tucker halpintsai 2

Upload: biju-balan-laly

Post on 06-Jul-2018

217 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/18/2019 Tucker Halpintsai 2

    1/34

    Stiffness Predictions for

    Unidirectional Short-Fiber Composites:Review and Evaluation

    Charles L. Tucker III

    Department of Mechanical and Industrial Engineering

    University of Illinois at Urbana-Champaign

    1206 W. Green St.

    Urbana, IL 61801

    Erwin Liang

    GE Corporate Research and Development

    Schenectady, NY 12301

    September 29, 1998

    To appear in Composites Science and Technology

    Abstract

    Micromechanics models for the stiffness of aligned short-fiber composites are reviewed

    and evaluated. These include the dilute model based on Eshelby’s equivalent inclusion, the

    self-consistent model for finite-length fibers, Mori-Tanaka type models, bounding models, the

    Halpin-Tsai equation and its extensions, and shear lag models. Several models are found

    to be equivalent to the Mori-Tanaka approach, which is also equivalent to the generalization

    of the Hashin-Shtrikman-Walpole lower bound. The models are evaluated by comparison to

    finite element calculations using periodic arrays of fibers, and to Ingber and Papathanasiou’s

    boundary element results for random arrays of aligned fibers. The finite element calculations

    provideE 

    1 1

      ,E 

    2 2

      , 

    1 2

      , and 

    2 3

      for a range of fiber aspect ratios and packing geometries, with

    other properties typical of injection-molded thermoplastic matrix composites. The Halpin-

    Tsai equations give reasonable estimates for stiffness, but the best predictions come from the

    Mori-Tanaka model and the bound interpolation model of Lielens et al.

    To whom correspondence should be addressed

  • 8/18/2019 Tucker Halpintsai 2

    2/34

    1 Introduction

    This paper reviews and evaluates models that predict the stiffness of short-fiber composites. The

    overall goal of the research is to improve processing-property predictions for injection-molded

    composites. The polymer processing communityhas made substantial progress in modeling process-

    induced fiber orientation, particularly in injection molding, and these results are now routinely usedto predict mechanical properties. The purpose of this paper is to review the relevant micromechan-

    ics literature, and to provide a critical evaluation of the available models. Real injection-molded

    composites invariably have misoriented fibers of highly variable length, but aligned-fiber proper-

    ties are always calculated as a prelude to modeling the more realistic situation. Hence, we focus

    here on composites having aligned fibers with uniform length and mechanical properties. The

    modeling of composites with distributions of fiber orientation and fiber length, and the treatment

    of multiple types of reinforcement, will be discussed in a subsequent paper 1.

    In selecting models for consideration, we impose the general requirements that each model

    must include the effects of fiber and matrix properties and the fiber volume fraction, include the

    effect of fiber aspect ratio, and predict a complete set of elastic constants for the composite. Any

    model not meeting these criteria was excluded from consideration.All of the models use the same basic assumptions:

     

    The fibers and the matrix are linearly elastic, the matrix is isotropic, and the fibers are either

    isotropic or transversely isotropic.

      The fibers are axisymmetric, identical in shape and size, and can be characterized by an

    aspect ratio ` = d   .

     

    The fibers and matrix are well bonded at their interface, and remain that way during defor-

    mation. Thus, we do not consider interfacial slip, fiber-matrix debonding or matrix micro-

    cracking.

    Section 2 presents some important preliminary concepts, emphasizing strain-concentration ten-

    sors and their relationship to composite stiffness. Section 3 then reviews the various theories.

    Section 4 compares and evaluates the available models. We use finite element computations of 

    periodic arrays of short fibers to provide reference properties, since it has not proved possible to

    create physical specimens with perfectly aligned fibers. A subsequent paper 1 will compare model

    predictions to experiments on well-characterized composites with misaligned fibers.

    2 Preliminaries

    2.1 Notation

    Vectors will be denoted by lower-case Roman letters, second-order tensors by lower-case Greek 

    letters, and fourth-order tensors by capital Roman letters. Whenever possible, vectors and tensors

    are written as boldface characters; indicial notation is used where necessary.

    A subscript or superscriptf 

      indicates a quantity associated with the fibers, andm 

      denotes

    a matrix quantity. Thus, the fibers have Young’s modulus E f

      and Poisson ratio  f

      , while the

    corresponding matrix properties are E m 

      and  m 

      .

    1

  • 8/18/2019 Tucker Halpintsai 2

    3/34

    The symbol I    represents the fourth-order unit tensor. C    and S    denote the stiffness and compli-

    ance tensors, respectively, and     and "    are the total stress and infinitesimal strain tensors. Hence,

    the constitutive equations for the fiber and matrix materials are

     

    f

    =  C 

    f

    f (1)

     

    =  C 

    (2)

    2.2 Average Stress and Strain

    Let x  denote the position vector. When a composite material is loaded, the pointwise stress field

      ( x  ) and the corresponding strain field "  ( x  )  will be non-uniform on the microscale. The solution

    of these non-uniform fields is a formidable problem. However, many useful results can be obtained

    in terms of the average stress and strain 2. We now define these averages.

    Consider a representative averaging volume V    . Choose V    large enough to contain many fibers,

    but small compared to any length scale over which the average loading or deformation of the

    composite varies. The volume-average stress     is defined as the average of the pointwise stress

      ( x  )

    over the volumeV 

      :

       

      ( x  ) d V 

    (3)

    The average strain "    is defined similarly.

    It is also convenient to define volume-average stresses and strains for the fiber and matrix

    phases. To obtain these, first partition the averaging volume V    into the volume occupied by the

    fibersV 

    f

      and the volume occupied by the matrixV 

      . We consider only two-phase composites, so

    that

    V  =  V 

    f

    +  V 

    (4)

    The fiber and matrix volume fractions are simply v f

    =  V 

    f

    = V   and v m 

    =  V 

    = V   and, since only

    fibers and matrix are present,v 

    +  v 

    f

    = 1 

      .

    The average fiber and matrix stresses are the averages over the corresponding volumes,

     

    f

     

    f

    f

      ( x  ) d V  a n d   

     

      ( x  ) d V  (5)

    The average strains for the fiber and matrix are defined similarly.

    The relationships between the fiber and matrix averages and the overall averages can be derived

    from the preceding definitions; they are

      =  v 

    f

     

    f

    +  v 

     

    m  (6)

    "  =  v 

    f

    f

    +  v 

    (7)

    An important related result is the  average strain theorem. Let the averaging volume V    be

    subjected to surface displacements u  0 ( x  )  consistent with a uniform strain "  0  . Then the average

    strain within the region is

    "  =  " 

    0 (8)

    This theorem is proved2 by substituting the definition of the strain tensor "    in terms of the dis-

    2

  • 8/18/2019 Tucker Halpintsai 2

    4/34

    placement vector u    into the definition of average strain "   , and applying Gauss’s theorem. The

    result is

    i j

     

    0

    i

    j

    +  n 

    i

    0

    j

    d S  (9)

    whereS 

      denotes the surface of V 

      andn 

      is a unit vector normal tod S 

      . The average strain withina volume

      is completely determined by the displacements on the surface of the volume, so dis-

    placements consistent with a uniform strain must produce the identical value of average strain. A

    corollary of this principle is that, if we define a perturbation strain "  C  ( x  )  as the difference between

    the local strain and the average,

    ( x  )   "  ( x  ) ?  "  (10)

    then the volume-average of  "  C  ( x  )  must equal zero:

    ( x  ) d V  = 0  (11)

    The corresponding theorem for average stress also holds. Thus, if surface tractions consistent with

    0  are exerted onS 

      then the average stress is

      =   

    0 (12)

    2.3 Average Properties and Strain Concentration

    The goal of micromechanics models is to predict the average elastic properties of the composite,

    but even these need careful definition. Here we follow the direct approach3. Subject the represen-

    tative volume V    to surface displacements consistent with a uniform strain "  0  ; the average stiffness

    of the composite is the tensor C    that maps this uniform strain to the average stress. Using eqn (8)we have

      =  C  "  (13)

    The average compliance S    is defined in the same way, applying tractions consistent with a uniform

    stress   0  on the surface of the averaging volume. Then, using eqn (12),

    "  =  S    (14)

    It should be clear that S  =  C  ?  1  . Other authors define the average stiffness and compliance through

    the integral of the strain energy over V    ; this is equivalent to the direct approach2,4.

    An important related concept, first introduced by Hill2, is the idea of strain- and stress-concentration

    tensors A    and B    . These are essentially the ratios between the average fiber strain (or stress) and

    the corresponding average in the composite. More precisely,

    f

    =  A  "  (15)

     

    f

    =  B    (16)

    3

  • 8/18/2019 Tucker Halpintsai 2

    5/34

    A and  B are fourth-order tensors and, in general, they must be found from a solution of the

    microscopic stress or strain fields. Different micromechanics models provide different ways to

    approximate A or B. Note that A and B have both the minor symmetries of a stiffness or compliance

    tensor, but lack the major symmetry. That is,

    i j k l

    =  A 

    j i k l

    =  A 

    i j l k (17)

    but in general,

    i j k l

    6=  A 

    k l i j

    (18)

    For later use it will be convenient to have an alternate strain concentration tensor ^A    that relates

    the average fiber strain to the average matrix strain,

    f

    ^

    A  " 

    m  (19)

    This is related to A    by

    A  = 

    ^

    h

    ( 1  ?  v 

    f

    ) I  +  v 

    f

    ^

    i

    ?  1

    (20)

    so the two forms are easily interchanged.

    Using equations now in hand, one can express the average composite stiffness in terms of the

    strain-concentration tensor  A and the fiber and matrix properties 2. Combining eqns (1), (2), (6),

    (7), (13), and (15), one obtains

    C  =  C 

    +  v 

    f

     

    f

    ?  C 

    A  (21)

    The dual equation for the compliance is

    S  =  S 

    +  v 

    f

     

    f

    ?  S 

    (22)

    Equations (21) and (22) are not independent, since S  = C ?  1

      . Hence, the strain-concentration

    tensor A and the stress-concentration tensor B are not independent either. The choice of which one

    to use in any instance is a matter of convenience.

    To illustrate the use of the strain-concentration and stress-concentration tensors, we note that

    the Voigt average corresponds to the assumption that the fiber and the matrix both experience the

    same, uniform strain. Then "  f =  "   , A  =  I   , and from eqn (21) the composite modulus is

    V o i g t

    =  C 

    +  v 

    f

     

    f

    ?  C 

    (23)

    =  v 

    f

    f

    +  v 

    Recall that the Voigt average is an upper bound on the composite modulus. The Reuss average

    assumes that the fiber and matrix both experience the same, uniform stress. This means that the

    stress-concentration tensor B equals the unit tensor  I, and from (22) the compliance is

    4

  • 8/18/2019 Tucker Halpintsai 2

    6/34

    εT 

    (a) (b) (c)

    εC (x)

    Figure 1: Eshelby’s inclusion problem. Starting from the stress-free state (a), the inclusion under-

    goes a stress-free transformation strain "  T   (b). Fitting the inclusion and matrix back together (c)

    produces the strain state "  C  ( x  )  in both the inclusion and the matrix.

    R e u s s

    =  S 

    +  v 

    f

     

    f

    ?  S 

    (24)

    =  v 

    f

    f

    +  v 

    This represents a lower bound on the stiffness of the composite.

    3 Theories

    3.1 Eshelby’s Equivalent Inclusion

    A fundamental result used in several different models is Eshelby’s equivalent inclusion5,6. Eshelby

    solved for the elastic stress field in and around an ellipsoidalparticle in an infinitematrix. By letting

    the particle be a prolate ellipsoid of revolution, one can use Eshelby’s result to model the stress

    and strain fields around a cylindrical fiber.

    Eshelby first posed and solved a different problem, that of a homogeneous inclusion (Fig. 1).

    Consider an infinite solid body with stiffness C  m   that is initially stress-free. All subsequent strains

    will be measured from this state. A particular small region of the body will be called the inclusion,

    and the rest of the body will be called the matrix. Suppose that the inclusion undergoes some type

    of transformation such that, if it were a separate body, it would acquire a uniform strain" 

      withno surface traction or stress. "  T   is called the transformation strain, or the eigenstrain. This strain

    might be acquired through a phase transformation, or by a combination of a temperature change

    and a different thermal expansion coefficient in the inclusion. In fact the inclusion is bonded to the

    matrix, so when the transformation occurs the whole body develops some complicated strain field

    ( x  ) relative to its shape before the transformation. Within the matrix the stress   m   is simply the

    stiffness times this strain,

    5

  • 8/18/2019 Tucker Halpintsai 2

    7/34

    Cm

    C f 

    εT 

    ε A

    (a) (b)

    σ I 

    Figure 2: Eshelby’s equivalent inclusion problem. The inclusion (a) with transformation strain

    has the same stress  I

      and strain as the inhomogeneity (b) when both bodies are subject to afar-field strain "  A 

     

    ( x  ) =  C 

    ( x  )

    (25)

    but within the inclusion the transformation strain does not contribute to the stress, so the inclusion

    stress is

     

    I

    =  C 

     

    ?  " 

    (26)

    The key result of Eshelby was to show that within an ellipsoidal inclusion the strain" 

      is uniform,and is related to the transformation strain by

    =  E  " 

    T  (27)

    E  is called Eshelby’s tensor , and it depends only on the inclusion aspect ratio and the matrix elastic

    constants. A detailed derivation and applications are given by Mura 7, and analytical expressions

    for Eshelby’s tensor for an ellipsoid of revolution in an isotropic matrix appear in many papers8–12.

    The strain field "  C  ( x  )  in the matrix is highly non-uniform 13, but this more complicated part of the

    solution can often be ignored.

    The second step in Eshelby’s approach is to demonstrate an equivalence between the homoge-

    neous inclusion problem and an inhomogeneous inclusion of the same shape. Consider two infinite

    bodies of matrix, as shown in Fig. 2. One has a homogeneous inclusion with some transformation

    strain "  T   ; the other has an inclusion with a different stiffness C  f  , but no transformation strain.

    Subject both bodies to a uniform applied strain "  A   at infinity. We wish to find the transformation

    strain "  T   that gives the two problems the same stress and strain distributions.

    For the first problem the inclusion stress is just eqn (26) with the applied strain added,

     

    I

    =  C 

     

    +  " 

    ?  " 

    (28)

    6

  • 8/18/2019 Tucker Halpintsai 2

    8/34

    while the second problem has no "  T   but a different stiffness, giving a stress of 

     

    I

    =  C 

    f

     

    +  " 

    (29)

    Equating these two expressions gives the transformation strain that makes the two problems equiv-

    alent. Using eqn (27) and some rearrangement, the result is

    h

     

    f

    ?  C 

    i

     

    f

    ?  C 

    A  (30)

    Note that "  T   is proportional to "  A   , which makes the stress in the equivalent inhomogeneity pro-

    portional to the applied strain.

    3.2 Dilute Eshelby Model

    One can use Eshelby’s result to find the stiffness of a composite with ellipsoidal fibers at dilute

    concentrations. Recall from eqn (21) that to find the stiffness one only has to find the strain-

    concentration tensor A. To do this, first note that for a dilute composite the average strain is iden-

    tical to the applied strain,

    "  =  " 

    A  (31)

    since this is the strain at infinity. Also, from Eshelby, the fiber strain is uniform, and is given by

    f

    =  " 

    +  " 

    C  (32)

    where the right-hand side is evaluated within the fiber. Now write the equivalence between the

    stresses in the homogeneous and the inhomogeneous inclusions, eqns (28) and (29),

    f

     

    +  " 

    =  C 

     

    +  " 

    ?  " 

    (33)

    then use eqns (27), (31) and (32) to eliminate "  T   , "  A   and "  C   from this equation, giving

    h

    I  +  E S 

     

    f

    ?  C 

    i

    f

    =  "  (34)

    Comparing this to eqn (15) shows that the strain-concentration tensor for Eshelby’s equivalent

    inclusion is

    E s h e l b y

    h

    I  +  E S 

     

    f

    ?  C 

    i

    ?  1

    (35)

    This can be used in eqn (21) to predict the moduli of aligned-fiber composites, a result first devel-

    oped by Russel14

    . Calculations using this model to explore the effects of particle aspect ratio onstiffness are presented by Chow15.

    While Eshelby’s solution treats only ellipsoidal fibers, the fibers in most short-fiber composites

    are much better approximated as right circular cylinders. The relationship between ellipsoidal and

    cylindrical particles was considered by Steif and Hoysan16, who developed a very accurate finite

    element technique for determining the stiffening effect of a single fiber of given shape. For very

    short particles, ` = d  = 4   , they found reasonable agreement for E 1 1

      by letting the cylinder and the

    ellipsoid have the same ` = d   . The ellipsoidal particle gave a slightly stiffer  composite, with the

    7

  • 8/18/2019 Tucker Halpintsai 2

    9/34

    difference between the two results increasing as the modulus ratio E f

    = E 

      increased. Henceforth

    we will use the cylinder aspect ratio in place of the ellipsoid aspect ratio in Eshelby-type models.

    Because Eshelby’s solution only applies to a single particle surrounded by an infinite matrix,

    E s h e l b y is independent of fiber volume fraction and the stiffness predicted by this model increases

    linearly with fiber volume fraction. Modulus predictions based on eqns (35) and (21) should be

    accurate only at low volume fractions, say up to v  f  of 1%. The more difficult problem is to findsome way to include interactions between fibers in the model, and so produce accurate results at

    higher volume fractions. We next consider approaches for doing that.

    3.3 Mori-Tanaka Model

    A family of models fornon-dilute compositematerials has evolved from a proposal originally made

    by Mori and Tanaka17. Benveniste18 has provided a particularly simple and clear explanation of 

    the Mori-Tanaka approach, which we use here to introduce the approach.

    We have already introduced the strain-concentration tensor in eqn (15). Suppose that a com-

    posite is to be made of a certain type of reinforcing particle, and that, for a single particle in an

    infinite matrix, we know the dilute strain-concentration tensor A  E s h e l b y  ,

    f

    =  A 

    E s h e l b y

    "  (36)

    The Mori-Tanaka assumption is that, when many identical particles are introduced in the compos-

    ite, the average fiber strain is given by

    f

    =  A 

    E s h e l b y

    m  (37)

    That is, within a concentrated composite each particle ‘sees’ a far-field strain equal to the average

    strain in the matrix. Using the alternate strain concentrator defined in eqn (19), the Mori-Tanaka

    assumption can be re-stated as^

    M T 

    =  A 

    E s h e l b y (38)

    Equation (20) then gives the Mori-Tanaka strain concentrator as

    M T 

    =  A 

    E s h e l b y

    h

    ( 1  ?  v 

    f

    ) I  +  v 

    f

    E s h e l b y

    i

    ?  1

    (39)

    This is the basic equation for implementing a Mori-Tanaka model.

    The Mori-Tanaka approach for modeling composites was first introduced by Wakashima, Ot-

    suka and Umekawa19 for modeling thermal expansions of composites with aligned ellipsoidal

    inclusions. (Mori and Tanaka’s paper 17 treats only the homogeneous inclusion problem, and says

    nothing about composites). Mori-Tanaka predictions for the longitudinal modulus of a short-fiber

    composite were first developed by Taya and Mura8 and Taya and Chou9, whose work also included

    the effects of cracks and of a second type of reinforcement. Weng20 generalized their method, and

    Tandon and Weng11 used the Mori-Tanaka approach to develop equations for the complete set of 

    elastic constants of a short-fiber composite. Tandon and Weng’s equations for the plane-strain bulk 

    modulus k 2 3

      and the major Poisson ratio  1 2

      must be solved iteratively. However, this iteration can

    be avoided by using an alterate formula for  1 2

      ; details are given in Appendix A.

    The usual development of the Mori-Tanaka model 8,9,11 differs somewhat from Benveniste’s

    8

  • 8/18/2019 Tucker Halpintsai 2

    10/34

    explanation. For an average applied stress     , the reference strain "  0  is defined as the strain in a

    homogeneous body of matrix at this stress,

      =  C 

    0 (40)

    Within the composite the average matrix strain differs from the reference strain by some perturba-tion ~ "  m   ,

    =  " 

    0

    m  (41)

    A fiber in the composite will have an additional strain perturbation ~ "  f  , such that

    f

    =  " 

    0

    f (42)

    while the equivalent inclusion will have this strain plus the transformation strain "  T   . The stress

    equivalence between the inclusion and the fiber then becomes

    f

     

    0

    f

    =  C 

     

    0

    f

    ?  " 

    (43)

    Compare this to the dilute version, eqn (33), noting that "  A   in the dilute problem is equivalent to

    ( " 

    0

    ) here. The development is completed by assuming that the extra fiber perturbation is

    related to the transformation strain by Eshelby’s tensor,

    f

    =  E  " 

    T  (44)

    Combining this with eqns (41) and (42) reveals that eqn (44) contains the essential Mori-Tanaka

    assumption: the fiber in a concentrated composite sees the average strain of the matrix.

    Some other micromechanics models are equivalent to the Mori-Tanaka approach, though this

    equivalence has not always been recognized. Chow21 considered Eshelby’s inclusion problem and

    conjectured that in a concentrated composite the inclusion strain would be the sum of two terms:the dilute result given by Eshelby (27) and the average strain in the matrix.

    ( " 

    )

    f

    =  E  " 

    + ( " 

    )

    m  (45)

    This can be combined with the definition of the average strain from eqn (7) to relate the inclusion

    strain ( "  C  ) f  to the transformation strain "  T   :

    ( " 

    )

    f

    = ( 1  ?  v 

    f

    ) E  " 

    T  (46)

    Chow then extended this result to an inhomogeneity following the usual arguments, eqns (28) to

    (35). This produces a strain-concentration tensor

    C h o w  

    h

    I  + ( 1  ?  v 

    f

    ) E S 

     

    f

    ?  C 

    i

    ?  1

    (47)

    which is equivalent to the Mori-Tanaka result (39). Chow was apparently unaware of the connec-

    tion between his approach and the Mori-Tanaka scheme, but he seems to have been the first to

    apply the Mori-Tanaka approach to predict the stiffness of short-fiber composites.

    A more recent development is the  equivalent poly-inclusion model of Ferrari 22. Rather than

    9

  • 8/18/2019 Tucker Halpintsai 2

    11/34

    use the strain-concentration tensor A    , Ferrari used an effective Eshelby tensor ^E    , defined as the

    tensor that relates inclusion strain to transformation strain at finite volume fraction:

    ( " 

    )

    f

    ^

    E  " 

    T  (48)

    Once ^E    has been defined, it is straightforward to derive a strain-concentration tensor A    and a

    composite modulus.

    Ferrari considered admissible forms for ^E    , given the requirements that ^E    must (a) produce a

    symmetric stiffness tensor C    , (b) approach Eshelby’s tensor E    as volume fraction approaches zero,

    and (c) give a composite stiffness that is independent of the matrix stiffness as volume fraction

    approaches unity. He proposed a simple form that satisfies these criteria,

    ^

    E  = ( 1  ?  v 

    f

    ) E 

    (49)

    The combination of eqns (48) and (49) is identical to Chow’s assumption (46) and, for aligned

    fibers of uniform length, Ferrari’s equivalent poly-inclusion model, Chow’s model, and the Mori-

    Tanaka model are identical. Important differences between the equivalent poly-inclusion model

    and the Mori-Tanaka model arise when the fibers are misoriented or have different lengths, a topic

    that will be addressed in a subsequent paper 1.

    3.4 Self-Consistent Models

    A second approach to account for finite fiber volume fraction is the self-consistent  method. This

    approach is generally credited to Hill 23 and Budiansky24, whose original work focused on spherical

    particles and continuous, aligned fibers. The application to short-fiber composites was developed

    by Laws and McLaughlin25 and by Chou, Nomura and Taya26.

    In the self-consistent scheme one finds the properties of a composite in which a single particle

    is embedded in an infinite matrix that has the average properties of the composite. For this reason,self-consistent models are also called embedding models.

    Again building on Eshelby’s result for a ellipsoidal particle, we can create a self-consistent

    version of eqn (35) by replacing the matrix stiffness and compliance tensors by the corresponding

    properties of the composite. This gives the self-consistent strain-concentration tensor as

    S C 

    h

    I  +  E S 

     

    f

    ?  C 

    i

    ?  1

    (50)

    Of course the properties C    and S    of the embedding ‘matrix’ are initially unknown. When the rein-

    forcing particle is a sphere or an infinite cylinder, the equations can be manipulated algebraically

    to find explicit expressions for the overall properties 23,24. For short fibers this has not proved pos-

    sible, but numerical solutions are easily obtained by an iterative scheme. One starts with an initialguess at the composite properties, evaluates E    and then A  S C   from eqn (50), and substitutes the

    result into eqn (21) to get an improved value for the composite stiffness. The procedure is repeated

    using this new value, and the iterations continue until the results for C    converge.

    An additional, but less obvious, change is that Eshelby’s tensor E    depends on the ‘matrix’

    properties, which are now transversely isotropic. Expressions for Eshelby’s tensor for an ellipsoid

    of revolution in a transversely isotropic matrix are given by Chou, Nomura and Taya27 and by Lin

    10

  • 8/18/2019 Tucker Halpintsai 2

    12/34

    and Mura28. With these expressions in hand one can use eqn (50) together with (21) to find the

    stiffness of the composite. This is the self-consistent approach used for short-fiber composites 25,26.

    A closely-related approach, called the ‘generalized self-consistent model,’ also uses an embed-

    ding approach. However, in these models the embedded object comprises both fiber and matrix

    material. When the composite has spherical reinforcing particles, the embedded object is a sphere

    of the reinforcement encased in a concentric spherical shell of matrix; this is in turn surroundedby an infinite body with the average composite properties. The generalized self-consistent model

    is sometimes referred to as a ‘double embedding’ approach. For continuous fibers the embedded

    object is a cylindrical fiber surrounded by a cylindrical shell of matrix. The first generalized self-

    consistent models were developed for spherical particles by Kerner29, and for cylindrical fibers

    by Hermans30. Both of these papers contain an error, which is discussed and corrected by Chris-

    tensen and Lo31. While the generalized self-consistent model is widely regarded as superior to the

    original self-consistent approach, no such model has been developed for short fibers.

    3.5 Bounding Models

    A rather different approach to modeling stiffness is based on finding upper and lower bounds for the

    composite moduli. All bounding methods are based on assuming an approximate field for either

    the stress or the strain in the composite. The unknown field is then found through a variational

    principle, by minimizing or maximizing some functional of the stress and strain. The resulting

    composite stiffness is not exact, but it can be guaranteed to be either greater than or less than the

    actual stiffness, depending on the variational principle. This rigorous bounding property is the

    attraction of bounding methods.

    Historically, the Voigt and Reuss averages were the first models to be recognized as providing

    rigorous upper and lower bounds 32. To derive the Voigt model, eqn (23), one assumes that the

    fiber and matrix have the same uniform strain, and then minimizes the potential energy. Since

    the potential energy will have an absolute minimum when the entire composite is in equilibrium,the potential energy under the uniform strain assumption must be greater than or equal to the exact

    result, and the calculated stiffness will be an upper bound on the actual stiffness. The Reuss model,

    eqn (24), is derived by assuming that the fiber and matrix have the same uniform stress, and then

    maximizing the complementary energy. Since the complementary energy must be maximum at

    equilibrium, the model provides a lower bound on the composite stiffness. Detailed derivations of 

    these bounds are provided by Wu and McCullough 33.

    The Voigt and Reuss bounds provide isotropic results (provided the fiber and matrix are them-

    selves isotropic), when in fact we expect aligned-fiber composites to be highly anisotropic. More

    importantly, when the fiber and matrix have substantially different stiffnesses then the Voigt and

    Reuss bounds are quite far apart, and provide little useful information about the actual composite

    stiffness. This latter point motivated Hashin and Shtrikman to develop a way to construct tighter

    bounds.

    Hashin and Shtrikman developed an alternate variational principle for heterogeneous materials34,35. Their method introduces a reference material, and bases the subsequent development on the

    differences between this reference material and the actual composite. Rather than requiring two

    variational principles, like the Voigt and Reuss bounds, their single variational principle gives both

    the upper and lower bounds by making appropriate choices of the reference material. For an upper

    11

  • 8/18/2019 Tucker Halpintsai 2

    13/34

    bound the reference material must be as stiff or stiffer than any phase in the composite (fiber or

    matrix), and for a lower bound the reference material must have a stiffness less than or equal to

    any phase. In most composites the fiber is stiffer than the matrix, so choosing the fiber as the

    reference material gives an upper bound and choosing the matrix as the reference material gives a

    lower bound. If the matrix is stiffer than the fiber, the bounds are reversed. The resulting bounds

    are tighter than the Voigt and Reuss bounds, which can be obtained from the Hashin-Shtrikmantheory by giving the reference material infinite or zero stiffness, respectively.

    Hashin and Shtrikman’s original bounds35 apply to isotropic composites with isotropic con-

    stituents. Frequently the bounds are regarded as applying to composites with spherical particles,

    though a fiber composite with 3-D random fiber orientation must also obey the bounds.

    Walpole re-derived the Hashin-Shtrikman bounds using classical energy principles 36, and ex-

    tended them to anisotropic materials37. Walpole also derived results for infinitely long fibers and

    infinitely thin disks in both aligned and 3-D random orientations38.

    The Hashin-Shtrikman-Walpole bounds were extended to short-fiber composites by Willis 39

    and by Wu and McCullough 33. These workers introduced a two-point correlation function into the

    bounding scheme, allowing aligned ellipsoidal particles to be treated. Based on these extensions,

    explicit formulae for aligned ellipsoids were developed by Weng40 and by Eduljee et al. 41,42.

    The general bounding formula, shown here in the format developed by Weng, gives the com-

    posite stiffnessC 

      as

    C  = 

    h

    f

    f

    f

    +  v 

    i h

    f

    f

    +  v 

    i

    ?  1

    (51)

    where the tensors Q  f  and Q  m   are defined as

    f

    h

    I  +  E 

    0

    0

    ( C 

    f

    ?  C 

    0

    )

    i

    ?  1

    a n d  Q 

    h

    I  +  E 

    0

    0

    ( C 

    ?  C 

    0

    )

    i

    ?  1

    (52)

    HereE 

    0

      is Eshelby’s tensor associated with the properties of the reference material, which hasstiffness C  0  and compliance S  0  .

    When the matrix is chosen as the reference material, eqn (51) gives a strain concentrator of 

    ^

    l o w e r

    h

    I  +  E 

    ( C 

    f

    ?  C 

    )

    i

    ?  1

    (53)

    This result is labeled here as the lower bound, on the presumption that the fiber is stiffer than the

    matrix. The composite stiffness is found by substituting ^A  l o w e r  into eqns (20) and (21). Eduljee

    and McCullough41,42 argue that the lower bound provides the most accurate estimate of composite

    properties, and recommend it as a model. Note that this lower bound prediction is identical to

    the Mori-Tanaka model, eqn (39)20,40. This correspondence lends theoretical support to the Mori-

    Tanaka approach, and guarantees that it will always obey the bounds.The other bound, found by using eqn (51) with the fiber as the reference material, has a strain

    concentrator of ^

    u p p e r

    h

    I  +  E 

    f

    f

    ( C 

    ?  C 

    f

    )

    i

    (54)

    Note that the Eshelby tensor E  f  is now computed for inclusions of matrix material surrounded by

    the fiber material. Equation (54) is labeled as the upper bound, presuming that the fiber is stiffer

    than the matrix. An identical result can be obtained from the Mori-Tanaka theory by assuming

    12

  • 8/18/2019 Tucker Halpintsai 2

    14/34

    that ellipsoidal particles of the matrix material are embedded in a continuous phase of the fiber

    material.

    If the matrix is stiffer than the fibers, then the right-hand sides of eqns (53) and (54) are un-

    changed but eqn (53) becomes the upper bound and eqn (54) becomes the lower bound. All of 

    the preceding bounding formulae have been given for two-component composites, but the theory

    readily accommodates multiple reinforcements.At fiber volume fractions close to unity, the matrix stiffness strongly influences the composite

    stiffness for the lower bound/Mori-Tanaka models, despite the tiny amount of it that is present.

    Packing considerations suggest that the only way to approach such high volume fractions is for

    the fiber phase to become continuous, and Lielens et al. 43 suggest that at very high fiber volume

    fractions the composite stiffness should be much closer to the upper bound, or equivalently to the

    Mori-Tanaka prediction using the fiber as the continuous phase. This insight prompted Lielens and

    co-workers to propose a model that interpolates between the upper and lower bounds, such that the

    lower bound dominates at low volume fractions and the upper bound dominates at high volume

    fractions (again presuming the fiber is the stiffer phase). They perform this interpolation on the

    inverse of the strain-concentration tensor ^A    , producing the predictive equation43

    ^

    L i e l e n s

    ( 1  ?  f  )

    ^

    l o w e r ?  1

    +  f 

    ^

    u p p e r ?  1

    ?  1

    (55)

    The interpolating factor f    depends on fiber volume fraction, and they propose

    f  = 

    f

    +  v 

    2

    f

    (56)

    This theory reproduces the lower bound and Mori-Tanaka results at low volume fractions, but is

    said to give improved results at reinforcement volume fractions in the 40 to 60% range.

    3.6 Halpin-Tsai Equations

    The Halpin-Tsai equations44,45 have long been popular for predicting the properties of short-fiber

    composites. A detailed review and derivation is provided by Halpin and Kardos46, from which we

    summarize the main points.

    The Halpin-Tsai equations were originally developed with continuous-fiber composites in mind,

    and were derived from the work of Hermans 30 and Hill47. Hermans developed the first generalized

    self-consistent model for a composite with continuous aligned fibers (see Section 3.4). Halpin and

    Tsai found that three of Hermans’ equations for stiffness could be expressed in a common form:

    1 +  v 

    f

    1  ?  v 

    f

    w i t h     = 

    ( P 

    f

    = P 

    ) ?  1 

    ( P 

    f

    = P 

    ) + 1 

    (57)

    Here P    represents any one of the composite moduli listed in Table 1, and P f

      and P m 

      are the

    corresponding moduli of the fibers and matrix, while     is a parameter that depends on the matrix

    Poisson ratio and on the particular elastic property being considered. Hermans derived expressions

    for the plane-strain bulk modulus k 2 3

      , and for the longitudinal and transverse shear moduli G 1 2

      and

    2 3

    . The 

      parameters for these properties are given in Table 1. Note that for an isotropic matrix

    13

  • 8/18/2019 Tucker Halpintsai 2

    15/34

    Table 1: Correspondence between Halpin-Tsai equation (57) and generalized self-consistent pre-

    dictions of Hermans30 and Kerner29. After Halpin and Kardos46.

    P P 

    f

      Comments

    2 3

    f

    1 ? 

    ?  2

    2

    1 + 

    plane strain bulk modulus, aligned fibers

    2 3

    f

    1 + 

    3 ? 

    ?  4

    2

    transverse shear modulus, aligned fibers

    1 2

    f

    1  longitudinal shear modulus, aligned fibers

    K K 

    f

    2 ( 1 ?  2

    )

    1 + 

    bulk modulus, particulates

    G G 

    f

    7 ?  5

    8 ?  1 0

    shear modulus, particulates

    2 ( 1 +  

    ) ( 1 ?  2

    )

    .

    Hill47 showed that for a continuous, aligned-fibercomposite the remaining stiffness parameters

    are given by

    1 1

    =  v 

    f

    f

    +  v 

    ?  4 

     

    f

    ?   

    1

    k

    f

    1

    k

    3

    5

    2

     

    2 3

    f

    f

    (58)

     

    1 2

    =  v 

    f

     

    f

    +  v 

     

     

    f

    ?   

    1

    k

    f

    1

    k

    3

    5

     

    2 3

    f

    f

    (59)

    This completes Hermans’ model for aligned-fiber composites; note that one must know k 2 3

      to find

    1 1

    and  1 2

      . We now know that Hermans’ result for G 2 3

      is incorrect, in that it does not satisfy all

    of the fiber/matrix continuity conditions3. It is, however, identical to a lower bound on G 2 3

      derived

    by Hashin48. Hermans’ remaining results are identical to Hashin and Rosen’s composite cylinders

    assemblage model49, so Hermans’k 

    2 3

      , and thus hisE 

    1 1

      and 

    1 2

      , are identical to the self-consistent

    results of Hill23.

    The Halpin-Tsai form (57) can also be used to express equations for particulate composites

    derived by Kerner29, who also used a generalized self-consistent model. Table 1 gives the details.

    Kerner’s result for shear modulus G    is also known to be incorrect, but reproduces the Hashin-

    Shtrikman-Walpole lower bound for isotropic composites, while Kerner’s result for bulk modulus

    K  is identical to Hashin’s composite spheres assemblage model50. See Christensen and Lo31 and

    Hashin3 for further discussion of Kerner’s and Hermans’ results.

    To transform these results into convenient forms for continuous-fiber composites, Halpin and

    Tsai made three additional ad hoc approximations:

      Equation (57) can be used directly to calculate selected engineering constants, with E 1 1

      or

    2 2

    replacing P    .

      The     parameters in Table 1 are insensitive to  m 

      , and can be approximated by constant

    values.

      The underlined terms in eqns (58) and (59) can be neglected.

    14

  • 8/18/2019 Tucker Halpintsai 2

    16/34

    Table 2: Traditional Halpin-Tsai parameters for short-fiber composites, used in eqn (57). For G 2 3

    see Table 1.

    P P 

    f

      Comments

    1 1

    f

    2 ( ` = d  ) longitudinal modulus

    2 2

    f

    2  transverse modulus

    1 2

    f

    1  longitudinal shear modulus

     

    1 2

    Poisson ratio, =v 

    f

     

    f

    +  v 

     

    In eqn (58) the underlined term is typically negligible, and dropping it gives the familiar rule

    of mixtures for E 1 1

      of a continuous-fiber composite. However, dropping the underlined term in

    eqn (59) and using a rule of mixtures for  1 2

      is not necessarily accurate if the fiber and matrix

    Poisson ratios differ. Halpin and Tsai argue for this latter approximation on the grounds that

    laminate stiffnesses are insensitive to   1 2  .In adapting their approach to short-fiber composites, Halpin and Tsai noted that     must lie

    between 0    and 1    . If    = 0   then eqn (57) reduces to the inverse rule of mixtures 46,

    f

    f

    (60)

    while for   =  1   the Halpin-Tsai form becomes the rule of mixtures,

    P  =  v 

    f

    f

    +  v 

    (61)

    Halpin and Tsai suggested that     was correlated with the geometry of the reinforcement and, when

    calculatingE 

    1 1

      , it should vary from some small value to infinity as a function of the fiber aspect

    ratio` = d 

      . By comparing model predictions with available 2-D finite element results, they found

    that  = 2 ( ` = d  )

      gave good predictions forE 

    1 1

      of short-fiber systems. Also, they suggested that

    other engineering constants of short-fiber composites were only weakly dependent on fiber aspect

    ratio, and could be approximated using the continuous-fiber formulae 45. The resulting equations

    are summarized in Table 2. The early references 44,45 do not mention G 2 3

      . When this property is

    needed the usual approach is to use the     value given in Table 1. While the Halpin-Tsai equations

    have been widely used for isotropic fiber materials, the underlying results of Hermans and Hill

    apply to transversely isotropic fibers, so the Halpin-Tsai equations can also be used in this case.

    The Halpin-Tsai equations are known to fit some data very well at low volume fractions, but to

    under-predict some stiffnesses at high volume fractions. This has prompted some modifications totheir model. Hewitt and de Malherbe 51 proposed making     a function of  v 

    f

      , and by curve fitting

    found that

      = 1 + 4 0  v 

    1 0

    f

    (62)

    gave good agreement with 2-D finite element results for G 1 2

      of continuous fiber composites.

    Nielsen and Lewis52,53 focused on the analogy between the stiffness G    of a composite and the

    viscosity     of a suspension of rigid particles in a Newtonian fluid, noting that one should find

    15

  • 8/18/2019 Tucker Halpintsai 2

    17/34

    =  

    =  G = G 

    when the reinforcement is rigid ( G f

    = G 

    ! 1  )  and the matrix is incompressible.

    They developed an equation in which the stiffness not only matches dilute theory at low volume

    fractions, but also displays G = G m 

    ! 1   as v f

      approaches a packing limit v f m a x

      . This leads to a

    modified Halpin-Tsai formP 

    1 +  v 

    f

    1  ?    ( v 

    f

    ) v 

    f

    (63)

    with     retaining its definition from eqn (57). Here the function   ( v f

    )  contains the maximum

    volume fraction v f m a x

      as a parameter.     is chosen to give the proper behavior at the upper and

    lower volume fraction limits, which leads to forms such as

      ( v 

    f

    ) = 1 +  

     

    1  ?  v 

    f m a x

    2

    f m a x

    f

    (64)

      ( v 

    f

    ) = 

    f

    1  ?  e x p 

     

    ?  v 

    f

    1  ?  ( v 

    f

    = v 

    f m a x

    )

    ! #

    (65)

    The Nielsen and Lewis model improves on the Halpin-Tsai predictions, compared to experimen-tal data forG 

      of particle-reinforced polymers 52 and to finite element calculations forG 

    1 2

      of 

    continuous-fiber composites53, using v f m a x

      values from 0.40 to 0.85.

    Recently Ingber and Papathanasiou 54 tested the Halpin-Tsai equation and its modifications

    against boundary element calculations of  E 1 1

      for aligned short fibers. They found the Nielsen

    modification to be better than the original Halpin-Tsai form. Hewitt and de Malherbe’s form could

    be adjusted to fit data for any single ` = d   , but was not useful for predictions over a range of aspect

    ratios. These results are discussed further in Section 4.

    3.7 Shear Lag Models

    Historically, shear lag models were the first micromechanics models for short-fiber composites55, as well as the first to examine behavior near the ends of broken fibers in a continuous-fiber

    composite56,57. Despite some serious theoretical flaws, shear lag models have enjoyed enduring

    popularity, perhaps due to their algebraic simplicity and their physical appeal.

    Classical shear lag models only predict the longitudinal modulusE 

    1 1

      , so they do not meet

    our criterion of predicting a complete set of elastic constants. However, we include them here

    because of their historical importance and their widespread use. One could obtain a complete

    stiffness model by using the shear lag prediction forE 

    1 1

      and some continuous-fiber model (such

    as Hermans’) for the remaining elastic constants. If the fiber is anisotropic then its axial modulus

    should be used in the shear lag equations.

    Following Cox55, the shear lag analysis focuses on a single fiber of length  ̀   and radius r f

      ,

    which is encased in a concentric cylindrical shell of matrix having radius R    . The fiber is aligned

    parallel to the z    axis, as shown in Fig. 3. Only the axial stress  1 1

      and axial strain " 1 1

      are of interest,

    and Poisson effects are neglected so that  f

    1 1

    =  E 

    f

    f

    1 1

      . The outer cylindrical surface of the matrix

    is subjected to displacement boundary conditions consistent with an average axial strain " 1 1

      , and

    one solves for the fiber stress  f

    1 1

    ( z  )  . (More rigorously,  f

    1 1

    ( z  )  is the average stress over the fiber

    16

  • 8/18/2019 Tucker Halpintsai 2

    18/34

    rf  R

     l

    z

    Figure 3: Idealized fiber and matrix geometry used in shear lag models.

    cross-section at z   .) Axial equilibrium of the fiber requires that

    d  

    f

    1 1

    d z 

    =  ? 

    2   

    r z

    f

    (66)

    where  r z

      is the axial shear stress at the fiber surface. The key assumption of shear lag theory is

    that  r z

      is proportional to the difference in displacement w    between the fiber surface and the outer

    matrix surface:

     

    r z

    ( z  ) = 

    2  r 

    f

    w  ( R ; z  ) ?  w  ( r 

    f

    ; z  ) (67)

    whereH 

      is a constant that depends on matrixpropertiesand fibervolume fraction. Solving eqn (66)

    for 

    f

    1 1

    ( z  )

      and applying boundary conditions of zero stress at the fiber ends gives an average fiber

    stress of 

     

    f

    1 1

    =  E 

    f

    1 1

    1  ? 

    t a n h ( ` =  2 )

    ( ` =  2 )

    #

    (68)

    with

     

    2

    2

    f

    f

    (69)

    It is convenient to rewrite this as an expression for the average fiber strain,

    f

    1 1

    =   

    `

    1 1

    (70)

    where    ̀ is a length-dependent ‘efficiency factor’,

     

    `

    1  ? 

    t a n h ( ` =  2 )

    ( ` =  2 )

    #

    (71)

    Note that    ̀ is a scalar analog of the strain-concentration tensor A    defined in eqn (15), and ( 1  =   )

    is a characteristic length for stress transfer between the fiber and the matrix.

    17

  • 8/18/2019 Tucker Halpintsai 2

    19/34

    Table 3: Values for K R 

      used in eqn (74) for shear lag models.

    Fiber packing K R 

    Cox 2  = p 

    3   = 3.628

    Composite Cylinders1 

      = 1.000Hexagonal =  2 

    3   = 0.907

    Square =  4   = 0.785

    Cox55 found the coefficientH 

      by solving a second idealized problem. The concentric cylinder

    geometry is maintained, but the outer cylindrical surface of the matrix is held stationary and the

    inner cylinder, which is now rigid, is subjected to a uniform axial displacement. An elasticity

    solution for the matrix layer then gives

    H  = 

    2  G 

    l n ( R = r 

    f

    )

    (72)

    Rosen56,57 simplified this part of the problem by assuming that the matrix shell was thin compared

    to the fiber radius, ( R  ?  r f

    )   r 

    f

      , obtaining

    H  = 

    2  G 

    ( R = r 

    f

    ) ?  1 

    (73)

    Rosen’s approximation gives an error in H    of 10% at v f

    = 0  6 0   , with much larger errors at lower

    volume fractions, and we will not consider it further.

    It remains to choose the radiusR 

      of the matrix cylinder, and the exact choice is important.

    Several choices have been used, all of which can be written in the form

    f

    f

    (74)

    where K R 

      is a constant that depends on the assumption used to find R    . Table 3 summarizes the

    choices for K R 

      . Cox55 assumed a hexagonal packing, and chose R    as the distance between centers

    of nearest-neighbor fibers (Fig. 4a). It seems more realistic to let R    equal half of the distance

    between nearest neighbors (Fig. 4b), a choice labeled ‘hexagonal’ in Table 3. Rosen56,57, and later

    Carman and Reifsnider58, chose r  2f

    = R 

    2

    =  v 

    f

      so that the concentric cylinder model in Fig. 3 would

    have the same fiber volume fraction as the composite. This is the same R    as the composite cylinders

    model of Hashin and Rosen49. More recently, Robinson and Robinson59,60 assumed a square array

    of fibers, and chose R    as half the distance between centers of nearest neighbors (Fig. 4c)61. Each

    of these choices gives a somewhat different dependence of  

     ̀ on fiber volume fraction, with larger

    values of K 

      producing lower values of E 

    1 1

      .

    Shear lag models are usually completed by combining the average fiber stress in eqn (68) with

    an average matrix stress to produce a modified rule of mixtures for the axial modulus:

    1 1

    =   

    `

    f

    f

    + ( 1  ?  v 

    f

    ) E 

    (75)

    18

  • 8/18/2019 Tucker Halpintsai 2

    20/34

    (a) (b) (c)

    RR

    R

    Figure 4: Fiber packing arrangements used to find R    in shear lag models. (a) Cox. (b) Hexagonal.

    (c) Square.

    However, the matrix stress in this formula is not consistent with the basic concepts of average

    stress and average strain. Note that eqn (7) must hold for "  1 1  , as for any other component of strain.Combining this with eqn (70) to find the average matrix strain, and following through to find the

    composite stiffness (with Poisson effects neglected), gives a result that is consistent with both the

    assumptions of shear lag theory and the basic concepts of average stress and strain:

    1 1

    =   

    `

    f

    f

    + ( 1  ?   

    `

    f

    ) E 

    (76)

    =  E 

    +  v 

    f

    ( E 

    f

    ?  E 

    )  

    `

    This equation is an exact scalar analog of the general tensorial stiffness formula, eqn (21). For the

    cases in this paper, the difference between eqns (75) and (76) is small, and we will use the classical

    shear lag result (75) when testing the models.

    A model by Fukuda and Kawata62 for the axial stiffness of aligned short-fiber composites is

    closely related to shear lag theory. They begin with a 2-D elasticity solution for the shear stress

    around a single slender fiber in an infinite matrix. The usual shear lag relation, eqn (66), is used

    to transform this into an equation for the fiber stress distribution, which is then approximated by

    a Fourier series. The coefficients of a truncated series are evaluated analytically using Galerkin’s

    method. This is a dilute theory, in which modulus varies linearly with fiber volume fraction.

    Like any shear lag theory, Fukuda and Kawata’s theory predicts thatE 

    1 1

      approaches the rule

    of mixtures result as the fiber aspect ratio approaches infinity. But for short fibers Fukuda and

    Kawata’s theory gives much lower E 1 1

      values than shear lag theory. In Fukuda and Kawata’s

    theory, the ratio of fiber strain to matrix strain is governed by the parameter ( ` = d  ) ( E m 

    = E 

    f

    )  . In

    contrast, for shear lag theory, eqn (71), the governing parameter is ` =  2 

      , which is proportional to( ` = d  )

    = E 

    f

    . Thus, for high modulus ratio and low aspect ratio, Fukuda and Kawata’s theory

    tends to underpredict E 1 1

      . For this reason we do not pursue their theory further.

    19

  • 8/18/2019 Tucker Halpintsai 2

    21/34

    Table 4: Models selected for comparison.

    Model Comments

    Halpin-Tsai eqn (57) and Table 2

    Nielsen eqns (63), (64), (57b) and Table 2Mori-Tanaka eqns (39), (35), and (21)

    Lielens eqns (55), (56), (53), (54), (20), and (21)

    Self-Consistent eqns (50) and (21)

    Shear Lag eqns (75), (71), (69), (72), (74), and Table 3

    4 Tests and Comparisons

    Obtaining reference data for unidirectional short-fiber composites presents a problem. Accurate

    experimental data is not available, since it has not proved possible to produce physical samples with

    perfectly aligned fibers. The best that can be done experimentally is to make samples with partially

    aligned fibers, though even in those samples the fibers may be clustered or bundled together in

    some unspecified way42. Any comparison between the properties of such samples and predictions

    necessarily includes both the model for aligned-fiber composites and the model for fiberorientation

    effects.

    In this paper we avoid this complication by using three-dimensional finite element models

    of aligned short-fiber composites, rather than experimental results, as the reference data. This

    necessitates the assumption of a periodic arrangement of the fibers, but all of the micromechanics

    models are sufficiently vague about the geometricarrangement of thefibers that they admit periodic

    geometries. We also compare the theories to some boundary element results for random arrays of 

    aligned fibers54

    .For clarity we limit our comparisons to the models listed in Table 4. For the shear lag model we

    show results only for the square array, noting that this choice for R    gives the highest stiffness. The

    models which are not shown are: the dilute Eshelby model, which is limited to small volume frac-

    tions; the Hashin-Shtrikman-Walpole lower bound, which is identical to the Mori-Tanaka model;

    and the upper bound, which is not claimed to be useful by itself.

    4.1 Finite Element Modeling

    Using the finite element method we analyzed two types of periodic, three-dimensional arrays of 

    fibers, which we call regular  and staggered  arrays. The representative volume elements (RVE’s)

    are shown in Fig. 5. The unit cell dimensions were chosen with b  =  a   , where     is a constant. Weused both   = 1   to obtain square packing, and   = 

    3   which gives hexagonal packing. For the

    regular fibers the distance between neighboring fiber ends (equal to 2  c  ?   ̀  in Fig. 5a) was set to

    0  5 3 8   ̀ for square packing and 0  1 3 6   ̀  for hexagonal packing. For the staggered arrays the distance

    along each fiber that is overlapped by neighboring fibers was set at a fixed percentage of the fiber

    length: 65% for square packing and 76% for hexagonal packing. These conditions, together with

    the fiber diameter and volume fraction, suffice to determine the dimensions a   , b    and c    for each

    20

  • 8/18/2019 Tucker Halpintsai 2

    22/34

    Table 5: Material properties used in finite element calculations.

    Property Fiber Matrix

    E  30 1

     

    0.20 0.38v 

    f

    0.20

    ` = d  1, 2, 4, 8, 16, 24, 48

    RVE. Note that a new RVE and its corresponding 3-D mesh are generated for each fiber aspect

    ratio.

    Stiffnesses of these RVE’s were calculated using ABAQUS 63. Twenty-node isoparametric ele-

    ments were used, and a sample mesh is shown in Fig. 6. The analysis was geometrically nonlinear

    but the applied strain was 0.5%, so the results are in the region of linear behavior. For axial or

    transverse loading, symmetry requires all faces of the RVE to remain plane. To determine E 1 1

      and

     

    1 2 we fixed the normal displacements of the back, left, and bottom faces of the RVE; requiredthe right and top faces to remain plane and parallel to the coordinate axes (using multi-point con-

    straints); and displaced the front face uniformly in the x 1

      direction. The tangential displacements

    on all faces were unconstrained. The average stress was computed from the reaction force in the

    loading direction, divided by the cross-sectional area of the RVE. Average strains were computed

    from the initial and deformed dimensions of the RVE. Analogous conditions were use to load the

    RVE in thex 

    2

      direction to determineE 

    2 2

      and 

    2 3

      . The longitudinal shear modulusG 

    1 2

      could

    in principle be determined using these same RVE’s, but that calculation requires a complicated

    application of periodic boundary conditions and we did not undertake it.

    All of the micromechanics theories reviewed here predict transversely isotropic properties.

    Transverse isotropy about the x 1

      axis implies that the tensile modulus is the same for any loading

    direction in the 2–3 plane. This not only requires that E 2 2

    =  E 

    3 3

      , but also that

    2 3

    2 2

    ( 1 +   

    2 3

    )

    (77)

    RVE’s with hexagonal packing should also be transversely isotropic and obey these same relation-

    ships. However, for square packing the properties are only guaranteed to be orthotropic. That is,

    calculations for square packing will always give E 2 2

    =  E 

    3 3

      , but the results will not necessarily

    obey eqn (77) nor will the transverse modulus necessarily be the same for other loading directions

    in the 2–3 plane. Here we simply report  2 3

      and E 2 2

      for loading in the x 2

      direction, and do not

    explore the other orthotropic constants for square packing.

    The material properties used in the finite element calculations (Table 5) are typical of fiber-reinforced engineering thermoplastics. All of the moduli are scaled by the matrix modulus.

    4.2 Results and Discussion

    Figure 7 compares the theoretical and finite element results for longitudinal modulus E 1 1

      . The

    strong influence of fiber aspect ratio onE 

    1 1

      is apparent, and all of the theories exhibit a similar

    21

  • 8/18/2019 Tucker Halpintsai 2

    23/34

    a

    b

    cl/2

    a

    b

    cl/2

    (a) (b)

    1

    3

    2

    Figure 5: Representative volume elements used in the finite element calculations. (a) Regular

    array; the bold lines show the RVE. (b) Staggered array.

    Figure 6: Example finite element mesh for a staggered, hexagonal array.

    22

  • 8/18/2019 Tucker Halpintsai 2

    24/34

    0.0

    1.0

    2.0

    3.0

    4.0

    5.0

    6.0

    7.0

    1 10 100 1000

    Halpin-Tsai

    Nielsen

    Mori-Tanaka

    Lielens

    Self-consistent

    Shear lag, square

    FE, Sqr. Reg.

    FE, Sqr. Stag.

    FE, Hex. Reg.

    FE, Hex. Stag.

       M  o   d  u   l  u  s   R  a   t   i  o ,

       E   1   1

       /   E  m

    Aspect Ratio, l/d

    Figure 7: Theoretical predictions and finite element results for E 1 1

      .

    S-shaped curve, asymptoting to the same rule-of-mixtures value at high aspect ratio. However, the

    various theories give quite different values for very short fibers, and rise at different rates.

    The different packing arrangements create some scatter in the finite element results, but the

    scatter is small for ` = d    8   . For ` = d    4   the scatter is significant. This is not surprising, since

    the properties of particulate-reinforced composites are known to be very sensitive to the packingarrangement. The high E 

    1 1

      values for the hexagonal staggered array probably occur because our

    rules for forming this particular type of RVE tend to create long ‘chains’ of nearly-touching par-

    ticles parallel to the x 1

      axis, with a high degree of axial overlap. While all of the finite element

    results are equally ‘true,’ we believe the lower finite element values are more representative of the

    actual packing and the actual stiffness of composites with very short fibers.

    Comparing models to finite element data for E 1 1

      , the Halpin-Tsai equation is accurate for very

    short fibers, but falls below the data for longer fibers. The Nielsen model improves on the Halpin-

    Tsai predictions for the very short fibers, but is still below the data for longer fibers. A better fit in

    the higher aspect ratio range is provided by the Mori-Tanaka and Lielens models, which are only

    slightly different from one another at this volume fraction. These models are good over most of 

    the data range. The self-consistent results are usually high, while the shear lag model is good forthe longer fibers but too low for very short fibers. This latter behavior is not surprising, since shear

    lag theory treats the fiber as a slender body. Using any of the other values for R    in the shear lag

    model shifts the curve to the right, moving the predictions away from the data.

    Results for transverse modulus E 2 2

      are shown in Fig. 8. The finite element data again have

    moderate scatter. Fiber aspect ratio has little effect on the transverse modulus, though some of 

    the packing geometries show a slight dip at low aspect ratio. Interestingly, the shape and location

    23

  • 8/18/2019 Tucker Halpintsai 2

    25/34

    0.0

    0.5

    1.0

    1.5

    2.0

    1 10 100 1000

    Halpin-TsaiNielsenMori-TanakaLielensSelf-consistent

    FE, Sqr. Reg.FE, Sqr. Stag.FE, Hex. Reg.FE, Hex. Stag.

       M  o   d  u   l  u  s   R  a   t   i  o ,

       E   2   2

       /   E  m

    Aspect Ratio, l/d

    Figure 8: Theoretical predictions and finite element results for E 2 2

      .

    of this dip are matched by the models that use the Eshelby tensor. Note that the Halpin-Tsai and

    Nielsen models contain no dependence on aspect ratio for E 2 2

      . Shear lag models do not predict

    2 2

    .

    Most of the models do a good job of predicting E 2 2

      , with the Mori-Tanaka and Lielens models

    being the most accurate. The Halpin-Tsai result is slightly higher than most of the data, while theNielsen model noticeably over-predicts this property. For comparison the upper bound result falls

    well above the data, with an asymptote of  E 2 2

    = E 

    = 3  5 9   at high aspect ratio.

    Data for the Poisson ratios  1 2

      and  2 3

      appear in Figs. 9 and 10. The Nielsen and Halpin-Tsai

    results for  1 2

      are identical, so only the Halpin-Tsai curve is shown. Both Poisson ratios show

    a moderate dependence on aspect ratio and some sensitivity to packing geometry. The shape of 

    this dependence is similar for all but the regular hexagon array and is matched qualitatively by

    several models, but the quantitative match is not as good. For  1 2

      the constant value provided by

    the Halpin-Tsai equations is at least as good a match to the data as the models that show some

    variation. However, the Halpin-Tsai and Nielsen models substantially over-predict  2 3

      , while the

    other models do very well on this property, especially at the higher aspect ratios. The error in the

    Halpin-Tsai value results from a combination of a slightly high prediction for E  2 2  (Fig. 8) and aslightly low prediction for G 

    2 3

      (not shown here), the effects combining through eqn (77).

    One weakness of the finite element calculations is that they require the assumption of a regular,

    periodic packing arrangement of the fibers. Calculations that do not have this limitation have

    been recently reported by Ingber and Papathanasiou 54. These workers used the boundary element

    method to calculate E 1 1

      for random arrays of aligned fibers. Each model typically contained 100

    fibers, and results from ten such models were averaged to produce each data point. We tested

    24

  • 8/18/2019 Tucker Halpintsai 2

    26/34

    0.20

    0.25

    0.30

    0.35

    0.40

    1 10 100 1000

    Halpin-TsaiMori-TanakaLielensSelf-consistentFE, Sqr. Reg.FE, Sqr. Stag.FE, Hex. Reg.FE, Hex. Stag.

       P  o   i  s  s  o  n   R  a   t   i  o ,      ν

           1        2

    Aspect Ratio, l/d

    Figure 9: Theoretical predictions and finite element results for  1 2

      .

    0.30

    0.40

    0.50

    0.60

    0.70

    1 10 100 1000

    Halpin-TsaiNielsenMori-TanakaLielensSelf-consistentFE, Sqr. Reg.FE, Sqr. Stag.

    FE, Hex. Reg.FE, Hex. Stag.

       P  o   i  s  s  o  n   R  a   t   i  o ,      ν        2        3

    Aspect Ratio, l/d

    Figure 10: Theoretical predictions and finite element results for  2 3

      .

    25

  • 8/18/2019 Tucker Halpintsai 2

    27/34

    0

    5

    10

    15

    20

    0.00 0.10 0.20 0.30

    Halpin-TsaiNielsenMori-TanakaLielensSelf ConsistentBE, randomFE, Sqr. Reg.FE, Sqr. Stag.

       M  o   d  u   l  u  s   R  a   t   i  o ,

       E   1   1   /

       E  m

    Fiber Volume Fraction, vf 

    Figure 11: Models compared to boundary element predictions of  E 1 1

      for random arrays of rigid

    cylinders by Ingber and Papathanasiou 54, and to finite element calculations with E f

    = E 

    = 1 0 

    6  ,

    all for` = d  = 1 0 

      .

    their results against the various theories, and also performed a limited number of finite element

    calculations for comparison purposes. The boundary element results are for rigid fibers (E 

    f

    = E 

    1  ) and an incompressible matrix (  m 

    = 0  5   ), but our finite element calculations and theoretical

    results use E f

    = E 

    = 1 0 

    6  and  m 

    = 0  4 9   to avoid numerical difficulties in some of the models.

    Figure 11 shows the results for E 1 1

      versus volume fraction for ` = d  = 1 0   . The boundary element

    data are most accurately matched by the Lielens and Nielsen models, though the Halpin-Tsai and

    Mori-Tanaka models are not bad. The self-consistent model predicts much higher stiffnesses than

    the other models and than the boundary element data. So far these results are consistent with our

    previous comparisons.

    What is surprising about Fig. 11 is that the finite element results fall so far above the boundary

    element results, and above the theories that work so well in other cases. Since the finite element

    data fall closer to the self-consistent model, it is tempting to think that they support the accuracy of 

    this model. But we believe it more likely that these results are revealing the sensitivity of stiffness

    to the packing arrangement of the fibers.

    Other researchers have noted that gathering short fibers into bundles or clusters tends to reduce

    1 1

    compared to evenly dispersed fibers 42. In the boundary element calculations of Ingber and

    Papathanasiou the inter-fiber spacing is random, and hence uneven, so there is a modest clustering

    effect. In contrast, our finite element models impose a uniform inter-fiber spacing, and so represent

    an unusually even dispersion of fibers. Our finite element models also maximize the axial overlap

    26

  • 8/18/2019 Tucker Halpintsai 2

    28/34

    between neighboring fibers. It seems that these geometric effects have the greatest influence on

    composite stiffness when the fibers are rigid. We believe that the boundary element calculations

    are more representative of real composite behavior than the finite element calculations in Fig. 11.

    Fortunately the influence of fiber packing is much smaller for the E f

    = E 

      ratios typical of 

    polymer-matrix composites. Note that the two different finite element results for v f

    = 0  2 0   and

    rigid fibers in Fig. 11 are far apart from one another, but in Fig. 7 where E  f = E  m  = 3 0   the samepacking geometries give nearly identical results at

    ` = d  = 8 

      . This lends support to the idea that

    fiber packing is important mainly when the fibers are extremely stiff compared to the matrix, and

    supports the finite element results in Figs. 7–10 as a meaningful test of the micromechanics theo-

    ries.

    5 Conclusions

    Our goal is to identify the best model for predicting the stiffness of aligned short-fiber composites.

    Among the models that we tested, the self-consistent approach tends to over-predict E 1 1

      at high

    volume fractions, though it gives good predictions for other elastic constants. The Halpin-Tsaimodel, long a standard for this problem, gives reasonable results for all the elastic constants except

     

    2 3

    , and its E 1 1

      values are low for moderate-to-high aspect ratios. Nielsen and Lewis’s modification

    of Halpin-Tsai improves the fit to Ingber and Papathanasiou’s boundary element data for E 1 1

      , but

    it does not substantially improve the fit to our E 1 1

      data and it substantially worsens the prediction

    of  E 2 2

      . The Mori-Tanaka and Lielens models give much better predictions than Halpin-Tsai for

     

    2 3

    , and slightly better predictions for all the other properties. Our finite element data does not

    allow us to choose between the Mori-Tanaka and Lielens models, since the differences between

    their predictions are small for the volume fractions we examined. Shear lag models can give good

    predictions forE 

    1 1

      for aspect ratios greater than 10, provided one makes the proper choice of R 

      ,

    but the predictions for shorter fibers are too low.

    Our results confirm that the Halpin-Tsai equations providereasonable estimates for the stiffness

    of short-fiber composites, but they indicate that the Mori-Tanaka model is more accurate. The

    bound interpolation model of Lielens et al. may improve on the Mori-Tanaka model for higher

    fiber volume fractions or modulus ratios, but for injection-molded composites the difference is

    small. We recommend the Mori-Tanaka model as the best choice for estimating the stiffness of 

    aligned short-fiber composites.

    Acknowledgments

    Funding to the University of Illinois was provided by The General Electric Company and GeneralMotors Corporation. This work was conducted in support of the Thermoplastic Engineering De-

    sign (TED) Venture, a Department of Commerce Advanced Technology Program administered by

    the National Institute of Standards and Technology. The authors are grateful to Mr. C. Matthew

    Dunbar of Hibbitt, Karlsson & Sorensen, Inc. for his assistance with mesh generation and the finite

    element analysis, and to Dr. T. D. Papathanasiou of the University of South Carolina for making

    available the detailed data from his recent paper.

    27

  • 8/18/2019 Tucker Halpintsai 2

    29/34

    References

    1. Tucker, C. L., O’Gara, J. F., Harris, J., and Inzinna, L., Stiffness predictions for misoriented

    short-fiber composites: Review and evaluation. Manuscript in preparation.

    2. Hill, R., Elastic properties of reinforced solids: Some theoretical principles. J. Mech. Phys.

    Solids, 1963, 11, 357–372.

    3. Hashin, Z., Analysis of composite materials—A survey.  ASME J. Applied Mech., 1983,  50,

    481–505.

    4. Hashin, Z., Theory of mechanical behavior of heterogeneous media.  Appl. Mech. Reviews,

    1964, 17, 1–9.

    5. Eshelby, J. D., The determination of the elastic field of an ellipsoidal inclusion, and related

    problems. Proc. Roy. Soc., 1957, A–241, 376–96.

    6. Eshelby, J. D., Elastic inclusions and inhomogeneities. In I. N. Sneddon and R. Hill (eds.),Progress in Solid Mechanics 2, North-Holland, Amsterdam, 1961 pp. 89–140, pp. 89–140.

    7. Mura, T., Micromechanics of Defects in Solids. Martinus Nijhoff, The Hague, 1982.

    8. Taya, M. and Mura, T., On stiffness and strength of an aligned short-fiber reinforced composite

    containing fiber-end cracks under uniaxial applied stress.  ASME J. Applied Mech., 1981,  48,

    361–367.

    9. Taya, M. and Chou, T.-W., On two kinds of ellipsoidal inhomogeneities in an infinite elastic

    body: An application to a hybrid composite. Int. J. Solids Structures, 1981, 17, 553–563.

    10. Taya, M., On stiffness and strength of an aligned short-fiber reinforced composite containingpenny-shaped cracks in the matrix. J. Compos. Mater., 1981, 15, 198–210.

    11. Tandon, G. P. and Weng, G. J., The effect of aspect ratio of inclusions on the elastic properties

    of unidirectionally aligned composites. Polym. Compos., 1984, 5, 327–333.

    12. M.Taya and Arsenault, R. J., Metal Matrix Composites: Thermomechanical Behavior . Perga-

    mon Press, 1989.

    13. Eshelby, J. D., The elastic field outside an ellipsoidal inclusion. Proc. Roy. Soc., 1959, A 252,

    561–569.

    14. Russel, W. B., On the effective moduli of composite materials: Effect of fiber length andgeometry at dilute concentrations. J. Appl. Math. Phys. (ZAMP), 1973, 24, 581–600.

    15. Chow, T. S., Elastic moduli of filled polymers: The effect of particle shape. J. Appl. Phys.,

    1977, 48, 4072–4075.

    16. Steif, P. S. and Hoysan, S. F., An energy method for calculating the stiffness of aligned short-

    fiber composites. Mech. Mater., 1987, 6, 197–210.

    28

  • 8/18/2019 Tucker Halpintsai 2

    30/34

    17. Mori, T. and Tanaka, K., Average stress in matrix and average elastic energy of materials with

    misfitting inclusions. Acta Metallurgica, 1973, 21, 571–574.

    18. Benveniste, Y., A new approach to the application of Mori-Tanaka’s theory in composite ma-

    terials. Mech. Mater., 1987, 6, 147–157.

    19. Wakashima, K., Otsuka, M., and Umekawa, S., Thermal expansion of heterogeneous solids

    containing aligned ellipsoidal inclusions. J. Compos. Mater., 1974, 8, 391–404.

    20. Weng, G. J., Some elastic properties of reinforced solids, with special reference to isotropic

    ones containing spherical inclusions. Int. J. Engng. Sci., 1984, 22, 845–856.

    21. Chow, T. S., Effect of particle shape at finite concentration on the elastic modulus of filled

    polymers. J. Polym Sci: Polym Phys Ed , 1978, 16, 959–965.

    22. Ferrari, M., Composite homogenization via the equivalent poly-inclusion approach.  Compos.

     Engr., 1994, 4, 37–45.

    23. Hill, R., A self-consistent mechanics of composite materials. J. Mech. Phys. Solids, 1965, 13,

    213–222.

    24. Budiansky, B., On the elastic moduli of some heterogeneous materials. J. Mech. Phys. Solids,

    1965, 13, 223–227.

    25. Laws, N. and McLaughlin, R., The effect of fibre length on the overall moduli of composite

    materials. J. Mech. Phys. Solids, 1979, 27, 1–13.

    26. Chou, T.-W., Nomura, S., and Taya, M., A self-consistent approach to the elastic stiffness of 

    short-fiber composites. J. Compos. Mater., 1980, 14, 178–188.

    27. Chou, T.-W., Nomura, S., and Taya, M., A self-consistent approach to the elastic stiffness of 

    short-fiber composites. In J. R. Vinson (ed.), Modern Developments in Composite Materials

    and Structures, ASME, New York, 1979 pp. 149–164, pp. 149–164.

    28. Lin, S. C. and Mura, T., Elastic fields of inclusions in anisotropic media (II). phys. stat. sol.

    (a), 1973, 15, 281–285.

    29. Kerner, E. H., The elastic and thermo-elastic properties of composite media. Proc. Phys. Soc.

     B, 1956, 69, 808–813.

    30. Hermans, J. J., The elastic properties of fiber reinforced materials when the fibers are aligned.

    Proc. Kon. Ned. Akad. v. Wetensch., 1967, B 65, 1–9.

    31. Christensen, R. M. and Lo, K. H., Solutions for effective shear properties in three phase sphere

    and cylinder models. J. Mech. Phys. Solids, 1979, 27, 315–330, erratum: v34, p639.

    32. Hill, R., The elastic behaviour of a crystalline aggregate. Proc. Phys. Soc. A, 1952,  65, 349–

    354.

    29

  • 8/18/2019 Tucker Halpintsai 2

    31/34

    33. Wu, C.-T. D. and McCullough, R. L., Constitutive relationships for heterogeneous materials.

    In G. S. Holister (ed.), Developments in Composites Materials–1, Applied Science, London,

    1977 pp. 118–186, pp. 118–186.

    34. Hashin, Z. and Shtrikman, S., On some variational principles in anisotropic and nonhomoge-

    neous elasticity. J. Mech. Phys. Solids, 1962, 10, 335–342.

    35. Hashin, Z. and Shtrikman, S., A variational approach to the theory of the elastic behavior of 

    multiphase materials. J. Mech. Phys. Solids, 1963, 11, 127–140.

    36. Walpole, L. J., On bounds for the overall elastic moduli for inhomogeneous systems—I.  J.

     Mech. Phys. Solids, 1966, 14, 151–162.

    37. Walpole, L. J., On bounds for the overall elastic moduli for inhomogeneous systems—II. J.

     Mech. Phys. Solids, 1966, 14, 289–301.

    38. Walpole, L. J., On the overall elastic moduli of composite materials. J. Mech. Phys. Solids,

    1969, 17, 235–251.

    39. Willis, J. R., Bounds and self-consistent estimates for the overall properties of anisotropic

    composites. J. Mech. Phys. Solids, 1977, 25, 185–202.

    40. Weng, G. J., Explicit eva