You are on page 1of 25

1.

17 Multiscale Mechanics of Composite Materials and Structures


Yehia A Bahei-El-Din, The British University in Egypt, El-Shorouk City, Egypt
r 2018 Elsevier Ltd. All rights reserved.

1.17.1 Introduction 426


1.17.2 Structural Composites Length Scales 426
1.17.3 Mechanics and Interactions of Composites Length Scales 427
1.17.3.1 Micro-Mechanics of Fibrous Composites 428
1.17.3.1.1 Matrix properties 429
1.17.3.1.2 Fiber properties 429
1.17.3.1.3 Unidirectional fibrous composites 430
1.17.3.1.4 Damage 431
1.17.3.2 Macro-Mechanics of Fibrous Composite Constructions 431
1.17.3.2.1 Fibrous laminates 431
1.17.3.2.2 Woven composites 433
1.17.3.3 Global Mechanics of Composite Structures 433
1.17.4 Modeling Options for Fibrous Composites 433
1.17.4.1 Averaging Models of Fibrous Composites 433
1.17.4.2 RVE of Fibrous Composites 434
1.17.4.3 RVE of Woven Composites 435
1.17.5 Applications 436
1.17.5.1 Unidirectional and Woven Composites 436
1.17.5.2 Composite Laminates 440
1.17.5.3 Composite Structures 444
1.17.6 Closing Remarks 448
References 449
Relevant Website 450

1.17.1 Introduction

Design of structural components fabricated from fibrous composites for a specific application involves several physical
phenomena and geometrical parameters, which might be related to multiple length scales. Multiple loads might also be
involved simultaneously while the composite structure is in service or under maintenance. This includes mechanical loads,
thermal variations, and electric fields. The latter has emerged as part of structural health monitoring techniques involving electro-
active filaments. Additionally, damage and/or inelastic deformation are inevitable in composite structures at multiple length
scales.
With this complex situation faced by designers of composite structures, reliable prediction of their behavior using analytical
and computational techniques would offer a significant support, not only during the design phase and material selection, but
also during fabrication and in planning for maintenance. The key to reliable prediction of the behavior of composite structures
is multiscale modeling, which considers physical phenomena and geometrical details at every length scale, as well as mechanics-
related interactions across length scales. For example, it is of interest to know how a localized failure at the fiber/matrix interface of
a fibrous composite, or localized plastic deformation in its matrix material would affect a structural component of complex
geometry and manufactured from multidirectional plies, or a woven architecture. The challenge is to develop a robust modeling
tool which works seamlessly and accurately across the length scales found in composite structures.
In this chapter, multiscale mechanics across length scales found in fibrous composite and woven structures is considered
within the context of stress and strain transformations, which can be caused by a variety of physical phenomena. This includes
thermal strains, plastic and viscoplastic deformations, piezoelectric and pyroelectric effects, and damage. We begin by describing
the length scales dealt with in this chapter, followed by a description of their mechanics and interactions. This includes micro-
mechanics of fibrous composites, macro-mechanics of fibrous composite constructions, and global mechanics of composite
structures. Treatment of the macro-mechanical length scale includes fibrous laminates and woven composites; both plane weave
and 3D weaves.
Specific models for unidirectional fibrous composites and woven architectures are outlined and utilized in several applications,
which illustrate how multiscale modeling can predicate the behavior of a variety of composite materials and structures seamlessly
across multiple length scales. Local transformation fields caused by a change of temperature, an electric field, and damage are
included in the multiscale modeling illustrations offered.

426 Comprehensive Composite Materials II, Volume 1 doi:10.1016/B978-0-12-803581-8.09892-1


Multiscale Mechanics of Composite Materials and Structures 427

1.17.2 Structural Composites Length Scales

Starting with the basic fiber and matrix building blocks, composite structures are assembled essentially from three levels, which
with regard to analysis define the length scales. Thinking in terms of a manufacturing sequence, the first length scale is that of
a microstructure of a fibrous composite monolayer, which describes the configuration and interaction between the fiber and the
matrix. The second length scale is a macrostructural one, which assembles several unidirectional fibrous monolayers, or plies, into
a multidirectional, multilayered plate or shell to sustain multiple loading regimes, which may be a combination of mechanical,
thermal, and possibly electrical loads. Finally, there is a structural length scale, which describes how the laminated assembly is
shaped into structural components with complex geometries to sustain service loads and perform defined functions.
Fig. 1 depicts the length scales described above. On the microstructural level, the architecture could be that of a unidirectional
composite reinforced with continuous fibers which are aligned into a unified direction, Fig. 1(a), and otherwise assume several
arrangements in the cross section depending on the fiber diameter and the manufacturing process. In other situations,
the continuous fibers are interlaced in two orthogonal directions to form a plane, fibrous composite weave, Fig. 1(b). On the
macrostructural level, a number of either of these monolayers is stacked to form a multilayered composite laminate. This may
take the form of a multidirectional, fibrous laminate, Fig. 1(c), or a plane weave laminate, Fig. 1(d). In 3D weaving processes, on
the other hand, the macrostructure is produced as a stack of non-interlaced, plane layers of orthogonal fibers held together with
continuous fibers in the “third” or “z” direction, Fig. 1(e).
Fig. 2 shows a few structural applications in which the above described macrostructural composite laminates are utilized. For
the purpose of manufacturing, the structural components are produced as net shapes from the composite laminates. For the
purpose of analysis, on the other hand, each length scale is treated on its own, taking into consideration the effects it causes on
both the higher and lower length scale. In particular, and in the treatment of multiscale mechanics in this chapter, the effect of fiber
and matrix properties, their volume content and the damage they may see, particularly at their interface is carried forward through
the macroscale and onto the structural scale. Similarly, the loads applied at the structural scale lead to stresses and deformation on
the microscale. The challenge is to model this two-way interaction between the three length scales seamlessly and efficiently, and
yet reliably.

1.17.3 Mechanics and Interactions of Composites Length Scales

In this section, mechanics of each length scale are outlined including interactions with adjacent length scales. For each length scale
idealized geometry and architectures are considered and governing equations are laid down. In this formulation, the fiber and
matrix are considered to be elastic but may sustain transformation, or eigen strains. The latter could be a result of a thermal change,
electro-elastic and/or electro-thermal coupling, or inelastic transformation.

Fig. 1 Length scales of structural composites: (a) microscale of a fibrous composite, (b) microscale of a plane weave, (c) macroscale of a
fibrous laminate, (d) macroscale of a plane weave laminate, (e) macroscale of a 3D-woven composite. Reproduced from Bahei-El-Din, Y.A.,
Rajendran, A.M., Zikry, M.A., 2004. A micromechanical model for damage progression in woven composite systems. International Journal of Solids
and Structures 41, 2307–2330.
428 Multiscale Mechanics of Composite Materials and Structures

Fig. 2 Examples of composite structures.

1.17.3.1 Micro-Mechanics of Fibrous Composites


Treatment of a composite reinforced with continuous, parallel filaments follows the early work of Budiansky1 and Hill2,3
on averaging models, and of Dvorak and Teply4, Michel et al.,5 Iwakuma and Nemat-Nasser,6 and Nemat-Nasser and Hori7
on periodic array models. In principle, these models assume some form of the microstructure, whether explicitly as in the
periodic array models, or implicitly in terms of a set of internal constraints as in averaging models, and estimate the local stresses
in the fiber (f) and matrix (m) phases in terms of their properties and volume fractions, vf and vm such that vf þ vm ¼ 1. This is
achieved by selecting a representative volume V of the fibrous composite and assuming the local stresses and strains to be
P
pointwise uniform over a number of subvolumes Vr, r¼ 1,Q, such that V¼ Vr. The volume fraction of a subvolume is denoted by
P
cr ¼Vr/V such that cr ¼ 1.
Adopting matrix notation in which matrices are denoted by boldface, italic letters, the local stresses and strains
are (6  1) vectors which list the independent components of the stress and strain tensors. Hence the local stresses and
strains are denoted, respectively, by rr ¼ (s11,s22,s33,s23,s31,s12) and er ¼ (e11,e22,e33,2e23,2e31,2e12). The lamina overall stresses
r and strains e, referred to the material principal axes xk , k¼1,2,3, are expressed as volume averages of their local counterparts,
rr and er, by

X X
r¼ c r rr ; e¼ c r er ð1Þ
r ¼ 1;Q r ¼ 1;Q

Let kr and lr denote transformation (eigen) stresses and strains residing in subvolumes of the representative volume V such that
they cannot be removed by mechanical unloading. Hence, the total fields found in a subvolume are given by the sum of the elastic
response caused by mechanical loads and the response due to the eigen fields8;

rr ¼ Lr er þ kr ; er ¼ Mr rr þ lr ; r ¼ 1; Q ð2Þ

where Lr, Mr ¼ L1


r are (6  6) matrices representing elastic stiffness and compliance, and kr ¼  Lrlr.
Multiscale Mechanics of Composite Materials and Structures 429

Considering transversely isotropic elastic symmetry, with x3 the axis of rotational symmetry, the elastic stiffness matrix is
written in terms of the Engineering moduli and Hill’s moduli9 as
2 31 2 3
1=ET nL =ET nL =EL 0 0 0 ðk þ mÞ ðk  mÞ ℓ 0 0 0
6 7 6
6 1=ET nL =EL 0 0 0 7 6 ðk þ mÞ ℓ 0 0 07
7
6 7 6 7
6 1=E 0 0 0 7 6 n 0 0 07
6 L 7
Lr ¼ 6 7 ¼6
6
7 ð3Þ
6 1=G L 0 0 7 6 p 0 07
7
6 7 6 7
6 SYM: 1=GL 0 7 4 05
4 5 SYM: p
1=GT m
Here, EL, GL, nL are longitudinal Young’s modulus, shear modulus and Poisson’s ratio, and ET, GT, nT ¼ ET/2GT  1 are their
transverse counterparts. The Engineering moduli and Hill’s moduli are related by9
 1
k ¼  1=GT  4=ET þ 4n2L =EL ; ℓ ¼ 2knL ð4Þ

n ¼ EL þ 4kn2L ¼ EL þ ℓ2 =k; m ¼ GT ; p ¼ GL ð5Þ

1.17.3.1.1 Matrix properties


With regard to elastic symmetry the matrix is assumed to be isotropic. Referring to Eq. (3), elastic moduli of the matrix are then
denoted by EL ¼ ET ¼ E, GL ¼ GT ¼ G, nL ¼ nT ¼ n ¼ E/2G  1. In this case Hill’s moduli are k ¼ G/(1  2n), ℓ ¼ 2kn, n¼ 2k(1  n),
G¼ m¼p.
The eigen stresses and strains, Eq. (2), define all fields, which are not recovered by mechanical unloading. For example, if the
matrix remains in the elastic domain, a uniform thermal change expressed as time derivative, y,_ causes the following eigen strain
and stress rates:
_
m_ 11 ¼ m_ 22 ¼ m_ 33 ¼ ya; m_ 23 ¼ m_ 31 ¼ m_ 12 ¼ 0 ð6Þ

l_ 11 ¼ l_ 22 ¼ l_ 33 ¼ ya
_ ð2n  1Þ=E; l_ 23 ¼ l_ 31 ¼ l_ 12 ¼ 0 ð7Þ
where a is the coefficient of thermal expansion.
If the matrix is elastic–plastic or elastic–viscoplastic, the eigen strain is given by the inelastic part of the deformation. Con-
sidering temperature-dependent plastic deformation and a Mises yield criterion,10 time rate of the eigen strain is given by
pffiffiffiffiffiffiffiffi rffiffiffi !
3=2 3 T
l_ ¼ _  Y ðyÞy n þ ay_
ðn  rÞ 0 _ ð8Þ
HðyÞ 2
Here, n is the unit normal to the yield surface at the current stress point, H(y) is the plastic tangent modulus of the stress–strain
curve under simple tension, Y is the yield stress in simple tension and Y0 its temperature rate, and a ¼(a,a,a,0,0,0) is the vector
of thermal expansion coefficients. For the Mises yield surface and kinematic hardening, the unit normal to the yield surface is
given by
1 _ _ _ _ _ _  _
nT ¼ pffiffiffiffiffiffiffiffi s 11 ; s 22 ; s 33 ; 2s 23 ; 2s 31 ; 2s 12 ; r ¼ s  a ð9Þ
2=3YðyÞ
where s is the deviatoric stress and a is the center of the yield surface in the deviatoric stress space.
For a thermo-viscoplastic matrix, a rate dependent theory with a Mises yield, or equilibrium, surface for non-proportional
loading can be utilized.11 In this case, the eigen strain rate is given by
rffiffiffiffiffiffiffiffiffiffiffiffi !
3
l_ ¼ kðyÞLpðyÞ n þ ay_ ð10Þ
2
where k(y) and p(y) are the material parameters and n is the unit normal to the yield surface at the current equilibrium stress point,
Eq. (9). In a deviatoric stress space, L is the effective overstress. This is a measure of the distance between the actual stress s and the
equilibrium stress s, which satisfies the yield criterion. Hence, L vanishes if the stress point lies on, or falls inside the yield surface.
For stress states which fall outside the yield surface,
 
3    1=2
L¼ s  s : s  s ð11Þ
2

1.17.3.1.2 Fiber properties


With regard to elastic symmetry the fiber is assumed to be transversely isotropic, Eq. (3). Considering a thermo-electro-mechanical
behavior,12,13 eigen strains and stresses are caused by an electric field intensity, E¼ (E1,E2,E3), and/or a uniform change of temperature y,
l ¼ dT E þ ay ð12Þ

k ¼  e E  Lay
T
ð13Þ
430 Multiscale Mechanics of Composite Materials and Structures

where d and e are the piezoelectric constants. For transverse isotropy with x3 the axis of rotational symmetry, they, and the coefficients of
thermal expansion take the following form14,15:
2 3 2 3
0 0 d31 0 0 e31
6 0 0 d31 7 6 0 0 e31 7
6 7 6 7
6 7 6 7
6 0 0 d 7 6 0 0 e 7
dT ¼ 6 7; eT ¼ 6 7; a ¼ ðaT ; aT ; aL ; 0; 0; 0Þ
33 33
6 0 d 7 6 0 e 7 ð14Þ
6 15 0 7 6 15 0 7
6 7 6 7
4 d15 0 0 5 4 e15 0 0 5
0 0 0 0 0 0
Here, aL and aT denote, respectively, the longitudinal and transverse coefficients of thermal expansion. The piezoelectric constants are
related by15
e31 ¼ 2kd31 þ ℓd33 ; e33 ¼ 2ℓd31 þ nd33 ; e15 ¼ pd15 ð15Þ
For completeness, the following dual equations are written for the electric displacement, D¼(D1, D2, D3) and electric field, E:
D ¼ κe E þ ee  qy ¼ κs E þ dr  qy ð16Þ

E ¼ κ1
s D þ gr þ hy ð17Þ
Such that
g ¼ κ1
s d; h ¼ κ1
s q ð18Þ
Here, g is the voltage constant, κs and κe ¼ κs  de are the permittivity matrices measured at constant stress and constant strain,
T

respectively, and q is the pyroelectric constant. The piezoelectric constants are interrelated by15
g31 ¼ d31 =κ33s ; g33 ¼ d33 =κ33s ; g15 ¼ d15 =κ11s ð19Þ
For a transversely isotropic fiber, the permittivity and the pyroelectric constant take the form 14,16

2 3
κ11 0 0
6 0 7
κ ¼ 4 0 κ11 5; q ¼ ð0; 0; q3 Þ ð20Þ
0 0 κ33

1.17.3.1.3 Unidirectional fibrous composites


In analogy with Eq. (2), constitutive behavior of a fibrous composite can be described in the material principal axes, xk , as
r ¼ Leþk; e ¼ Mrþl ð21Þ
1
The overall stiffness L and compliance M ¼ L are given in terms of their phase counterparts by8
X X
L¼ c r Lr Ar ; M ¼ c r Mr Br ð22Þ
r ¼ 1;Q r ¼ 1;Q

where Ar, Br are the strain and stress concentration factors, such that
er ¼ Ar e; rr ¼ Br r ð23Þ
Substituting in Eq. (1), the following identities are found:
X X
L¼ cr Ar ¼ Ι; cr Br ¼ I ð24Þ
r ¼ 1;Q r ¼ 1;Q

where I is the identity matrix. In the absence of eigen fields, Eqs. (21) and (23) lead to the following identities:
Ar M ¼ Mr Br ; Br L ¼ Lr Ar ð25Þ
The ply eigen strains and stresses are given in terms of their phase counterparts by8
X X
k¼ c r ATr kr ; l ¼ c r BTr lr ð26Þ
r ¼ 1;Q r ¼ 1;Q

The local stresses in the fiber and matrix, where physical phenomena originate, are of interest. These are given by superposition
of the overall effect, Eq. (23), and the local effects caused by the eigen strains and stresses,8
X X
er ¼ Ar eþ Drs ls ; rr ¼ Br rþ F rs ks ; r ¼ 1; Q ð27Þ
s ¼ 1;Q s ¼ 1;Q

Here, Drs and Frs are the strain and stress influence functions. They follow the following identities17:
X X
cs Dsr ¼ 0; cs F sr ¼ 0; r ¼ 1; Q ð28Þ
s ¼ 1;Q s ¼ 1;Q

Considering plane stress states in the x2 x3 -plane, the lamina stresses in the principal material axes are reduced to
^ ¼ ðs22 ; s33 ; s23 Þ and the corresponding strain is denoted by ^e ¼ ðe22 ; e33 ; 2e23 Þ. These are extracted from the full (6  1) stress and
r
Multiscale Mechanics of Composite Materials and Structures 431

strain vectors as
^ ¼ ^I T r; ^e ¼ ^I T e;
r ^I ¼ ði2 ; i3 ; i4 Þ ð29Þ
^ ^
where ik is a (6  1) vector with a unit entry in the kth position. In this case the stiffness matrix, L, and compliance, M, are given in
terms of Hill’s moduli by18
2 3
4km 2mℓ 0
^ 1¼L
 ^¼ 1 6 7
M 4 2mℓ EL k þ mn 0 5 ð30Þ
kþm
0 0 pðk þ mÞ
^ l.
Similarly, Eq. (29) can be utilized to reduce the full eigen stress and strain vectors, k, l, to their plane stress counterparts, k, ^
If it is desired to refer the lamina plane stresses and corresponding strains to an x2x3-plane, which is rotated with respect to the
material principal axes such that the angle between the x3 -axis and the x1-axis is j, they, and the overall stiffness and compliance
matrices can be found as18
^ ^e ¼ RT^e; L
^ ¼ N T r;
r ^
^ ¼ N T LN; ^ ¼^
M ^
L1 ¼ RT MR ð31Þ
2 3
1
sin2 j cos2 j  sin2j
6 2 7
6 7
RT ¼ N 1 ¼ 6
6 cos j sin j
2 2 1
sin2j
7
7 ð32Þ
4 2 5
sin2j sin2j cos2j

1.17.3.1.4 Damage
Damage in a fibrous composite is modeled by introducing auxiliary eigen fields in the subvolumes to evacuate the stresses, which are
caused by the applied loads and satisfy certain failure criteria.19 If such eigen stress field is denoted by kAUX
s and introduced in subvolumes
Vs, s¼ 1,O, O r Q, it can be found by limiting the total stress components affected by the failure criterion to zero. Hence, Eq. (27),
X
rs ¼ Bs rþ F sq kAUX
q ¼ 0; s ¼ 1; O ð33Þ
q ¼ 1;O

and the auxiliary eigen stress field is thus determine. The lamina eigen stress and strain can then be found from Eq. (26).

1.17.3.2 Macro-Mechanics of Fibrous Composite Constructions


Unidirectional fibrous composites are utilized to form a variety of composite constructions to sustain service loads. This includes
nonwoven and woven architectures. In this section, governing equations, which describe the behavior of fibrous laminates and
certain woven architectures are presented.

1.17.3.2.1 Fibrous laminates


Multidirectional laminates consisting of a fully bonded stack of n fiber reinforced plies are considered. If the ply thickness is
P P
denoted ti, i¼ 1,n, leading to a total thickness h ¼ i ¼ 1;n ti , the ply volume fraction is given by ci ¼ ti/h such that i ¼ 1;n ci ¼ 1. For
description of stresses and strains an overall coordinate system xj, j¼1,2,3, is selected such that the x1x2-plane coincides with
mid-plane of the laminate, and the x3-axis is in the thickness direction.
The overall mechanical loads consist of membrane force, N ¼ ½N1 ; N2 ; N12 , and bending moment, M ¼ ½M1 ; M2 ; M12 , Fig. 3.
These cause stresses and strains, which vary pointwise across the laminate thickness. Hence, the in-plane stresses and strains
averaged over the thickness of the individual plies can be written as
Z Z
1 zi þti =2 1 zi þti =2
r^i ¼ ^ðzÞdz; ^ei ¼
r ^eðzÞdz ð34Þ
ti zi ti =2 ti zi ti =2
where z¼x3 is the Cartesian coordinate in the direction perpendicular to the laminated plate (Fig. 3), with the origin located on
the mid-plane of the laminate, and zi is the x3 coordinate of the mid-plane of the ply. Resultants of the ply stresses are then equal
to the applied membrane forces and bending moments;
Z h=2 !
X Z zi þti =2 X
N¼ ^ðzÞdz ¼
r ^ðzÞdz ¼
r ^i
ti r ð35Þ
h=2 i ¼ 1;n zi ti =2 i ¼ 1;n

Z Z !
h=2 X zi þti =2
M¼ rðzÞdz ¼
z^ rðzÞdz
z^ ð36Þ
h=2 i ¼ 1;n zi ti =2

For thin laminates the transverse shear deformations are negligible and the in-plane strains are assumed to vary linearly across
the thickness with the z coordinate. Hence
^eðzÞ ¼ eo þ zK o ð37Þ
432 Multiscale Mechanics of Composite Materials and Structures

Fig. 3 Geometry of a fibrous laminate. Reproduced from Bahei-El-Din, Y.A., Micheal, A., 2012. Micromechanical modeling of multifunctional
composites. In: Proceedings of the International Mechanical Engineering Congress & Exposition. ASME 2012 IMECE, November 9–15, 2012,
Houston, TX. Paper No. 89304. New York: American Society of Mechanical Engineers.

where eo ¼ ½eo11 ; eo22 ; 2eo12  is the strain at mid-plane of the laminate, and Ko ¼ [K1,K2,K12] is the curvature with respect to the
mid-plane.
Following the formulation for thin laminates and its modification to account for eigen stresses and strains, which originate in
the individual plies,20 the mid-plane strain and curvature of the laminate are written as
0 0
eo ¼ A0 N þ B0 M þ f o ; K o ¼ C0 N þ D0 M þ go ð38Þ
where f 0 o and g0 o are the eigen strain and curvature, respectively. The coefficient matrices in the first two terms of Eq. (38) are a
function of the elastic properties of the individual plies, their volume fractions and the laminate layup;

A0 ¼ ðI  B0 BÞA1 ; B0 ¼  A1 BD0 ; C0 ¼  D0 BA1 ; D0 ¼ ½D  BA1 B1 ð39Þ

X X X 1 
A¼ ^i ;
ti L B¼ ^i ; D ¼
ðti zi ÞL ti ti2 þ z2i L^i ð40Þ
i ¼ 1;n i ¼ 1;n i ¼ 1;n
12

The ply stresses are found in analogy of Eq. (27) as the superposition of the stress caused by the overall membrane forces and
bending moments, and that caused by the eigen stresses,20
X
^i ðzi Þ ¼ Pi N þ Qi M þ
r U ij ^
kj ð41Þ
j ¼ 1;n

The coefficient matrices, Pi and Qi, denote stress distribution factors, and Uij denotes stress influence functions. They vary point
wise along the laminate thickness, and are a function of elastic moduli and volume fractions of the plies, and the laminate layup20;
^i ðA0 þ zi C0 Þ;
Pi ¼ L ^i ðB0 þ zi D0 Þ
Qi ¼ L ð42Þ
 
U ij ¼ δij I  tj Pi  tj zj Qi ð43Þ

where δij is the Kronecker’s tensor. Equilibrium of the lamina forces provides the following identities:
X X
ti Pi ¼ I; ðti zi ÞPi ¼ 0 ð44Þ
i ¼ 1;n i ¼ 1;n

X X
ti Qi ¼ 0; ðti zi ÞQi ¼ I ð45Þ
i ¼ 1;n i ¼ 1;n

Finally, the eigen strain, f 0 o , and curvature, g o , at the laminate mid-plane are caused by the ply eigen stresses, ^
0
ki , i¼ 1,n. They are
20
given by Bahei-El-Din et al. as

f 0 o ¼  B0 g  A0 f o ; g 0 o ¼  C0 f o  D0 go ð46Þ
X X
fo ¼ ti ^
ki ; go ¼ ðti zi Þ^
ki ð47Þ
i ¼ 1;n i ¼ 1;n
Multiscale Mechanics of Composite Materials and Structures 433

1.17.3.2.2 Woven composites


Among the numerous variety of woven architectures which utilize bundles of fibrous composites, 2D, simple weave (Fig. 1(d)) and
3D-woven constructions (Fig. 1(e)) are considered. In both cases, fiber bundles are used to construct a regular pattern, which is
interlaced in two orthogonal directions in 2D-woven composites, and non-interlaced in three orthogonal directions in 3D-woven
composites. Considering overall stress and strain states which are uniform, the overall response can be determined from a small
domain of the woven construction as considered in Section 1.17.4.
In some applications, 2D-woven composites are stacked to form a laminated plate (Fig. 1(d)). In this case, the formulation of
Section 1.17.3.2 applies to the laminate while the individual, woven plies are treated as described in Section 1.17.4 for the selected
unit cell.

1.17.3.3 Global Mechanics of Composite Structures


Composite structures of complex geometries are handled within the framework of the finite element method. At this structural
length scale, and depending on the geometry and architecture of the composites utilized, the underlying phenomena in the
microscale and their effect on the macroscale are considered in evaluating the structural response.
Consistent with the transformation field approach utilized so far in the micro- and macro-length scales, a composite structure is
treated as elastic with superimposed deformations stemming from the non-mechanical fields. The latter are represented in subvolumes of
a unidirectional, fibrous composite (Section 1.17.3.1.3) or a woven composite by a set of eigen stresses, kr, and strains, lr, r¼ 1,Q, and
this leads to overall, in-plane eigen stresses, ^
k, and strains, l^. When a composite structure is subdivided in the framework of finite
elements into laminated or woven composite elements, the overall eigen stresses are used to update the structural nodal loads. In essence,
this is a wider application of thermal loads, which are typically included in quasi-static and implicit dynamic finite element procedures to
account for other non-mechanical effects such as thermo-electro-mechanical coupling and damage. This permits the use of existing finite
element codes to determine the response of composite structures under these effects without the need to develop new codes.
Hence, if a composite structure is modeled with a finite element mesh consisting of M elements, which sustain eigen stresses km,
m¼1,M, the nodal forces, qm, which equilibrate the eigen stresses are given by the principle of virtual work,
Z
qm ¼ Gm km dVm ð48Þ
Vm
where Cm represents spatial derivatives of the displacement field in element m, and Vm is the volume of the element. In explicit,
dynamic finite element procedures, on the other hand, the elemental stresses rather than nodal forces are updated according to the
underlying constitutive behavior. When eigen stresses exist, updates of the element stresses are given by superposition of stresses
caused by the mechanical loads applied to the structure, and those caused by the eigen stress field.

1.17.4 Modeling Options for Fibrous Composites

In this section we provide formulations for specific models of fibrous and woven composites, which will be utilized in imple-
menting the multiscale scheme of Section 1.17.3.

1.17.4.1 Averaging Models of Fibrous Composites


These models utilize Eshelby’s solution21 of an ellipsoidal inhomogeneity in an infinite matrix under remotely applied uniform
fields to estimate the average stresses and strains in the matrix and the fiber reinforcement. Referring to Section 1.17.3.1, the
number of subvolumes of a representative volume element (RVE) of volume V in this case is Q ¼ 2. In this context, we list the
overall moduli of a fibrous composite given by the self-consistent model,1,3 and the Mori–Tanaka model.22 In both models,
Eshelby’s solution for continuous, parallel cylindrical fibers embedded in an infinite, transversely isotropic matrix is utilized to
determine the stress and strain concentration factors (Eq. (23)) and hence determine the overall moduli (Eq. (22)).
In the self-consistent model, the matrix assumes the overall moduli of the composite and the fiber concentration factors
are thus determined in terms of the yet unknown overall moduli. This leads to partially coupled equations for the overall
moduli. Considering Hill’s moduli,9 the overall moduli m, p, and k are found from the following uncoupled equations2,3:

cf kf cm km cf mm cm mf
þ ¼2 þ ð49Þ
kf þ m km þ m mm  m mf  m
1 cf cm 1 cf cm
¼ þ ; ¼ þ ð50Þ
2p p  pm p  pf kþm kf þ m km þ m
Here, cf and cm are volume fractions of the fiber and matrix, and mr, pr, and kr, r¼f,m are Hill’s moduli of the fiber and
matrix. The remaining two moduli, ℓ, n, are found from the universal connections2,3
k  kf k  km ℓ  cf ℓf  cm ℓm kf  km
¼ ¼ ¼ ð51Þ
ℓ  ℓf ℓ  ℓm n  cf nf  cm nm ℓf  ℓm
In the Mori-Tanaka model, Eshelby’s solution is utilized in two steps to find explicit forms for the fiber and matrix con-
centration factors.23 First, the fibers are embedded into the matrix to determine partial strain concentration factors, which define
434 Multiscale Mechanics of Composite Materials and Structures

the fiber uniform strain in terms of that of the matrix. This connection facilitates computation of the concentration factors of the
fiber and matrix when embedded in the composite aggregate. The result was utilized by Chen et al.24 to determine the Mori-Tanaka
overall moduli of a fibrous composite in the following explicit form:
 
2cf pm pf þ cm pm pf þ p2m mm mf ðkm þ 2mm Þ þ km mm ðcf mf þ cm mm Þ
p¼ ; m¼ ð52Þ
2cf pm þ cm ðpf þ pm Þ km mm þ ðkm þ 2mm Þðcf mm þ cm mf Þ
cf kf ðkm þ mm Þ þ cm km ðkf þ mm Þ cf ℓf ðkm þ mm Þ þ cm ℓm ðkf þ mm Þ
k¼ ; ℓ¼ ð53Þ
cf ðkm þ mm Þ þ cm ðkf þ mm Þ cf ðkm þ mm Þ þ cm ðkf þ mm Þ
ℓf  ℓm
n ¼ cm nm þ cf nf þ ð1  cf ℓf  cm ℓm Þ ð54Þ
kf  km
Having determined the overall moduli of a fibrous composite, the stress and strain concentration factors can then be found
from25
   
Ar ¼ ðLr  Ls Þ1 LLs =cr ; Br ¼ ðMr  Ms Þ1 MMs =cr ; r; s ¼ f ; m ð55Þ
The stress and strain influence functions (Eq. (27)) are also given in closed form as8
Drf ¼ ðI  Ar ÞðLf  Lm Þ1 Lf ; Drm ¼  ðI  Ar ÞðLf  Lm Þ1 Lm ; r ¼ f; m ð56Þ

Frf ¼ ðI  Br ÞðMf  Mm Þ1 Mf ; F rm ¼  ðI  Br ÞðMf  Mm Þ1 Mm ; r ¼ f; m ð57Þ

1.17.4.2 RVE of Fibrous Composites


If microgeometry of a fibrous composite subjected to overall uniform fields is approximated by a periodic dispersion of fibers
embedded in the matrix, a unit cell or RVE V can be selected for evaluation of the local fields and the overall response.4,6,7,26–29
The unit cell V is then divided into subvolumes, Vr, r¼ 1,Q, such that each subvolume belongs to either the matrix or the fiber, and
Q42. Assuming uniform stresses and strains within each subvolume, the treatment in Section 1.17.3.1.3 for the unidirectional
composite aggregates apply provided that the overall fields, r, e are defined, and the stress and strain concentration factors, Ar, Br,
(Eq. (23)), as well as the influence functions, Drs, Frs (Eq. (27)), r,s ¼1,Q, are determined.
The finite element method is utilized to evaluate the above quantities. First, the representative volume is supported against
rigid-body motion and subjected to nodal forces, which equilibrate the overall stresses, and is subjected to displacement boundary
conditions which reflect periodicity of the local fields, and to generalized plane strain boundary conditions. Considering for
example a periodic hexagonal dispersion of the fibers,4 Fig. 4, a unit cell, or a RVE, can be selected and subdivided into elements as
shown in Fig. 5. Assuming a linear displacement field for an equivalent, macroscopically homogeneous volume, the nodal forces
applied at the independent degrees of freedom, Fig. 5, are derived as30
H H Hξ H 1
p1 ¼  s22 ; p2 ¼  s23 ; p3 ¼  s33 þ s23  s31 ð58Þ
ξ ξ 2 2ξ 2
Hξ H 1
p4 ¼ s33 þ s23  s31 ; p5 ¼ s11 ; p6 ¼ s12 ð59Þ
2 2ξ 2
pffiffiffi
where ξ ¼ 4 3, and H is the length of the unit cell in the axial direction, Fig. 5.
Referring to Eq. (23), the kth column, k ¼ 1,6, of the stress concentration factors, Br, r,s ¼ 1,Q, is given by the stresses caused in
Vr, r¼1,Q, by an overall stress component sk ¼ 1. Similarly, Eq. (27) suggests that the kth column, k¼ 1,6, of the influence
matrix Drs, r,s¼1,Q, represents the strains caused in Vr, r¼1,Q, by a transformation strain mk ¼ 1 applied to Vs, while the unit

Fig. 4 Hexagonal periodic array idealization of fibrous composites. Reproduced from Bahei-El-Din, Y.A., Botrous, A.G., 2003. Analysis of
progressive fiber debonding in elastic laminates. International Journal of Solids and Structures 40, 7035–7053.
Multiscale Mechanics of Composite Materials and Structures 435

Fig. 5 Unit cell of a hexagonal array model of fibrous composites. Reproduced from Bahei-El-Din, Y.A., Botrous, A.G., 2003. Analysis of
progressive fiber debonding in elastic laminates. International Journal of Solids and Structures 40, 7035–7053.

Fig. 6 Idealization of woven composites: (a) plane weave and (b) 3D weave. Reproduced from Bahei-El-Din, Y.A., Rajendran, A.M., Zikry, M.A.,
2004. A micromechanical model for damage progression in woven composite systems. International Journal of Solids and Structures 41,
2307–2330.

cell is fully constrained. This translates to equivalent forces applied at the nodes of Vs by Eq. (48), where the eigen stress is given by
ks ¼  Lsls, and Ls is the elastic stiffness. Hence, for uniform fields, the nodal forces are
qs ¼  Vs CTs Ls ls ð60Þ

1.17.4.3 RVE of Woven Composites


Considering actual microgeometry of 2D and 3D-woven composites, Fig. 1(d) and (e), it is seen that in a 2D plane weave the warp
and weft fiber bundles are interlaced in two orthogonal directions, while in 3D-woven composites they are not interlaced and tied
together by a fiber bundle in the z-direction. These regular woven patterns can be idealized as shown in Fig. 6. In this case, and
under uniform overall stresses and strains, the local fields are periodic and a unit cell can be identified to represent the overall
response, Fig. 7.15,31
The unit cell is a right, parallelepiped domain, which contains an entire fiber in each of the warp and weft directions. In the
third direction, the unit cell contains a full z-fiber for a 3D weaves, and a full layer of 2D-woven ply. Treatment of the unit cell for
fibrous composite given in Section 1.17.4.2 for evaluation of the stress and strain concentration factors and influence functions in
the subvolumes then applies provided that the overall stress is translated into nodal loads at the independent degrees of freedom.
For this purpose, generalized plane strain conditions are introduced at opposite faces of the unit cells of Fig. 7.31 For example, let r
and s denote two nodal points located at the negative and positive faces of the unit cell that are perpendicular to the x1-axis such
that x2 ¼ x2 and x3 ¼ x3 . The displacement boundary condition for this pair of nodes is ui  ui ¼ u
ðrÞ ðsÞ ðrÞ ðsÞ ðrÞ ðsÞ
i , i¼ 1,2,3, where ui is the
436 Multiscale Mechanics of Composite Materials and Structures

Fig. 7 Unit cell of woven composites: (a) plane weave and (b) 3D weave. Reproduced from Bahei-El-Din, Y.A., Rajendran, A.M., Zikry, M.A.,
2004. A micromechanical model for damage progression in woven composite systems. International Journal of Solids and Structures 41,
2307–2330.

displacement in the direction of the xi-axis, and u is an independent relative displacement. Similar boundary conditions are
applied to the opposite sides of the unit cell that are perpendicular to the x2 and x3 axes. In addition, displacement boundary
conditions which eliminate rigid-body translations and rotations are introduced as shown for example in Fig. 7(b).
Assuming a linear spatial variation for the overall displacement field in a homogenized unit cell, the nodal forces are given in
terms of the overall stresses by31

p1 ¼ ℓ2 ℓ3 s11 ; p2 ¼ ℓ1 ℓ3 s22 ; p3 ¼ ℓ1 ℓ2 s33 ð61Þ

p4 ¼ ℓ1 ℓ2 s23 ; p5 ¼ ℓ2 ℓ3 s31 ; p6 ¼ ℓ1 ℓ3 s12 ð62Þ

where ℓ1 , ℓ2 , ℓ3 are the side lengths of the unit cell, which are parallel, respectively, to the x1, x2 and x3 axes.

1.17.5 Applications

1.17.5.1 Unidirectional and Woven Composites


Three applications are offered at the micro-length scale, one for a fibrous composite, and two for a 3D-woven composite. They
illustrate how the multiscale mechanics approach can be utilized to model damage in the matrix and/or its interface with the fiber,
and electromechanical coupling found in some fibers.
First, a glass ceramic matrix (CAS), reinforced with a silicon-carbide fiber at 39% and subjected to in-plane loading is
considered.19 Assuming Ef ¼ 170 GPa, vf ¼ 0.25, elastic moduli of the matrix were back calculated by matching the overall moduli
given by the Mori–Tanaka averaging model (Eqs. (52)–(54)) with those measured by Wooh and Daniel.32 This provided Em ¼98
GPa, vm ¼0.135. Damage was limited to the fiber/matrix interface where debonding occurs if the radial stress reaches 50 MPa, and
sliding occurs if the longitudinal shear stress reaches 80 MPa. A coefficient of friction of 0.3 was assumed at the interface. Details of
the failure criteria utilized at the fiber/matrix interface considering averaging models and periodic array models of fibrous
composites can be found in Bahei-El-Din and Botrous.19
The response of a single ply subjected to off-axis, normal stress is shown in Fig. 8. Predictions from a unit cell based on a
hexagonal array model (Fig. 5) and the Mori–Tanaka averaging model are shown for loading in the fiber direction, 0 degree, the
transverse direction, 90 degree, and at several intermediate angles. Under the fiber and matrix moduli and volume fractions given
above, the unit cell model provides EL ¼ 132 GPa and ET ¼ 126 GPa. Accordingly, the loading direction has a small effect on the
overall elastic stiffness for this composite system, and the changes seen in Fig. 8 in the overall stiffness beyond the onset of
interface failure reflect only the effect of damage on the overall response. This is clearly seen in Fig. 9 where the damage strain,
computed as the difference between the overall strain and the elastic strain, is plotted versus the applied stress.
The second application is for a T300-carbon/epoxy, 3D-woven composite,31 Fig. 1(e), which is modeled with the unit cell
shown in Fig. 7(b). The fiber bundles in the warp, weft and z-directions are treated as fibrous composites. The fiber volume fraction
in the bundles is 0.57. Mori–Tanaka estimates of the overall elastic properties of the fiber bundles are found as E1 ¼ 134.64 GPa,
E2 ¼E3 ¼7.25 GPa, v23 ¼ 0.403, v12 ¼ v13 ¼ 0.236, G12 ¼ G13 ¼ 2.9 GPa, G23 ¼ 2.58 GPa.
Damage is considered in the form of internal fiber bundle sliding, interface failure by peeling and/or sliding, and matrix failure.
The affected stresses were assumed to fall down to zero following a softening behavior of the stress–strain response.31
Multiscale Mechanics of Composite Materials and Structures 437

Fig. 8 Predictions of the stress–strain response of a SiC/CAS, unidirectional composite subjected to off-axis loading. Reproduced from Bahei-El-
Din, Y.A., Botrous, A.G., 2003. Analysis of progressive fiber debonding in elastic laminates. International Journal of Solids and Structures 40,
7035–7053.

Fig. 9 Predictions of the stress–damage strain response of a SiC/CAS, unidirectional composite subjected to off-axis loading. Reproduced from
Bahei-El-Din, Y.A., Botrous, A.G., 2003. Analysis of progressive fiber debonding in elastic laminates. International Journal of Solids and Structures
40, 7035–7053.

Figs. 10–13 show the computed overall response of the 3D-woven, carbon/epoxy composite under uniaxial normal strain, and
pure shear strain components. Symbols indicate the onset of local damage in one or more subvolumes of the RVE. The dashed
lines represent the response computed under frictionless failure conditions, whereas the solid lines represent the response con-
sidering friction. The three magnitudes shown in Figs. 11–13 for the coefficient of frictions m are assigned to longitudinal and
transverse sliding within the fiber bundle, and interface sliding, respectively. The results obtained for two subdivisions of the unit
cell are shown. Under a strain-controlled, uniaxial tensile load applied in the warp direction, Fig. 10, the response is dominated by
the warp fiber bundle. The response in this case is not affected either by the number of subvolumes considered in the RVE, or by
438 Multiscale Mechanics of Composite Materials and Structures

Fig. 10 Computed response of a 3D woven, T300-carbon/epoxy composite subjected to in-plane tension. Reproduced from Bahei-El-Din, Y.A.,
Rajendran, A.M., Zikry, M.A., 2004. A micromechanical model for damage progression in woven composite systems. International Journal of Solids
and Structures 41, 2307–2330.

Fig. 11 Computed response of a 3D woven, T300-carbon/epoxy composite subjected to out-of-plane compression. Reproduced from Bahei-El-
Din, Y.A., Rajendran, A.M., Zikry, M.A., 2004. A micromechanical model for damage progression in woven composite systems. International
Journal of Solids and Structures 41, 2307–2330.

internal friction. In contrast, the overall response of the 3D-woven composite under strain-controlled, compression in the direction
of the z-fiber (Fig. 11) is more sensitive to these parameters, and local damage has a significant effect on the overall behavior.
Under in-plane shear straining, Fig. 12, several damage mechanisms are triggered progressively in subvolumes of the unit cell.
Under out-of-plane shear straining, on the other hand, Fig. 13, failure is triggered mainly by interface sliding of the fiber bundles in
the warp, weft and z-directions.
These results indicate that the multiscale modeling offered captures very well the main features of the measured response for
3D-woven composites. In particular, the softening behavior seen in the estimated overall stress–strain response due to damage is
typical of the behavior measured in woven systems.33–35
Multiscale Mechanics of Composite Materials and Structures 439

Fig. 12 Computed response of a 3D woven, T300-carbon/epoxy composite subjected to in-plane shear. Reproduced from Bahei-El-Din, Y.A.,
Rajendran, A.M., Zikry, M.A., 2004. A micromechanical model for damage progression in woven composite systems. International Journal of Solids
and Structures 41, 2307–2330.

Fig. 13 Computed response of a 3D woven, T300-carbon/epoxy composite subjected to out-of-plane shear. Reproduced from Bahei-El-Din, Y.A.,
Rajendran, A.M., Zikry, M.A., 2004. A micromechanical model for damage progression in woven composite systems. International Journal of Solids
and Structures 41, 2307–2330.

Finally, a 3D-woven composite constructed from E-glass in the weft-direction and the z-direction, and PZT filament in the warp
direction is considered.15 Overall properties of a fiber bundle were found with the Mori–Tanaka model. Damage was considered in
the form of internal sliding within the fiber bundles, interface sliding, and matrix failure as detailed in Bahei-El-Din.15
The unit cell of Fig. 7(b) was loaded by overall, uniaxial strain components, e11, e22, e33,  e33, which were applied individually
while the unit cell is otherwise fully constrained. The electric displacement computed in the warp PZT fiber is shown in
Figs. 14–16. In the undamaged state, the electric displacement varies linearly with the applied load. Deviations in the electric
displacement are seen to occur as damage develops and progresses.
440 Multiscale Mechanics of Composite Materials and Structures

Fig. 14 Electric displacement computed in a 3D-woven, E-glass/PZT/epoxy composite subjected to overall in-plane uniaxial strain in the direction
parallel to the measured electric displacement. Reproduced from Bahei-El-Din, Y.A., 2009. Modeling electromechanical coupling in woven
composites exhibiting damage. Proceedings of the Institution of Mechanical Engineers, Part G: J Aerospace Engineering 223, 485–495.

Fig. 15 Electric displacement computed in a 3D-woven, E-glass/PZT/epoxy composite subjected to overall in-plane uniaxial strain in the direction
perpendicular to the measured electric displacement. Reproduced from Bahei-El-Din, Y.A., 2009. Modeling electromechanical coupling in woven
composites exhibiting damage. Proceedings of the Institution of Mechanical Engineers, Part G: J Aerospace Engineering 223, 485–495.

1.17.5.2 Composite Laminates


The overall response of a quasi-isotropic, (0/745/90) symmetric laminate caused by an overall stress applied in the direction of
the 0-degree ply is predicted while the fiber/matrix interface is exhibiting damage by debonding and/or sliding.19 The matrix is
glass ceramic (CAS), and the fiber is silicon-carbide fiber at volume content of 39%. Elastic properties and the interface strength for
peeling and sliding damage modes are found in Section 1.17.5.1. The total and damage strains corresponding to the applied
overall stress are shown in Figs. 17 and 18. The results predicted by both the unit cell model, Fig. 5, and by the averaging,
Mori–Tanaka model are shown. Initial failure is shown to occur in the 90-degree ply by peeling at the fiber/matrix interface.
Interface debonding follows in the off-axis, 745-degree plies, also by debonding rather than sliding. The predicted progression of
damage at the fiber/matrix interface is shown in Figs. 19 and 20.
Multiscale Mechanics of Composite Materials and Structures 441

Fig. 16 Electric displacement computed in a 3D-woven, E-glass/PZT/epoxy composite subjected to overall out-of-plane uniaxial strain.
Reproduced from Bahei-El-Din, Y.A., 2009. Modeling electromechanical coupling in woven composites exhibiting damage. Proceedings of the
Institution of Mechanical Engineers, Part G: J Aerospace Engineering 223, 485–495.

Fig. 17 Predictions of the stress–strain response of a SiC/CAS, (0/745/90)s composite laminate subjected to normal stress in direction of the 0-
degree ply. Reproduced from Bahei-El-Din, Y.A., Botrous, A.G., 2003. Analysis of progressive fiber debonding in elastic laminates. International
Journal of Solids and Structures 40, 7035–7053.

In another application involving laminates exhibiting damage, the mechanical behavior of fibrous laminates tested by Soden
et al.36,37 was predicted for two composite systems; AS4 carbon/epoxy and Silenka E-glass/epoxy.20 The Mori–Tanaka averaging
model was utilized for the unidirectional composite plies. The predicted and measured stress–strain responses for two laminates,
(90/745/0)s, AS4-carbon/3501-6-epoxy, and (745)s, E-glass/DY-063-epoxy are shown in Figs. 21 and 22. Failure envelopes of the
quasi-isotropic laminate in the biaxial stress plane are compared in Fig. 23.
Considering laminates with electro-active fibers, the effect of embedding PZT fibers in a composite on the permittivity
is modeled by Bahei-El-Din and Micheal.13 Utilizing PZT-5A fibers at a volume content of 55%, and a DY-063 epoxy matrix, a
(0/90)s laminate, and a (0/745/90)s laminate are considered, and the fiber in the 0-degree ply is subjected to an electric field E3 .
The resulting electric displacement D3 is found using the unit cell of the periodic hexagonal array model, Fig. 5, by averaging over
the fiber subvolumes. Calculated as κ33 ¼ D3/E3, the “apparent” permittivity of the PZT fiber changes from the inherent magnitude
442 Multiscale Mechanics of Composite Materials and Structures

Fig. 18 Predictions of the stress–damage strain response of a SiC/CAS, (0/745/90)s composite laminate subjected to normal stress in direction
of the 0-degree ply. Reproduced from Bahei-El-Din, Y.A., Botrous, A.G., 2003. Analysis of progressive fiber debonding in elastic laminates.
International Journal of Solids and Structures 40, 7035–7053.

Fig. 19 Predictions of the debonding configuration in the 90-degree ply of a SiC/CAS, (0/745/90)s composite laminate subjected to normal
stress in direction of the 0-degree ply. Reproduced from Bahei-El-Din, Y.A., Botrous, A.G., 2003. Analysis of progressive fiber debonding in elastic
laminates. International Journal of Solids and Structures 40, 7035–7053.

of 15.1 (109 C/V m) to 13.1 for the (0/90)s laminate, and to 10.9 for the (0/745/90)s laminate due to the stresses introduced by
mutual constraints as the electric field is applied to the electro-active fiber.
Capabilities of the multiscale model to capture thermo-electro-mechanical coupling are illustrated for the above (0/745/90)s
laminate13 under overall tensile stress of 25 MPa applied in the direction of the 0-degree ply and a temperature change,
which varies across the thickness with a differential temperature of 2001C. The average ply strain in the direction of the
mechanical load is shown in Fig. 24. Also shown is the electric displacement found in the longitudinal direction of the fibers and
the ply strain.
Multiscale Mechanics of Composite Materials and Structures 443

Fig. 20 Predictions of the debonding configuration in the 745 degree plies of a SiC/CAS, (0/745/90)s laminate subjected to normal stress in
direction of the 0-degree ply. Reproduced from Bahei-El-Din, Y.A., Botrous, A.G., 2003. Analysis of progressive fiber debonding in elastic
laminates. International Journal of Solids and Structures 40, 7035–7053.

Fig. 21 Comparison of measured and computed stress–strain response of a symmetric, quasi-isotropic, AS4-carbon/3501-6-epoxy laminate.
Reproduced from Bahei-El-Din, Y.A., Khire, R., Hajela, P., 2010. Multiscale transformation field analysis of progressive damage in fibrous
laminates. International Journal for Multiscale Computational Engineering 8, 69–80.

Finally, a sandwich plate comprised of a cellular foam core and PZT-5A/DY-063-epoxy, (0/745) face sheets is subjected to a
bending moment of 10 kN m, Fig. 25, and the electric displacement generated in the fibers is calculated with the multiscale
model.13 Unit cell of the periodic array model, Fig. 5, was utilized.
444 Multiscale Mechanics of Composite Materials and Structures

Fig. 22 Comparison of measured and computed stress–strain response of a symmetric 745 degree, E-glass/DY-063-epoxy laminate.
Reproduced from Bahei-El-Din, Y.A., Khire, R., Hajela, P., 2010. Multiscale transformation field analysis of progressive damage in fibrous
laminates. International Journal for Multiscale Computational Engineering 8, 69–80.

Fig. 23 Comparison of measured and computed failure envelope of a symmetric, quasi-isotropic, AS4-carbon/3501-6-epoxy laminate.
Reproduced from Bahei-El-Din, Y.A., Khire, R., Hajela, P., 2010. Multiscale transformation field analysis of progressive damage in fibrous
laminates. International Journal for Multiscale Computational Engineering 8, 69–80.

1.17.5.3 Composite Structures


Two applications for multiscale analysis of composite structures are offered. In both applications the formulation of Section 1.17.3
is applied within a commercial finite element code, where the micro- and macro-length scales (Sections 1.17.3.1 and 1.17.3.2) are
introduced through a user-defined routine.
In the first application, the composite under consideration is a carbon/epoxy, 3D-woven composite subjected to impact load.31
The fibers are electrically inactive and damage is taken into consideration. In this case, LS-Dyna explicit dynamic code38 was
utilized and a user-defined material subroutine was developed for the 3D-woven composite unit cell, Fig. 7(b). The problem
analyzed is that of a composite plate under out-of-plane high-velocity impact. The impactor is a steel plate with surface area that is
identical to that of the target plate, and as such the plate particles, which are far removed from the boundary, essentially move in
Multiscale Mechanics of Composite Materials and Structures 445

Fig. 24 Strain and electric displacement response of a PZT-5A/DY-063-epoxy, (0/745/90)s composite laminate under thermomechanical load.
Reproduced from Bahei-El-Din, Y.A., Micheal, A., 2012. Micromechanical modeling of multifunctional composites. In: Proceedings of the
International Mechanical Engineering Congress & Exposition. ASME 2012 IMECE, November 9–15, 2012, Houston, TX. Paper No. 89304. New
York: American Society of Mechanical Engineers.

Fig. 25 Strain and electric displacement response of a sandwich plate, with cellular foam core and PZT-5A/DY-063-epoxy, (0/745) face sheets,
under bending moment. Reproduced from Bahei-El-Din, Y.A., Micheal, A., 2012. Micromechanical modeling of multifunctional composites. In:
Proceedings of the International Mechanical Engineering Congress & Exposition. ASME 2012 IMECE, November 9–15, 2012, Houston, TX. Paper
No. 89304. New York: American Society of Mechanical Engineers.

the direction of the incident velocity vector. Accordingly, both the target plate and the impactor were modeled as 3D solids as
shown in Fig. 26, while only axial displacements in the x3-direction were permitted. The impact velocity applied is 200 m/s.
Figs. 27 and 28 show response of the woven composite plate under impact computed for a damaged and an undam
aged material in terms of spatial distribution of the peak axial stress, computed along the thickness of the target plate, and the
446 Multiscale Mechanics of Composite Materials and Structures

Fig. 26 Finite element model for impact of a 3D-woven composite plate under uniaxial strain condition. Reproduced from Bahei-El-Din, Y.A.,
Rajendran, A.M., Zikry, M.A., 2004. A micromechanical model for damage progression in woven composite systems. International Journal of Solids
and Structures 41, 2307–2330.

Fig. 27 Peak axial compressive stress predicted along the thickness of a 3D-woven, T300-carbon/epoxy plate subjected to impact under uniaxial
strain condition. Reproduced from Bahei-El-Din, Y.A., Rajendran, A.M., Zikry, M.A., 2004. A micromechanical model for damage progression in
woven composite systems. International Journal of Solids and Structures 41, 2307–2330.

Fig. 28 Strain energy predictions for a 3D-woven, T300-carbon/epoxy plate subjected to impact under uniaxial strain condition. Reproduced from
Bahei-El-Din, Y.A., Rajendran, A.M., Zikry, M.A., 2004. A micromechanical model for damage progression in woven composite systems.
International Journal of Solids and Structures 41, 2307–2330.
Multiscale Mechanics of Composite Materials and Structures 447

Fig. 29 Contours of overall stress, s11 (MPa), caused in a PZT-5A/DY-063-epoxy composite plate with a hole by a uniform load of 50 MPa
applied in direction of the x1-axis. Reproduced from Bahei-El-Din, Y.A., Micheal, A., 2013. Multiscale analysis of multifunctional composite
structures. In: Proceedings of the International Mechanical Engineering Congress & Exposition. ASME 2013 IMECE, November 13–21, 2013, San
Diego, CA. Paper No. 62427. New York: American Society of Mechanical Engineers.

Fig. 30 Electric displacement, D3 (102 C/m2) caused in fibers of the 0-degree plies of a PZT-5A/DY-063-epoxy composite plate with a hole by a
uniform load of 50 MPa applied in the direction of the x1-axis. Reproduced from Bahei-El-Din, Y.A., Micheal, A., 2013. Multiscale analysis of
multifunctional composite structures. In: Proceedings of the International Mechanical Engineering Congress & Exposition. ASME 2013 IMECE,
November 13–21, 2013, San Diego, CA. Paper No. 62427. New York: American Society of Mechanical Engineers.
448 Multiscale Mechanics of Composite Materials and Structures

Fig. 31 Apparent permittivity, κ3 (102 C/V m), computed in fibers of the 0-degree plies of a PZT-5A/DY-063-epoxy composite plate with a hole
subjected to a uniform load of 50 MPa in direction of the x1-axis and an electric field, E3 ¼1.0x106 V/m, superimposed on s11 stress
contours (MPa). Reproduced from Bahei-El-Din, Y.A., Micheal, A., 2013. Multiscale analysis of multifunctional composite structures. In:
Proceedings of the International Mechanical Engineering Congress & Exposition. ASME 2013 IMECE, November 13–21, 2013, San Diego, CA.
Paper No. 62427. New York: American Society of Mechanical Engineers.

strain energy. Energy dissipation due to damage is clearly captured by the proposed multiscale analysis. This behavior compares
remarkably well with that measured in impact experiments of woven systems.39–41
In the second application, a (0/745/90)s laminated plate with a central hole is considered under an electric field applied to the
PZT-5A fibers, and under overall mechanical load.42 The matrix is a DY-063 epoxy, and the fiber volume fraction is 0.55.
The multiscale analysis was applied for the periodic hexagonal array unit cell, Fig. 5, using the ABAQUS finite element code
through its user-defined routine, UEXPAN to model eigen strains and stresses caused by an electric filed applied to the electro-
active fibers, or read the electric displacement in response to mechanical load. With the plate unconstrained and subjected to an
overall tensile stress of 50 MPa in the direction of the 0-degree fiber, the stress distribution shown in Fig. 29 is found. These
represent overall stresses caused on the laminated elements. In a post-processor exercise, element stresses found from the finite
element solution are used to determine the local fields in the fiber and matrix within all plies, as well as the electric displacement
generated in the PZT fiber. The latter is shown in Fig. 30 for the 0-degree ply. It is seen that magnitudes of the electric displacement
reflect the stress concentration caused by discontinuity of the plate geometry and the stress gradient in general.
While free of mechanical load, the same laminated plate was constrained as shown in Fig. 31 and subjected to an electric field
E3 ¼1.0  106 V/m in the PZT fiber of the 0-degree ply. Following a finite element analysis with the ABAQUS program and the
multiscale modeling interface through UEXPAN routine, post-processing of the results using the coupling equations of Sections
1.17.3.1.3 and 1.17.3.2.1 provides the ply and phase stresses, as well as the electric displacement in all electro-active fibers. For the
0-degree ply this includes the direct electric effect, κ33E3, and the effect of the local stress generated as a result of the mutual constraints
of the phases and plies. When this result is divided by the applied electric field, the apparent permittivity is obtained, Fig. 31.42

1.17.6 Closing Remarks

The method discussed in this chapter for multiscale modeling of composite materials and structures offers a seamless approach to
include all length scales found in these systems into a single analysis tool. The method centers on modeling the effects of physical
phenomena, which may or may not be directly related to mechanical load through the introduction of local transformation
Multiscale Mechanics of Composite Materials and Structures 449

stresses and strains. This includes inelastic deformations, as well as several coupled effects, as outlined by Nye,12 such as thermo-
elastic, electro-mechanical, and electro-thermal coupling. Damage, which is traditionally modeled by reduction of the material
stiffness, is efficiently modeled by the introduction of transformation fields.
The use of any micromechanical model for unidirectional, fibrous composites in this multiscale analysis approach provides
flexibility to obtain a quick, but rather crude estimate of the response of composite structures or a more detailed and accurate
result. Examples are given where both averaging models and unit cells extracted from periodic array models were used. In the
latter, the computational cost could be quite high and even prohibitive for large composite structures modeled with the finite
element method coupled with the multiscale model. In this case, limiting the multiscale analysis within the finite element solution
to the areas of interest in the structural mesh, while using “smeared” properties for the rest of the mesh would be helpful. This
and/or structural partitioning in the finite element sense may be combined to reduce the computational cost and yet obtain
reliable predictions of the structural response.

References

1. Budiansky, B., 1965. On the elastic moduli of some heterogeneous materials. Journal of the Mechanics and Physics of Solids 13, 223–227.
2. Hill, R., 1965. Theory of mechanical properties of fibre-strengthened materials – III. Self-consistent model. Journal of the Mechanics and Physics of Solids 13, 189–198.
3. Hill, R., 1965. A self-consistent mechanics of composite materials. Journal of the Mechanics and Physics of Solids 13, 213–222.
4. Dvorak, G.J., Teply, J.L., 1985. Periodic hexagonal array models for plasticity analysis of composite materials. In: Sawczuk, A., Bianchi, V. (Eds.), Plasticity Today:
Modelling, Methods and Applications. In: Olsazak, W. (Ed.), Memorial Volume. Amsterdam: Elsevier Science Publishers, pp. 623–642.
5. Michel, J.C., Moulinec, H., Suquet, P., 1999. Effective properties of composite materials with periodic microstructure: A computational approach. Computer Methods in
Applied Mechanics and Engineering 172, 109–143.
6. Iwakuma, T., Nemat-Nasser, S., 1983. Composites with periodic microstructure. Composite Structures 16, 13–19.
7. Nemat-Nasser, S., Hori, M., 1993. Micromechanics: Overall Properties of Heterogeneous Solids. Amsterdam: Elsevier Science Publishers.
8. Dvorak, G.J., 1992. Transformation field analysis of inelastic composite materials. Proceedings of the Roya Society London A 437, 311–327.
9. Hill, R., 1964. Theory of mechanical properties of fibrestrengthened materials: I. Elastic behaviour. Journal of the Mechanics and Physics of Solids 12, 199–212.
10. Bahei-El-Din, Y.A., Dvorak, G.J., 2000. Micromechanics of inelastic composite materials. In: Kelly, A., Zweben, C. (Eds.-in-Chief), Comprehensive Composite Materials. In:
Chou, T.-W. (Ed.), General Theory of Composites, vol. 1. Amsterdam: Elsevier Science B.V (Chapter 15).
11. Bahei-El-Din, Y.A., Shah, R.S., Dvorak, G.J., 1991. Numerical analysis of the rate-dependent behavior of high temperature fibrous composites. In: Singhal, S.N., Jones, W.
F., Herakovich, C.T. (Eds.), Mechanics of Composites at Elevated and Cryogenic Temperatures, AMD-vol. 118. American Society for Mechanical Engineers, pp. 67–78.
12. Nye, J.F., 1985. Physical Properties of Crystals. London: Oxford University Press.
13. Bahei-El-Din, Y.A., Micheal, A., 2012. Micromechanical modeling of multifunctional composites. In: Proceedings of the International Mechanical Engineering Congress &
Exposition. ASME 2012 IMECE, November 9–15, 2012, Houston, TX. Paper No. 89304. New York: American Society of Mechanical Engineers.
14. Chen, T., 1994. Micromechanical estimates of the overall thermoelectroelastic moduli of multiphase fibrous composites. International Journal of Solids and Structures 31,
3099–3111.
15. Bahei-El-Din, Y.A., 2009. Modeling electromechanical coupling in woven composites exhibiting damage. Proceedings of the Institution of Mechanical Engineers, Part G: J
Aerospace Engineering 223, 485–495.
16. Kumar, A., Chakraborty, D., 2009. Effective properties of thermo-electro-mechanically coupled piezoelectric fiber reinforced composites. Material and Design 30,
1216–1222.
17. Dvorak, G.J., Benveniste, Y., 1992. On transformation strains and uniform fields in multiphase elastic media. Proceedings of the Roya Society London A 437, 291–310.
18. Bahei-El-Din, Y.A., 1992. Uniform fields, yielding, and thermal hardening in fibrous composite laminates. International Journal of Plasticity 8, 867–892.
19. Bahei-El-Din, Y.A., Botrous, A.G., 2003. Analysis of progressive fiber debonding in elastic laminates. International Journal of Solids and Structures 40, 7035–7053.
20. Bahei-El-Din, Y.A., Khire, R., Hajela, P., 2010. Multiscale transformation field analysis of progressive damage in fibrous laminates. International Journal for Multiscale
Computational Engineering 8, 69–80.
21. Eshelby, J.D., 1957. The determination of the elastic field of an ellipsoidal inclusion, and related problems. Proceedings of the Royal Society London A 241, 376–396.
22. Mori, T., Tanaka, K., 1973. Average stress in matrix and average elastic energy of materials with misfitting inclusions. Acta Metallurgica 21, 571–574.
23. Benveniste, Y., 1987. A new approach to the application of Mori–Tanaka theory in composite materials. Mechanics of Materials 6, 147–157.
24. Chen, T., Dvorak, G.J., Benveniste, Y., 1992. Mori–Tanaka estimates of the overall elastic moduli of certain composite materials. Journal of Applied Mechanics 59,
539–546.
25. Hill, R., 1963. Elastic properties of reinforced solids: Some theoretical principles. Journal of the Mechanics and Physics of Solids 11, 357–372.
26. Buryachenko, V., 1996. The overall elastoplastic behavior of multiphase materials. Acta Mechanica 119, 93–117.
27. Moulinec, H., Suquet, P., 1994. A fast numerical method for computing the linear and nonlinear mechanical properties of composites. Comptes Rendus de I’Academie des
Sciences, Serie II 318, 1417–1423.
28. Teply, J.L., Dvorak, G.J., 1988. Bounds on overall instantaneous properties of elastic–plastic composites. Journal of the Mechanics and Physics of Solids 36, 29–58.
29. Walker, K.P., Freed, A.D., Jordan, E.H., 1994. Thermoviscoplastic analysis of fibrous periodic composites by use of triangular subvolumes. Composites Science and
Technology 50, 71–84.
30. Bahei-El-Din, Y.A., Ibrahim, I.A., Botrous, A.G., 1998. Micromechanical analysis of inelastic laminates. In: Bahei-El-Din, Y.A., Dvorak, G.J. (Eds.) IUTAM Symposium on
Transformation Problems in Composite and Active Materials, pp. 45–60. Dordrecht: Kluwer Academic Publishers.
31. Bahei-El-Din, Y.A., Rajendran, A.M., Zikry, M.A., 2004. A micromechanical model for damage progression in woven composite systems. International Journal of Solids and
Structures 41, 2307–2330.
32. Wooh, S.C., Daniel, I.M., 1994. Real-time ultrasonic monitoring of fiber-matrix debonding in ceramic-matrix composite. Mechanics of Materials 17, 379–388.
33. Khan, M.Z., Simpson, G., 2000. Mechanical properties of a glass reinforced plastic naval composite material under increasing compressive strain rates. Material Letters 45,
167–174.
34. Kim, J.K., Sham, M.L., 2000. Impact and delamination failure of woven-fabric composites. Composites Science and Technology 60, 745–761.
35. McGee, J.D., Nemat-Nasser, S., 2001. Dynamic bi-axial testing of woven composites. Materials Science and Engineering A 317, 135–139.
36. Soden, P.D., Hinton, M.J., Kaddour, A.S., 1998. Lamina properties, lay-up configurations and loading conditions for a range of fibre-reinforced composite laminates.
Composites Science and Technology 58, 1011–1022.
37. Soden, P.D., Hinton, M.J., Kaddour, A.S., 2002. Biaxial test results for strength and deformation of a range of E-glass and carbon fibre reinforced composite laminates:
Failure exercise benchmark data. Composites Science and Technology 62, 1489–1514.
450 Multiscale Mechanics of Composite Materials and Structures

38. LSTC, 2003. LS-Dyna 9.6. CA: Livermore Software Technology Corporation, Livermore.
39. Boteler, J., Rajendran, A.M., Grove, D., 1999. Shock wave profile in polymer matrix composite. In: Furnish, M.D., Chhabildas, L.C., Hixson, R.S. (Eds.) Proceedings of the
APS Conference on Shock Compression of Condensed Matter, pp. 563–566. College Park, MD: American Physical Society.
40. Dandekar, D.P., Beaulieu, P.A., 1995. Compressive and tensile strengths of glass reinforced polyester under shock wave propagation. In: Rajapakse, Y.D.S., Vinson, J.R.
(Eds.), High Strain Rate Effects on Polymers, Metal and Ceramic Matrix Composites and Other Advanced Materials. AD. New York, NY: ASME Press, pp. 63–70.
41. Espinosa, H.D., Dwivedi, S., Lu, H.C., 2000. Modeling impact induced delamination of woven fiber reinforced composites with contact/cohesive laws. Computer Methods in
Applied Mechanics and Engineering 183, 259–290.
42. Bahei-El-Din, Y.A., Micheal, A., 2013. Multiscale analysis of multifunctional composite structures. In: Proceedings of the International Mechanical Engineering Congress &
Exposition. ASME 2013 IMECE, November 13–21, 2013, San Diego, CA. Paper No. 62427. New York: American Society of Mechanical Engineers.

Relevant Website

www.bue.edu.eg
The British University in Egypt.

You might also like