You are on page 1of 24

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/283029972

Numerical Modeling of Combined Matrix


Cracking and Delamination in Composite
Laminates Using Cohesive Elements

Article in Applied Composite Materials · October 2015


DOI: 10.1007/s10443-015-9465-0

CITATIONS READS

7 149

4 authors, including:

Deepak Kumar Rene Roy


Maharishi Markandeshwar University, Mullana Gyeongsang National University
8 PUBLICATIONS 8 CITATIONS 11 PUBLICATIONS 104 CITATIONS

SEE PROFILE SEE PROFILE

Jin-Ho Choi
Sungkyunkwan University
353 PUBLICATIONS 5,989 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Composite Repair View project

Percutaneous Coronary Intervention View project

All content following this page was uploaded by Deepak Kumar on 30 December 2015.

The user has requested enhancement of the downloaded file.


Appl Compos Mater
DOI 10.1007/s10443-015-9465-0

Numerical Modeling of Combined Matrix Cracking


and Delamination in Composite Laminates Using
Cohesive Elements

Deepak Kumar 1 & Rene Roy 1 & Jin-Hwe Kweon 1 &


Jin-ho Choi 2

# Springer Science+Business Media Dordrecht 2015

Abstract Sub-laminate damage in the form of matrix cracking and delamination was
simulated by using interface cohesive elements in the finite element (FE) software
ABAQUS. Interface cohesive elements were inserted parallel to the fiber orientation in
the transverse ply with equal spacing (matrix cracking) and between the interfaces
(delamination). Matrix cracking initiation in the cohesive elements was based on stress
traction separation laws and propagated under mixed-mode loading. We expanded the
work of Shi et al. (Appl. Compos. Mater. 21, 57–70 2014) to include delamination and
simulated additional [45/−45/0/90]s and [02/90n]s {n=1,2,3} CFRP laminates and a
[0/903]s GFRP laminate. Delamination damage was quantified numerically in terms of
damage dissipative energy. We observed that transverse matrix cracks can propagate to
the ply interface and initiate delamination. We also observed for [0/90n/0] laminates
that as the number of 90° ply increases past n=2, the crack density decreases. The
predicted crack density evolution compared well with experimental results and the
equivalent constraint model (ECM) theory. Empirical relationships were established
between crack density and applied stress by linear curve fitting. The reduction of
laminate elastic modulus due to cracking was also computed numerically and it is in
accordance with reported experimental measurements.

Keywords Fracture . Crack density . Finite element analysis (FEA) . Composite materials

* Jin-Hwe Kweon
jhkweon@gnu.ac.kr

1
Department of Aerospace and System Engineering, Research Center for Aircraft Parts Technology,
Gyeongsang National University, Jinju, Gyeongnam 660-701, South Korea
2
School of Mechanical Engineering, Research Center for Aircraft Parts Technology, Gyeongsang
National University, Jinju, Gyeongnam 660-701, South Korea
Appl Compos Mater

1 Introduction

Composite materials have been extensively used in aerospace and automobile industries
because of their high specific strength and stiffness properties. Currently, polymer composites
reinforced with carbon or glass fibers have taken the place of the conventional metal alloys in
the aerospace industries [1]. Composite laminate fracture, under static or fatigue loading, will
involve a sequential accumulation of damage in the form of matrix-dominated cracking [2].
Crossman and Wang [3] observed that as the thickness of the transverse ply increases, the
matrix crack density decreases. Therefore, the thickness of the ply is also an important factor
for transverse matrix cracking. Wang et al. [4] used a 3-D finite element model to evaluate the
energy released as the delamination grows. In their study, they described two types of
delamination: i) the delamination starting from the intersection of the transverse cracks and
the free edges of the laminates and ii) the delamination starting from the intersection of
transverse cracks and splits in the 00 layers. They found good agreement with the experimental
results. However, a common problem associated with laminated materials is the presence of
sub-critical ply cracks. These cracks depend on the applied stress, boundary conditions and
material properties. Wang [5] provided an overview of a fracture mechanics approach to
understand sub-laminate cracks in composite laminates. He also studied the effect of the
thickness of the 900 layers to understand the influence of the thickness on the transverse
cracking. Composite failure is a complex process involving intra- and inter-laminar damage
that results in stiffness loss and load-carrying capacity loss as the damage becomes more
extensive [6]. The matrix tensile and compressive damage occur under intra-laminar damage in
single laminas; this damage also includes de-bonding between fiber and resin interface. As the
applied pressure increases, tensile or compressive fiber breakage occurs, resulting in the final
failure of laminated composites [7, 8]. In this type of failure, the key driving damage is due to
transverse loading, which is transverse matrix cracking and fiber splitting. Extensive literature
is available on these damage modes occurring due to the stress concentrations at the crack tip.
These damage modes are developed due to inter-laminar damage (delamination) between the
plies, resulting in fiber breakage and a total loss of load carrying capability [9–13]. Thus, these
damage modes are very crucial and must be investigated for better understanding of the
damage initiation and damage evolution of composite laminates. Such a study will help in
designing composite laminates with better resistance and tolerance.
Literature exists on composite material failure and damage based on stress criteria or
damage/fracture mechanics. For example, in [14, 15], stress/strain-based failure criteria such
as Tsai-Wu or Tsai-Hill are used to illustrate the failure envelope of any given multi-directional
laminate subjected to multi-axial loading. Using these failure criteria, the damage mechanisms
of different modes were not clearly identified. Hashin developed an effective method to model
matrix cracking as a plane problem [14, 15], and various researchers [16–19] developed the
method further. Zhang et al. [20] suggested the Equivalent Constraint Model (ECM) to predict
the crack density as a function of applied load and stiffness reduction based on a 2-D shear-lag
analysis [20, 21]. Smith et al. [22] compared the transverse cracking phenomenon in [0/90]s
and [90/0]s CFRP laminates. They observed experimentally that cracking initiation starts at a
lower strain in [90m/0m]s laminates, but full transverse cracking develops first in [0m/90m]s
laminates. Berthelot [23] derived an analytical model to evaluate the stress distribution in
cross-ply laminates subjected to tensile loading and containing transverse cracks and delam-
ination, starting from the transverse crack tips. They also performed a one-dimensional
analysis of the delamination model. Kachanov [24] and Rabotonov [25] described the use of
Appl Compos Mater

continuum damage mechanics, which is a popular approach to model failures in


composite laminates [26–29]. Barbero et al. [30] described a mechanistic model for
transverse damage initiation as well as damage growth and proposed an analytical
model that provides good agreement with experimental results. Alternatively, modeling
of failure in long fiber reinforced composites by X-FEM (extended finite element
method) and the cohesive zone model were implemented by Bouhala et al. [31].
They modeled the cracks within a finite element without the requirement of either re-
meshing or aligning the mesh with the interface by using Heaviside functions for the
jump in the displacement (at the interface). Shi et al. [32] studied the low-velocity
impact damage on composites, including the intra-and inter-laminar damage criteria
along with nonlinear shear behavior, accurately predicting the matrix cracking and
delamination at different impact energy levels. Many methods were successful in
predicting the extent of the damaged area; however, few methods in the literature can
predict the process of matrix cracking within a damaged region. Shi et al. [33]
developed a numerical model using a cohesive element for the matrix cracking evolu-
tion. In their study, they found good agreement between the theoretical and experimen-
tal data for crack densities compared to their numerical model. Low-velocity impact
damage on composite laminates and its simulation is a topic still under study recently
[34, 35]. Carraro and Quaresimin [36] presented an analytical model which is capable
of estimating the elastic properties of a multidirectional symmetric laminate with an
arbitrary lay-up and cracks in the plies considering the crack mutual interactions. The
effect of cohesive length on the damage and toughening mechanisms were studied by
Sills and Thouless [37]. They extensively studied the laminate strength in the terms of
effective modulus and traction which can be used at any load before failure, and at any
point along the interfaces. Van der Meer et al. [38] tried to bridge the gap between
classical understanding of transverse cracking in cross ply laminates and recent com-
putational methods for the modeling of progressive failure of laminate. They modeled
with single element across the thickness of the ply and reproduced the in situ effect for
the ply strength provided the interface stiffness properly selected. Cohesive zone (CZ)
fracture analysis techniques are also used to predict the initiation of crack growth from
the interface corner of an adhesively bonded butt joint [39].
In this paper, transverse matrix cracking and delamination in composite laminates are
simulated numerically by inserting interface cohesive elements in the finite element model.
The transverse plies are modeled by inserting cohesive elements in each lamina between
adjacent finite elements along the fiber direction. Furthermore, cohesive elements are inserted
between ply interfaces (00/900) to simulate delamination. The ultimate crack density at
saturation level implemented in the finite element model (highest crack density possible) is
chosen according to experimental observations [5] and a mesh convergence analysis. A mesh
size and cohesive element spacing are thus selected on the basis of the convergence analysis,
and a uniform crack spacing in each ply is assumed. The finite element model is built for
composite laminates with various stacking sequences of [±θm/90n]s and [0/90n/0]. The numer-
ically predicted crack density evolution is compared with theoretical (ECM) [21] and exper-
imental results [5]. The Equivalent Constraint Model is chosen for theoretical study [21]
because of its equally spaced crack assumption. A composite laminate with a stacking
sequence [+45/−45/0/90]s is selected for the delamination study. Matrix cracking as well as
delamination are included in the model. A simulation is also performed for a glass fiber
reinforced plastic (GFRP) composite laminate. We observed that the transverse matrix crack
Appl Compos Mater

density depends on the level of applied stress and the laminate configuration. The longitudinal
elastic modulus reduction ratios due to transverse matrix cracking are also obtained against the
applied laminate stress for the various laminate configurations.

2 Equivalent Constraint Model (ECM)

To predict the transverse matrix cracking in multi-directional laminates under multi-axial in-
plane loading, a theoretical approach (ECM) is adopted [20, 21]. The assumptions and
simplifications are discussed next. The first assumption is that the cracks in a damaged lamina
are uniformly spaced, which is essential for solving a problem by analysis of a representative
volume element (RVE). A representative ECM with damaged lamina is shown in Fig. 1. Here,
the layer k denotes the damaged lamina, and all of the plies above and below the kth ply are
replaced with homogeneous layers (layers I and II). These layers are governed by the
equivalent constraint effect. The stiffness properties of ECM can be computed using the
laminate plate theory (LPT), which provides the stress and strain relationship.
Representative segments of RVE were reduced to 1/4th the size due to the symmetry of each
[±θm/90n]s laminate, as shown in Fig. 2. Matrix cracking in the transverse ply (900) is expected
to start first. Stresses are computed from the stiffness of the constraint homogeneous layers and
the modified stiffness of the cracked ply. To evaluate the stresses in the damaged ply, the total
strain in the individual lamina and the laminate (implying continuity) were assumed to be
equivalent as follows:
ðk Þ
εi ¼ εi k ¼ 1; 2 ð1Þ

Here, εi ðk Þ and εi represent the total strain vectors of the kth layer and the laminate,
respectively. Thus, the average constitutive equations of a damaged ply are given by
 
ðk Þ 0 ðk Þ
σi ¼ Qkij ε j þ ε j k ¼ 1; 2i; j ¼ 1; 2; 6 ð2Þ

where σi ðk Þ represents the total stress vector of the constraint layers (k=1) and the damaged
900 ply (k=2), ε0j ðk Þ is the residual thermal strain vector of the kth layer, and Qkij is the stiffness
of the constraining layers (k=1) and the modified reduced stiffness of the damaged 900 ply
(k=2). The reduced stiffness matrix of the damaged ply can be derived using the in-situ
damage effective function (∀ ij). The in-situ function depends on the crack density, which is

Fig. 1 A schematic of the equivalent constraint model (ECM) of damaged laminate a Laminate structure b ECM
model [20, 21]
Appl Compos Mater

Fig. 2 Schematic diagram of composite laminates with transverse matrix cracking

given by 2D shear lag analysis [20]. The laminated stress can be written using the classical
laminate plate theory [1] as
 
1 ð2Þ ð1Þ
σi ¼ σi þ φσi ð3Þ
ð1 þ φÞ
where φ is the thickness ratio of the constraining layer over the thickness of the 900 layer. The
constitutive relation of the cracked laminate is obtained by combining Eqs. (2) and (3) as
follows:
 p
σi ¼ Qi j ε j −ε j ð4Þ

where Qi j is the in-plane stiffness matrix of the damaged laminate. The term εpj is a permanent
strain vector, and defined as

p 1 X 2
0ðk Þ
0ðk Þ
ε j ¼ −S i j hk Qkjl S lm σm ð5Þ
ðh1 þ h2 Þ k¼1

where S i j is the in-plane compliance matrix, hk is the thickness of the kth layer, Qkjl represents
the stiffness matrix of the kth layer, S0(k) th
lm is the residual compliance matrix of the k layer, and
Appl Compos Mater

σ0m ðk Þ represents the residual stress of the kth layer. It is observed that the laminate acquires a
permanent strain εpj due to the interaction effect of damage and residual stresses [21]. For the
transverse crack, we consider the composite laminates with stacking sequence [±θm/90n]s with
a finite gauge length (2l) and width (w). Therefore, the potential energy (PE) is written as

PE ¼ U −2ðh1 þ h2 Þw2lσi εi ð6Þ

Here, U represents the total strain energy of the composite laminate element. Using Eq. (2),
the total strain energy is computed as
  
1X
2
0ðk Þ 0ðk Þ
U¼ 2lw2hk Qkij εi þ εi εj þ εj ð7Þ
2 k¼1

where Qkij is the in-plane stiffness matrix of the kth layer, εi and ε j are the total strain vectors of
the laminate, and ε0i ðk Þ and ε0j ðk Þ are the residual thermal strain vectors of the kth layer. The
energy release rate (ERR) is obtained as the first partial derivative of the potential energy with
respect to the crack surface area (A) at a constant applied laminate stress:
 
∂PE 
G¼−  ð8Þ
∂A σi

Rearranging Eqs. (4)−(7) and substituting them into Eq. (8) gives the energy release rate
associated with matrix cracking, which can be derived [21] and expressed as
ð2Þ      
∂Qi j 0ð2Þ p 0ð2Þ p 0ð2Þ p
Gðσ; Da Þ ¼ −h2 S il S jm σl σm þ 2S il σl ε j þ ε j þ εi þ εi εj þ εj ð9Þ
∂Da

where Da denotes the total crack surface area per unit length and width of the laminate, h2 is
the thickness of the 900 plies, and S il represents the compliance matrix. The second and third
terms of the right hand side of Eq. (9) are represented by residual stresses and their interaction
with damage. The total crack surface area (Da) is expressed as
Da ¼ 2h2 C d ð10Þ

In the above equation,Cd denotes the average crack density. The energy release rate can be
derived if the in-plane stiffness matrix of the cracked laminate is known for a given crack
density. In general, under quasi-static uni-axial tension loading, the resistance of a composite
laminate to transverse matrix cracking increases with crack density. However, in reality, this is
due to the decreasing of the driving force for crack formation and propagation, because of the
shielding effect due to presence of cracks themselves. The resistance curve (R-curve concept
of fracture mechanics) can be used to predict the resistance to propagation of transverse
cracking, as in the following [40–42]:
Gðσ; Da Þ ¼ GR ðDa Þ ð11Þ

where GR represents the laminate resistance to the multiple transverse ply cracking. The R-
curve is dependent on the 900 ply thickness but independent of the stiffness of the constraining
layers [41]. A simple mathematical expression for GR can be derived by curve fitting as [42]

GR ¼ GIC þ G0 1−e−RD ð12Þ


Appl Compos Mater

where D and GIC are the crack density function and the critical energy release rate associated
with matrix cracking mode I, respectively, and G0 and R are considered as material and
laminate constants, respectively, that capture the resistance to the crack growth with increasing
applied load. The resistance curve (GR vs. D) [21] can be estimated by following previous
literature [42], and the fracture parameters are computed from curve fitting of the plot GR vs. D
[21] and Eq. (12). Typical crack patterns observed in laminates [0/90]s are shown in Fig. 3.

3 Computational Damage Model

The finite element code ABAQUS 6.10 is utilized to predict the transverse matrix crack
density as a function of the tensile applied stress using interface cohesive elements. A traction
separation law is used for the growth of the matrix cracking as well as delamination under
mixed-mode loading. The FE model is created in ABAQUS, and appropriate boundary
conditions are applied, as discussed in Section 3.2.

3.1 Sub-Laminate Crack Modeling Method Using Interface Cohesive Elements

Various researchers [32] have studied damage due to impacts in which they modeled inter-
laminar cracking using cohesive elements. For example, Camanho and Davila [43] and
Camanho et al. [44] defined a stress failure criterion to predict damage initiation, while the
delamination propagation was based on the fracture mechanics concepts, where an interface
element was inserted between each ply of the composite laminate. A stress failure criterion is
used to determine the delamination onset, as expressed by
 σ 2  σ 2  σ  2
n s t
þ þ ¼1 ð13Þ
N S T
where σi (i=n, s, t) represents the traction stress vector in the normal n direction, shear
direction s, and shear direction t, respectively. The N, S, and T parameters are defined as the
corresponding inter-laminar normal and two shear strengths. By using the ABAQUS manual
[45] as a reference, the traction stress σi can be expressed in terms of the stiffness Ki in modes
I, II, and III and the opening and/or sliding displacements δi [46] as follows:
σi ¼ K i δi i ¼ n; s; t ð14Þ

Once the damage initiation criterion has been reached, the material stiffness is gradually
degraded in terms of the damage variable Dm. The damage variable values range from 0, i.e.,
when damage initiates, to 1, i.e., when complete delamination/matrix crack has occurred at the
interface elements. The energy release rates associated with modes I (n), II (s), and III (t) for
the different failure criteria under mixed-mode are implemented. All three traction components

Fig. 3 Typical crack pattern observed in a [0/90]s laminate


Appl Compos Mater

σI, σII and σIII and all three separation components δI,δII and δIII are active when the interface/
matrix crack of the laminated material is under mixed-mode failure. The penalty stiffness is the
same for all modes considered in the model.
The mixed-mode ratios (β) are defined between pairs of mode components,
δII δIII
β ¼ β δII ¼ and β δIII ¼ ð15Þ
δI δI
Regardless of the definition used, mixed-mode ratios are parameters characterizing the mixed-
mode state. Furthermore, these parameters allow for a simplification of the analysis by assuming
that de-cohesion progresses at constant mixed-mode ratios. It is further assumed that the modes
are uncoupled, even though they occur simultaneously; that is, the stress-separation relationship
for each uncoupled mode is expressed separately for each mode (Eq. (14)). Next, a mixed-mode
separation at the interface is expressed as the L2 norm of the mode separations, i.e.,
vffiffiffiffiffiffiffiffiffiffiffiffi
u M
uX
δm ¼ t δ2i ð16Þ
i¼1

where M is the number of modes involved, and δm is the mixed-mode separation.


The mixed-mode (mode I and II) damage initiation is predicted by rewriting Eq. (13) for
two modes in terms of separation by using Eq. (14) as
!2 !2
δI δII
þ ¼1 ð17Þ
δ0I δ0II

where δI and δII are the separation components for modes I and II, respectively. δ0I andδ0II are
the separation components for modes I and II at onset, respectively. Using Eq. (15) in Eq. (16),
δI and δII are obtained as
δm
δI ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð18Þ
1 þ β2

δm
δII ¼ β pffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð19Þ
1 þ β2
Substituting the above two values of δI and δII into Eq. (17), which represents damage
initiation, leads to
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u
u
2
2 1 þ β2
δm ¼ t δ0II δ0I
2
0

2 ð20Þ
δ0II þ β 2 δ0I

A possible choice to predict fracture is to use an energy release rate (ERR) power criterion
under the mixed-mode condition, which is given by
3  
X Gi αi
¼1 ð21Þ
i¼1
Gic
Appl Compos Mater

where Gic is the single mode critical ERR, αi is the exponent for modes I and II, and Gi is the
single mode component ERR at the moment of mixed mode fracture and is calculated as

1 δc −δi
Gi ¼ Gic − K i δ0i δci ci 0 ð22Þ
2 δi −δi

where Gic is the single mode critical ERR, δ0i is the separation at damage onset under single
mode loading, and δci is the separation at fracture under single mode loading. A Benzeggagh–
Kenane (B-K) fracture energy based criterion [43] can be used to define the mixed-mode
displacement for complete failure, δcm, as
8
> 2
< 0 ½GIC þ ðGIIC −GIC Þζ η  δn > 0
δm ¼
c kδ m qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð23Þ
>
: c
2 c
2
δs þ δt δn ≤0

where ηis the B-K power law parameter that can be determined using a least-squares fit from a
β2
set of mixed-mode bending experimental data; ζ ¼ 1þβ 2
, with ζ taking values between zero
and one, and parameter β is the mixed-mode ratio. When ζ= 0, the crack is mode I driven,
while as ζ→1, the fracture is mode II dominated (and this is also the case when the exponent
η= 0). The damage variable for mixed-mode conditions can be expressed as

δcm δm −δ0m
Dm ¼ c 0
ð24Þ
δm δm −δm

where the onset and fracture separations δ0m and δcm are calculated from Eqs. (20) and (23),
respectively. Fracture modes I, II, and III are shown in Fig. 4 for the linear traction separation
law. Initially, the linear elastic response is represented by the stiffness terms Ki (i=n, s, t) until
the normal and shear strengths are reached. Figure 4c shows the stress transfer mechanism for
the cohesive zone. Beyond these strength values, the stiffness will start to be reduced linearly
according to the damage evolution variable Dm defined in Eq. (24) and finally complete the
damage when the maximum separation is achieved. This damage modeling approach is
implemented in the FE model, as described in Section 3.2.

3.2 Finite Element Model with Matrix Cracking and Delamination

The 3-D model geometry is created in ABAQUS (Figs. 5, 6 and 7) using eight-node linear
brick elements (C3D8R). The thickness of each ply is 0.132 mm for the following CFRP
laminates: [0/90n/0], n=1,2,3,4, [±25/90n]s, n=2,4, and [45/−45/0/90]s. The glass epoxy
composite laminate studied had a stacking sequence of [0/903]s with a ply thickness
0.144 mm. Cohesive elements (COH3D8) of zero-thickness (Fig. 5) are inserted between
neighboring elements parallel to the fiber orientation within the 900 plies as well as between
the ply interface (00/900). The mesh size of the model is a crucial issue that dictates the
numerical efficiency and the accuracy with which the transverse matrix crack formation and
crack density can be modeled. To determine the material mesh size, convergence tests can be
completed to establish the accuracy of prediction compared with the measurements, as
performed by Shi et al. [33] (Fig. 8). They observed that mesh refinement can slightly improve
the accuracy and suggested that an initial mesh size resulting in a crack spacing of 0.5 mm was
Appl Compos Mater

Fig. 4 Matrix cracking represented by cohesive element a Crack modes and co-ordinate used b A schematic of
the assumed crack traction-opening or sliding displacement c Stress transfer for cohesive zone

sufficient for predicting the crack density. A model that is too refined, which has a large
number of solid and cohesive elements, will require more computational time to solve.
Conversely, a model that is too coarse would under estimate the stresses and introduce
errors in the numerical prediction of the matrix crack density. Here, the selected mesh
density was based on convergence analysis on matrix crack density (Fig. 8) and a crack
density of 2 cracks/mm was assumed. For the delamination study, laminates with
configuration [45/−45/0/90]s are considered. Global and local co-ordinates are defined
to apply an orientation of each lamina. Figure 6 shows a 3-D finite element model for a
[0/90]s lay-up that contains 3160 elements. To apply the load, one end of the laminate
is fixed and an axial displacement is applied at the other end. The model can
experience excessive element distortion because of localized stiffness reduction due to
internal damage. Distortion will lead to additional difficulties in the numerical conver-
gence, resulting in a slow running or even aborted solution. Distortion control was
activated in ABAQUS, and the damage variables were limited to a maximum value of
0.99, which helps to maintain some residual stiffness. The mixed-mode traction sepa-
ration law is used to define the evolution of transverse matrix cracking, as discussed in

Fig. 5 A representative figure showing cohesive element inserted in transverse ply 900 and between the interface
00/900
Appl Compos Mater

Fig. 6 Geometry of the simulation


model [0/90]s

Section 3.1. If the damage criteria are not satisfied, then separation (matrix cracking)
does not occur and the adjacent elements will be correctly connected. Otherwise, matrix
cracking will initiate, and then the stiffness of the cohesive elements will be gradually degraded
following the linear softening law described by Eq. (14). The present analysis neglects cracks that
could develop in other ±θm off-axis plies but includes delamination
The simulated CFRP composite laminates are made of an epoxy resin (Cycom®934,
Cytec Engineered Materials) reinforced with unidirectional T300 carbon fibers. Details
of the material properties are listed in Table 1 [21]. The properties of the cohesive
elements are also presented in Table 2, and they include the elastic stiffness, strength,
and fracture toughness [47–49]. The accuracy of the analysis strongly depends on the
stiffness of the interface elements [50]. High stiffness may cause inter-penetration of
crack faces and also lead to numerical difficulties. To obtain a relatively high stiffness,
Daudeville et al. [51] proposed the normalization of the interface stiffness in terms of a
small thickness t (10−2 mm) in the resin-rich zone of the composite laminate. Here, the
resin-rich zone represents the ply interface thickness in composite laminates. A number
of researchers have suggested different values for the interface stiffness, and the
selected values include 107 N/mm3 [52], 5.7×107 N/mm3 [53], and 108 N/mm3 [54].
Zou et al. [55] suggested a value for the stiffness between 104 and 107 times the value
of the strength of the interface per unit length. In the present study, the interface
stiffness is considered as 106 N/mm3 for the matrix crack, which has been demonstrated
to provide a reasonable prediction for carbon/epoxy laminates [54]. Damage evolution
under mixed-mode loading was predicted by the Benzeggagh-Kenane fracture energy
law in which the factor of η=1.45 is based on experimental data [43]. For the GFRP
material system, the interface cohesive material parameters were assumed to be the
same as for the CFRP material system. The material properties of GFRP are presented
in Table 3 [6].

Fig. 7 FE model [0/90]s


Appl Compos Mater

Fig. 8 Convergence analysis on crack density for two different mesh size (0.5 mm and 0.25 mm) for [0/90]s

4 Results and Discussions

4.1 Effect of Transverse Matrix Crack

Experimental results from [5] are compared for the laminates with the stacking sequences of
[0/90n/0], n=1,2,3,4, [02/90n]s, n=1,2,3, [+25/−25/902]s and [+25/−25/904]s via the numeri-
cally computed crack density values versus the applied laminate stress. Zhang et al. [21]
presented a theoretical prediction of the transverse matrix crack density, which is also
compared with the predicted numerical values. In their approach to predicting the crack
density, they considered different fracture parameters as in Eq. (12) and chose values by curve
fitting of the plot of GR vs. D[21] as follows:

GIC ¼ 190 J =m2 ; G0 ¼ 125 J =m2 ; R ¼ 6:5

These fracture parameters were used to theoretically evaluate the crack density, which
agrees well with the experimental results. To validate this theoretical model, Shi et al. [33]
predicted the crack density numerically but without considering delamination or multi-axial
loading. We also included the delamination in the model but observed that there is not much
impact in [0/90]s laminates; delamination modeling will be considered in the following section.
The predicted transverse matrix crack density is plotted versus the applied laminate stress for
all of the laminates (Figs. 9, 10, 11, 12, 13, 14, 15 and 16). Here, we computed the crack
densities by observing full separation of the cohesive elements in the FE results.
In the first configuration of plies ([0/90]s), the crack density increases very rapidly after its
initiation at an applied stress of 550 MPa. Crack propagation is observed to slow after reaching 10
cracks cm−1. Zhang et al. [21] theoretically predicted a crack density value of 16 cracks cm−1,
corresponding to an applied stress of 844 MPa; the experimental results indicate a maximum crack

Table 1 Material properties for a T300/934 unidirectional laminate [21]

Material properties E11 =144.8GPa, E22 =E33 =11.38GPa,ϑ12= 0.3,ϑ23 =ϑ12 =0.42
G12 =6.48GPa, G23 =G13 =3.45GPa, ρ=1600 kg/m3
Appl Compos Mater

Table 2 Material parameters for the cohesive elements [47–49]

Direction, n Direction, s Direction, ta

Normalized elastic modulus (MPa/mm) 106 106 106


Interface strength (MPa) 51.7 40 40
Fracture toughness (J/m2) 190 790 790b
a
n normal, s shear, t tearing, see Fig. 4a
b
The value in the t direction may vary from that of the s direction, but in this study the formation of cracks is
mainly affected by modes I and II

density of 15.3 cracks cm−1 [5]. The current FE model predicted 16 cracks cm−1, corresponding to
an applied stress of 829 MPa, which is in good agreement with the theoretical and experimentally
measured crack density value. The theoretical model and the FE model predicted results for
laminates with stacking sequence [+25/−25/902]s and [+25/−25/904]s are shown in Figs. 10 and
11, respectively. Numerically computed results for the [+25/−25/902]s laminate are in good agree-
ment with the experimental and theoretical results [21], but for the [+25/−25/904]s laminate, the
crack density is under-predicted compared to the experimental values [5]. For the [+25/−25/902]s
CFRP laminate, the maximum crack density predicted using FEM is 6 cracks cm−1, compared to an
experimental value of 5.3 cracks cm−1 (Fig. 10). The same thickness ratio (φ) occurs for laminates
[0/90]s and [+25/−25/902]s, which is used to compute the crack densities theoretically [21]. This
thickness ratio has a great impact on the fracture parameters, so both lay-ups used the same fracture
parameters (GIC, G0, and R). Next, Fig. 11 shows the comparison of the predicted numerical,
theoretical and experimental results for the [+25/−25/904]s laminate, where the thickness ratio is
φ=0.5. To calculate the theoretical crack density in Fig. 10, the fracture parameters used in Eq. (12)
were altered to obtain a better fit (plot of GR vs. D[21]), as

GIC ¼ 228 J =m2 ; G0 ¼ 178 J =m2 ; R ¼ 6:2

Zhang et al. [21] described that the critical energy release rate GIC and the R-curve values
(G0 and R) differ for various lay-up configurations, explaining that crack initiation and accumulation
are dependent on the thickness ratio (φ). The present FE model predicted a maximum crack density
of 5 cracks cm−1, which is near the experimentally measured value of 4.27 cracks cm−1 (the
fractional value is due to the experimentally measured value being in cracks/inch [3]).
One more configuration of laminates with stacking sequence of [02/90n], n=1,2,3 is
simulated to validate the developed method. The maximum crack densities are found to be
18 cracks cm−1 (n=1, experimental 17.44 cracks cm−1), 18 cracks cm−1 (n=2, experimental
18.07 cracks cm−1), and 14 cracks cm−1 (n=3, experimental 14.44 cracks cm−1), correspond-
ing to an applied load 930.69 MPa, 965.16 MPa, and 688.71 MPa, as shown in Figs. 12, 13,
and 14, respectively. Transverse crack densities are calculated for [0/90n/0], n=1,2,3,4, by
varying the 900 ply quantity and thickness. We observed that as the thickness of the 900 ply
increases, the applied laminate stress required for crack initiation decreases (Fig. 15). For the

Table 3 Material properties for a E-Glass Epoxy unidirectional laminate [6]

Material properties E11 =41.7GPa, E22 =E33 =13GPa,ϑ12= 0.3,ϑ23 =ϑ13 =0.3a
G12 =3.4GPa, G23 =G13 =1.3GPa, ρ=1800 kg/m3
a
Assumed value
Appl Compos Mater

Fig. 9 Numerically predicted crack densities compared with experimental [5] and theoretical [21] results for the
[0/90]s CFRP laminates

GFRP laminates, the stacking sequence [0/903]s is considered. Crack initiation starts at
approximately 40 MPa, with a maximum of 7 cracks cm−1 found, corresponding to
276 MPa, in comparison with the experimental results for which the maximum crack density
is 7.09 cracks cm−1 at a similar stress level (Fig. 16). An empirical relationship is established
between the crack density (ρc) and the laminate stress (σa) by a linear curve fitting of the FE
result plots of the crack density vs. laminate stress (Eq. 25, Table 5).
Crack density ðρc Þ ¼ A þ Bσa ð25Þ

For the thickness variation of 900 plies in [0/90n/0] CFRP laminates (Fig. 15), we observe
that the stress level at crack initiation (σ0 (ρc =0)) decreases as the thickness of the 900 plies is
increased: 585.5 MPa ([0/90]s), 374.4 MPa ([0/903/0]), and 265.0 MPa ([0/904/0]). The crack
density slope for the CFRP laminates ranges in value from 15.40 to 42.73 (MPa/(crack/cm)),
representing the applied stress increment required to form one new crack per centimeter (B−1).

Fig. 10 Numerically predicted crack densities compared with experimental [3] and theoretical [21] results for the
[25/−25/902]s CFRP laminates
Appl Compos Mater

Fig. 11 Numerically predicted crack densities compared with experimental [3] and theoretical [21] for the [25/
−25/904]s CFRP laminates

4.2 Degradation of the Longitudinal Elastic Modulus

The longitudinal elastic modulus reduction ratios depend on the obtained crack densities,
corresponding to different laminate configurations. The effective longitudinal elastic moduli of
the laminates (0° direction) were first calculated using classical laminate theory [1]; these
calculated values are listed in Table 4. Here, we defined the longitudinal elastic modulus
reduction ratio as the ratio of the degraded stiffness (E) to the original stiffness (E0). We
computed the degraded stiffness by dividing the degraded stress value by the longitudinal
strain. The degraded stress is calculated by dividing the reaction force (computed from FEM)
by the cross-sectional area of the laminate. Figures 9, 10, 11, 12, 13, 14, 15 and 16 also show
the evolution of the longitudinal elastic modulus reduction ratio as a function of the applied
laminate stress. The elastic modulus is displayed on the graph as a reduction ratio (E/E0) from
the undamaged state. Figure 9 shows that for the [0/90]s laminate configuration, the longitu-
dinal elastic modulus exhibits only a slight degradation as the matrix crack density increases,

Fig. 12 Numerically predicted crack densities compared with experimental [5] results for the [02/90]s CFRP
laminates
Appl Compos Mater

Fig. 13 Numerically predicted crack densities compared with experimental [5] results for the [02/902]s CFRP
laminates

reaching a value of 0.94. A clear downward trend in the E/E0 data is observed for the [+25/
−25/902]s (Fig. 10) and [+25/−25/904]s (Fig. 11) CFRP laminates; the ratio reaches a minimum
value of approximately 0.84. The [02/90n] (n=1,2,3) lay-ups also show measurable degrada-
tion, with a decrease in the longitudinal elastic modulus reduction ratio value down to 0.82~
0.89 observed (Figs. 12, 13 and 14). The GFRP laminate [0/903]s also exhibits a clear
downward trend in the longitudinal reduction ratio value; the minimum ratio is at a much
lower value of 0.45. This result is consistent with the fact that for GFRP, the modulus of a 90°
ply is closer to the one for a 0°ply: 13 GPa vs. 41.7 GPa (GFRP, Table 3), compared to
11.38 GPa vs. 144.8 GPa (CFRP, Table 1). A damaged 90° ply should have more effect on the
modulus reduction for the GFRP laminate.
The longitudinal modulus reduction ratio range (E/E0 =0.82~0.99) for CFRP lami-
nates was compared with the values in the existing literature [56, 57] (E/E0 =0.84~0.99)
and was found to be in good agreement. In addition, in the case of GFRP laminates, the

Fig. 14 Numerically predicted crack densities compared with experimental [5] results for the [02/903]s CFRP
laminates
Appl Compos Mater

Fig. 15 Numerically predicted crack densities compared with experimental [5] results for the [0/90n/0]s CFRP
laminates

longitudinal modulus reduction ratio range (E/E0 =0.45~0.98) is in agreement with the
values in the existing literature (Hashin, 1987; Henaff-Gardin et al. 1996) (E/E0 =0.54~
0.99). One can observe from the results presented in Figs. 9, 10, 11, 12, 13 14 and 15
that the longitudinal elastic modulus reduction ratio decreases up to 6 % for the [0/90]s
CFRP laminates, but other configurations of CFRP laminates exhibit up to 12 % to 18 %
degradation in the longitudinal elastic modulus value.

4.3 Delamination Model

To study the delamination behavior, we simulated laminates with the stacking sequence of [45/
−45/0/90]s under in-plane tensile loading for two cases; one with only matrix cracking
modeling (1st model), and another one with delamination and matrix cracking modeling
(2nd model). The selected model geometry is 50 mm×50 mm to obtain a better crack density

Fig. 16 Numerically predicted crack densities compared with experimental [6] results for the [0/903]s GFRP
laminates
Appl Compos Mater

Table 4 Longitudinal elastic modulus for various laminates [1]

Laminates configurations Longitudinal elastic modulus (E0) (GPa)

[0/90]s CFRP 78.5


[±25/902]s CFRP 55.6
[±25/904]s CFRP 41.3
[02/90]s CFRP 101
[02/902]s CFRP 78.5
[02/903]s CFRP 65.1
[0/903]s GFRP 20.3

resolution. To check the delamination behavior, cohesive elements are inserted between the ply
interfaces (+45/−45, −45/0, and 0/90) and the matrix cracks are modeled only in the 90° plies.
Numerically predicted delamination is shown in Fig. 17. We observed that the matrix crack
density increases after its initiation at an applied stress of 134.1 MPa in the 1st model. The
maximum transverse matrix crack density is observed as 5 cracks cm−1, corresponding to a
stress of 595.4 MPa (1st model, Fig. 18). However, we observed that delamination is the first
sub-laminate failure in the second model. This result is consistent with the existing experi-
mental data reported [5]. We also observed that the matrix crack density increases after its
initiation at an applied stress of 126.7 MPa, i.e., 5.5 % lower than the 1st model. The
maximum transverse matrix crack density is observed as 5.2 cracks cm−1, corresponding to
a stress of 581.8 MPa (2nd model, Fig. 19). Figs. 17 and 18 also show the observed crack
density due to transverse matrix cracks, which gradually increase with applied stress. The
damage dissipation energies, EDMD, are computed using the ABAQUS software as:

Zt Z 

Dm
E DM D ¼ σu : εel dVdτ ð26Þ
2ð1−Dm Þ
0V

Fig. 17 Delamination in the [45/−45/0/90]s CFRP laminates


Appl Compos Mater

Fig. 18 Numerically predicted crack densities and damage dissipated energy for the [45/−45/0/90]s CFRP
laminates (1st model)

where Dm represents the damage variable, σu is the stress tensor, Ḋm is the damage rate (starts
at zero (undamaged) and increases to a maximum value of no more than one (fully damaged))
and εel is the linear elastic strain tensor. Here ‘:’ represents a tensor product operation.
Figure 18 also shows the damage dissipation energy vs. applied stress plot for the 1st model
(matrix cracking only). Here, delamination damage is calculated numerically in terms of
damage dissipative energy [58]. We observed that as the matrix crack density reaches a high
level, the damage dissipation energy increases at a higher rate. We also see that for zero crack
density (no full cracks), some damage dissipated energy exists; this dissipated energy is due to
partial cracks because no full transverse matrix crack exists below 134.1 MPa of applied
laminate stress. Similarly, Fig. 19 also shows the damage dissipation energy vs. applied stress
plot for the 2nd model (delamination and matrix cracking). We also observe that as the matrix
crack density reaches a high level, the damage dissipation energy increases at a higher rate.
The difference in damage dissipative energy is very large between the 1st and 2nd models:

Fig. 19 Numerically predicted crack densities and damage dissipated energy for the [45/−45/0/90]s CFRP
laminates (2nd model)
Appl Compos Mater

approximately 1.8 vs. 33.8 J/m2 in the plateau region of the curve. We think that this difference
is due to delamination occurring at the ply interfaces (+45/−45, −45/0, and 0/90), which have a
much larger surface area than the cracks alone (Table 5).
Finite element modeling provides a reasonable prediction of the initiation and accumulation
of transverse cracks, especially for the cross-ply laminates [0/90]s (Fig. 9). Experimental
observation [5] indicated that different types of internal cracks existed in the laminates
examined, i.e., straight cracks, partial angle cracks, and curved cracks. At higher applied load,
closer to ultimate failure, some local delamination at the intra-laminar matrix crack tip
developed. These types of damage mechanisms do dissipate energy and delay laminate
fracture [3]. In the future, more detailed analysis of the degraded mechanical properties should
be presented.

5 Conclusion

We extended the work of Shi et al. (2014) [31] and studied using FEM the effect of
thickness variation and laminate configuration on transverse crack densities and also on
longitudinal elastic modulus degradation due to transverse matrix cracking. Transverse
matrix cracking, as well as the interface between plies, was simulated using equally
spaced cohesive elements. For matrix cracking modeling, these cohesive elements are
inserted in the transverse ply (900) parallel to the fiber direction. For the delamination
study, cohesive elements are inserted between the ply interfaces. A theoretical model
(ECM) presented by Zhang et al. [21] was implemented for predicting transverse matrix
crack density while considering equally spaced cracks. Based on experimental obser-
vation at the saturation level, a crack spacing length was implemented using FEM. A
crack spacing of 0.5 mm was used for positioning of the cohesive interface elements in
the transverse ply according to the mesh convergence analysis [31]. Laminates of
various stacking sequences ([0/90]s, [+25/−25/902]s and [+25/−25/904]s) were simulated

Table 5 Coefficient of empirical relationships

Material Laminate Configuration Crack density


ρc =A+Bσa

σ0 (ρc =0) B R2
(MPa) crack 
=cm
M Pa

CFRP [0/90]s 585.5 0.0649 0.956


[25/−25/902]s 206.3 0.0443 0.935
[25/−25/904]s 106.1 0.0393 0.880
[02/90]s 500.0 0.0438 0.988
[02/902]s 155.9 0.0234 0.951
[02/903]s 160.2 0.0267 0.983
[0/903/0] 374.4 0.0478 0.912
[0/904/0] 265.0 0.0483 0.943
GFRP [0/903]s −11.9 (=0) 0.0287 0.778
Appl Compos Mater

using cohesive elements. The obtained transverse matrix crack densities were found to
be in good agreement with experimental results [3, 5]. In addition, we also analyzed the
effect of the transverse ply (900) thickness proportion on the crack density for [0/90n/0]
laminates. We observed that as the thickness of the transverse ply increases, the crack
density decreases. The present method was also validated for a GFRP laminate; good
agreement is found with existing experimental results [6]. Delamination damage was
quantified numerically in terms of damage dissipative energy. The longitudinal elastic
modulus reduction ratio was also computed in the range of transverse crack occurrence.
The longitudinal elastic modulus was observed to decrease up to 6 % for the [0/90]s
CFRP laminate compared to an experimental measurement of 5 % [55, 56]. The other
configurations of CFRP laminates exhibit up to 12 % to 18 % degradation in the
longitudinal elastic modulus value compared to a representative experimental measure-
ment of 16 % [55, 56]. Empirical relations were deducted for crack density vs. applied
stress by linear curve fitting. Further work will be required to account for multi-axial
loading and its impact on the stiffness/strength properties and on laminate fatigue life
prediction.

Acknowledgments This work was supported by the Human Resources Development Program (No.
20114010203070) of the Korea Institute of Energy Technology Evaluation and Planning (KETEP) Grant funded
by the Korea government Ministry of Trade, Industry and Energy. The research was supported by ‘Software
Convergence Technology Development Program’, through the Ministry of Science, ICT and Future Planning
(S1002-13-1004).

References

1. Jones, R.M.: Mechanics of Composite Materials. Second Edition, Taylor and Francis Publication, Boca
Rotan (2008)
2. Berthelot, J.M., Le Corre, J.-F.: A model for transverse cracking and delamination in cross-ply laminates.
Compos. Sci. Technol. 60, 1055–1066 (2000)
3. Crossman, F.W., Wang, S.D.: The dependence of transverse cracking and delamination on ply thickness in
Graphite/Epoxy laminates. Damage Composite Material op. ct. 118–39 (1984)
4. Wang, A.S.D., Kishore, N.N., Li, C.A.: Crack development in graphite-epoxy cross-ply laminates under uni-
axial tension. Compos. Sci. Technol. 24, 1–31 (1985)
5. Wang, A.S.D.: Fracture mechanics of sub-laminate cracks in composite materials. Compos. Technol. Rev. 6,
45–62 (1984)
6. Highsmith, A.L., Reifsnider, K.L.: Stiffness reduction mechanism in composite laminates. Damage in
Composite Materials. ASTM, STP 775, 103–117 (1982)
7. Berbinau, P., Soutis, C., Guz, I.A.: Compressive failure of 00 unidirectional carbon-fiber-reinforced plastics
(CFRP) laminates by fiber micro buckling. Compos. Sci. Technol. 59, 1451–1455 (1999)
8. Anderson, T.L.: Fracture Mechanics- Fundamentals and Applications. CRC Press, New York (1995)
9. Kashtalyan, M., Soutis, C.: The effect of delamination induced by transverse cracks and splits on stiffness
properties of composite laminates. Compos. A: Appl. Sci. Manuf. 31, 107–119 (2000)
10. Kashtalyan, M., Soutis, C.: Stiffness degradation in cross-ply laminates damaged by transverse cracking and
splitting. Compos. A: Appl. Sci. Manuf. 31, 335–351 (2000)
11. Kashtalyan, M., Soutis, C.: Analysis of local delamination in composite laminates with angle-ply matrix
cracks. Int. J. Solids Struct. 39, 1515–1537 (2002)
12. Kashtalyan, M., Soutis, C.: Mechanisms of internal damage and their effect on the behavior and properties of
cross-ply composite laminates. Int. Appl. Mech. 38(6), 641–657 (2002)
13. Zhang, J., Soutis, C., Fan, J.: Strain energy release rate associated with local delamination in cracked
composite laminates. Composites 25(9), 851–862 (1994)
14. Hashin, Z., Rotem, A.: A fatigue failure criterion for fiber reinforced materials. J. Compos. Mater. 7, 448–
464 (1973)
Appl Compos Mater

15. Hashin, Z.: Failure criteria for uni-directional fiber composites. J. Appl. Mech. 47(1), 329–334 (1980)
16. Nairn, J.A.: The strain energy release rate of composite micro-cracking: a variational approach. J. Compos.
Mater. 23(11), 1106–1129 (1989)
17. Varna, J., Berglund, L.A.: Multiple transverse cracking and stiffness reduction in cross-ply laminates. J.
Compos. Technol. Res. 13(2), 97–106 (1991)
18. Varna, J., Berglund, L.A.: A model for prediction of the transverse cracking strain in cross-ply laminates. J.
Reinf. Plast. Compos. 11(7), 708–728 (1992)
19. Berglund, L.A., Varna, J.: Thermo-elastic properties of composite laminates with transverse cracks. J.
Compos. Technol. Res. 16(1), 77–87 (1994)
20. Zhang, J., Fan, J., Soutis, C.: Analysis of multiple matrix cracking in [±θm/90n]s composite laminates. Part1:
In-plane stiffness properties. Composites 23(5), 291–298 (1992)
21. Zhang, J., Fan, J., Soutis, C.: Analysis of multiple matrix cracking in [±θm/90n]s composite laminates. Part 2:
Development of transverse ply cracks. Composites 23(5), 299–304 (1992)
22. Smith, P.A., Boniface, L., Glass, N.F.C.: A comparison of transverse cracking phenomena in [0/90]s and [90/
0]s CFRP laminates. Appl. Compos. Mater. 5, 11–23 (1998)
23. Berthelot, J.-M.: Transverse cracking and delamination in cross-ply glass-fiber and carbon fiber reinforced
plastics laminates: static and fatigue loading. Appl. Mech. Rev. 56(1), 111–148 (2003)
24. Kachanov, L.M.: On the creep rupture time. Izv AN SSSR Otd Tekhn Nauk 8, 26–31 (1958)
25. Rabotnov, Y.N.: On the Equations of State for Creep. Progress in Appl. Mech. Prager Anniversary Volume.
Macmillan, New York (1963)
26. Donadon, M.V., Iannucci, L., Falzon, B.G., Hodgkinson, J.M., Almeida, S.F.M.: A progressive failure model
for composite laminates subjected to low-velocity impact damage. Comput. Struct. 86, 1232–1252 (2008)
27. Faggiani, A., Falzon, B.G.: Predicting low-velocity impact damage on a stiffened composite panel. Compos.
A: Appl. Sci. Manuf. 41, 737–749 (2010)
28. Iannucci, L., Ankersen, J.: An energy based damage model for thin laminated composites. Compos. Sci.
Technol. 66, 934–951 (2006)
29. Yokoyama, N.O., Donadon, M.V., Almeida, S.F.M.: A numerical study on the impact resistance of
composite shells using an energy based failure model. Compos. Struct. 93, 142–152 (2010)
30. Barbero, E.J., Cortes, D.H.: A Mechanistic model for transverse damage initiation, evolution, and stiffness
reduction in laminated composites. Compos. Part B. Eng. 41(2), 124–132 (2010)
31. Bouhala, L., Makradi, A., Belouettar, S., Kiefer-Kamal, H., Freres, P.: Modeling of failure in long fibers
reinforced composites by X-FEM and cohesive zone model. Compos. Part B. Eng. 55, 352–361 (2013)
32. Shi, Y., Swait, T., Soutis, C.: Modeling damage evolution in composite laminates subjected to low-velocity
impact. Compos. Struct. 94, 2902–2913 (2012)
33. Shi, Y., Pinna, C., Soutis, C.: Interface cohesive element to model matrix crack evolution in composite
laminates. Appl. Compos. Mater. 21, 57–70 (2014)
34. Caputo, F., De Luca, A., Lamanna, G., Lopresto, V., Riccio, A.: Numerical investigation of onset and
evolution of LVI damages in Carbon–Epoxy plates. Compos. Part B. Eng. 68, 385–391 (2015)
35. Riccio, A., De Luca, A., Di Felice, G., Caputo, F.: Modeling the simulation of impact induced damage onset
and evolution in composites. Compos. Part B. Eng. 66, 340–347 (2014)
36. Carraro, P.A., Quaresimin, M.: A stiffness degradation model for cracked multidirectional laminates with
cracks in multiple layers. Int. J. Solids Struct. 58, 34–51 (2015)
37. Sills, R.B., Thouless, M.D.: Cohesive-length scales for damage and toughening mechanisms. Int. J. Solids
Struct. 55, 32–43 (2015)
38. Van der Meer, F.P., Davila, C.G.: Cohesive modeling of transverse cracking in laminates under in-plane
loading with a single layer of elements per ply. Int. J. Solids Struct. 50, 3308–3318 (2013)
39. Reedy Jr., E.D.: Cohesive zone finite element analysis of crack initiation from a butt joint’s interface corner.
Int. J. Solids Struct. 51, 4336–4344 (2014)
40. Han, Y., Hahn, H.T., Croman, R.B.: A simplified analysis of the transverse ply cracking in cross-ply
laminates. Compos. Sci. Technol. 31, 165–177 (1988)
41. Hahn, H.T., Han, Y.M., Kim, R.Y.: Resistance curves for ply cracking in composite laminates. Proc 33rd. Int.
SAMPE. Symp., 1101–8 (1998)
42. Fan, J., Zhang, J.: In-situ damage evolution and micro/macro transition for laminated composites. Compos.
Sci. Technol. 47(2), 107–118 (1993)
43. Camanho, P.P., Dávila, C.G.: Mixed-mode decohesion finite elements for the simulation of delamination in
composite materials., Tech. Rep. NASA/TM-2002-211737 (2002)
44. Camanho, P.P., Dávila, C.G., de Moura, M.F.: Numerical simulation of the mixed-mode progressive
delamination in composite materials. J. Compos. Mater. 37(16), 1415–1438 (2003)
45. ABAQUS.: ABAQUS Version 6.10, Dassault systems. Providence (2010)
46. Barbero, E.J.: Finite Element Analysis of Composite Materials Using Abaqus. CRC Press (2010)
Appl Compos Mater

47. Turon, A.: Simulation of Delamination in Composites Under Quasi-Static and Fatigue Loading Using
Cohesive Zone Models. PhD Dissertation Universitat de Girona (2006)
48. Ankersen, J., Davies, G.A.O.: Interface elements–advantages and limitations in CPRP delamination model-
ing. In: 17th International Conference on Composite Materials. Edinburgh, UK (2009)
49. Pinho, S.T., Iannucci, L., Robinson, P.: Fracture toughness of the tensile and compressive fiber failure modes
in laminated composites. Compos. Sci. Technol. 66(13), 2069–2079 (2006)
50. Turon, A., Dávila, C.G., Camanho, P.P., Costa, J.: An engineering solution for mesh size effects in the
simulation of delamination using cohesive zone models. Eng. Fract. Mech. 74, 1665–1682 (2007)
51. Daudeville, L., Allix, O., Ladevèze, P.: Delamination analysis by damage mechanics: Some applications.
Compos. Eng. 5(1), 17–24 (1995)
52. Gonçalves, J.P.M., de Moura, M.F.S.F., de Castro, P.M.S.T., Marques, A.T.: Interface element including
point-to-surface constraints for three-dimensional problems with damage propagation. Eng. Comput. 17(1),
28–47 (2000)
53. Mi, Y., Crisfield, M.A.: Analytical Derivation of Load/Displacement Relationships for Mixed-Mode
Delamination and Comparison with Finite Element Results. Imperial College, Department of Aeronautics,
London (1996)
54. Schelleckens, J.C.J., de Borst, R.: On the numerical integration of interface elements. Int. J. Numer. Methods
Eng. 36, 43–66 (1993)
55. Zou, Z., Reid, S.R., Li, S., Soden, P.D.: Modeling inter-laminar and intra-laminar damage in filament wound
pipes under quasi-static indentation. J. Compos. Mater. 36, 477–499 (2002)
56. Hashin, Z.: Analysis of orthogonally cracked laminates under tension. Trans. ASME J. Appl. Mech. 54, 272–
279 (1987)
57. Henaff-Gardin, C., Lafarie-Frenot, M.C., Gamby, D.: Doubly period matrix cracking in composite laminates
Part 1: general in-plane loading. Compos. Struct. 36, 113–130 (1996)
58. Turon, A., Camanho, P.P., Costa, J., Renart, J.: Accurate simulation of delamination growth under mixed-
mode loading using cohesive elements: definition of interlaminar strengths and elastic stiffness. Compos.
Struct. 92, 1857–1864 (2010)

View publication stats

You might also like