects of both procedures are revisited here, but extended to account for the distribution of

    local eigenstrain or eigenstress fields, which enter the constitutive equations to accommodate

    creep or relaxation effects.

    The present article is organized as follows. Section 2 outlines formulation of incremental

    form of local constitutive equations describing the linear viscoelastic response on the basis of

    generalized Maxwell chain model. Their introduction into a specific form of homogenization

    methods treating spatially periodic microstructures is addressed in Section 3. An alternate

    procedure to predict the time dependent response of random composites via the H-S princi-

    ples is presented in Section 4. The results for the selected material system are summarized

    in Section 5.

    2 Local constitutive equations

    The present section provides a brief overview of local constitutive equations describing the

    viscoelastic response of homogeneous and isotropic materials. Such equations can be formu-

    lated either in the integral form, or in the differential form. The differential form, that is more

    convenient for numerical implementation, can be derived by converting the integral equations

    into a rate-type form and by subsequent integration under certain simplifying assumptions.

    Nevertheless, both forms of constitutive equations requires, in general, an introduction of

    either creep J(t, τ) (compliance) or relaxation R(t, τ) function. Recall that the compliance

    function of a linear viscoelastic material represents the strain at time t due to a unit stress

    applied at time τ and kept constant, while the relaxation function represents the stress at

    time t due to a unit strain applied at time τ and held constant. To facilitate the numerical

    solution these functions are typically approximated by the degenerate (Dirichlet) kernels in

    4

  • the form

    J(t, τ) =M∑µ=1

    1

    Dµ(τ){1− exp [yµ(τ)− yµ(t)]} ,

    R(t, τ) =M∑µ=1

    Eµ(τ) exp [yµ(τ)− yµ(t)], (1)

    where yµ(t) = (t/Θµ). Retardation times Θµ must satisfy certain rules necessary for the

    success of calculation [?]. Functions Dµ and Eµ are usually obtained by fitting the creep

    or relaxation functions via Eqs. (1) using the method of least squares. Also note that rep-

    resentation (1) corresponds to the well known Maxwell and Kelvin chain models displayed

    in Fig. 2. Both rheological models are equivalent and can be used to deliver the searched

    constitutive equations. As an example, consider the Maxwell chain model and write the local

    stress in the form

    σ(x) =M∑µ=1

    σµ(x), (2)

    where σµ, called hidden stress, represents the stress in the µth Maxwell unit, which satisfies

    the differential constitutive equation

    σ̇µ(x) + ẏµσµ(x) = Eµ(x)L̂(x)(�̇(x)− �̇0(x)), ẏµ(t) =Eµ(t)

    ηµ(t). (3)

    Lµ(x, t) = Eµ(x, t)L̂(x) is the instantaneous stiffness matrix of a linear elastic isotropic

    material at the material point x. The initial strain vector ∆�0i may represent many different

    physical phenomena including thermal strains, shrinkage, swelling, plastic strains, etc. A

    simple finite difference based integration scheme can be devised to integrate Eq. (3). In

    doing so, first subdivide the time axis into intervals of length ∆ti. Next, suppose that at

    the beginning of the ith interval 〈ti−1, ti〉, the stress vector σµ(x, ti−1), µ = 1, 2, . . . ,M , is

    known. Further assume that functions Eµ(x, τ) = Eµ(x, ti−∆ti/2) remain constant within a

    given time interval. Then, integrating Eq. (3) with respect to time over ∆t and using Eq. (2)

    5

  • . . . . . .

    1 2 3 M

    σ

    σ

    η

    µE

    ηMµ

    8

    µEηµ

    . . .

    .

    M

    2

    1

    σ

    σ

    η 01

    (a) (b)

    Fig. 2. Rheological models: (a) Maxwell chain, (b) Kelvin chain.

    leads to the following incremental form of the local constitutive equation [?]

    ∆σi(x) = Li(x) (∆�i(x)−∆µi(x)) , (4)

    where the current increment of local eigenstrain reads

    ∆µi(x) = ∆�0i (x) + ∆�̂i(x). (5)

    The vector ∆�̂i(x) in Eq. (5) then represents the time dependent deformation and ∆�0i (x) is

    the increment of initial strain. When admitting only thermal and creep effects their respective

    forms are

    ∆�0i (x) = m(x)∆θi, ∆�̂i(x) =

    [L̂(x)

    ]−1Êi(x, t)

    M∑µ=1

    (1− e−∆yµ

    )σµ(x, ti−1), (6)

    where m(x) lists the coefficients of thermal expansion. The stiffness Êi for the ith interval

    is determined by

    Êi(x, t) =M∑µ=1

    Eµ(x, ti −∆ti/2)(1− e−∆yµ

    )/∆yµ. (7)

    6

  • 0 10 20 30 40 50

    t-τ [days]

    0

    500

    1000

    1500

    2000

    R(t

    -τ) [

    MP

    a]

    R - Findley’s relation

    R - Dirichlet series expansion

    Fig. 3. Relaxation function.

    Finally recall that the stress σµ(x, ti) at the end of the ith interval depends solely on the

    stress σµ(x, ti−1) found at the beginning of the ith interval such that

    σµ(x, ti) = σµ(x, ti −∆t)e−∆yµ + Eµ(x, ti −∆ti/2)λµL̂(x)(∆�i(x, ti)−∆�0i (x, ti)

    ).(8)

    Providing the material is age dependent, the coefficients Eµ should be determined at every

    time step. In the present formulation, however, the time dependent material properties of

    the epoxy matrix derived experimentally from a set of well cured specimens are assumed,

    [?], so that the material aging can be neglected. If this is the case a simple power law like

    formula for the unit creep rate

    J(t, τ) = a+ b(t− τ)n, R(t, τ) = 1a+ b(t− τ)n

    , (9)

    can be employed to fit the experimental data. For the selected epoxy resin the model data are

    a = 0.04744, b = 0.002142, n = 0.3526 with (t− τ) given in minutes, [?]. The corresponding

    relaxation function R(t, τ) appears in Fig. 3. Ten elements of the Dirichlet series expansion

    Eq. (1) uniformly distributed in log(t − τ) over the period of hundred days were assumed.

    The fit of the relaxation function Eq. (92) by Eq. (12) is plotted in Fig. 3.

    7

  • Eqs. (4) – (8) will now be used within the framework of individual solution techniques to

    introduce the time dependent behavior of the polymer systems into composites. We begin

    with simulations of viscoelastic processes in random composites by introducing a statistically

    equivalent unit cell in Section 3 and then by following an alternate approach based on H-S

    principles in Section 4.

    3 Description of viscoelastic behavior via periodic fields

    As already outlined in the introductory section, most of heterogeneous materials and compos-

    ite systems in particular exhibit random distribution of reinforcing constituents. Analyzing

    large samples of such materials, however, is computationally very expensive and cumbersome.

    Although exceptions exist with reference to elastic problems, such an approach inevitably

    becomes infeasible when allowing inelastic or time dependent deformations. A viable substi-

    tute, which yet provides notion of the local nature of variables of interest such as stresses

    and strains, relies on existence of the representative volume element (RVE) defined in terms

    of a periodic unit cell with a certain number of particles, which possesses similar statisti-

    cal properties as the original material, and therefore it can be considered as its reasonable

    approximation. This task requires formulation of an objective function, which relates the

    material’s statistics of the real microstructure and the corresponding unit cell. Its minimum

    then provides locations of individual reinforcements together with an optimal ratio of the

    unit cell dimensions. Solution systems based on evolution strategies were found appealing

    in the search for optimal unit cells. A thorough discussion on this subject with reference

    to evaluation of the overall thermoelastic properties is given in [?]. The interested reader

    may also consult the work by [?] for a comprehensive overview of various useful algorithms

    based on evolution strategies applied to this and similar problems commonly encountered in

    engineering practice.

    The section now proceeds by considering an optimal unit cell similar to one of Fig. 4 that

    8

  • (a) (b)

    Fig. 4. Periodic unit cells: (a) 5-fibers PUC, (b) 10-fibers PUC.

    statistically resembles the real microstructure, Fig. 1(b). Further suppose that the prescribed

    loading conditions produce a uniform distribution of macroscopic strain E or stress Σ fields.

    In either case, an increment of the local displacement field u(x) admits the following decom-

    position

    ∆u(x) = ∆E x + ∆u∗(x), (10)

    where ∆u∗(x) represents a fluctuation of the local displacement due to the presence of

    heterogeneities and is considered being periodic, [?,?, and references therein]. The local

    strain increment then assumes the form

    ∆�(x) = ∆E + ∆�∗(x), (11)

    where the fluctuation part ∆�∗(x) must vanish upon volume averaging. This requirement is

    fulfilled for the present boundary conditions since

    〈∆�∗ij(x)

    〉=

    1

    ∫Ω

    ∆�∗ij(x) dΩ =1

    2Ω

    ∫∂Ω

    (∆u∗i (x)nj + ∆u

    ∗j(x)ni

    )d(∂Ω) = 0, (12)

    due to periodicity of ∆u∗(x) (the same displacement on opposite sides of the unit cell). In

    9

  • Eq. (12) Ω and ∂Ω represent the volume and the boundary of the unit cell, respectively.

    The goal now becomes the evaluation of local fields within the unit cell and then their aver-

    aging to arrive at the desired macroscopic response. To proceed, we first write the principle

    of virtual work (Hill’s lemma) in the form

    〈δ�T∆σ

    〉= δET∆Σ. (13)

    Next, substituting Eq. (4) into the above expression yields

    δET〈Li(x)

    (∆E + ∆�∗(x)−∆�0i (x)−∆�̂i(x)

    )〉+〈δ�∗(x)TLi(x)

    (∆E + ∆�∗(x)−∆�0i (x)−∆�̂i(x)

    )〉= δETΣ, (14)

    where subscript i now links Eq. (14) with the current time increment. Since δE and δ�∗(x)

    are independent, the preceding equation can be split into two equalities

    δET(〈Li(x)〉∆E +

    〈Li(x)

    (∆�∗i (x)−∆�0i (x)−∆�̂i(x)

    )〉)− δET∆Σ = 0

    (15)〈δ(�∗)T(x)Li(x)

    〉∆E +

    〈δ(�∗)T(x)Li(x)

    (∆�∗i (x)−∆�0i (x)−∆�̂i(x)

    )〉= 0.

    A standard finite element approach is usually adopted to solve the above system of equations.

    We proceed by introducing a set of C0 continuous shape functions such that ∆u(x) =

    N(x)∆r; ∆r stores the components of the increment of nodal displacements. The vector of

    local strains then receives, after employing the geometrical equations, its familiar form

    ∆�∗(x) = B(x)∆r, δ�∗T = (δr)TBT(x). (16)

    Inserting these formulae in Eqs. (15) provides the following incremental form of algebraic

    equations K11 K12K21 K22

    i

    ∆E∆ri

    =

    ∆Σ + ∆F0

    ∆f 0

    i

    . (17)

    10

  • Individual matrices listed in Eq. (17) are written as

    K11i =1

    ∫Ω

    Li(x) dΩ,

    K12i = KT21i =

    1

    ∫Ω

    Li(x)B(x) dΩ, (18)

    K22i =1

    ∫Ω

    BT(x)Li(x)B(x) dΩ,

    and components of the right-hand side vector are

    ∆F 0i =1

    ∫Ω

    Li(x)(∆�0i (x) + ∆�̂i(x)

    )dΩ,

    ∆f 0i =1

    ∫Ω

    BT(x)Li(x)(∆�0i (x) + ∆�̂i(x)

    )dΩ. (19)

    Finally, eliminating the fluctuating displacements vector ∆ri from Eq. (17) readily provides

    the incremental form of the macroscopic constitutive law

    ∆Σi = Di∆Ei + ∆Λ0i , (20)

    where

    Di =(K11 −K12K−122 KT12

    )i, ∆Λ0i = −∆f 0i +

    (K12K

    −122 ∆f

    0)i. (21)

    When prescribing the overall strain only the system of equations (17) reduces to

    (K22)i∆ri = −(K21)i∆Ei + ∆f 0i . (22)

    Eqs. (17) and (22) can now be used to run either the creep or relaxation experiments. The

    results are discussed in Section 5.

    11

  • 4 Overall viscoelastic response via Hashin-Shtrikman principles

    Suppose that no information about the nature of local fields is needed, but instead a notion

    about the phase volume averages of field variables is sufficient in simulating the behavior

    of composite materials. Then, the sometimes time consuming implementation of finite ele-

    ments in conjunction with the unit cell analysis can be replaced by more simple averaging

    techniques. Such an approach is examined in this section.

    To introduce the subject consider an ergodic heterogeneous material system with statisti-

    cally homogeneous distribution of reinforcements. Here, we limit our attention to two-phase

    fibrous composites with fibers aligned along the x3 direction. The morphology of such a sys-

    tem can be conveniently described by the two-point probability function Srs. As suggested

    by Drugan & Willis [?] this function, when combined with the Hashin-Shtrikman varia-

    tional principles provides sufficient information for deriving bounds on effective properties

    of random composites. The material system under present study has been examined in very

    details in [?] with reference to evaluation of effective thermoelastic properties of statistically

    homogeneous materials using both the primary and dual variational principles of Hashin

    and Shtrikman. The remainder of this section outlines implementation of this approach to

    solve the viscoelastic problem. It is organized as follows: a review of basic statistical descrip-

    tors (Section 4.1), viscoelastic formulation based on the primary H-S principle to run the

    relaxation tests (Section 4.2) and viscoelastic formulation based on the dual H-S principle

    to study the creep behavior (Section 4.3).

    4.1 Review of basic statistical descriptors

    To reflect a random character of heterogeneous media it is convenient to introduce the

    concept of an ensemble – the collection of a large number of systems which are different

    in their microscopical details but identical in their macroscopic details. In the context of

    12

  • quantification of the microstructure morphology, an ensemble represents the collection of

    material micrographs taken from different samples of the material. To describe a random

    microstructure we introduce a characteristic function χr(x, α), which is equal to one when

    point x lies in the phase r within the sample α and equal to zero otherwise

    χr(x, α) =

    1 x ∈ Dr(α)0 otherwise. (23)

    The symbol Dr(α) denotes here the domain occupied by r-th phase in the sample α. For

    a two-phase fibrous composite, r = f,m, characteristic functions χf (x, α) and χm(x, α) are

    related by

    χm(x, α) + χf (x, α) = 1. (24)

    With the aid of function χr, the general n-point probability function Sr1,...,rn is given by [?].

    Sr1,...,rn(x1, . . . ,xn) = χr1(x1, α) · · ·χrn(xn, α). (25)

    Thus, Sr1,...,rn gives the probability of finding n points x1, . . . ,xn randomly thrown into the

    media located in the phases r1, . . . , rn. We limit our attention to functions of the order of

    one and two.

    Analysis of random composites usually relies on various statistical assumptions such as er-

    godic hypothesis, spatial homogeneity or isotropy, which may simplify the computational

    effort to a great extent. In particular, the ergodic hypothesis demands all states available to

    an ensemble of the systems to be available to every member of the system in the ensemble

    as well [?]. Then, the spatial or volume average of function χr(x, α) given by

    〈χr(x, α)〉 = limV→∞

    1

    V

    ∫V

    χr(x + y, α)dy, (26)

    13

  • is independent of α and identical to the ensemble average,

    χr(x) = Sr = 〈χr(x)〉 = cr, (27)

    where cr is the volume fraction of the rth phase. Note that the above assumption is usually

    accepted as an hypothesis subject to experimental verification. The statistical homogeneity

    assumption means that value of the ensemble average is independent of the position of

    coordinate system origin. Then, for example, the two-point matrix probability function reads

    Smm(x1,x2) = Smm(x12), (28)

    where xij = xj − xi. In the context of a representative volume element (RVE: a material

    element which effectively samples all microstructural configurations) the one-point proba-

    bility function Sr and the two-point probability function Srs are the same in any RVE (a

    micrograph of the material sample) irrespective of its position. Thus only one such sample is

    needed for their evaluation. When constructing the RVE we add an additional requirement

    with respect to its minimum size. Apart from the above statement we shall require the size

    of the RVE to be at least such that there exist two points within the RVE which are statisti-

    cally independent. Then, it appears acceptable to consider a periodicity of the selected RVE.

    This becomes particularly important when developing an efficient procedure for evaluation

    of Sr and Srs. Note that for an ergodic and periodic microstructure the two-point probability

    function Srs receives the following form

    Srs(x) =1

    ∫Ω

    χr(y)χs(x + y)dy, (29)

    where Ω is the size of the RVE (the micrograph area). It is worthwhile to mention that only

    the Fourier transform of function Srs given by

    S̃rs(ξ) =1

    Ωχ̃r(ξ)χ̃s(ξ), (30)

    14

  • is needed in H-S variational formulation [?]. Note that · now stands for the complex conjugate.

    When introducing a binary image of the actual microstructure may evaluate Eq. (30) very

    efficiently employing the discrete Fourier transform. See also [?] for further discussion on

    this subject. Eqs. (27) and (30) will be now implemented within the framework of the H-

    S variational principles. We present only the most important aspects of the formulation

    relevant to this study. For a personal review we refer the reader to original papers by [?,?].

    Additional references to related work are, e.g., [?,?] and [?].

    4.2 Extended primary Hashin-Shtrikman variational principle

    Suppose that the composite body is loaded by prescribed surface displacements resulting in

    a uniform macroscopic strain E and by the local eigenstresses λ(x) = −L(x)µ(x). With

    reference to the primary Hashin-Shtrikman variational principle the following representation

    of local stresses is equivalent

    σ(x) = L(x)�(x) + λ(x), σ(x) = L0�(x) + τ (x), (31)

    where L(x) is the local stiffness matrix and L0 is the stiffness matrix of a certain homogeneous

    reference medium. The vector τ is called the polarization stress. It can be shown that τ

    satisfying the equilibrium conditions, constitutive equations and boundary conditions is the

    one, which minimizes the functional

    Uτ =1

    2

    ∫Ω

    (ETΣ− (τ − λ)T(L− L0)−1(τ − λ)− 2τTE − �′Tτ − λTL−1λ

    )dΩ, (32)

    where the fluctuation part �′

    of the local strain � is provide by (see [?])

    �′(x) = �(x)−E =∫Ω

    �∗0(x− x′) (τ (x′)− 〈τ 〉) dΩ(x′). (33)

    15

  • �∗0(x− x′) is termed the fundamental solution that relates τ to a strain derived under zero

    displacement boundary conditions and for which the stress is self-equilibrated.

    Substituting Eq. (33) into Eq. (32) and then taking variation with respect to τ provides a

    system of algebraic equation for the searched polarization stress τ . To facilitate the solution

    we further restrict our attention to a piecewise uniform distribution of polarization stress

    τ r(x) = τ r and the eigenstress vector λr(x) = λr within a given phase r. Then, with

    reference to Eq. (23), the trial field for τ and eigenstress λ at any point x located in the

    sample α are provided by

    τ (x, α) =n∑r=1

    τ rχr(x, α), λ(x, α) =n∑r=1

    λrχr(x, α). (34)

    These expressions when plugged into functional Uτ Eq. (32) open a way for evaluation of

    its ensemble average Uτ , e.g., [?,?]. Performing variation of Uτ with respect to τ r finally

    supplies a set of equations for unknown phase averages of polarization stress τ r

    n∑s=1

    [δrs(Lr − L0)−1cr −Ars

    ]τ s = Ecr + (Lr − L0)−1λrcr, r = 1, . . . , n, (35)

    where the microstructure-dependent matrices Ars do not depend on x and are provided by,

    see [?],

    Ars =∫Ω

    �∗0(x− x′) [Srs(x− x′)− crcs] dΩ(x′) =1

    (2π)2

    ∫Ω

    �̃∗0(ξ′)S̃ ′rs(ξ

    ′)dξ′. (36)

    Formal inversion of Eq. (35) yields the overall constitutive equation in the form

    σ = LE + λ, (37)

    where the spatially average overall stiffness matrix L and the macroscopic eigenstress vector

    λ are provided by

    16

  • L= L0 +n∑r=1

    n∑s=1

    crTrscs, (38)

    λ=n∑r=1

    n∑s=1

    crTrscs(Ls − L0)−1λs. (39)

    Trs then represents individual blocks of the inverse matrix to the left hand side of system (35).

    For a two-phase composite medium they can be written in the form

    A = Amm, Kr = Lr − L0,Trs = Kr

    (cfKm + cmKf − cfcmA−1

    ) [Kf + Km −Kr + δrs(1− cr)A−1

    ], (40)

    To introduce the thermal and viscoelastic effects we first recall Eq. (4). Then, in analogy

    with Eqs. (20) and (22) the incremental form of Eq. (37) becomes

    ∆σi = Li∆Ei + ∆λi, (41)

    and the current increment of λ and the instantaneous stiffness matrix Lm of the matrix

    phase attain the forms

    ∆λi = −n∑r=1

    n∑s=1

    crTrscs(Ls − L0)−1i (Ls)i(∆�0s + ∆�̂s

    )i, s = m→ (Lm)i = ÊiL̂m, (42)

    whereas the fiber phase is assumed elastic. Eqs (41) and (42) drive the solution of a vis-

    coelastic problem under strain control loading conditions suitable for modeling the stress

    relaxation.

    4.3 Extended dual Hashin-Shtrikman variational principle

    Consider a composite body under prescribed surface tractions which produce a uniform

    macroscopic stress Σ. In addition, a distribution of local eigenstrains µ(x) can be introduced

    in the present formulation. In analogy with Eq. (33) we write the local constitutive equations

    17

  • in the form

    �(x) = M(x)σ(x) + µ(x), �(x) = M0σ(x) + γ(x), (43)

    where M(x) now stands for the local compliance matrix and M0 is the compliance matrix of a

    homogeneous reference medium. The polarization strain γ, when satisfying the compatibility

    condition and the constitutive equation, provides the minimum of the extended dual H-S

    functional

    Uγ =1

    2

    ∫Ω

    (ΣTE − (γ − µ)T(M−M0)−1(γ − µ) + 2γTΣ + σ′Tγ

    )dΩ, (44)

    where the fluctuation part σ′

    of the local stress σ written in terms of polarization strain γ

    is, see [?],

    σ′(x) = σ(x)−Σ = [σ∗0 (γ − 〈γ〉)]−M−10 (γ(x)− 〈γ〉) . (45)

    The operator [σ∗0 (γ − 〈γ〉)] can be identified with the operator [�∗0 (τ − 〈τ 〉)] when replacing

    γ for τ and σ∗0 for �∗0 and suitably modifying the boundary term to reflect the traction

    boundary conditions. Specific forms of tensors �∗0 and σ∗0 can be found, e.g., in [?].

    Assuming only a piecewise uniform distribution of phase polarization strain γr(x) = γr and

    the phase eigenstress vector µr(x) = µr provides the trial fields for γ and µ at any point x

    located in the sample α in the form, recall Eq. (34),

    γ(x, α) =n∑r=1

    γrχr(x, α), µ(x, α) =n∑r=1

    µrχr(x, α). (46)

    We now proceed, in analogy with Section 4.2, by introducing Eqs. (45) and (46) into func-

    tional (44), evaluating an ensemble average of Uγ and finally taking variation with respect

    to γr to get

    18

  • n∑s=1

    {δrs[(Mr −M0)−1 + M−10

    ]cr −Brs −M−10 crcs

    }γs

    = Σcr + (Mr −M0)−1crµr, r = 1, 2, . . . , n. (47)

    Eq. (47) represents a system of algebraic equations to be solved for unknown components

    of phase averages of the polarization strain γr. To continue, recall Eq. (36) and write the

    microstructure dependent matrices Brs in the form

    Brs =1

    (2π)2

    ∫Ω

    σ̃∗0(ξ′)S̃ ′rs(ξ

    ′)dξ′. (48)

    Finally, after symbolic inversion of Eq. (47) we arrive at the macroscopic constitutive law,

    see also Eqs. (37) – (39),

    � =MΣ + µ, (49)

    where the effective compliance matrixM and the macroscopic eigenstrain vector µ read

    M= M0 +n∑r=1

    n∑s=1

    crRrscs, (50)

    µ=n∑r=1

    n∑s=1

    crRrscs(Ms −M0)−1µs. (51)

    For a two-phase composite medium the matrices Rrs assume the form

    B = cfcmM−10 −Bmm, Nr =

    [(Mr −M0)−1 + M−10

    ]−1,

    Rrs = Nr(cfN2 + cmN1 − cfcmB−1

    ) [Nf + Nm −Nr + δrs(1− cr)B−1

    ]. (52)

    When solving a viscoelastic problem we replace Eq. (49) by its incremental counterpart such

    that

    ∆�i =Mi∆Σi + ∆µi, (53)

    19

  • and finally the current increment of µ and the instantaneous compliance matrix Mm of the

    matrix phase are expressed as

    ∆µi =n∑r=1

    n∑s=1

    crRrscs(Ms −M0)−1i(∆�0s + ∆�̂s

    )i, s = m→ (Mm)i =

    1

    ÊiM̂m, (54)

    Eqs. (53) and (54) can be used, while assuming the time independent elastic behavior of the

    fiber phase, to find the solution to a stress control viscoelastic problem such as creep. The

    results for a selected class of problems are presented in the next section.

    5 Numerical examples

    As an example, consider a two-phase fibrous composite system displayed in Fig. 1(b). As

    suggested in [?] this medium can be regarded as ergodic and statistically homogeneous thus

    serving as a suitable candidate for the numerical analysis. Both phases are homogeneous

    with constant material properties listed in Table 1. The Cartesian coordinate system with

    phase EA ET GT νA a b n

    [GPa] [GPa] [GPa] [GPa]−1 [GPa]−1

    fiber 386 7.6 2.6 0.41

    matrix 2.1 2.1 0.75 0.40 0.0474 0.00214 0.3526Table 1Material properties of T30/Epoxy system.

    the x3-axis directed along the fiber direction is selected. Therefore, the macroscopic stress

    and strain fields written in contracted notation are Σ = {Σ11,Σ22,Σ12,Σ33} T and E =

    {E11, E22, E12, E33} T. Recall that the generalized plane strain conditions are assumed. In

    the present study, the numerical results are derived for loading applied within the transverse

    plane section of the composite aggregate only. Both the strain and stress control loading

    conditions are considered in simulations. Fig. 5 illustrates the time variation of the applied

    load.

    20

  • 11

    t [min]

    Σ[MPa]

    15

    11.25

    360 720 t [min]

    Σ[MPa]

    12

    15

    11.25

    360 720

    (a) (b)

    11

    t [min]

    E[%]

    360 720

    0.5

    0.375

    t [min]

    [%]E 12

    360 720

    0.5

    0.375

    (a) (b)

    Fig. 5. Applied loading: (a,b) creep test, (c,d) relaxation test.

    The first set of figures (Figs. 6-9) represents the composite response derived from the unit cell

    model. The macroscopic creep behavior is considered first. Fig. 6 shows the time variation of

    the overall strain developed under pure tensile and shear stress loading conditions, respec-

    tively (Fig. 5a,b). Similar response resulting from the strain loading conditions is plotted in

    Fig. 7. Individual results suggest that at least 5-fibers Unit cell should be used in numerical

    simulations to provide sufficiently accurate response of the actual composite, Fig. 1(b). Note

    that the same characteristic of the present material system was discovered when studying

    a pure-elastic behavior of such composites, [?]. However, unlike the results derived assum-

    ing pure elasticity, the viscoelastic response of the hexagonal array model slightly deviates

    from that found using the statistically equivalent unit cells. Nevertheless, a very low con-

    trast in material parameters in the transverse direction promotes the hexagonal array model

    as a suitable substitute for more complicated unit cells when the fiber volume fraction is

    sufficiently large.

    21

  • 0.0 200.0 400.0 600.0 800.0

    t [min]

    2.5e-03

    3.0e-03

    3.5e-03

    4.0e-03

    4.5e-03

    5.0e-03O

    vera

    ll st

    rain

    E11

    2-fibers Unit cell

    5-fibers Unit cell

    10-fibers Unit cell

    20-fibers Unit cell

    Hexagonal array

    0.0 200.0 400.0 600.0 800.0

    t [min]

    8.0e-03

    1.0e-02

    1.2e-02

    1.4e-02

    1.6e-02

    1.8e-02

    Ove

    rall

    stra

    in E

    12

    2-fibers Unit cell

    5-fibers Unit cell

    10-fibers Unit cell

    20-fibers Unite cell

    Hexagonal array

    (a) (b)

    Fig. 6. Overall response – Unit cell model: creep test.

    0.0 200.0 400.0 600.0 800.0

    t [min]

    1.0e+01

    1.5e+01

    2.0e+01

    2.5e+01

    3.0e+01

    Ove

    rall

    stre

    ss Σ

    11 [M

    Pa]

    2-fibers Unit cell

    5-fibers Unit cell

    10-fibers Unit cell

    20-fibers Unit cell

    Hexagonal array

    0.0 200.0 400.0 600.0 800.0

    t [min]

    3.0e+00

    3.5e+00

    4.0e+00

    4.5e+00

    5.0e+00

    5.5e+00

    Ove

    rall

    stre

    ss Σ

    12 [M

    Pa]

    2-fibers Unit cell

    5-fibers Unit cell

    10-fibers Unit cell

    20-fibers Unit cell

    Hexagonal array

    (c) (d)

    Fig. 7. Overall response – Unit cell model: relaxation test.

    Quite different conclusion, however, can be drawn from Fig. 8. The macroscopic response

    plotted in Fig. 8 provides evidence that increasing the material contrast leads to a noticeable

    difference in the material response derived from the hexagonal array model and statistically

    optimal unit cells. To arrive at such a result we simply replaced the actual transversally

    isotropic fiber with the isotropic one to attain a higher contrast between the material prop-

    erties of the fiber and matrix phases. Also note a possible change in the smallest RVE to

    be considered for numerical simulations (10-fibers UC). Additional support for using the

    22

  • 0.0 200.0 400.0 600.0 800.0

    t [min]

    1.5e-03

    2.0e-03

    2.5e-03

    3.0e-03

    3.5e-03

    4.0e-03

    Ove

    rall

    stra

    in E

    11

    5-fibers Unit cell

    10-fibers Unit cell

    20-fibers Unit cell

    Hexagonal array

    Fig. 8. Overall response – Unit cell model: creep test assuming isotropic fiber.

    optimal unit cells instead of the hexagonal array model is provided by Fig. 9 suggesting

    an anisotropic character of the present medium. Such a result cannot be attained by sim-

    ple periodic unit cells. The present approach, which draws on the existence of a unit cell

    statistically equivalent to the actual composite system, is therefore preferable.

    0.0 200.0 400.0 600.0 800.0

    t [min]

    1.5e+01

    2.0e+01

    2.5e+01

    3.0e+01

    Str

    ess

    [MP

    a]

    Σ22 due to ε22Σ11 due to ε11

    Overall stress

    Matrix stress

    Fiber stress

    0.0 200.0 400.0 600.0 800.0

    t [min]

    1.0e-03

    2.0e-03

    3.0e-03

    4.0e-03

    5.0e-03

    6.0e-03

    7.0e-03

    Str

    ain

    ε11 due to Σ11ε22 due to Σ22

    Overall strain

    Matrix strain

    Fiber strain

    (a) (b)

    Fig. 9. Overall and local response – Unit cell model: (a) relaxation test, (b) creep test.

    The remaining set of results (Figs. 10–14) was found when employing the Hashin-Shtrikman

    variational principles. Recall that the primary H-S principle might be invoked to simulate

    the relaxation response while the dual H-S principle should be called to study the creep

    behavior. In both instances, however, the resulting representation of the viscoelastic response

    is governed by the selection of a homogeneous comparison medium (L0,M0). To this end, we

    draw the readers attention to Fig. 10. First, recall Table 1 and notice that the matrix moduli

    23

  • 0.0 200.0 400.0 600.0 800.0

    t [min]

    1.0e+01

    1.5e+01

    2.0e+01

    2.5e+01

    3.0e+01O

    vera

    ll st

    ress

    S11

    [MP

    a]

    H-S Upper bound

    H-S Estimate

    H-S Lower bound

    Unit cell

    0.0 200.0 400.0 600.0 800.0

    t [min]

    2.5e-03

    3.0e-03

    3.5e-03

    4.0e-03

    4.5e-03

    5.0e-03

    Ove

    rall

    stra

    in E

    11

    H-S Upper bound

    H-S Estimate

    H-S Lower bound

    Unit cell

    (a) (b)

    Fig. 10. Overall response – hexagonal packing: (a) relaxation test, (b) creep test.

    are indeed weaker than those of fiber. Therefore, according to [?] we select the matrix phase

    to fill individual entries of L0 and M0 providing we search for a lower bound on the relaxation

    response (Fig. 10(a)) and an upper bound on the creep response (Fig. 10(b)), respectively.

    The fiber phase is then selected to yield estimates on opposite bounds. A number of other

    results, contained within the H-S bounds, can be derived when mixing individual phases to

    set L0 and M0 (e.g., L0 =12(Lf + Lm)).

    (a) (b)

    Fig. 11. Idealized binary images: (a)–resolution 488x358 pixels, (b)–resolution 244x179 pixels.

    Fig. 10 gives some idea of this approach assuming hexagonal packing of fibers. One notewor-

    thy feature of the hexagonal array model is the correspondence of the periodic hexagonal

    unit cell with the Mori-Tanaka averaging technique. It is well known that the Mori-Tanaka

    24

  • estimates of the overall response of composites with weak matrices correspond to the lower

    and upper bounds derived from the primary and dual H-S variational principles, respectively.

    From this point of view one may also judge the results derived from the periodic hexagonal

    unit cell (solid lines in Fig. 10a,b). Therefore, in order to make the results found from the H-S

    variational principles comparable with those derived previously using the optimal unit cells,

    we selected the matrix phase to create the 4x4 homogeneous stiffness L0 and compliance

    M0 matrices. Also note that these matrices are kept constant throughout the integration

    process. This assumption is acceptable since the instantaneous moduli of the matrix phase

    do not vary considerably for the selected time-stepping procedure.

    0.0 200.0 400.0 600.0 800.0

    t [min]

    2.5e-03

    3.0e-03

    3.5e-03

    4.0e-03

    4.5e-03

    5.0e-03

    Ove

    rall

    stra

    in E

    11

    976 x 716 pixels

    488 x 358 pixels

    244 x 358 pixels

    122 x 84 pixels

    0.0 200.0 400.0 600.0 800.0

    t [min]

    8.0e-03

    1.0e-02

    1.2e-02

    1.4e-02

    1.6e-02

    1.8e-02

    Ove

    rall

    stra

    in E

    12

    976 x 716 pixels

    488 x 358 pixels

    244 x 179 pixels

    122 x 84 pixels

    (a) (b)

    Fig. 12. Overall response – Hashin-Shtrikman principle: creep test.

    0.0 200.0 400.0 600.0 800.0

    t [min]

    1.0e+01

    1.5e+01

    2.0e+01

    2.5e+01

    3.0e+01

    Ove

    rall

    stre

    ss Σ

    11 [M

    Pa]

    976 x 716 pixels

    488 x 358 pixels

    244 x 179 pixels

    122 x 84 pixels

    0.0 200.0 400.0 600.0 800.0

    t [min]

    3.0e+00

    3.5e+00

    4.0e+00

    4.5e+00

    5.0e+00

    5.5e+00

    Ove

    rall

    stre

    ss Σ

    12 [M

    Pa]

    976 x 716 pixels

    488 x 358 pixels

    244 x 179 pixels

    122 x 84 pixels

    (a) (b)

    Fig. 13. Overall response – Hashin-Shtrikman principle: relaxation test.

    25

  • 200.0 400.0 600.0 800.0

    t [min]

    1.5e+01

    2.0e+01

    2.5e+01

    3.0e+01S

    tres

    s [M

    Pa]

    Σ11 due to ε11Σ22 due to ε22

    Overall stress

    Matrix stress

    Fiber stress

    0.0 200.0 400.0 600.0 800.0

    t [min]

    1.0e-03

    2.0e-03

    3.0e-03

    4.0e-03

    5.0e-03

    6.0e-03

    7.0e-03

    Str

    ain

    ε11 due to Σ11ε22 due to Σ22

    Overall strain

    Matrix strain

    Fiber strain

    (a) (b)

    Fig. 14. Overall and local response – Hashin-Shtrikman principle: (a) relaxation test, (b) creep test.

    0.0 200.0 400.0 600.0 800.0

    t [min]

    1.5e+01

    2.0e+01

    2.5e+01

    3.0e+01

    Str

    ess

    [MP

    a]

    UCH-S: LB

    Overall stress

    Matrix stress

    Fiber stress

    0.0 200.0 400.0 600.0 800.0

    t [min]

    1.0e-03

    2.0e-03

    3.0e-03

    4.0e-03

    5.0e-03

    6.0e-03

    7.0e-03

    Str

    ain

    Unit cellH-S: UB

    Overall strain

    Matrix strain

    Fiber strain

    (a) (b)

    Fig. 15. Overall and local response – UC vs. H-S: (a) relaxation test, (b) creep test.

    As outlined in Section 4.1, evaluation of the overall response of random composites requires

    first selection of a RVE. Binary images of the RVE of the present material system, Fig. 1(b),

    which complies with certain requirements discussed in Section 4.1, are displayed in Fig. 11.

    The RVEs of Fig. 11 were employed to evaluate the microstructure dependent matrices in

    Eqs. (36) and (48). The effect of bitmap resolution on the overall response was examined

    first. Results appear in Figs. 12 and 13. Evidently, even low resolution of 244x179 pixels ( ≈

    55 pixels per fiber ) provides sufficiently accurate results. This should also hold for combined

    loading. Such a result is quite encouraging, particularly if one would like to increase the

    26

  • size of the RVE. Fig. 14 further confirms ability of this approach to model an anisotropic

    character of the present material system already suggested by the FEM analysis of optimal

    periodic unit cells.

    Finally, we bring some comparison between the UC analysis and the H-S variational prin-

    ciples plotted in Fig. 15. As expected, recall our previous discussion on H-S bounds, the

    relaxation data obtained from 5-fibers periodic UC correlates fairly well with the H-S lower

    bound and similarly the creep response of UC is found close to the H-S upper bound. Thus

    the applicability of both approaches to simulate the viscoelastic behavior of statistically

    homogeneous material systems such as the one under present study is confirmed.

    6 Conclusion

    Most of the real composite material systems experience a random distribution of reinforce-

    ments. In this contribution, two generally accepted approaches in the micromechanics were

    reviewed and applied to the analysis of random statistically homogeneous composite systems

    undergoing viscoelastic deformation. Formulation of macroscopic constitutive equations was

    first outlined within the framework of periodic fields. Random nature of microstructure

    configuration was accounted for through various unit cells found such as to represent the

    material statistics of real composites. An effect of number of reinforcements in the unit cell

    needed for an accurate representation of macroscopic response was examined. In view of the

    results presented in [?] and plots displayed in Figs. 6-9 it appears necessary to introduce at

    least five fibers in the optimal unit cell to arrive at sufficiently accurate predictions of the

    overall behavior of the present material system. In addition, the present results also proved

    necessity for an accurate modeling of microstructure configuration particularly when the

    contrast between the phase material parameters becomes large.

    The second part of this contribution revisited the well known Hashin-Shtrikman variational

    27

  • principles further extended to reflect the presence of eigenstresses and eigenstrains in the

    formulation of macroscopic equations. An efficient procedure for evaluation of required mate-

    rials statistics, already examined in [?], was implemented in connection with binary images of

    a real microstructure. The results plotted in Figs. 12-13 illustrate insensitivity of the solution

    procedure to a selected bitmap resolution, which may further increase expected efficiency of

    this approach. Final comparison with the FEM analysis shown in Fig. 15 assesses the appli-

    cability and supports the use of this method, at least in the range of elastic or viscoelastic

    response.

    7 Acknowledgement

    Financial support was provided by the research project CEZ:MSM 210000003.

    28

    IntroductionLocal constitutive equationsDescription of viscoelastic behavior via periodic fieldsOverall viscoelastic response via Hashin-Shtrikman principlesReview of basic statistical descriptorsExtended primary Hashin-Shtrikman variational principleExtended dual Hashin-Shtrikman variational principle

    Numerical examplesConclusionAcknowledgement