higher order thry

Upload: sai-prasad-dasari

Post on 07-Apr-2018

223 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/4/2019 Higher Order Thry

    1/56

    Higher-order theory for functionally graded materials

    J. Aboudia, M.-J. Pinderab,*, S.M. Arnoldc

    aTel-Aviv University, Ramat-Aviv 69978, IsraelbUniversity of Virginia, Charlottesville, VA 22903, USA

    cNASA Glenn Research Center, Cleveland, OH 44135, USA

    Received 18 June 1999; accepted 12 August 1999

    Abstract

    This paper presents the full generalization of the Cartesian coordinate-based higher-order theory for functionally graded materials

    developed by the authors during the past several years. This theory circumvents the problematic use of the standard micromechanicalapproach, based on the concept of a representative volume element, commonly employed in the analysis of functionally graded composites

    by explicitly coupling the local (microstructural) and global (macrostructural) responses. The theoretical framework is based on volumetric

    averaging of the various field quantities, together with imposition of boundary and interfacial conditions in an average sense between the

    subvolumes used to characterize the composites functionally graded microstructure. The generalization outlined herein involves extension

    of the theoretical framework to enable the analysis of materials characterized by spatially variable microstructures in three directions.

    Specialization of the generalized theoretical framework to previously published versions of the higher-order theory for materials functionally

    graded in one and two directions is demonstrated. In the applications part of the paper we summarize the major findings obtained with the

    one-directional and two-directional versions of the higher-order theory. The results illustrate both the fundamental issues related to the

    influence of microstructure on microscopic and macroscopic quantities governing the response of composites and the technologically

    important applications. A major issue addressed herein is the applicability of the classical homogenization schemes in the analysis of

    functionally graded materials. The technologically important applications illustrate the utility of functionally graded microstructures in

    tailoring the response of structural components in a variety of applications involving uniform and gradient thermomechanical loading.

    1999 Elsevier Science Ltd. All rights reserved.

    Keywords: Higher-order theory; Functionally graded materials; Microstructural tailoring/optimization

    1. Introduction

    Functionally graded materials (FGMs) are a new genera-

    tion of engineered materials wherein the microstructural

    details are spatially varied through non-uniform distribution

    of the reinforcement phase(s), by using reinforcement with

    different properties, sizes and shapes, as well as by inter-

    changing the roles of reinforcement and matrix phases in a

    continuous manner [1]. The result is a microstructure thatproduces continuously or discretely changing thermal and

    mechanical properties at the macroscopic or continuum

    scale. Examples illustrating different types of functionally

    graded microstructures are presented in Fig. 1 (where the

    terminology to describe the gradation type is used to reflect

    the spatial variation of macroscopic properties). This new

    concept of engineering the materials microstructure marks

    the beginning of a revolution both in the materials science

    and mechanics of materials areas as it allows one, for the

    first time, to fully integrate the material and structural

    considerations into the final design of structural compo-

    nents. The rapidly growing interest in FGMs, originally

    conceived in Japan, is evidenced by the large number of

    recent conferences and special issues of technical journals

    devoted to the analysis, design and fabrication of these

    materials [27].

    Functionally graded materials are ideal candidates forapplications involving severe thermal gradients, ranging

    from thermal structures in advanced aircraft and aerospace

    engines to computer circuit boards. In one such application,

    a ceramic-rich region of a functionally graded composite is

    exposed to hot temperature while a metallic-rich region is

    exposed to cold temperature, with a gradual microstructural

    transition in the direction of the temperature gradient,

    Fig. 1(a). By adjusting the microstructural transition appro-

    priately, optimum temperature, deformation and stress

    distributions can be realized. This concept has been success-

    fully employed to enhance the thermal fatigue resistance

    Composites: Part B 30 (1999) 777832

    1359-8368/99/$ - see front matter 1999 Elsevier Science Ltd. All rights reserved.

    PII: S1359-8368(99) 00053-0

    www.elsevier.com/locate/compositesb

    * Corresponding author. Tel.: 1-804-924-1040; fax: 1-804-982-2951.

    E-mail address: [email protected] (M.-J. Pindera)

  • 8/4/2019 Higher Order Thry

    2/56

    and life of ceramic thermal barrier coatings [8,9]. Micro-

    structural grading through non-uniform reinforcementspacing, Fig. 1(b), or through the use of different types of

    reinforcement, Fig. 1(c), can also be effectively used to

    reduce the mismatch in the thermomechanical properties

    between differently oriented, adjacent plies in a laminated

    plate. Thus, the reduction of thermally induced interlaminar

    stresses at the free edge of a laminate (which result from the

    large property mismatch between the adjacent plies) can be

    realized by using the functional grading concept to smooth

    out the transition between dissimilar plies. Along similar

    lines, joining of dissimilar materials can be made more effi-

    cient through the use of functionally graded joints [10 12].

    Other benefits to be realized from the use of functionally

    graded architectures include fracture toughness enhance-ment in ceramic matrix composites through tailored inter-

    faces [13], and/or introduction of a second phase that creates

    compressive stress fields in critical, crack-prone regions.

    Many more applications of FGMs can be found in the afore-

    mentioned conference proceedings and special issues of

    technical journals, including the most recent focus on the

    solar energy conversion devices [14], dental implants [15],

    and naturally occurring biological FGMs [16,17].

    Owing to the many variables that control the design of

    functionally graded microstructures, full exploitation of the

    FGMs potential requires the development of appropriate

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832778

    Nomenclature

    dpa ; h

    qb ; l

    rg dimensions of the subcell (ab g) in the (p,q,r)th generic cell

    eabgijl;m;n coefficients of the total strain series expansion in the subcell (ab g)

    kabgi coefficients of heat conductivity of the material in the subcell ( abg)

    p; q; r indices used to identify the cell (p; q; r)q

    abgi components of the heat flux vector in the subcell ( ab g)

    vpqra;b;g volume of the subcell (abg) in the (p,q,r)th generic cell

    uabgi displacement components in the subcell (abg)

    xa1 ; x

    b2 ; x

    g3 local subcell coordinates

    Cabgijkl elements of the stiffness tensor of the material in the subcell (abg)

    1abgij local strain components in the subcell (abg)

    1inabgij local inelastic strain components in the subcell (a bg)

    Np, Nq, Nr number of cells in the x1, x2 and x3 directions, respectively

    Qabgil;m;n

    average values of the subcell heat flux component qabgi

    when l

    m

    n

    0; higher-order heat flux terms for

    other values of l,m,n

    Pn Legendre polynomial of order n

    Rabgijl;m;n inelastic strain distribution components in the subcell (abg)

    Sabgijl;m;n average values of the subcell (a bg) stress components s

    abgij when l m n 0; higher-order stress compo-

    nents for other values of l,m,n

    T(abg) temperature field in the subcell (abg)

    Tabgl;m;n temperature at the center of the subcell (abg) when l m n 0; coefficients associated with higher-order

    terms in the temperature field expansion within the subcell ( abg) for other values of l,m,n

    Wabgil;m;n displacement components at the center of the subcell (abg) when l m n 0; coefficients associated with

    higher-order terms in the displacement field expansion within the subcell (abg) for other values of l,m,n

    a ,b,g indices used to identify the subcell (abg)

    m(abg) shear modulus of an isotropic material in the subcell (abg)

    sabgij local stress components in the subcell (abg)

    sTabgij local thermal stress components in the subcell (abg)

    tabgijl;m;n coefficients of the stress series expansion in the subcell (ab g)

    tTabgijl;m;n coefficients of the thermal stress series expansion in the subcell (abg)

    Gabgij elements of the thermal tensor of the material in the subcell (abg)

  • 8/4/2019 Higher Order Thry

    3/56

    modeling strategies for their response to combined thermo-

    mechanical loads. Presently, most computational strategies

    for the response of FGMs do not explicitly couple the mate-

    rials heterogeneous microstructure with the structural

    global analysis. Rather, local effective or macroscopic prop-

    erties at a given point within the FGM are first obtained

    through homogenization based on a chosen micromechanics

    scheme, and subsequently used in a global thermomechani-

    cal analysis, Fig. 2. In this approach, the local micromecha-nical analysis is carried out in an independent fashion from

    the global macromechanical analysis, essentially decou-

    pling the influence of the surrounding spatially inhomoge-

    neous microstructure on the local response of the equivalent

    (homogenized) continuum point. This is what is meant by

    the absence of micro-macrostructural coupling in this type

    of an analysis, as will be elaborated in the next section. The

    exclusion of the possibility of coupling between local and

    global effects often leads to potentially erroneous results in

    the presence of macroscopically non-uniform material prop-

    erties and large field variable gradients. This is particularly

    true when the temperature gradient is large with respect to

    the dimension of the inclusion phase, or the characteristic

    dimension of the inclusion phase is large relative to the

    global dimensions of the composite and the number of

    uniformly or non-uniformly distributed inclusions is rela-

    tively small. Perhaps the most important objection to

    using the standard micromechanics approach based on the

    concept of a representative volume element (RVE) in

    the analysis of FGMs is the lack of a theoretical basis for

    the definition of an RVE, which clearly cannot be unique in

    the presence of continuously changing properties due to

    non-uniform inclusion spacing [1820].

    As a result of the limitation of the standard micromecha-

    nics approaches in FGM applications, a new higher-order

    micromechanical theory (HOTFGM), which explicitly

    couples the local (microstructural) and global (macrostruc-

    tural) effects, has been developed and applied to function-

    ally graded composites. The theoretical framework is based

    on volumetric averaging of the various field quantities,

    satisfaction of the field equations in a volumetric sense,and imposition in an average sense of boundary and inter-

    facial conditions between the subvolumes used to character-

    ize the composites functionally graded microstructure. The

    need for development of such a theory has been demon-

    strated by comparison with results obtained using a standard

    micromechanics approach that neglects the micro-macro-

    structural coupling effects.

    The original formulation of HOTFGM was developed for

    the thermoelastic analysis of metal matrix composite plates

    functionally graded in the thickness direction, constrained

    from deforming in the plane of the plate, and subjected to a

    through-thickness temperature gradient [2123]. Thisversion of the higher-order theory is known as HOTFGM-

    1D as it reflects the one-directional grading character of the

    materials microstructure. Extensive comparison between

    the predictions of the one-directional version of the

    higher-order theory and the results generated using the

    finite-element analysis [24] and the boundary-element

    analysis [25] established the theory as a viable tool for the

    analysis of functionally graded composites. Subsequent

    incorporation of inelastic constitutive models with tempera-

    ture-dependent parameters for the response of the constitu-

    ent phases into HOTFGM-1Ds theoretical framework

    made possible the analysis of through-thickness function-

    ally graded metal matrix composites over a wide tempera-ture range [26]. Relaxation of the constraint on the inplane

    deformation through a partial homogenization procedure in

    the non-functionally graded (periodic) directions extended

    the theorys applicability to practical structural problems by

    allowing simulation of generalized plane strain inplane

    boundary conditions [27].

    The analytical framework of HOTFGM, which results in

    a closed-form system of equations for the unknown micro-

    variables that govern the distribution of internal tempera-

    ture, displacement and stress fields at each point within the

    heterogeneous material, greatly facilitates the incorporation

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832 779

    Fig. 1. Examples of different types of functionally graded microstructures.

  • 8/4/2019 Higher Order Thry

    4/56

    of HOTFGM into an optimization algorithm as demon-

    strated by Aboudi et al. [28] using HOTFGM-1D coupled

    to the commercially available optimizer DOT [29]. In that

    investigation, optimum through-thickness fiber distributions

    were determined that minimized the inplane force and

    moment resultants in unidirectionally and bidirectionally

    reinforced MMC plates subjected to a through-thickness

    thermal gradient.The recently developed two-directional version of the

    higher-order theory, HOTFGM-2D, allows coupled micro-

    macrostructural analysis of orthogonal composite laminates,

    with finite dimensions in the plane containing arbitrarily

    spaced inclusions or continuous fibers, subjected to

    combined thermomechanical loading [30,31]. In particular,

    the technologically important interlaminar stress fields in

    laminated composites in the vicinity of the free edge were

    analyzed and the advantages of functionally graded micro-

    structures in reducing the free-edge stress concentrations

    were demonstrated. In the most recent application of the

    two-directional theory to functionally graded thermal

    barrier coatings subjected to thermal gradient cyclic load-

    ing, the creep induced stress redistribution was shown to

    depend on the level of microstructural refinement, further

    illustrating the shortcomings of the homogenization-based

    analysis [32,33]. Summaries of selected developments of

    the higher-order theory during the periods 1992 through

    1995 and 1995 through 1997 presented in the above refer-ences have been provided by Pindera et al. [34,35].

    As the development of the higher-order theory was driven

    not only by our desire to address fundamental issues related

    to microstructural effects in composites but also by various

    technologically important applications, the notation

    employed in constructing the theory was not always

    uniform. In particular, the notation employed in construct-

    ing the one-directional version was influenced by the nota-

    tion used in developing the first-order homogenization

    theory known in the literature as the generalized method

    of cells [36,37]. In contrast, the extension of the theory

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832780

    Fig. 2. Homogenization-based micromechanical analysis of functionally graded materials.

  • 8/4/2019 Higher Order Thry

    5/56

    that enables the analysis of functionally graded microstruc-

    tures in two directions necessitated the adaptation of a more

    general notation. This most recent notation makes further

    generalization of the higher-order theory to three dimen-

    sions relatively straightforward.

    One of the objectives of this review article, therefore, is to

    present the full generalization of the higher-order theory to

    three dimensions from which the two-directional and one-

    directional versions are obtained as special cases using a

    unified notation. In addition, the major results that illustrate

    both the fundamental issues related to the influence ofmicrostructure on macroscopic and microscopic quantities

    governing the response of composites and the technologi-

    cally important applications are summarized. A major

    theme that the authors wish to highlight is the issue of the

    applicability of the classical homogenization schemes in the

    analysis of functionally graded materials. The technologi-

    cally important applications summarized herein, on the

    other hand, illustrate the utility of functionally graded

    microstructures in tailoring the response of structural

    components under spatially uniform and non-uniform ther-

    momechanical load histories.

    2. Applicability of classical homogenization schemes

    An alternative solution methodology for the thermal and

    mechanical fields in functionally graded materials involves

    a combination of the standard micromechanics and conti-

    nuum approaches in which local (micromechanical) and

    global (macromechanical) analyses are carried out inde-

    pendently of each other, i.e. in a decoupled manner. In

    this approach it is assumed that the heterogeneous micro-

    structure of an FGM can be replaced by an equivalent

    continuum with a set of macroscopic properties that vary

    with spatial coordinates in a manner commensurate with

    the materials heterogeneity, Fig. 2. The macroscopic prop-

    erties at each point of the equivalent continuum are gener-

    ated by assuming that it is possible to define an RVE in the

    presence of spatially variable microstructure and then

    carrying out an RVE-based micromechanical analysis

    using a chosen micromechanics model. That is, the macro-

    scopic properties are generated on the premise that no

    coupling exists between local and global responses asthese properties are calculated without explicitly taking

    into account the influence of the adjacent spatially variable

    microstructural details. Subsequently, the continuum or

    macrolevel results are obtained by solving the given

    boundary-value problem of a homogenized medium with

    spatially variable equivalent material properties calculated

    in the manner described above. As stated previously, this

    decoupled two-step procedure is the standard micro-macro-

    mechanical approach currently employed by most

    researchers working in the area of functionally graded

    materials. With the knowledge of the macroscopic thermo-

    mechanical fields, the stresses in the individual phases at acontinuum point may then be estimated by applying

    average thermomechanical field quantities at the selected

    point (or region), treating it as an RVE within the frame-

    work of the chosen micromechanical model. A problem

    arises, however, when the microstructural scale is large

    relative to the thermomechanical field gradients so as to

    invalidate the basic assumptions on which the RVE

    concept is based. The situation becomes more complicated

    when the microstructure changes continuously as in this

    case, the micromechanical concept of an RVE, and thus

    material property, is not well defined.

    The various micromechanical approaches used to calcu-

    late effective properties of composites include use of simpleReuss and Voigt hypotheses, self-consistent schemes and

    their generalizations, differential schemes, the Mori

    Tanaka method, concentric cylinder models, bounding tech-

    niques and approximate (e.g. the generalized method of

    cells) or numerical (e.g. finite-element or finite-difference)

    analyses of periodic arrays of inclusions or fibers in the

    surrounding matrix phase [38]. The central assumption in

    applying these well-established techniques is the existence

    of an RVE and the ability to apply homogeneous boundary

    conditions to such an element, Fig. 3. These homogeneous

    boundary conditions can be specified either in terms of

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832 781

    Fig. 3. Calculation of macroscopic moduli based on the concept of a repre-

    sentative volume element.

  • 8/4/2019 Higher Order Thry

    6/56

    surface displacements

    uiS 10ijxj 1

    or in terms of prescribed surface tractions

    Ti

    S

    s0ijnj

    2

    where ni is the unit outward normal vector on the boundary

    surface S of the composite, xi are the Cartesian coordinates

    of the surface, 10ij and s0ij are constants, and repeated index

    implies summation. For a heterogeneous medium the

    constants 10ij and s0ij are the volume averaged strains and

    stresses under the prescribed boundary conditions given

    by Eqs. (1) and (2), respectively. This result is a conse-

    quence of the following relations:

    1ij 1

    V

    V

    1ijxk dV 1

    V

    S

    1

    2uinj ujni dS 3

    sij 1

    V

    V

    sijxk dV 1

    V

    S

    1

    2Tixj Tjxi dS 4

    where Vis the volume enclosed by the surface S. The above

    relations hold provided that: the displacements ui are contin-

    uous; the tractions Ti are continuous at all interfaces of the

    heterogeneous medium; and body forces vanish.

    In practice, the average strains and stresses that result

    from the application of homogeneous boundary conditions

    are calculated for an RVE whose macroscopic behavior is

    indistinguishable from the behavior of the composite-at-

    large. By applying the homogeneous boundary conditions

    to the bounding surface of the RVE, which are the same as

    the boundary conditions applied to the entire composite, its

    average behavior can be calculated. This average behavior,

    in turn, defines the composites macroscopic properties. To

    qualify as an RVE, the volume of the element used to calcu-

    late average composite behavior must meet two criteria.

    First, it must be sufficiently small with respect to the dimen-

    sions of the composite-at-large in order to be considered a

    material point in the equivalent homogeneous continuum

    (i.e. hp H; see Fig. 4). Second, it must be sufficiently

    large with respect to the inclusion phase (i.e. dp h) so

    that to the first order the elastic strain energy induced by

    both sets of homogeneous boundary conditions is the same,rendering the effective elastic properties independent of the

    manner in which boundary conditions are applied [18]. In

    this case, the microstructure of the RVE can be replaced by

    a fictitious homogeneous material with effective or homo-

    genized moduli or compliances, Cijkl or Sijkl, respectively,

    that connect the average stress and strain quantities as

    follows:

    sij Cijkl 1kl 5

    1ij Sijkl skl 6

    where Sijkl C1ijkl ; Fig. 3. In the case of periodic fiber arrays,the repeating unit cell is interpreted as an RVE provided that

    the homogeneous boundary conditions are replaced by

    either symmetry conditions on the deformation of the unit

    cell or periodic boundary conditions, depending on the type

    of loading. It should be mentioned that the use of periodic

    boundary conditions in conjunction with a multi-scale

    asymptotic expansion of the displacement field in the

    repeating unit cell forms the basis of the so-called homo-

    genization methods for estimating the macroscopic proper-

    ties of periodic composites [39,40].

    Clearly, the range of applicability of the aforementioned

    micromechanical approaches is limited to composites rein-

    forced by numerous fibers with very small diameters such

    as graphite or carbon fibers. In such composites, a typical

    RVE contains a sufficiently large number of fibers while

    occupying a very small volume of the entire composite,

    allowing one to disregard boundary-layer effects near the

    bounding surfaces of the RVE upon application of either

    type of homogeneous boundary conditions. As a result,even in the presence of highly inhomogeneous deformation

    gradients within the composite-at-large, the field quantities

    within the RVE will not vary significantly, thereby permit-

    ting the definition of a material property at a point in the

    equivalent homogeneous continuum. In contrast, in compo-

    sites with few and relatively large-diameter fibers (with

    respect to the thickness of a single ply), the variation of

    the quantities of interest within the RVE (assuming that it

    can be defined) invalidates the basic assumptions on which

    the concept of effective properties is based. These local

    variations of the field quantities within the RVE may

    give rise to unexpected phenomena rooted in the local-global coupling which is neglected in the traditional micro-

    mechanical homogenization schemes. For instance, differ-

    ent thermal conductivities of the individual phases together

    with their directional arrangement may produce thermal

    gradients in the individual phases which are quite different

    from the thermal gradients in the homogeneous composite

    with equivalent effective properties subjected to identical

    boundary conditions, Fig. 4. This, in turn, may alter the

    local conductivity characteristics and produce unexpected

    effects such as localized hot spots for instance. The size

    of the RVE in relation to the thickness of the composite

    and the temperature gradient obviously will play an impor-

    tant role in the above scenario.The preceding discussion raises questions about the

    applicability of the traditional micromechanical approach

    based on the concept of an RVE in the presence of large

    thermal gradients and coarse or spatially variable micro-

    structure. In light of this discussion, the current practice of

    decoupling the local response from the global response by

    calculating pointwise effective thermoelastic properties of

    functionally graded materials without regard to whether the

    actual microstructure admits the presence of an RVE, and

    subsequently using these properties in the global analysis of

    the heterogeneous material, remains to be justified. These

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832782

  • 8/4/2019 Higher Order Thry

    7/56

    issues were discussed qualitatively as early as 1974 by

    Pagano [41] with regard to mechanical loading of macro-

    scopically homogeneous composites. Most recently, discus-

    sions regarding the admissibility of the classical RVEconcept in the homogenization analysis of random heteroge-

    neous media can be found in the works of Huet [42], Ostoja-

    Starzewski and Schulte [43], Ostoja-Starzewski et al. [44]

    and Ostoja-Starzewski [45]. To account for the materials

    random microstructure, these authors used the concept of a

    material window containing different numbers of inhomo-

    geneously distributed inclusions in order to study the effect

    of the window size on the local macroscopic properties.

    In order to resolve the issue of the applicability of the

    classical (in the sense of Hill) RVE concept in the analy-

    sis of FGMs, a model is required that explicitly couples

    the microstructural and macrostructural analyses. The

    higher-order micromechanical theory, developed by the

    authors during the past several years and presented in

    the following section, is a step in this direction for appli-cations involving, but not limited to: composites with

    uniformly or non-uniformly spaced; large-diameter fibers

    subjected to through-thickness thermal gradients; finite

    cross-section metal matrix composite laminates subjected

    to combined thermomechanical loading; and thermal

    barrier coatings. The generalization of this higher-order

    theory for materials functionally graded in three orthogo-

    nal directions is outlined next. Finally, we mention that

    the present approach is based on deterministic considera-

    tions, in contrast with the statistical approaches mentioned

    above.

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832 783

    Fig. 4. Applicability of the representative volume element for heterogeneous materials in the presence of a thermal gradient.

  • 8/4/2019 Higher Order Thry

    8/56

    3. Generalized three-directional higher-order theory

    3.1. Model overview

    The generalized higher-order theory is based on the

    geometric model of a heterogeneous parallelepiped occupy-

    ing the region 0 x1 D; 0 x2 H; 0 x3 L

    (Fig. 5). The loading applied on the bounding surface of

    the parallelepiped may involve an arbitrary temperature or

    heat flux distribution and mechanical effects represented by

    a combination of surface displacements and/or tractions

    consistent with global equilibrium requirements. Thecomposite may be reinforced by arbitrarily spaced planar

    arrays of continuous fibers (with arbitrary in-plane spacing)

    oriented along any of the three axes x1, x2, x3, or finite-length

    inclusions that are arranged arbitrarily in the three planes.

    The microstructure of the heterogeneous composite is

    discretized into Np, Nq and Nrcells in the intervals 0 x1

    D; 0 x2 H; 0 x3 L; respectively. The generic cell

    (p,q,r) used to construct the composite, highlighted in Fig. 5,

    consists of eight subcells designated by the triplet (abg),

    Fig. 6, where each index a ,b,g takes on the values 1 or 2

    which indicate the relative position of the given subcell

    along the x1, x2 and x3 axis, respectively. The indices p, q

    and r, whose ranges are p 1; 2; ;Np; q 1; 2; ;Nq andr 1; 2; ;Nr; identify the generic cell in the three planes.The dimensions of the generic cell along the x1, x2 and x3axes, d

    p1 ; d

    p2 ; h

    q1 ; h

    q2 ; and l

    r1 ; l

    r2 ; can vary in an arbitrary

    fashion such that

    D Np

    p1dp1 dp2 H

    Nqq1

    hq1 hq2

    L Nrr1

    lr1 lr2

    Given the applied thermomechanical loading, an approx-

    imate solution for the temperature and displacement fields is

    constructed based on volumetric averaging of the field equa-

    tions together with the imposition of boundary and continu-

    ity conditions in an average sense between the subvolumes

    used to characterize the materials microstructure. This is

    accomplished by approximating the temperature and displa-

    cement fields in each subcell of a generic cell using a quad-

    ratic expansion in the local coordinates xa; xb; xg

    centered at the subcells mid-point. A higher-order repre-

    sentation of the temperature and displacement fields is

    necessary in order to capture the local effects created bythe thermomechanical field gradients, the microstructure

    of the composite and the finite dimensions in the function-

    ally graded directions. This is in sharp contrast with

    previous treatments involving fully periodic composite

    media which employed linear expansions [38]. We note

    that it is possible to specialize this temperature and displa-

    cement field representation to model composites that are

    functionally graded in one direction with the remaining

    directions periodic, or composites that are functionally

    graded in a plane with periodicity in the out-of-plane direc-

    tion. In either case, the temperature and displacement fields

    in the periodic directions are approximated using a linear

    expansion in local coordinates. In the former case, the one-directional version of the theory is obtained, while the latter

    case yields the two-directional theory. Both specialized

    cases will be elaborated upon in Section 4.

    The unknown coefficients associated with each term in

    the temperature and displacement field expansions are

    obtained by constructing systems of equations that satisfy

    the requirements of a standard boundary-value problem for

    the given field variable approximations. That is, the zeroth,

    first, and second moments of the heat and equilibrium equa-

    tions are satisfied in a volumetric sense. Similarly, the ther-

    mal and heat flux, as well as the displacement and traction,

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832784

    Fig. 5. Schematic of a composite functionally graded in three directions.

  • 8/4/2019 Higher Order Thry

    9/56

    continuity conditions (the mechanical analogues) within agiven generic cell, and between a given generic cell and its

    neighbors, are imposed in an average sense across the inter-

    facial planes. The solution for these coefficients in the

    generalized higher-order theory follows the general frame-

    work for the solution of the corresponding two-directional

    thermoelastic and thermoinelastic problems discussed

    previously [30,31]. Therefore, only a summary of the

    governing equations for the temperature and displacement

    fields in the individual subcells within the rows and columns

    of cells considered in solving the outlined boundary-value

    problem is given in the following sections. Detailed deriva-

    tion of these equations can be found in the above references.

    3.2. Thermal analysis

    Let the functionally graded parallelepiped be subjected to

    steady-state temperature or heat flux distributions on its

    bounding surfaces. Under steady-state heat conduction,

    the heat flux field in the material occupying the subcell

    (abg) of the (p,q,r)th generic cell, in the region xa1 12

    dpa ; x

    b2

    12

    hqb ; x

    g3

    12

    lrg ; must satisfy:

    21qabg1 22q

    abg2 23q

    abg3 0 7

    where 21 2=2 xa1 ; 22 2=2 xb2 ; 23 2=2 xg3 : The

    components qabg

    i of the heat flux vector in this subcellare obtained from the Fouriers heat conduction law foranisotropic materials,

    qabgi kabgij 2jTabg; i;j 1; 2; 3 8

    where kabgij are the coefficients of heat conductivity of the

    material in the subcell (abg), with kabgij kabgi dij (no sum

    on i) for orthotropic materials, and no summation is implied

    by repeated Greek letters in the above and henceforth.

    The temperature distribution in the subcell (a bg) of the

    (p,q,r)th generic cell, measured with respect to a reference

    temperature Tref, is denoted by T(abg). We approximate this

    temperature field by a second-order expansion in the local

    coordinates xa1 ; xb2 ; x

    g3 as follows:

    Tabg Tabg000 xa1 Tabg100 xb2 Tabg010 xg3 Tabg001

    1

    23 xa21

    dp2a

    4

    2 3T

    abg200

    1

    23 x

    b22

    hq2b

    4

    2 3T

    abg020

    1

    23 x

    g23

    lr2g

    4

    2 3T

    abg002 9

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832 785

    Fig. 6. Generic cell (p,q,r) of a composite functionally graded in three directions showing the dimensions and the designation convention of the internal subcells

    (a b g).

  • 8/4/2019 Higher Order Thry

    10/56

    where Tabg000 which is the temperature at the center of the

    subcell, and Tabglmn l; m; n 0; 1; or 2 with l m n 2

    are unknown coefficients which are determined from condi-

    tions that will be outlined subsequently. It should be noted

    that no terms of the form xi x

    j i j appear in the tempera-ture field representation, as well as in the displacement field

    representation given in the sequel, due to the averaging

    procedure employed in our higher-order theory.

    Given the seven unknown quantities associated with each

    subcell (i.e. Tabg000 ; ; T

    abg002 ) and eight subcells within each

    generic cell, 56NpNqNr unknown quantities must be deter-

    mined for a composite with Np, Nq, and Nrrows and columns

    of cells containing arbitrarily specified materials. These

    quantities are determined by first satisfying the heat conduc-

    tion equation, as well as the first and second moment of this

    equation, in each subcell in a volumetric sense in view of the

    temperature field approximation given by Eq. (9). Subse-

    quently, continuity of heat flux and temperature is imposed

    in an average sense at the interfaces separating adjacent

    subcells, as well as neighboring cells. Fulfillment of thesefield equations and continuity conditions, together with the

    imposed thermal boundary conditions on the bounding

    surfaces of the composite, provides the necessary 56NpNqNrequations for the 56NpNqNr unknown coefficients in the

    temperature field expansion.

    We begin the outline of steps to generate the required

    number of equations by first considering an arbitrary

    (p,q,r)th generic cell in the interior of the composite (i.e.

    p 2; ;Np1; q 2; ;Nq1; and r 2; ;Nr1: Thisproduces 56Np 2Nq 2Nr 2 equations. The addi-tional equations are obtained by considering the boundary

    cells (i.e. p 1;Np

    ;q 1

    ;Nq and r 1

    ;Nr). For thesecells, most of the relations obtained from considering the

    interior cells will also hold, with the exception of some of

    the interfacial continuity conditions between adjacent

    interior cells which will be replaced by the specified

    boundary conditions.

    In the course of satisfying the steady-state heat equation

    in a volumetric sense, it is convenient to define the follow-

    ing flux quantities:

    Qabgil;m;n

    1

    npqrabg

    dpa=2d

    pa=2

    hqb=2h

    qb=2

    lrg=2l

    rg=2

    xa1 l xb2 m

    xg3 nqabgi d xa1 d xb2 d xg3 10

    where l; m; n 0; 1; or 2 with l m n 2; and vpqrabg d

    pa h

    qb l

    rg is the volume of the subcell (a bg) in the (p,q,r)th

    generic cell. For l; m; n 0; Qabgi0;0;0 is the volume averagevalue of the heat flux component q

    abgi in the subcell,

    whereas for other values of (l,m,n) Eq. (10) defines

    higher-order heat fluxes. These flux quantities can be eval-

    uated explicitly in terms of the coefficients Tabglmn by

    performing the required volume integration using Eqs. (8)

    and (9) in Eq. (10). This yields the following non-vanishing

    zeroth-order and first-order heat fluxes in terms of the

    unknown coefficients in the temperature field expansion:

    Qabg10;0;0 kabg1 Tabg100 11

    Qabg11;0;0 kabg1

    dp2a

    4T

    abg200 12

    Qabg

    20;0;0 kabg

    2 Tabg

    010 13

    Qabg20;1;0 kabg2

    hq2b

    4T

    abg020 14

    Qabg30;0;0 kabg3 Tabg001 15

    Qabg30;0;1 kabg3

    lr2g

    4T

    abg002 16

    Subsequently, satisfaction of the zeroth, first and second

    moment of the steady-state heat equation results in the

    following eight relationships among the first-order heat

    fluxes Qabgil;m;n

    in the different subcells (abg) of the

    (p,q,r)th generic cell, after some involved algebraic manip-

    ulations (see [21,22]):

    Qabg11;0;0=d2a Qabg20;1;0=h2bQabg30;0;1=l2gp;q;r 0 17where the triplet (abg) assumes all permutations of the

    integers 1 and 2.

    The continuity of heat fluxes at the subcell interfaces and

    between individual generic cells, imposed in an average

    sense, is ensured by:

    12Q1bg11;0;0=d1 Q2bg10;0;0 6Q2bg11;0;0=d2p;q;r

    Q2bg

    10;0;0 6Q2bg

    11;0;0=d2p1;q;r

    0 18

    Q1bg10;0;0 12 Q2bg10;0;0 3Q

    2bg11;0;0=d2p;q;r

    12Q2bg10;0;0 6Q2bg11;0;0=d2p1;q;r 0 19

    12Qa1g20;1;0=h1 Qa2g20;0;0 6Qa2g20;1;0=h2p;q;r

    Qa2g20;0;0 6Qa2g20;1;0=h2p;q1;r 0 20

    Qa1g20;0;0 12 Qa2g20;0;0 3Q

    a2g20;1;0=h2p;q;r

    12 Qa2g20;0;0 6Qa2g20;1;0=h2p;q1;r 0 21

    12Qab130;0;1=l1 Qab230;0;0 6Qab230;0;1=l2p;q;r

    Qab230;0;0 6Qab230;0;1=l2p;q;r1 0 22

    Qab130;0;0 12 Qab230;0;0 3Q

    ab230;0;1=l2p;q;r

    12Qab230;0;0 6Qab230;0;1=l2p;q;r1 0 23

    Eqs. (18)(23) provide us with 24 additional relations

    among the zeroth-order and first-order heat fluxes. These

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832786

  • 8/4/2019 Higher Order Thry

    11/56

    relations, together with Eq. (17), can be expressed in terms

    of the unknown coefficients Tabglmn by making use of

    Eqs. (11)(16), providing a total of 32 of the required 56

    equations necessary for the determination of these

    coefficients in the interior (p,q,r)th generic cell.

    An additional set of 24 equations necessary to deter-

    mine the unknown coefficients in the temperature field

    expansion is subsequently generated by the thermal

    continuity conditions imposed on an average basis at

    each subcell and cell interface. Imposing the interfacial

    thermal continuity conditions we obtain the following

    relations:

    T1bg000 12 d1T1bg100

    14

    d21 T

    1bg200 p;q;r

    T2bg000 12 d2T2bg100

    14

    d22T

    2bg200 p;q;r 24

    T2bg000 12 d2T2bg100

    14

    d22 T

    2bg200 p;q;r

    T1bg

    000 1

    2 d1T1bg

    100 1

    4 d

    2

    1T1bg

    200 p1;q;r

    25Ta1g000 12 h1T

    a1g010

    14

    h21T

    a1g020 p;q;r

    Ta2g000 12 h2Ta2g010

    14

    h22T

    a2g020 p;q;r 26

    Ta2g000 12 h2Ta2g010

    14

    h22T

    a2g020 p;q;r

    Ta1g000 12 h1Ta1g010

    14

    h21T

    a1g020 p;q1;r 27

    Tab1000 12 l1Tab1001

    14

    l21T

    ab1002 p;q;r

    Tab2

    000 12 l2T

    ab2

    001 14 l

    2

    2Tab2

    002 p;q;r

    28

    Tab2000 12 l2Tab2001

    14

    l22T

    ab2002 p;q;r

    Tab1000 12 l1Tab1001

    14

    l21T

    ab1002 p;q;r1 29

    which comprise the required additional 24 relations.

    The steady-state heat equations, Eq. (17), together with

    the heat flux and thermal continuity equations, Eqs. (18)

    (23) and Eqs. (24)(29), respectively, form 56 linear

    algebraic equations which govern the 56 field variables

    Tabglmn in the eight subcells (abg) of an interior generic

    cell

    p; q; r

    ; p

    2; ;Np1; q

    2; ;Nq1; r

    2; ;Nr1: For the boundary cells p 1;Np; q 1;Nqand r 1;Nr a different treatment must be applied. For

    p 1; for instance, the above relations are operative, butrelations (18) and (19), on the other hand, which follow

    from the continuity of heat flux between a given generic

    cell and the preceding one are not applicable. They are

    replaced by the condition that the heat flux at the interface

    between subcell (1bg) and (2bg) of the cell (1,q,r) is

    continuous, as well as the applied temperature relation at

    the bottom surface x1 0 (see Fig. 5). For the cell p Np;the previous equations are applicable except Eq. (25) which

    are obviously not operative. These equations are replaced by

    the specific temperature applied at the top surface x1 D:The boundary conditions at these surfaces are

    T1bg1;q;r Tbottomx2;x3; x11 12 d

    11 30

    T2bgNp ;q;r Ttopx2;x3; x21 12 d

    Np2 31

    where q 1; ;Nq; r 1; ;Nr:Similar reasoning holds for the cells q 1 and q Nq:

    That is, the temperature in the cells (p,1,r) and (p,Nq,r) at the

    x2 0 and x2 H(i.e. left and right) surfaces, respectively,must equal the applied temperature,

    Ta1gp;1;r Tleftx1;x3; x12 12 h

    11 32

    Ta2gp;Nq ;r Trightx1;x3; x22 12 h

    Nq2 33

    where p 1; ;Np; r 1; ;Nr:Finally, the temperature in the cells (p,q,1) and (p,q,Nr) at

    the x3 0 and x3 L (i.e. front and back) surfaces, respec-tively, must equal the applied temperature,

    Tab1p;q;1 Tfrontx1;x2; x13 12 l

    11 34

    Tab2p;q;Nr Tbackx1;x2; x23 12 l

    Nr2 35

    where p 1; ;Np; q 1; ;Nq: Alternatively, it ispossible to impose mixed-boundary conditions involving

    temperature and heat flux at different portions of the

    boundary.

    The governing equations for the interior and boundary

    cells form a system of 56NpNqNr algebraic equations in the

    unknown coefficients Tabglmn : Their solution determines the

    temperature distribution within the functionally gradedcomposite subjected to the boundary conditions (30)(35).

    The final form of this system of equations is symbolically

    represented below

    kT t 36where the structural thermal conductivity matrix k contains

    information on the geometry and thermal conductivities of

    the individual subcells (abg) in the NpNqNr cells spanning

    the x1, x2 and x3 directions, the thermal coefficient vector T

    contains the unknown coefficients that describe the tempera-

    ture field in each subcell, i.e.

    T T111

    111 ; ; T222

    NpNqNrwhere

    Tabgpqr T000; T100; T010; T001; T200; T020; T002abgpqrand the thermal force vector t contains information on the

    boundary conditions.

    3.3. Mechanical analysis

    Given the temperature field generated by the applied

    surface temperatures and/or heat fluxes obtained in the

    preceding section, we proceed to determine the resulting

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832 787

  • 8/4/2019 Higher Order Thry

    12/56

    displacement and stress fields. This is carried out for arbi-

    trary mechanical loading, consistent with global equili-

    brium requirements, applied to the surfaces of the

    composite.

    The stress field in the subcell (a bg) of the (p,q,r)th

    generic cell generated by the given temperature field must

    satisfy the equilibrium equations

    21sabg1j 22s

    abg2j 23s

    abg3j 0; j 1; 2; 3 37

    where the operator 2ii 1; 2; 3 has been definedpreviously. The components of the stress tensor, assuming

    that the material occupying the subcell (abg) of the

    (p,q,r)th generic cell is orthotropic, are related to the

    strain components through the familiar generalized

    Hookes law:

    sabgij Cijkl1abgkl 1inabgkl Gabgij Tabg 38

    where Cijkl are the elements of the stiffness tensor, 1in

    abg

    klare the inelastic strain components, and the elements

    Gabgij of the so-called thermal tensor are the products

    of the stiffness tensor and the thermal expansion coeffi-

    cients. In this paper, we consider either elastic ortho-

    tropic materials or inelastic materials which are

    isotropic in both elastic and inelastic domains. Hence,

    Eq. (38) reduces to

    sabgij Cijkl 1abgkl 2mabg1inabgij sTabgij 39

    where m (abg) is the elastic shear modulus of the mate-

    rial filling the given subcell (a bg), and the term

    sTabgij ; henceforth referred to as thermal stress, standsfor the thermal contribution G

    abgij T

    abg: The compo-

    nents of the strain tensor in the individual subcells

    are, in turn, obtained from the strain-displacement rela-

    tions

    1abgij 12 2iu

    abgj 2ju

    abgi ; i;j 1; 2; 3 40

    The displacement field in the subcell (abg) of the

    (p,q,r)th generic cell is approximated by a second-order

    expansion in the local coordinates xa1 ; xb2 ; and x

    g3 as

    follows:

    uabg1 Wabg1000 xa1 Wabg1100 xb2 Wabg1010 xg3 Wabg1001

    1

    23 xa21

    dp2a

    4

    2 3W

    abg1200

    1

    23 x

    b22

    hq2b

    4

    2 3W

    abg1020

    1

    23 x

    g23

    lr2g

    4

    2 3W

    abg1002 41

    uabg2 Wabg2000 xa1 Wabg2100 xb2 Wabg2010 xg3 Wabg2001

    1

    23 xa21

    dp2a

    4

    2 3W

    abg2200

    12

    3 xb22 h

    q2b4

    2 3Wabg2020

    1

    23 x

    g23

    lr2g

    4

    2 3W

    abg2002 42

    uabg3 Wabg3000 xa1 Wabg3100 xb2 Wabg3010 xg3 Wabg3001

    1

    23 xa21

    dp2a4

    2 3W

    abg3200

    1

    23 x

    b22

    hq2b4

    2 3W

    abg3020

    1

    23 x

    g23

    lr2g

    4

    2 3W

    abg3002 43

    where Wabgi000 ; which are the displacements at the center of

    the subcell, and the higher-order terms Wabgilmn i 1; 2; 3

    must be determined from conditions similar to those

    employed in the thermal problem. In this case, there are

    168 unknown quantities in a generic cell (p,q,r). The deter-

    mination of these quantities parallels that of the thermalproblem. Here, the heat conduction equation is replaced

    by the three equilibrium equations, and the continuity of

    tractions and displacements at the various interfaces

    replaces the continuity of heat fluxes and temperature.

    Finally, the boundary conditions involve the appropriate

    mechanical quantities. As in the thermal problem, we start

    with the internal cells and subsequently modify the govern-

    ing equations to accommodate the boundary cells p 1;Np; q 1;Nq; and r 1;Nr:

    In the perfectly elastic case, the quadratic displacement

    expansion, Eqs. (41) (43), produces linear variations in

    strains and stresses at each point within a given subcell. In

    the presence of inelastic effects, however, a linear strainfield generated by Eqs. (41)(43) does not imply the linear-

    ity of the stress field due to the path-dependent deformation.

    Thus the displacement field microvariables must depend

    implicitly on the inelastic strain distributions, giving rise

    to a higher-order stress field than the linear strain field

    generated from the assumed displacement field representa-

    tion. In the presence of inelastic effects, this higher-order

    stress field is represented by a higher-order Legendre poly-

    nomial expansion in the local coordinates. Therefore, the

    strain field generated from the assumed displacement

    field, and the resulting mechanical and thermal stress fields,

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832788

  • 8/4/2019 Higher Order Thry

    13/56

    must also be expressed in terms of Legendre polynomials:

    1abgij

    l0

    m0

    n0

    1 2l1 2m1 2n

    p

    eabgijl;m;nPlza1 Pmzb2 Pnzg3 44

    sabgij

    l0

    m0

    n0

    1 2l1 2m1 2n

    p

    tabgijl;m;nPlza1 Pmzb2 Pnzg3 45

    sTabgij

    l0

    m0

    n0

    1 2l1 2m1 2n

    p

    tTabgijl;m;nPlza1 Pmzb2 Pnzg3 46

    where the non-dimensionalized variables z

    i s, defined in theinterval 1 zi 1; are given in terms of the local

    subcell coordinates as za1 xa1 =dpa =2; zb2 xb2 =hqb =2; and zg3 xg3 =lrg =2: For the given displace-

    ment field representation, Eqs. (41)(43), the upper limits

    on the summations in Eq. (44) become 1, while for the given

    temperature distribution, Eq. (9), the upper limits on the

    summations in Eq. (46) become 2. Alternatively, the

    upper limits on the summations in Eq. (45) are chosen so

    that an accurate representation of the stress field (which

    depends on the amount of inelastic flow) is obtained within

    each subcell. This will be discussed in Section 3.5 and

    Appendix A where the solution of the resulting equations

    for the unknown coefficients appearing in the displacementexpansions is outlined. The coefficients e

    abgijl;m;n; t

    abgijl;m;n;

    tTabgijl;m;n in the above expansions are determined as described

    below.

    The strain coefficients eabgijl;m;n are explicitly determined in

    terms of the displacement field microvariables of Eqs. (41)

    (43), using orthogonal properties of Legendre polynomials.

    Similarly, the thermal stress coefficients eTabgijl;m;n can be

    expressed in terms of the temperature field microvariables

    Tabglmn of Eq. (9). For example,

    eabg110;0;0 Wabg1100

    and

    tTabg110;0;0 Gabg11 Tabg000

    The complete sets of non-zero strain and thermal stress

    coefficients, eabgijl;m;n and t

    Tabgijl;m;n; respectively, are given in

    Appendix A.

    The stress coefficients tabgijl;m;n are expressed in terms of

    the strain coefficients, the thermal stress coefficients and the

    unknown inelastic strain distributions, by first substituting

    the Legendre polynomial representations for 1(abg), s(abg),

    and sT(abg) into the constitutive equations, Eq. (39), and

    then utilizing the orthogonality of Legendre polynomials:

    tabgijl;m;n Cijkleabgkll;m;n tTabgijl;m;n Rabgijl;m;n 47

    The Rabgijl;m;n terms represent inelastic strain distributions

    calculated in the following manner,

    Rabgijl;m;n mabgLlmn 1 1

    1 1

    1 1

    1inabgij Plza1

    Pmzb2 Pnzg3 dza1 dzb2 dzg3 48

    where

    Llmn 141 2l1 2m1 2n

    pIt should be noted that the choice of the Legendre poly-

    nomials in the above expansions is motivated by the simpli-

    city of their orthogonal properties (i.e. the weight function is

    equal to one). It is possible, of course, to use any other

    orthogonal set of polynomials.In the course of satisfying the equilibrium equations in a

    volumetric sense, it is convenient to define the following

    stress quantities:

    Sabgijl;m;n

    1

    npqrabg

    dpa =2d

    pa =2

    hqb

    =2

    hqb

    =2

    lrg =2 l

    rg =2

    xa1 l xb2 m

    xg3 nsabgij d xa1 d xb2 d xg3 49

    For l m n 0; Eq. (49) provides volume average stres-ses in the subcell, whereas for other values of ( l,m,n) higher-

    order stresses are obtained that are needed to describe the

    governing field equations of the higher-order continuum.

    These stress quantities can be evaluated explicitly in terms

    of the unknown coefficients Wabgilmn by performing the

    required volume integration upon substituting Eqs. (39),

    (40) and (41)(43) in Eq. (49). This yields the following

    non-vanishing zeroth-order and first-order stress compo-

    nents in terms of the unknown coefficients in the displace-

    ment field expansion:

    Sabg110;0;0 Cabg11 Wabg1100 Cabg12 Wabg2010 Cabg13 Wabg3001

    Gabg1 T

    abg000 R

    abg110;0;0 (50)

    Sabg111;0;0

    1

    4d

    p2a C

    abg11 W

    abg1200

    1

    12d

    p2a G

    abg11 T

    abg100

    1

    2

    3p dpa Rabg111;0;0 51

    Sabg110;1;0

    1

    4h

    q2b C

    abg12 W

    abg2020

    1

    12h

    q2b G

    abg11 T

    abg010

    1

    23p hqb Rabg110;1;0 52

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832 789

  • 8/4/2019 Higher Order Thry

    14/56

    Sabg110;0;1

    1

    4lr2g C

    abg13 W

    abg3002

    1

    12lr2g G

    abg11 T

    abg001

    1

    2

    3p lrg Rabg110;0;1 53

    with similar expressions for Sabg220;0;0; S

    abg221;0;0; S

    abg220;1;0;

    Sabg

    220;0;1; and Sabg

    330;0;0; Sabg

    331;0;0; Sabg

    330;1;0; Sabg

    330;0;1 and

    Sabg120;0;0 Cabg66 Wabg1010 Wabg2100Rabg120;0;0 54

    Sabg121;0;0

    1

    4d

    p2a C

    abg66 W

    abg2200

    1

    2

    3p dpa Rabg121;0;0 55

    Sabg120;1;0

    1

    4h

    q2b C

    abg66 W

    abg1020

    1

    2

    3p hqb Rabg120;1;0 56

    Sabg130;0;0 Cabg55 Wabg1001 Wabg3100Rabg130;0;0 57

    Sabg131;0;0 14 dp2a C

    abg55 W

    abg3200 1

    2

    3p dpa Rabg131;0;0 58

    Sabg130;0;1

    1

    4lr2g C

    abg55 W

    abg1002

    1

    2

    3p lrg Rabg130;0;1 59

    Sabg230;0;0 Cabg44 Wabg2001 Wabg3010Rabg230;0;0 60

    Sabg230;1;0

    1

    4h

    q2b C

    abg44 W

    abg3020

    1

    2

    3p hqb Rabg230;1;0 61

    Sabg23

    0;0;1

    1

    4

    lr2g C

    abg44 W

    abg2002

    1

    23p lrg R

    abg23

    0;0;1

    62

    where contracted notation has been employed for the stiff-

    ness elements Cabgijkl :

    Subsequently, satisfaction of the zeroth, first, and second

    moments of the equilibrium equations results in the follow-

    ing 24 relations among the volume-averaged first-order

    stresses Sabgijl;m;n in the different subcells (abg) of the

    (p,q,r)th generic cell, after lengthy algebraic manipulations

    (see Refs. [21,22]):

    Sabg1j1;0;0=d2a Sabg2j0;1;0=h2b Sabg3j0;0;1=l2gp;q;r 0 63where j

    1; 2; 3 and, as in the case of Eq. (17), the triplet

    (abg) assumes all permutations of the integers 1 and 2.The continuity of tractions at the subcell interfaces and

    between adjacent cells, imposed in an average sense, is

    ensured by the following relations:

    12S1bg1j1;0;0=d1 S2bg1j0;0;0 6S2bg1j1;0;0=d2p;q;r

    S2bg1j0;0;0 6S2bg1j1;0;0=d2p1;q;r 0 64

    S1bg1j0;0;0 12 S2bg1j0;0;0 3S

    2bg1j1;0;0=d2p;q;r

    12S2bg1j0;0;0 6S2bg1j1;0;0=d2p1;q;r 0 65

    12Sa1g2j0;1;0=h1 Sa2g2j0;0;0 6Sa2g2j0;1;0=h2p;q;r

    Sa2g2j0;0;0 6Sa2g2j0;1;0=h2p;q1;r 0 66

    Sa1g2j0;0;0 12 Sa2g2j0;0;0 3S

    a2g2j0;1;0=h2p;q;r

    12 S

    a2g2j0;0;0 6S

    a2g2j0;1;0=h2

    p;q1;r 0 67

    12Sab13j0;0;1=l1 Sab23j0;0;0 6Sab23j0;0;1=l2p;q;r

    Sab23j0;0;0 6Sab23j0;0;1=l2p;q;r1 0 68

    Sab13j0;0;0 12 Sab23j0;0;0 3S

    ab23j0;0;1=l2p;q;r

    12Sab23j0;0;0 6Sab23j0;0;1=l2p;q;r1 0 69

    where j 1; 2, and 3.Eqs. (64)(69) provide us with 72 additional relations

    among the zeroth-order and first-order stresses. These rela-tions, together with Eq. (63), can be expressed in terms of

    the unknown coefficients Wabgilmn by making use of Eqs.

    (50)(62), providing a total of 96 of the required 168 equa-

    tions necessary for the determination of these coefficients in

    the (p,q,r)th generic cell.

    The additional 72 relations necessary to determine the

    unknown coefficients in the displacement field expansion

    are subsequently obtained by imposing displacement conti-

    nuity conditions on an average basis at each subcell and cell

    interface. This produces,

    W1bgj000 12 d1W1bg

    j100 14

    d21 W

    1bgj200p;q;r

    W2bgj000 12 d2W2bg

    j100 14

    d22 W

    2bgj200p;q;r 70

    W2bgj000 12 d2W2bg

    j100 14

    d22 W

    2bgj200p;q;r

    W1bgj000 12 d1W1bg

    j100 14

    d21 W

    1bgj200p1;q;r 71

    Wa1gj000 12 h1Wa1g

    j010 14

    h21W

    a1gj020p;q;r

    Wa2gj000 12 h2Wa2g

    j010 14

    h22W

    a2gj020p;q;r 72

    W

    a2gj

    000

    12

    h2Wa2g

    j

    010

    14

    h22W

    a2gj

    020

    p;q;r

    Wa1gj000 12 h1Wa1g

    j010 14

    h21W

    a1gj020p;q1;r 73

    Wab1j000 12 l1Wab1

    j001 14

    l21W

    ab1j002p;q;r

    Wab2j000 12 l2Wab2

    j001 14

    l22W

    ab2j002p;q;r 74

    Wab2j000 12 l2Wab2

    j001 14

    l22W

    ab2j002p;q;r

    Wab1j000 12 l1Wab1

    j001 14

    l21W

    ab1j002p;q;r1 75

    which comprise the required additional 72 relations.

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832790

  • 8/4/2019 Higher Order Thry

    15/56

    The equilibrium equations, Eq. (63), together with the

    traction and displacement continuity equations Eqs. (64)

    (69) and Eqs. (70)(75), respectively, form 168 equations in

    the 168 Wabgilmn unknowns which govern the equilibrium of a

    subcell (abg) within the (p,q,r)th generic cell in the inter-

    ior. As in the thermal problem, a different treatment must be

    adopted for the boundary cells (1,q,r), (Np,q,r), and (p,1,r),

    (p,Nq,r), and (p,q,1), (p,q,Nr). For (1,q,r), the above relations

    are operative, except Eqs. (64) and (65), which follow from

    the continuity of tractions between a given generic cell and

    the preceding one. These 12 equations must be replaced by

    the conditions of continuity of tractions at the interior inter-

    faces of the cell (1,q,r) and by the applied tractions at x1 0: For the cell (Np,q,r), the previously derived governing

    equations are operative except for the 12 relations given

    by Eqs. (71), which are obviously not applicable. These

    are replaced by the imposed traction conditions at the

    surface x1 D: Similar arguments hold for boundary cells(p,1,r), (p,Nq,r), and (p,q,1), (p,q,Nr).

    The traction vector in the cells (1,q,r) and (Np,q,r) at thebounding surfaces must equal the applied surface loads,

    s1bg1i

    1;q;r tbottomx2;x3; x11 12 d11 76

    s2bg1i

    Np ;q;r ttopx2;x3; x21 12 dNp2 77

    where q 1; ;Nq; r 1; ;Nr; and tbottom and ttop describethe spatial variation of these loads at the bottom and top

    surfaces. Similar conditions hold for the traction vector in

    the cells (p,1,r) and (p,Nq,r) at the left and right surfaces,

    respectively,

    sa1g

    2i p;1;r

    tleftx1;x3; x12 12 h

    11 78

    sa2g2i

    p;Nq ;r trightx1;x3; x22 12 hNq2 79

    where p 1; ;Np; r 1; ;Nr; and tleft and tright describethe spatial variation of these loads at the left and right

    surfaces. Similar traction boundary conditions can be writ-

    ten for the front and back faces. Alternatively, if the front or

    back surfaces are rigidly clamped (say), then

    uab1i

    p;q;1 0; x13 12 l11 80

    uab2i

    p;q;Nr

    0; x2

    3

    1

    2lNr2

    81

    where p 1; ;Np; q 1; ;Nq: For other types of bound-ary conditions, Eqs. (76)(81) should be modified accord-

    ingly.

    Consequently, the governing equations for the interior

    and boundary cells form a system of 168NpNqNr algebraic

    equations in the unknown coefficients Wabgilmn : The final

    form of this system of equations is symbolically represented

    by

    KU f g 82where the structural stiffness matrix K contains information

    on the geometry and thermomechanical properties of the

    individual subcells (abg) within the cells comprising the

    functionally graded composite, the displacement coefficient

    vector U contains the unknown coefficients that describe the

    displacement field in each subcell, i.e.

    U

    U

    111111 ; ; U

    222NpNqNr

    where

    Uabg

    pqr Wi000; Wi100; Wi010; Wi001;

    Wi200; Wi020; Wi002abgpqr i 1; 2; 3and the mechanical force vector f contains information on

    the boundary conditions and the thermal loading effects

    generated by the applied temperature. In addition, the

    inelastic force vector g appearing on the right-hand side of

    Eq. (82) contains the inelastic effects given in terms of the

    integrals of the inelastic strain distributions that are repre-

    sented by the coefficients Rabgijl;m;n

    : These integrals depend

    implicitly on the elements of the displacement coefficient

    vector U, requiring an incremental solution of Eq. (82) at

    each point along the loading path.

    3.4. Higher-order theory vs finite-element analysis

    Despite some similarities between the finite-element

    method and the outlined coupled higher-order theory, the

    higher-order theory contains many unique features not

    found in the traditional displacement-based finite-element

    analyses. The common features include discretization of the

    microstructure into subvolumes wherein the displacement

    field is approximated by certain functions (in this casesecond-order expansions in the local subcell coordinates).

    However, the present approach is based on the volumetric

    averaging of the field equations (heat conduction and equi-

    librium) in the individual subcells which, in turn, involves

    the satisfaction of the zeroth, first, and second moments of

    these equations in a volumetric sense. In contrast, the tradi-

    tional finite-element analysis involves the minimization of

    either an energy expression or a weak form of the field

    equations taken over a subvolume (but not their higher

    moments). The satisfaction of the higher moments of the

    field equations in the present approach naturally gives rise to

    the higher moments of the heat flux and stress quantities

    given in Eqs. (10) and (49). The higher-order quantitiesappear in the higher-order moments of the field equations

    that must be satisfied. The volume-averaged zeroth and

    first-order quantities are similar to the force and moment

    resultants introduced in the standard plate theories except

    here we average throughout each subcells volume rather

    than through the entire plates thickness. Such higher-order

    quantities are not employed in the traditional finite-

    element analysis based on the second-order displacement

    expansion.

    Additional differences between the present approach and

    the traditional finite-element analysis involve the manner in

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832 791

  • 8/4/2019 Higher Order Thry

    16/56

    which the continuity of displacements and tractions are

    satisfied between the various subvolumes. In the present

    formulation, both displacement and traction continuity

    between the individual subcells is satisfied in an integral

    (average) sense, in contrast with the traditional finite-

    element analysis based on a second-order displacement

    expansion where only the displacement continuity is speci-

    fied between adjacent elements. Moreover, in the present

    approach, exact displacement field continuity between

    adjacent domains is not required, whereas finite-element

    analysis requires satisfaction of exact displacement con-

    tinuity at the nodes (even in the case of non-conformable

    elements).

    Finally, in the present theory, the governing field

    equations and the displacement and traction continuity

    conditions are always satisfied in a volumetric sense

    irrespective of the subvolume discretization. Naturally,

    more detailed (accurate) results are obtained upon further

    discretization, but further refinement is not necessary to

    ensure that the traction continuity is satisfied in an integralsense. This is in sharp contrast with the traditional dis-

    placement-based finite-element analysis where refinement

    typically leads to better satisfaction of the traction continu-

    ity in a point-wise fashion, and thus to more accurate point-

    wise results. The present theory makes possible consistent

    analysis of different microstructures at different levels of

    volume discretization, and the fully analytical nature of

    the approach allows us to do this efficiently in a volumetric

    sense, although not as accurately as the point-wise finite-

    element analysis.

    3.5. Solution of the governing equations

    The choice of an appropriate technique for the solution of

    Eq. (82) depends on the inelastic constitutive model

    employed to calculate the inelastic strain distributions in

    each subcell from which the coefficients Rabgijl;m;n can be

    generated. For instance, if the classical incremental plasti-

    city (PrandtlReuss) equations are employed to model the

    inelastic response of the matrix phase, then Mendelsons

    iterative method of successive elastic solutions is an appro-

    priate technique for the determination of the plastic strains

    needed in the solution of Eq. (82) at each increment of the

    applied load [46]. This method has been employed by

    Pindera et al. [47,48], in investigating the thermoplasticresponse of unidirectional metal matrix composites

    subjected to axisymmetric loading for those situations

    where rate effects can be neglected. An advantage of this

    solution technique is its efficiency and relative quick

    convergence even for relatively large load increments

    [49]. If, on the contrary, a unified viscoplastic constitutive

    theory is employed to model the inelastic response of the

    matrix phase, then either an implicit or an explicit technique

    can be employed to integrate the viscoplastic rate equations

    at each increment of the applied load. The integration of

    viscoplastic constitutive equations, however, may require

    a substantial computational effort due to the potentially

    stiff behavior of this class of equations.

    In the examples presented in Section 7, three constitutive

    models are employed to describe the inelastic response of

    the matrix phase, namely the classical incremental plasticity

    theory, the BodnerPartom unified viscoplasticity theory,

    and the power-law creep model, briefly outlined in

    Appendix A. The plasticity theory is employed to efficiently

    model the inelastic constitutive response of the matrix phase

    when rate effects can be neglected, whereas the more

    computationally intensive BodnerPartom theory and the

    power-law creep model are employed for those situations

    where rate-dependent deformation must be taken into

    account. It should be noted that the present formulation is

    sufficiently general to accommodate other types of unified

    viscoplastic theories.

    4. Specialization of the higher-order theory

    The generalized (i.e. three-directional) higher-order

    theory can easily be specialized for materials functionally

    graded in a given plane or along a specific direction. In the

    former case we obtain the two-directional version of the

    higher-order theory referred to in previous communications

    as HOTFGM-2D, and in the latter case we obtain the one-

    directional version referred to as HOTFGM-1D. Herein, we

    briefly show how the generalized higher-order theory

    presented in the preceding section can be specialized to

    the above two cases.

    4.1. Two-directional higher-order theory

    To maintain consistency with our previous communica-

    tions, we consider a composite functionally graded in the

    x2x3 plane (see Fig. 7). Such a composite has a finite thick-

    ness H and finite length L in the functionally graded direc-

    tions and extends to infinity in the periodic x1 direction. In

    the x2x3 plane, the composite is reinforced by an arbitrary

    distribution of infinitely long fibers oriented along the x1axis or finite-length inclusions that have periodic spacing

    in the direction of the x1 axis. Both aligned and arbitrary

    fiber or inclusion architectures in the x2x3 plane, Fig. 7(a)

    and (b), respectively, are admissible. The microstructure of

    the heterogeneous composite is discretized into Nq and Nr

    cells in the intervals 0 x2 H; 0 x3 L in the x2x3plane. The indices q and r of the generic cell p; q; r thatidentify its location in the x2x3 plane have the same ranges

    as in the generalized higher-order theory, while the range of

    the index p is infinite due to the microstructures periodicity

    in the x1 direction. This index can therefore be either

    suppressed or retained with the understanding that its

    value is indeterminate. The dimensions of the generic cell

    along the x1 axes, d1, d2, are fixed for a given configuration

    whereas the dimensions along the x2 and x3 axes, hq1 ; h

    q2 ;

    and lr1 ; lr2 ; can vary in an arbitrary fashion as previously

    discussed.

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832792

  • 8/4/2019 Higher Order Thry

    17/56

    The temperature distribution T(abg)

    in the subcell (ab g)of the (p,q,r)th cell is approximated as follows:

    Tabg Tabg000 xb2 Tabg010 xg3 Tabg001

    1

    23 xa21

    d2a

    4

    2 3T

    abg200

    1

    23 x

    b22

    hq2b

    4

    2 3T

    abg020

    1

    23 x

    g23

    lr2g

    4

    2 3T

    abg002 83

    In contrast with the generalized theory, the above represen-tation does not contain a linear term in the local coordinates

    xa1 ; i.e.

    Tabg100 0 84

    (see Eq. (9)). This follows directly from the assumed peri-

    odicity in the x1 direction and symmetry with respect to the

    lines xa1 0 for a 1 and 2. In the two-directional theory,therefore, the temperature field within the subcell (abg) is

    described by six unknown coefficients instead of seven.

    Thus for a composite with Nq rows and Nr columns,

    48NqNr unknown thermal coefficients must be determined

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832 793

    Fig. 7. Composite with non-periodic fiber distributions in the x2x3 plane: (a) aligned inclusion architecture, (b) arbitrary inclusion architecture.

  • 8/4/2019 Higher Order Thry

    18/56

    in order to be able to describe the temperature field in every

    subcell (abg) of every generic cell p; q; r:The determination of these coefficients follows the same

    methodology as outlined previously. The system of

    equations for the unknown coefficients within the interior

    subcells can be obtained from the generalized solution by

    using Eq. (84) and suppressing those equations which

    involve the variations associated with the index p appearing

    in the superscripts, i.e. p 1; q; r; p; q; r and p; q; r;p 1; q; r: Thus the satisfaction of the zeroth, first, andsecond moment of the steady-state heat equation, Eq. (7),

    results in eight relations among the first-order heat fluxes

    Qabgil;m;n in the different subcells (abg) of the (p,q,r)th

    generic cell which are the same as Eq. (17). Twenty addi-

    tional relations are obtained among the zeroth and first-order

    heat fluxes upon satisfaction of the continuity of heat fluxes

    at the subcell interfaces and individual cells, whereas the

    same number of additional equations is obtained upon

    imposing thermal continuity at the subcell interfaces and

    individual cells. The boundary subcells are dealt with inthe same manner as discussed previously. The above

    relations provide the required 48NqNr equations for the

    48NqNr unknown coefficients Tabg000 ; T

    abg010 ; T

    abg001 ; T

    abg200 ;

    Tabg020 ; T

    abg002 ; and can be expressed in the same form as

    Eq. (36).

    Next, the displacement field in the subcell (abg) of the

    (p,q,r)th generic cell is approximated by a second-order

    expansion in the local coordinates xa1 ; xb2 ; and x

    g3 as

    follows, taking into account symmetry considerations:

    uabg1 Wabg1000 xa1 Wabg1100 85

    uabg2 Wabg2000 xb2 Wabg2010 xg3 Wabg2001

    1

    23 xa21

    d2a

    4

    2 3W

    abg2200

    1

    23 x

    b22

    hq2b

    4

    2 3W

    abg2020

    1

    23 x

    g23

    lr2g

    4

    2 3W

    abg2002 86

    uabg3 Wabg3000 xb2 Wabg3010 xg3 Wabg3001

    1

    23 xa21

    d2a

    4

    2 3W

    abg3200

    1

    23 x

    b22

    hq2b

    4

    2 3W

    abg3020

    1

    23 x

    g23

    lr2g

    4

    2 3W

    abg3002 87

    The first equation does not contain the linear terms in the

    local coordinates xb2 and x

    g3 ; i.e.

    Wabg1010 Wabg1001 0 88

    This follows from the assumed periodicity in the x1 direction

    and symmetry with respect to xa1 0 a 1; 2: Further,the presence of the constant term W

    abg

    1000 in the firstequation, that represents subcell center x1 displacements,

    produces uniform composite strain 111 upon application of

    a partial homogenization scheme described in the next

    section. This partial homogenization, which couples the

    present higher-order theory and an RVE-based theory,

    leads to an overall behavior of a composite, functionally

    graded in the x2 and x3 directions, that can be described as

    a generalized plane strain in the periodic x1 direction. Plane

    strain behavior in the periodic direction is obtained by

    setting the constant term to zero [27]. The absence of the

    higher-order terms in the first equation, i.e.

    Wabg

    1200 Wabg

    1020 Wabg

    1002 0 89is a direct consequence of the periodicity in the out-of-plane

    direction. The last two equations, on the other hand, do not

    contain linear terms in the local coordinate xa1 ; i.e.

    Wabg2100 Wabg3100 0 90

    in order to ensure symmetry of the deformation field in the

    out-of-plane direction. In the two-directional theory, there-

    fore, the displacement field within the subcell (abg) is

    described by 13 unknown coefficients instead of 21, noting

    that the constant term Wabg1000 in the first equation will be

    replaced by the applied strain111 through homogenization.Thus for a composite with Nq rows and Nr columns, 104NqNr

    unknown coefficients must be determined in order to be able

    to describe the displacement field in every subcell (abg) of

    every generic cell (p,q,r).

    As in the thermal case, the determination of these coef-

    ficients follows the same methodology as outlined

    previously. The system of equations for the unknown

    coefficients within the interior subcells can be obtained

    from the generalized solution by using Eqs. (88) (90),

    and suppressing those equations which involve the varia-

    tions associated with the index p appearing in the super-

    scripts. Thus the satisfaction of the zeroth, first, and

    second moment of the equilibrium equations, Eq. (37),results in 16 relations among the volume averaged first-

    order stresses Sabgijl;m;n in the different subcells (abg) of

    the (p,q,r)th generic cell which are the same as Eq. (63).

    Forty-four additional relations are obtained among the

    zeroth and first-order stresses upon satisfaction of the

    continuity of tractions at the subcell interfaces and indi-

    vidual cells. The remaining 44 relations necessary to

    determine the unknown coefficients in the displacement

    field expansion are obtained by imposing a partial homo-

    genization procedure in the periodic x1 direction together

    with displacement continuity conditions applied on the

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832794

  • 8/4/2019 Higher Order Thry

    19/56

    average basis at each subcell and cell interface in the

    functionally graded x2 and x3 directions.

    The equations resulting from the partial homogenization

    procedure are:

    d1W1bg1100 d2W2bg1100p;q;r d1 d2 111 91

    where 111 is the unknown uniform far-field strain in the

    x2 x3 plane that is either specified or determined. In the

    latter case, the additional equation for the unknown 111 is

    obtained from the imposed s11: These equations replace

    Eq. (70) in the generalized higher-order theory solution

    for j 1: On the contrary, the remaining equations obtainedfrom the imposition of the continuity of displacements at

    each subcell interface of the (p,q,r)th generic cell, as well as

    between adjacent cells, in the functionally graded directions

    are obtained by specializing Eqs. (71)(75) in the manner

    described previously.

    Finally, the boundary subcells are also dealt with in the

    same manner as discussed previously. The above relations

    provide the required 104NqNr equations for the 104NqNrunknown coefficients

    Wabg1100; W

    abg2000; W

    abg2010; W

    abg2001; W

    abg2200; W

    abg2020; W

    abg2002

    and

    Wabg3000; W

    abg3010; W

    abg3001; W

    abg3200; W

    abg3020; W

    abg3002

    and can be expressed in the same form as Eq. (82).

    4.2. One-directional higher-order theory

    To maintain consistency with our original communica-

    tions, we consider a functionally graded composite in the x1direction. Such a composite has a finite depth D in the

    functionally graded direction and extends to infinity in

    the x2x3 plane, Fig. 8. The composite is reinforced by

    periodic arrays of fibers in the direction of the x2 axis or

    the x3 axis, or both. In the direction of the x1 axis, the fiber

    spacing between adjacent arrays may vary. The reinforcing

    fibers can be either continuous or finite-length, Fig. 8(a) or

    (b), respectively.The microstructure of the heterogeneous composite is

    discretized into Np cells in the interval 0 x1 D: The

    index p of the generic cell (p,q,r) that identifies its location

    along the x1 axis has the same range as in the generalized

    higher-order theory, while the range of the indices q and ris

    infinite due to the microstructures periodicity in the x2 and

    x3 directions. These indices can therefore be either

    suppressed or retained with the understanding that their

    values are indeterminate. The dimensions of the generic

    cell along the x2 and x3 axes, h1, h2, and l1, l2, are fixed for

    a given configuration whereas the dimensions along the x1

    axis, dp1 ; d

    p2 ; can vary in an arbitrary fashion as previously

    discussed.

    The temperature distribution T(abg) in the subcell

    (abg) of the (p,q,r)th generic cell is approximated as

    follows:

    Tabg

    Tabg

    000 xa

    1 Tabg

    100 1

    2 3 xa

    2

    1

    dp2a

    42 3Tabg200

    1

    23 x

    b22

    h2b

    4

    2 3T

    abg020

    1

    23 x

    g23

    l2g

    4

    2 3T

    abg002 (92)

    The above representation does not contain linear terms in

    the local coordinates xb2 and x

    g3 ; i.e.

    Tabg

    010

    T

    abg

    001

    0

    93

    because of the assumed periodicity in these directions. In

    the one-directional theory, therefore, the temperature field

    within the subcell (abg) is described by five unknown

    coefficients instead of seven. Thus for a composite with

    Np planar slices in the functionally graded direction

    containing inclusions with periodic spacing in the x2

    x3 plane, 40Np unknown thermal coefficients must be

    determined in order to be able to describe the tempera-

    ture field in every subcell (abg) of every generic cell

    (p,q,r).

    The system of equations for the unknown coefficients

    within the interior subcells can be obtained from thegeneralized solution by using Eq. (93) and suppressing

    those equations which involve the variations associated

    with the indices q and r appearing in the superscripts,

    i.e. (p,q 1,r), (p,q,r) and (p,q,r), (p,q 1,r) for the q

    variation, and p; q; r 1; (p,q,r) a nd (p,q,r), p; q; r1 for the r variation. Thus the satisfaction of the zeroth,first, and second moment of the steady-state heat equation,

    Eq. (7), results in eight relations among the first-order

    heat fluxes Qabgil;m;n in the different subcells (abg) of the

    (p,q,r)th generic cell which are the same as Eq. (17).

    Sixteen additional relations are obtained among the zeroth

    and first-order heat fluxes upon satisfaction of the conti-

    nuity of heat fluxes at the subcell interfaces and individualcells, whereas the same number of additional equations is

    obtained upon imposing thermal continuity at the subcell

    interfaces and individual cells. The boundary subcells are

    dealt with in the same manner as discussed previously.

    The above relations provide the required 40Np equations

    for the 40Np unknown coefficients Tabg000 ; T

    abg100 ; T

    abg200 ;

    Tabg020 ; T

    abg002 ; and can be expressed in the same form as

    Eq. (36).

    Next, the displacement field in the subcell (a bg) of the

    (p,q,r)th generic cell is approximated by a second-order

    expansion in the local coordinates xa1 ; xb2 ; and x

    g3 as

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832 795

  • 8/4/2019 Higher Order Thry

    20/56

    follows, taking into account symmetry considerations:

    uabg1

    Wabg1000

    xa1

    Wabg1100

    1

    23 xa2

    1

    dp2a

    42 3Wabg1200

    1

    23 x

    b21

    h2b

    4

    2 3W

    abg1020

    1

    23 x

    g23

    l2g

    4

    2 3W

    abg1002 (94)

    uabg2 Wabg2000 xb2 Wabg2010 95

    uabg3 Wabg3000 xg3 Wabg3001 96

    The first equation does not contain linear terms in the local

    coordinates xb2 and x

    g3 ; i.e.

    Wabg

    1010 Wabg

    1001 0 97in order to ensure symmetry of the deformation field in the

    periodic directions. The second equation does not contain

    linear terms in the local coordinates xa1 and xg3 ; i.e.

    Wabg2100 Wabg2001 0 98

    while the third equation does not contain linear terms in the

    local coordinates xa1 and xb2 ; i.e.

    Wabg3100 Wabg3010 0 99

    This follows from the assumed periodicity in the x2 and x3

    J. Aboudi et al. / Composites: Part B 30 (1999) 777832796

    Fig. 8. Composite with non-periodic fiber distributions in the x1 direction: (a) unidirectionally reinforced material, (b) particulate inclusion reinforced material.

    RCS denotes the representative cross section.

  • 8/4/2019 Higher Order Thry

    21/56

    directions and symmetry with respect to xb2 0 b 1; 2

    and xg3 0 g 1; 2: The constant terms Wabg2000 and

    Wabg3000 in these equations, that represent subcell center

    displacements in the x2 and x3 directions, respectively,

    produce uniform composite strains 122 and 133 upon appli-

    cation of a partial homogenization scheme. This partial

    homogenization leads to an overall behavior of a composite

    that can be described as a generalized plane strain in the

    periodic directions. Plane strain behavior in the periodic

    directions is obtained by setting the constant terms to zero

    [21]. The absence of the higher-order terms in the last two

    equations, i.e.

    Wabg2200 Wabg2020 Wabg2002 0 100

    and

    Wabg3200 Wabg3020 Wabg3002 0 101

    is a direct consequence of the periodicity in the x2 and x3

    directions. In the one-directional theory, therefore, thedisplacement field within the subcell (abg) is described

    by seven unknown coefficients instead of 21, noting that

    the constant terms Wabg2000 and W

    abg3000 in the last two equa-

    tions will be replaced by the applied strains 122 and 122through homogenization. Thus for a composite with Npplanar slices in the functionally graded direction, 56Npunknown coefficients must be determined in order to be

    able to describe the displacement field in every subcell

    (abg) of every generic cell (p,q,r).

    The system of equations for the unknown coefficients

    within the interior subcells is obtained from the generalized

    solution by using Eqs. (97)(101) and suppressing those

    equations which involve the variations associated with the

    indices q and rappearing in the superscripts. Thus the satis-

    faction of the zeroth, first, a