You are on page 1of 45

Accepted Manuscript

Dynamic Delamination in Laminated Fiber Reinforced Composites: A Contin-


uum Damage Mechanics Approach

Amir Shojaei, Guoqiang Li, P.J. Tan, Jacob Fish

PII: S0020-7683(15)00293-0
DOI: http://dx.doi.org/10.1016/j.ijsolstr.2015.06.029
Reference: SAS 8832

To appear in: International Journal of Solids and Structures

Received Date: 2 January 2015


Revised Date: 30 April 2015

Please cite this article as: Shojaei, A., Li, G., Tan, P.J., Fish, J., Dynamic Delamination in Laminated Fiber
Reinforced Composites: A Continuum Damage Mechanics Approach, International Journal of Solids and
Structures (2015), doi: http://dx.doi.org/10.1016/j.ijsolstr.2015.06.029

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Dynamic Delamination in Laminated Fiber Reinforced Composites: A
Continuum Damage Mechanics Approach

Amir Shojaei1*†, Guoqiang Li1,2, PJ Tan3, and Jacob Fish4


1
Department of Mechanical & Industrial Engineering, Louisiana State University, Baton Rouge, LA 70803, USA
2
Department of Mechanical Engineering, Southern University, Baton Rouge, LA 70813, USA
3
Department of Mechanical Engineering, University College London, Torrington Place, London WC1E 7JE, UK
4
Department of Civil Engineering and Engineering Mechanics, Columbia University, NY, USA
*Corresponding author. Tel.: 001-225-933-1078; Fax: 001-225-578-5924;

Currently at Halliburton Energy Services, Houston, TX
Emails: a.shojaei.mech.eng@gmail.com, and amir.shojaei@halliburton.com (A. Shojaei); lguoqi1@lsu.edu (G.
Li), pj.tan@ucl.ac.uk (PJ. Tan), and fishj@columbia.edu (J. Fish)

Abstract

Delamination causes a significant reduction in the load carrying capacity of Fiber Reinforced

Polymer (FRP) composites, which is a major concern to the aerospace and automotive

industries. High performance FRPs are often subjected to dynamic loadings of different

energy densities in their service life when strain rate, stress triaxiality, temperature, and mode

of fracture can have a significant knock-down effect on the interfacial strength of plies in a

laminate. This paper develops a predictive tool, in a Continuum Damage Mechanics (CDM)

framework, to assess delamination damage in FRPs under dynamic loading; the model takes

into account the effects of dynamic energy density, mixed-mode fractures and temperature.

The CDM model is formulated based on the Fracture Mechanics (FM) of decohesion and an

advantage of the proposed model is that nearly all of the material parameters can be obtained

directly through calibration to experimental data rather than numerical curve fitting of

simulation results. The developed scheme is coded into a commercial FEA package

(ABAQUS) through user-defined subroutines and the fidelity of the model is assessed by

comparison with existing experimental data in the literature. The validated model is used to

investigate delamination damage in a laminated FRP subject to projectile impact loading,

where it will be shown that the extent of delamination damage through the thickness of the

FRP structure is dependent upon different wave propagation scenarios. The proposed model

provides a design platform for damage assessment caused by dynamic delamination and may

be a useful tool for designing FRP composites with a greater impact tolerance.

Keywords: Dynamic Delamination, Fiber Reinforced Polymers, Continuum Damage Mechanics

1
1. Introduction

The common failure modes encountered in FRPs are delamination, transverse matrix

cracking, fiber fracture, and fiber-matrix interfacial debonding (Voyiadjis et al. , Chaboche et

al. 2001, Le Quang and He 2008, Horstemeyer and Bammann 2010, Azizi et al. 2011, Doghri

et al. 2011, Kruch and Chaboche 2011, Voyiadjis et al. 2011, Brassart et al. 2012, Li and

Shojaei 2012, Shojaei et al. 2012, Voyiadjis et al. 2012a, Voyiadjis et al. 2012b, Voyiadjis et

al. 2012c, Shojaei and Li 2013, Shojaei et al. 2014, Shojaei et al. 2014). Delamination is

particularly pertinent to composite lay-ups with resin-rich interlaminar layers that have not

been reinforced in the thickness direction. Due to the lower mechanical strength of the

interlaminar layers, compared to their adjacent reinforced plies, interlaminar delamination can

occur in any of the three basic modes shown schematically in Fig. 1. In reality, the state of the

stress in the interlaminar layer is three-dimensional (3D) and delamination occurs under

mixed-mode condition. Of particular concern to designers is delamination caused by local

impact loading and free-edge stresses. A wealth of literature on the modeling of delamination

damage in FRPs have already existed; only a selection of recent pertinent ones will be

reviewed here. Readers are referred to (Hutchinson 1982, Wisnom 2012, Liu et al. 2013, Park

and Paulino 2013) for a more complete review on recent developments in cohesive modeling

methodologies.

Figure 1 Schematic of the three basic modes of loading, viz. (a) Mode I (b) Mode II and (c) Mode III, for

inter-ply delamination.

Different approaches are followed to simulate delamination in composites, and they can be

broadly classified into three categories, viz. (a) Virtual Crack Closure Techniques (VCCT),

2
(b) Cohesive Zone Models (CZMs) (Li and Chandra 2003, Zhang and Paulino 2005, Xu and

Lu 2013), and (c) CDM models. The VCCT technique is based on Irwin’s assumption that

when a crack extends by a small amount, the energy released in the process is equal to the

work required to close the crack to its original length (Krueger et al. 1999). Whilst this

method is computationally efficient, there are numerical difficulties associated with it, say,

utilizing re-meshing techniques, identification of the delamination front after each crack ‘pop-

up’ and prescribing the delamination path. Cohesive zone models, on the other hand, treat the

fracture formation as a gradual phenomenon in which the separation of the crack surfaces

takes place across an extended crack tip, or cohesive zone, resisted by cohesive tractions

(Needleman 1987a, Needleman and Tvergaard 1987b). Different types of cohesive elements

have been proposed to model cohesive fracture with the finite element method, ranging from

zero-thickness volumetric elements connecting solid elements (De Moura et al. 1997), finite-

thickness volumetric elements connecting shell elements (Reedy et al. 1996), to line elements

(Chen et al. 1999). (Kubair and Geubelle 2003) have argued that intrinsic cohesive

delamination models may result in numerical instabilities and give rise to unrealistic results as

the crack opening velocity becomes negative at the cohesive zone tip. A decohesion element,

capable of dealing with crack propagation under mixed-mode loading, has been proposed and

its efficacy has been demonstrated by (Benzeggagh and Kenane 1996, Kenane and

Benzeggagh 1997, Camanho et al. 2003). The main disadvantages of using cohesive elements

in FEA are (1) difficulties associated with obtaining converged solutions (Alfano and

Crisfield 2001), and (2) the a priori identification and prescription of likely fracture path(s).

(Corigliano 1993) discussed the difficulties concerning the use of interface models in

numerical analyses. (Schellekens and De Borst 1993) simulated the free edge delamination of

uniaxially stressed layered specimens, using non-linear FEA. CDM models have been

developed to model delamination process in FRPs by a number of researchers, including

(Allix and Ladevèze 1992, Li et al. 1998, Zou et al. 2002, Li et al. 2006, Maimi et al. 2008).

The CDM framework outperforms the conventional VCCT and CZM by (1) CDM utilizes

regular solid elements whilst CZM or VCCT requires that specific types of element to be used

3
for meshing the fracture path, (2) CDM can be coupled with plasticity theories to provide a

more realistic representation of the fracture mechanisms in ductile materials, and (3) fracture

path in CDM approach evolves naturally based upon damage dissipative energies while in the

case of CZM and VCCT the fracture path needs to be prescribed.

The topic of dynamic crack growth has been investigated by many researchers including

(Glennie 1971, Glennie 1971a, Freund and Hutchinson 1985, Tvergaard and Hutchinson

1996, Landis et al. 2000, Chen and Ghosh 2012). In general, higher crack velocity results in

higher strain rate hardening effect within the fracture process zone, leading to increased

toughness of the interface (Wei and Hutchinson 1997). Dynamic crack growth along the

interface of a FRP composite has been investigated experimentally and numerically by

(Tzaferopoulos and Panagiotopoulos 1993, Brunner 2000, Corigliano and Allix 2000,

Espinosa et al. 2000, Hashagen and de Borst 2000, Sprenger et al. 2000, Coker et al. 2003,

Khan and Khraisheh 2004, Mariani and Corigliano 2005, Khan and Farrokh 2006, Ahmed

and Sluys 2014). (Slepyan 2010) has studied the non-uniform dynamic crack growth in a

homogeneous isotropic elastic medium subjected to the action of remote oscillatory loads and

(Remmers et al. 2008) carried out FE analyses, using cohesive segments, of the dynamic

decohesion. Response of an infinite orthotropic material with a semi-infinite crack under

impact loads has been investigated by (Rubio-Gonzalez and Mason 2000) and X-FEM has

been utilized by (Grégoire et al. 2007) to study the dynamic crack growth problem.

Asymptotic crack tip stress-strain fields have been developed by (Zhu and Hwang 2008) for

the case of dynamic fractures. Rate-dependent models for damage and plastic deformation of

brittle and ductile materials under dynamic loading have also been extensively investigated

(Zuo et al. 2010, Shojaei et al. 2013). The stability problem of a dynamically propagating

crack has been studied by (Obrezanova et al. 2002, Obrezanova et al. 2002). Bazant and co-

workers have studied the size effect of cohesive cracks (Zi and Bažant 2003, Caner and

Bažant 2009). Despite the developments, nearly all the delamination models for FRPs in the

literature consider only a subset of possible failure mechanisms. Their approach is typically

4
limited to using traction separation laws (cohesive laws) or CDM models that have been

developed based on quasi-static experimental data (Ouyang and Li 2009b, Ouyang and Li

2009c). One of the problems associated with these simplified approaches is that they do not

account for the complete state of the applied stress when investigating the delamination

mechanism. In other words, in dynamic problems with different energy densities, the applied

deviatoric and hydrostatic stress waves may be significant and need to be accounted for in

order to accurately predict fracture (Bao and Wierzbicki 2004, Hooputra et al. 2004, Bao and

Wierzbicki 2005, Dey et al. 2007, Crowell et al. 2012). It is conventional to classify dynamic

problems into the low and high energy density cases where different deformation and damage

mechanisms may operate in each case (Shojaei et al. 2013). In the case of high dynamic

energy density problems, the induced hydrostatic stress may be several times the material

strength and the temperature, due to the dissipative mechanisms, may reach the melting

temperature of the material (Eftis and Nemes 1991, Eftis and Nemes 1996, Eftis et al. 2003).

Furthermore, when the compressive shock waves induced by the loading is reflected from any

free surfaces as tensile waves, delamination damage can occur by spalling (Eftis et al. 2003;

Shojaei, et al. 2013). In the case of lower dynamic energy densities, the shear stress is the

dominant driving force for the delamination damage (Chen and Ghosh 2012) even though

hydrostatic stress could still have a significant effect on the delamination process. In this

paper, a comprehensive delamination model is developed within the CDM framework that

accounts for the influence of shear stress and hydrostatic stress effects, as well as the fracture

mode-mixity effects, on the delamination of FRPs. Several attempts have been made to

correlate the micromechanics of damage mechanisms in composite materials to the continuum

level, including homogenization techniques (Jain and Ghosh 2008, Fish 2013, Shojaei and Li

2013, Shojaei and Fish 2015). Instead of dealing with microscale and macroscale bridging

complications, the present work focuses on developing continuum level damage models with

enhanced failure criteria that can capture progressive delamination in FRPs and eliminate the

mesh dependency of ill-posed constitutive relations. The proposed model is a

phenomenological mechanism based CDM model, and the thermodynamic consistency is not

5
discussed herein. The main contributions of the present work include, (i) a rate and

temperature sensitive CDM model is formulated for the delamination analysis of FRPs, (ii)

the developed model is capable of capturing mode-I, mode-II and mixed-mode delaminations,

and (iii) the dynamic loading effect on the progressive delamination is considered.

The present paper is structured as follows. The kinematics of the problem is set out in Section

2 and the fracture mechanics of decohesion is presented in Section 3. The CDM model for

delamination is developed in Section 4; and in Section 5 the effects of dynamic energy

density are incorporated into the CDM model. Section 6 concerns the computational

implementation of the proposed model, and the simulations and results are discussed in

Section 7.

2. Kinematics

Both inelastic deformation and damage mechanisms are concurrently active during the

delamination process of laminated composites. In this work it is assumed that the inelastic

deformation effect on the non-linear mechanical response of each reinforced ply during the

delamination process is negligible. Thus, only damage mechanisms in the interface remain the

main source of non-linearity in softening and post-softening behavior during the delamination

process. Furthermore, in order to simplify the formulation of the delamination process, the

plasticity and damage mechanisms within each reinforced ply are assumed to have negligible

effect on the interlaminar delamination process. Hence, the only damage mechanism,

considered herein, is due to the delamination of the interface that connects two adjacent

elastic plies. The developed delamination model may be used in conjunction with the

plasticity/damage constitutive relations of the reinforced plies, to perform a fully coupled

analysis though this is beyond the scope of the present study.

It is assumed that the infinitesimal deformation mechanism governs the delamination process

within the fracture process zone. The additive decomposition of the total strain rate,  , into


its elastic,  , and damage, 

, strain rate components holds as follows

6
 = 

+ 

. (1)

Appendix A elaborates the physics behind Eq. (1). The constitutive relations for 

and



will need to be specified. Nucleation and growth of microcracks lead to a ‘deterioration’

of the macroscopic properties of the interlaminar layer either in the form of added elastic

flexibility (diminution of the elastic stiffness) or increased material anisotropy (Lemaitre and

Dufailly 1987). Hence, it is convenient to write the constitutive equation for the damaged

material in the form of a generalised Hooke’s Law as (Bouvard et al. 2009)


 =   (2)

where  is the fourth-order time-dependent damaged elastic stiffness tensor (to be defined

later based on the damage parameters  ) and


 is the Cauchy stress tensor. The inelastic

damage strain tensor 



= ∫ 

, defined as the residual strain upon complete unloading, is

given by



=  − 

(3)

where 

= ∫ 

 is the damaged elastic strain tensor. It remains to specify an appropriate

evolution law for the damage parameters and to formulate the added compliance tensor using

a CDM approach to be discussed later. It is worthwhile noticing that when a delaminated

interface is unloaded, the magnitude of the residual strains may not be visible at macroscopic

scale; although, decohesion and debonding processes will induce the residual strains at

microscale within the fracture process zone. In other words, one may consider the localized

strains at the crack tip to justify the existence of the damage strain rate tensor 

.

3. Delamination and Fracture

The strength and thickness of the interface layer that connects two reinforced plies are

determined by many factors including depth of the diffused resin, the intermolecular van der

Waals bonds, environmental effects and temperature. Figure 2(a) shows a schematic of the

7
atomistic level decohesion and progressive plastic deformation in the fracture process zone as

the crack-tip advances. The effect of cohesive layer thickness on traction-separation relations

of composite layups is experimentally studied by Ji et al. (2013). In most of practical

applications the thickness of cohesive layer is negligible compared to the thickness of plies.

Thus, it is not unreasonable to assume that interlaminar delamination in fiber-reinforced

composites occurs over a surface of vanishing thickness, allowing the use of a cohesive

formulation. The cohesive-zone methodology is particularly useful when the crack path in

known a priori – which in this study is along the two connecting plies. Figure 2(b) shows a

schematic of the typical decohesion relation where separation of adjacent plies is related to

the magnitude of the cohesive traction across the interface layer. The work of separation,

which is the work needed to create a unit area of fully developed crack, is given by

(Barenblatt 1959)


Γ = 


(4)

with
and  the stress and relative displacement across the fracture process zone. In the

current application, the three important parameters in the CZM are peak separation stress
∗,

critical separation distance  and work of separation or fracture energy given by Eq. (4).

When the peak separation stress


∗ is higher than the tensile yield strength of the interface, a

significant plastic zone develops (Tvergaard and Hutchinson 1992, Hutchinson and Evans

2000). However, if the size of this plastic zone is small compared to other characteristic

lengths in the model, it is acceptable to consider the plastic dissipation energy Γ as a

material parameter that can be measured experimentally (Tvergaard and Hutchinson 1992,

Hutchinson and Evans 2000); otherwise, another fracture framework is needed to describe the

effect of plastic dissipation on the delamination response.

8
Figure 2 (a) Fracture energy dependence on cohesive and plastic dissipative works, and (b) Normal or shear

traction-separation curve for Embedded Process Zone (EPZ) at the crack tip. The EPZ under normal

opening mode-I is depicted in which  denotes the limiting fracture energy,  is the fracture energy

without any plasticity dissipation and  indicates the radius of the plastically affected area.

In order to simplify the formulation of cohesive behavior, it is assumed that small-scale

plastic deformation prevails in the vicinity of the crack-tip so that Γ → 0 and Γ dominates

the decohesion response. Two cohesive behaviors are assumed to be in operation at the

interface; they are modeled by the normal (mode-I) and shear (mode-II and –III) cohesive

laws. Based on experimental observations the normal traction


# and normal separation #

curve is constituted from two distinguishable sections including linear elastic, and

fracture/post-fracture softening response. The following phenomenological model is

introduced to simulate the decohesion in normal mode:

%# # for # < #

# = $ ∗ # − # .
- 0

*1 −  , for # > #
# − #
(5)


where 12 is the non-linearity exponent; %# the normal stiffness of cohesion; # (=


∗ /%# )

and # (=256 /


∗) are the initiation and critical displacements, respectively, and 56 is the

critical energy release rate for mode-I fracture. In a similar vein, the relationship between

shear traction 7 and sliding 8 is

9
%8 8 for 8 < 8
7=9 ∗ 8 − 8
-: 0
7 *1 −  , for 8 > 8
8 − 8
(6)

where 1; is the non-linearity exponent; %8 the shear stiffness; 8 (=7 ∗ /%8 ) and 8 (=2566 /

7 ∗ ) are the initiation and critical displacements, respectively, and 566 is the critical energy

release rate for mode-II fracture. Mixed-mode fracture is considered in Section 4.2 by

introducing the mode-mixity effect on the normal and tangential cohesive responses. Due to

the fact that the interlaminar layer is constructed from a thin resin rich layer, it is not

unreasonable to assume that the interlaminar layer is isotropic. In other words, the

interlaminar layer is made of bulk resin material without any reinforcement and it can be

considered isotropic.

4. Modeling the Delamination using Continuum Damage Mechanics

In a CDM framework, the density of micro-flaws within the interlaminar layer is represented

by a damage parameter (Kachanov 1958, Murakami 1988, Haddag et al. 2007, Haddag et al.

2009, Voyiadjis et al. 2012a). Following (Sayers and Kachanov 1991), the state of anisotropic

damage is incorporated within a second rank damage tensor, < . Figure 1 shows the

Cartesian coordinate axes adopted to describe the geometry of the problem. If assuming that

delamination damage occurs only by normal opening (mode-I) and/or shear sliding (mode-II

or mode-III), then the active damage parameters in the interlaminar layer are the through-

thickness normal damage == and the shear damages =; and =2 . Hence, the damage tensor

reduces to

0 0 2=
< = > 0 0 ;= ?.
2= ;= ==
(7)

Since the interlaminar layer between two adjacent plies is assumed to be isotropic (see

Section 3), it follows immediately that only one shear damage model needs to be defined to

10
calculate 2= and ;= . To avoid complications associated with size scale effects on the

decohesion process, the thickness of the interlaminar layer needs to be large compared to the

size of the plastic zone at the tip of the delamination crack. The damage parameters will be

defined analytically based on traction-separation laws in the following subsections.

4.1. CDM framework

Two approaches exist to link the damaged and undamaged material properties, viz.

equivalence of strain energy densities and principle of strain equivalence between the

damaged and fictitious effective configurations (Voyiadjis et al. 2012a, Voyiadjis et al.

2012c). Based on the principle of strain equivalence, the relationships between the

interlaminar material properties and damage parameters are given by

@= = @A B1 − ==C,

D = D2= = D;= = D̅ B1 − 2= C,


(8)

where @A and D̅ are the elastic tensile and shear moduli of an undamaged interface,

respectively. A simple procedure is proposed here to calibrate the normal and shear damage

parameters in Eq. (8) against known normal and shear traction separation laws that are widely

available in the literature for different FRPs. Upon the initiation of damage, the parameters in

Eq. (8) are computed directly from known traction-separation relationships as follows

== = 1 − H
FG
G G
,

2= = ;= = 1 − H  ,
I
(9)
J J

where # and 8 are the normal and tangential displacements of the interface; whilst
# K# L

and 7K8 L are computed directly from Eqs. (5) and (6), respectively. It is worth noting that the

coupling effects between normal and shear damages are ignored in Eq. (9) and mode-I and –II

fractures are treated separately; however, after the shear and normal moduli of the interface

are updated, other properties may be affected because of the relationships between the elastic

constants.

11
4.2. Mixed-mode delamination

In mixed-mode fractures, the normal and shear separations are concurrently active. For any

combination of in-plane crack deformation modes, the relative amount of mode–I and –II

loading within the cracked edge zone may be represented by a phase angle M based on the

local stress state given by (Freund and Suresh 2003)

|7 |
M = lim Rarctan W YZ
Q→
#
(10)

where |7| = [; \ \ is the effective shear stress in the interface where \ K=
 − =
 L
= 2

is the deviatoric part of the applied Cauchy stress


 . Depending on the material, the phase

angle typically ranges between – ^/2 ≤ M ≤ ^/2 (Stringfellow and Freund 1993).

Hutchinson and Evans (2000) have shown that in the case of highly negative or positive phase

angles, i.e. M < −^/6 or M > ^/4, a tenfold increase in fracture resistance compared to the

model-I toughness is possible (Hutchinson and Evans 2000). To account for mixed-mode

fracture of the interlaminar layer, the phase angle must be incorporated into the traction-

separation relationships of the interlaminar layer given in Eq. (5) (for pure model-I, i.e. M =

0) and into Eq. (6) (for pure mode-II fracture, i.e. M = +^/2 ).

In the present study, mixed-mode delamination is treated phenomenologically by introducing

equivalent normal and tangential traction-separation laws for – M2 < M < where – M2 is a
b
;

lower bound value dependent on the materials (Freund and Suresh 2003). Ji et al. (Ji et al.

2011, Ji et al. 2012a, Ji et al. 2012b, Ji et al. 2013) have shown experimentally that fracture

occurs predominantly in mode–II as M → ^/2 and in mode–I as M → 0. Therefore, it is

proposed that the equivalent normal and shear traction-separation laws be obtained by a

straightforward interpolation of observed experimental results at M = 0, , and . Hence, the


b b
d ;

12
normal strength (
∗), normal initiation (# ) and normal critical (# L displacements in Eqs.

(5) are replaced by the following modified values

A ∗ = K
efee −
∗ LM +
∗ ,
d ∗
b

#̅ = ghKefeeL − # iM + # , and


d 
b
(11)

#̅ = ghKefeeL − # iM + # ,


d

b

where
efee

, hKefeeL

, and hKefeeL

are the mixed-mode-I/II normal strength, initiation and

critical displacements, respectively, to be obtained from the experimental results of (Ji et al.

2012b). In a similar vein, the shear traction separation parameters in Eq. (6) are replaced with

the following

7̅ ∗ = K7 ∗ − 7efee
∗ L
jM − k + 7efee
d b ∗
b d
,

8̅ = b g8 − lKefeeL i jM − k + lKefeeL


d  b 
d
, (12)

8̅ = b g8 − lKefeeL i jM − k + lKefeeL


d  b 
d
,

where 7efee

, lKefeeL

, and lKefeeL

are the mixed-mode (I-II) shear strength, initiation and

critical displacement, respectively. Eqs. (11) and (12) provide a simple approach to account

for the effects of phase angle on the fracture resistance of the interlaminar layer and it will be

shown to capture the mixed-mode fracture response fairly accurately.

It is worth noting that the phase angle for mixed-mode fractures may depend on the dynamic

loading rates as well. The rate dependency of the phase angle will be considered by the

authors in a forthcoming paper.

4.3. Damage criterion

To complete the CDM formulation, a damage initiation criterion must be specified. The

criterion used will depend on whether the interlaminar layer is initially intact or whether there

is a pre-existing crack of a structural length scale. A mixture of stress-based, mF , and fracture

mechanics-based, mn , damage criterion will be adopted here. If the interlaminar layer is

13
initially intact, then a stress-based damage initiation criterion o , based on a quadratic

interaction of the interlaminar stresses, is employed as follows


q== ; 7̂2; ; 7̂2= ;
o = mF = p ∗ r + p ∗ r + p ∗ r ≤ 1

7 7
(13)

where
q==, 7̂2; and 7̂2= are the normal (out-of-plane) and in-plane shear stresses, respectively.

For a partially fractured interlaminar layer, the stress-based criterion in Eq. (13) may lead to

numerical instabilities and/or difficulties associated with resolving the stress field near the

crack tip. An alternative criterion based on a quadratic interaction of the energy release rates

is used – this is similar to one proposed by Reeder (1992) – as follows

nt 2
ntt 2
nttt 2
nt ntt nt nttt ntt nttt
o = mn< = p r +p r +p r +p × r+p × r+p × r
5tu 5ttu 5tttu 5tu 5ttu 5tu 5tttu 5ttu 5tttu
(14)

≤w

where wK< 1L is a material-dependent parameter that characterises damage initiation; 56 ,

566 and 5666 are the critical energy release rates in mode I, II and III, respectively; and n

(i=I,II, and III) are energy release rates computed based on the traction-displacement

responses of the interlaminar medium given by

n6 = ∫
# dy==,

n66 = ∫ 7 dy2= , and (15)

n666 = ∫ 7 dy;=,

where y== , y;= and y2= are the normal and tangential displacements; and
# and 7 are the

traction-separation laws of Eqs. (5) and (6). The computational strategy will be described in

Section 6.

14
It is worthwhile noting that the main objective of introducing two separate damage initiation

and propagation criteria in this work is to minimize the stress singularity effects that may

have a profound effect on mesh dependency of the results. The damage mechanisms are

multiscale phenomena that start from micro-cracks/-voids and eventually they form

macroscale cracks upon their coalescence. A major challenge in FEA implementation of

CDM models is the element removal process in which elements with excessive damage are

removed from analysis. The element removal may result in stress singularity points, viz.

stress approaches to infinity upon mesh refinement steps. The transition between the stress,

i.e. Eq. (13), and fracture energy, i.e. Eq. (14), based criteria can be either smooth or sudden.

In the case of smooth transition, two damage thresholds may be utilized where at lower

damage levels, e.g. less than 0.05, the stress based criterion controls the damage

initiation/propagation while at higher levels of damages, e.g. higher than 0.25, the fracture

based criterion is used. In other words at lower damage levels, it is assumed that the softening

due to damage is negligible and the stress singularity effect is not dominant and the stress

based criterion can effectively capture the damage initiation and propagation. While in the

case of higher damage levels the fracture based criterion is utilized to capture the damage

initiation and propagation that is less sensitive to stress singularity effects. In the midrange

damage level, viz. 0.05 to 0.25, a mixed criterion can be designed to provide a smooth

transition between the two criteria. Obviously the lower and higher damage levels are

material parameters that can be estimated using bi-material elasticity problem solution in

which the elastic moduli difference is calculated for inducing a stress singularity at the

boundary, e.g. see (Williams, and Zak, 1962, Sator and Becker 2011).

5. Stress Triaxiality, Strain Rate, and Temperature Effects

Delamination in FRPs is affected by the physical and chemical composition of the distinct

phases, the operating temperature, chemical environment, rate of loading, rate of cracking,

and other factors. It is postulated that the dynamic energy density of a system has a significant

influence on the delamination process and is related to the level of stress triaxiality λ (∶=

15
Σ/|τ| where Σ is the hydrostatic part of the applied stress, and |7| is the effective shear stress).

A schematic showing how critical displacement varies with λ is given in Figure 3 (Bao and

Wierzbicki 2005, Shojaei et al. 2013). Different loading regimes can be defined depending on

λ as follows:

- Regime (I) - { < −{2 : High level of compressive hydrostatic stress prevents micro-flaws

from nucleating and growing in this regime.

- Regime (II) - −{2 < { < 0: The hydrostatic stress, whilst still negative, is sufficiently

reduced to allow microcracking. The dominant damage mechanism is shear-driven

microcracking with friction. The microcrack growth rate equation by Deng and Nemat-Nasser

(1992) and Shojaei et al. (2013) takes into account the effects of frictional sliding.

- Regime (III) - 0 < { < {; : The hydrostatic stress is tensile in this regime, allowing the

microcrack surfaces to separate and propagate with negligible sliding friction. Shear forces

remain the dominant mechanism that drives the damage process because of the weak shear

strength.

- Regime (IV) - {; < { < {= : High level of tensile hydrostatic stress facilitates void

nucleation and growth in this regime.

- Regime (V) - { > {= : Compressive stress wave is reflected from the free surface as a tensile

wave. If the magnitude of the reflected tensile wave is higher than the tensile strength of the

interlaminar layer, failure by spalling occurs as shown schematically in Fig. 3. Tensile

hydrostatic stress controls the damage process and appropriate equations of states must be

introduced. This regime of loading is not considered in the present paper.

16
Figure 3 Effect of the applied dynamic energy density (stress triaxility level) on the critical failure

displacement.

The effects of stress triaxiality are incorporated into the cohesive laws by modifying the

critical normal #̅ and shear 8̅ displacements in Eqs. (11) and (12). In regimes (II) and

(III), the new critical normal #̿ and shear 8̿  displacements, double over-bar denotes post

adjustment for the effects of stress triaxiality, are given by

#̿ = #̅ 1 − ln j k‚ for −|{2 | < { < {;


€|€. |
€:|€.|
(16)

where subscript “#” is either “n” or “t”. If stress triaxiality falls between {; < { < {ƒ„ , a

linear correlation is adopted as follows

#̿ = #̅ K1 − L for {; < { < {ƒ„


€f€:
€ …†‡ˆˆ f€:
(17)

where “#” is the same as above. The proposed phenomenological approach enables dynamic

energy density to be incorporated into the delamination model and it will be shown later to

compare well with experimental data. From the mathematical point of view, Eqs. (16) and

(17) result in very high critical strain (no failure) in the case of { → −|{2 | and the critical

strain is reduced by marching from −|{2 | to {ƒ„ (resulting in higher damage levels). For

17
more experimental justifications, the interested reader may refer to Johnson and Cook (1985)

and Steinberg (1996).

Strain rate can have a significant influence on the ultimate tensile and shear stresses (Huang et

al. 2009, Amini et al. 2010). Likewise, elevated temperature may also give rise to material

instability; in particular, under dynamic loading of high dynamic energy content (Deng and

Nemat-Nasser 1992). The effects of strain rate and elevated temperature are accounted for by

adjusting the interfacial mechanical strength in Eqs. (11) and (12). Following (Johnson and

Cook 1985), the rate-dependent interfacial strengths are given by

‰ ∗ =
A∗ × 1 + Š log j k‚ × Ž1 −  ∗  ‘
|Π|
Œ 

7̿ = 7̅∗ × 1 + Š log j Œ k‚ × Ž1 −  ∗ ‘
|Π|
(18)



where || = [   is the applied strain rate;  is a reference strain rate;
A∗ and 7̅∗ are,
;
=

respectively, the normal and shear interface strengths measured at reference strain rate  ; T*

is the homologous temperature to be defined in Eq. (20); and C and “ are material parameters

widely available in the literature (Johnson and Cook 1985, Steinberg 1996). One may note

that Johnson-Cook model has originally been developed for metals, and this model

provides an efficient way for including the rate and temperature effects into the

constitutive modeling framework. In this work the performance of the JC model has

experimentally investigated and it is shown that this model can be used for capturing

rate dependent cohesive strength of FRPs. In a similar vein, the quasi-static normal and

”# and shear %
shear stiffness of the interface must also be adjusted to dynamic normal % ”8

stiffness as follows

”# = %
% •# × 1 + Š ′ log j|Œ |k‚ × 1 −  ∗  ‚
Œ 

(19)

18
”8 = %
% •8 × 1 + Š ′ log j|Œ |k‚ × 1 −  ∗  ‚
Œ

•# =
A∗ /#̅  and %
where % •8 = 7̅∗ /8̅  are, respectively, the reference normal and shear

stiffness measured at the reference strain rate  . The homologous temperature, T*, in Eqs.

(18) and (19) is defined as follows (Johnson and Cook 1985)

0 m™š  < –„#


 = $–
∗ –f–J‡G
0 m™š 8„# <  <  8
—˜ˆJ f–J‡G
1 m™š  >  8
(20)

where  (K) is the current temperature, 8„# (K) is the transition temperature and  8 is

the melting temperature. Extensive testing is needed to accurately establish the coefficients Š ′

and “′ for various delamination scenarios. Because experimental data regarding temperature

effect is not currently available, only the strain rate effect is considered in this work. The

temperature/strain rate effect on the delamination process is currently under experimental

investigation by the authors and the results will be reported in a forthcoming paper.

6. Implementation Aspects

Figure 4 shows a flow chart of the numerical implementation into a commercial FEA package

(ABAQUS) through user-defined field variables (VUSDFLD), and material (VUMAT)

subroutines. The dynamic implicit integration scheme, see (ABAQUS, 2011), with refined

mesh, together with enhanced hourglass meshes, are utilized. The developed damage

variables are updated incrementally at each material point as field variables and the updated

values are utilized to reduce the elastic tensile and shear moduli. Once the damage parameter

reaches to its critical value in one of the integration points, the corresponding stiffness is set

to a small value to artificially remove the respective element from the model. This element

deletion technique keeps the mass of the system unchanged. As discussed by Lemaitre and

Dufailly (1987) the critical elastic modulus should be defined based on the material system.

19
In dynamic problems with progressive damage, stiffness reduction of elements (due to the

damage) may result in strain-softening behavior, leading to strain localization, which is

known as ill-posed constitutive relations (de Borst and Sluys 1991, Sluys and de Borst 1994,

Shojaei et al. 2013). This latter phenomenon results in strong mesh dependencies in the FEA

results where the dissipated energies decrease with repeated mesh refinement. There are a few

approaches, proposed in the literature, to avoid pathological mesh dependency of the CDM

results, such as non-local continua, see (Shojaei et al., 2013) for more details. In the case of

non-local methods, a gradient enhanced theory is required to remove the ill-posedness of the

problem via integration schemes that involve characteristic lengths, see (Voyiadjis et al.

2014) for example. Due to the fact that the proposed CDM model is a local theory, an

alternative approach is adopted to compensate for the ill-posedness of the boundary value

problem, where the softening part of the CDM constitutive law is defined based upon a stress-

displacement relation. The energy dissipated during the damage process is specified per unit

area, not per unit volume. Thus, the damage dissipation energies (consistent with critical

energy release rate parameters in fracture mechanics) are used to compute the displacement at

which full material damage occurs. This formulation ensures that the correct amount of

energy is dissipated and alleviates the mesh dependency (ABAQUS 1011). Consequently, the

constitutive law is expressed as a stress-displacement relation where the displacements, see

y== , y2= and y;= in Eq. (15), are computed based on energy descriptions, i.e. damage

dissipation energy, instead of the ill-posed stress-strain constitutive relations. Appendix A

discusses more in detail the computational difficulties facing ill-posed constitutive relations.

The mesh convergence study is provided in Appendix B to show the efficiency of the

proposed approach for alleviating the mesh dependency of the results.

20
Figure 4 FEA implementation flowchart for the delamination damage model

7. Results and Discussions

Experimental investigations into Mode-I delamination were carried out by (de Morais et al.

2002, Ji et al. 2010, Ji et al. 2013) using Double Cantilever Beam (DCB) specimens. End-

Notched Flexure (ENF) specimens were used to investigate the mode-II interlaminar fractures

by (Pereira et al. 2004, Ji et al. 2011, Ouyang et al. 2011, Ji et al. 2012a). Mode-III

delamination was studied using Edge Crack Torsion (ECT) samples for carbon/epoxy

laminates by (de Morais et al. 2009). Mixed mode I/II fracture for single-leg bending (SLB)

specimens was studied by (Ouyang et al. 2011, Ji et al. 2012a). In this section, the

performance of the developed delamination model is compared to the experimental data

previously reported by (Ji et al. 2011, Ji et al. 2012a, Ji et al. 2012b, Ji et al. 2013).

The adhesive, LOCTITE Hysol 9460, that was used has a high peel strength, good impact

resistance, and good fatigue resistance. According to the manufacturer the adhesive agent has

an elastic modulus of 2.76 GPa, tensile strength of 30.3 MPa and failure elongation of 3.5%.

Prefabricated plain woven glass fabric reinforced epoxy laminates are used to manufacture

the Double Cantilever Beam (DCB) specimen for mode-I experiments, and the End Notched

Flexure (ENF) specimen for mode-II experiments. Single Leg Bending (SLB) samples made

from steel plates bonded with Hysol 9460 were used for mixed mode-I/II tests. The DCB

21
specimen is made of two 3.1 mm thick slabs, with a width and length of 25.4 mm and 254.0

mm, respectively, and an initial crack length of 25mm. Each slab is made of 17 layers of

continuous glass woven fabric with epoxy resin (Ji et al. 2013). The ENF specimens, made of

the same material system, are 20mm wide and 150mm long, with a total thickness of 3 mm

for each laminate and an initial crack length of 35 mm (Ji et al. 2011). Table 1 summarizes

the material properties of the DCB, and ENF samples. For the SLB samples, the steel plate

used for the upper adherent is 6.35 mm thick, 25.4 mm in width and has a length of 304.80

mm; whilst the lower adherent has a length of 279.40 mm (Ji et al. 2012b). For each test,

performed on an MTS 810 machine, the local deformations are captured using a Sony XCD-

CR90 high resolution CCD camera. The test data will be used to validate the results of the

current numerical simulations.

Table 2 lists the material parameters for the shear and normal traction-separation relationships

used in different mode-mixity scenarios. It is worth noting that only two parameters 12 and 1;

are obtained from numerical curve fitting techniques. Fig. 5 shows a 3D representation of

how the shear traction separation relationship varies with phase angle. The experimental data

for the ENF specimens (M = 90œ ) and for the SLB specimens (M = 45œ ) are plotted in the

same figure for comparison.

Table 1 Material parameters for epoxy laminates used in DCB and ENF tests

E11 E22=E33 G12=G13 GIc GIIc


v12=v13
(GPa) (GPa) (GPa) (kJ.m-2 ) (kJ.m-2)
18.6 15.17 7.38 0.29 0.268 0.4747

Table 2 Traction-separation material parameters obtained from DCB, ENF and SLB

ž Ÿ ž ∗
Ÿ∗ f   ¡¢ ¡£ ¡¢K f  L
experiments
∗ ∗
 f  
(DCB) (ENF) (SLB) (SLB) (DCB) (ENF) (SLB)
(MPa) (MPa) (MPa) (MPa) (µm) (µm) (µm)

¡£K f  L ¡¤¥ ¡¢ ¡¤¥ ¡¤¥ ¦§ * ¦¨ *


22 8.5 28 11 35 310 75
£ ¢ £
(SLB) (ENF) (DCB) (DCB) (ENF) (Curve (Curve
(µm) (µm) (µm) (µm) (µm) Fitting) Fitting)
60 600 35 90 600 1.5 1.5
* non-linearity exponent

22
Figure 6 shows a 3D representation for the normal traction-separation responses for various

phase angles. The DCB data is for M = 0œ and SLB test data is for M = 45œ . Good

agreements between simulations and experiments are observed for both normal and shear

traction-separations. The effects of stress triaxiality on the critical normal and shear

displacements are given in Fig. 7. Simulation results in this graph confirm experimental

observations that high negative stress triaxialities delay failures while highly positive stress

triaxialities result in accelerated failure mechanisms (Bao and Wierzbicki 2004). According to

the experimental observation the following values for the boundaries between different

regimes in Fig. 3 are chosen: {2 = −1/3, {; = 2 and {ƒ„ = 10, see for example (Bao and

Wierzbicki 2004). One may notice that {= is ignored herein and a linear correlation is used in

Eq. (17) from {; to {ƒ„ . Low to medium range strain rate tests were carried out to capture

normal (DCB test) and shear (ENF test) interfacial strengths. As shown in Fig. 8, the strain

rate dependency of the interfacial strengths can be effectively captured using Eq. (18). The

reference strain rate here is set to  = 1 sec-1 and


∗ = 35.75 MPa and 7∗ = 11.1 MPa are

measured at this strain rate. The effect of temperature rise is not considered in the current

simulations, as it requires extensive experimental studies. The simulation results for the

normal = and shear 2= (=;= ) damage parameters are given in Figs. (9) and (10),

respectively, which shows how the extent of delamination damage varies with phase angle.

For sake of clarity different colors are used in Figs. 5, 6, 9, 10 in which the meaning of each

color is discussed in the respective figure captions.

23
Figure 5 Experimental and simulation results for shear traction-separation in which ENF and

SLB test data are compared to the simulated results. Experimental data are from (Ji et al. 2011,

Ji et al. 2012b). Colors black and blue show the simulation results for the mixed-mode and mode-

II, respectively. Other colors show the transition between mode-II and mixed-mode.

Figure 6 Experimental and simulation results for normal traction-separation in which DCB and

SLB test data are compared to simulated results. Experiment data are from (Ji et al. 2012b, Ji et

al. 2013). Colors green and blue show the simulation results for the mode-I and mixed-mode,

respectively. Color spectrum illustrates the transition between mode-I and mode-II.

24
To investigate the performance of the developed framework for dynamic fracture of the

laminated FRPs, the delamination CDM model is introduced into a commercial FEA package,

ABAQUS, through user-defined subroutines VUMAT. At first stage, the DCB and ENF tests

are modeled to compare the FEA and experimental outputs for the sake of validation and

conducting the mesh convergence studies. The dimensions of the DCB and ENF samples

have already been mentioned, and an interlaminar layer with thickness of 0.5 mm is

considered for both DCB and ENF models. The C3D8R reduced integration elements with

enhanced hourglass are used in both models. Plies are meshed by the same element type and

as mentioned before they are modeled as elastic medium for sake of simplicity (see paragraph

before Eq. (1) for more details). Fig. 11 compares the traction separation response of DCB

and ENF models to their respective experimental data. It is worth noting that the FEA

example herein provides a performance demonstration for the proposed CDM model. It is

now well-understood that eight-node, one-point quadrature element, e.g. C3D8R, which has

reduced integration and hourglass control, has very poor behavior in bending-dominated

problems. The interested reader may refer to (Fish and Belytschko, 2007) for more details on

alternative enhanced finite element formulations which are highly recommended to use in

complex problems involving bending.

Figure 7 Effect of stress triaxiality on the dynamic delamination response of the material system

25
Figure 8 Comparison of experimental data and results of numerical simulations for the

variations of the normal and shear interfacial strengths with strain rate

Figure 9 Normal damage parameter for various phase angles. The color spectrum shows the

transition from mode-I to mode-II.

Figure 10 Shear damage parameter for various phase angles. The color spectrum shows the

transition from mode-I to mode-II.

26
Figure 11 Experimental and FEA results for (a) DCB, (b) ENF tests. The experimental setups are

shown in red boxes in each graph, refer to (Ji et al. 2013) for details of experiments.

The performance of the developed scheme in the case of a dynamic loading condition is

investigated next. A laminated FRP under projectile impact is studied. The same material

properties, considered for DCB and ENF models, are assumed for the plies and interface. The

simulation result of progressive delamination damage in a DCB specimen is depicted in Fig.

12 in which the numerically measured load (F) and experimentally characterized delamination

lengths (Lf) correlate well. A special orthotropic laminate configuration is constructed as

shown in Fig. 13. Each laminated substrate, consists of 20 plies, is 2 mm thick and

interlaminar layers with 0.5 mm thickness are considered between them. The laminated FRP

is assumed to be clamped all around external edges. The element type and size should be

chosen in accordance to the mesh convergence studies for the DCB and ENF models. The

mesh convergence study has been conducted for the DCB sample in Appendix B and the

element size related to the converged results are utilized for the case of projectile impact

simulation. A rigid spherical projectile with 10 mm radius and 0.2 kg mass is impacted onto

the center of the system with initial velocity of 100 m/s. The compressive and tensile wave

propagations in the system are studied and the dynamic delamination is investigated in each

interface. Fig. 14(b) to 14(d) illustrate stress wave propagations across the thickness and

transverse directions. General contact between rigid spherical surface of the projectile and the

composite plate is prescribed where frictionless tangential and hard normal (no penetration)

contact properties are assigned to the model. In the case of Explicit Integration methods in

27
which the displacement equilibrium solution at time “t + ∆t” is based on using the conditions

of equilibrium at time “t”; a stable integration requires that ∆t is smaller than a critical value

∆t ¬Q . In the case of ∆t > ­t ¬Q the integration is unstable that means errors resulting from the

numerical integration or round off are too high that affects the response calculations. The

critical time step ∆t ¬Q is correlated to the mass, stiffness and damping properties of the

structure. An approximation to the stability limit can be expressed in terms of the smallest

transit time of a dilatational wave across any of the elements in the mesh ∆t ¬Q ≅ D°±h /c²

where D°±h is the smallest element dimension and c² is the dilatational wave speed.

Figure 12 DCB simulation results (a) von-Mises stress distribution, and (b) Progressive

delamination damage in DCB specimen. The Load (F) -Delamination lengths (Lf) data correlate

well with the observed experiments. The color spectrum shown represents various damaged

states in which the red color indicates fully damaged (damage variable=1.0) and the blue color

shows the non-damaged state (damage variable=0.0).

28
Figure 13 Schematic for a special orthotropic composite lay-up configuration with the

interlaminar layer shown by solid lines between plies

Figure 14 A laminate impacted by a rigid spherical projectile with V0=100 (m/s), (a) schematic of

a quarter (X and Y cut planes are shown) of the FEA model and boundary conditions (B.C.), (b)

reflection of the projectile after impacting the laminate, (c) propagation of stress waves at time

step 1e-4, and (d) propagation of compressive wave through the thickness at time step 1e-3 .

Fig. 15(a) shows the stress triaxiality distribution on the center line, see Fig. 14(a), of the

laminate for different time steps. The local coordinate system (1,2,3), shown in Fig. 14(b), is

used to plot Fig. 15. It is worth noting that upon incidence of the compressive and tensile

29
waves with interlaminar layers, a portion of the wave is transmitted to the next ply while

another portion is reflected back. The noises in the stress triaxiality distributions in Fig. 15(a)

are due to these interactions. Fig. 15(b) depicts the state of the normal damage distribution in

the interlaminar layers. It is obvious from Fig. 15(b) that due to the compressive wave

propagation the normal delamination damage is quite small in interlaminar layers near the

impacted face. Upon reflection of the compressive wave from the free surface (bottom of the

laminate), the effect of normal delamination becomes more significant and results in mode-I

dominated delamination of plies near the free edges. Fig. 15(c) depicts the state of the shear

damage for different interlaminar layers. One may argue that the thickness of the interlaminar

layer in laminated FRPs might be less than 0.5 mm. The effect of the interlaminar layer

thickness on the delamination responses can be taken into account through sensitivity

analysis.

Figure 15 (a) Distribution of the stress triaxiality across the thickness of the laminate on the

center line for various step times, (b) normal damage distribution in different interlaminar

layers, and (c) shear damage distribution in different interlaminar layers.

30
The actual strain hardening/softening rate dependency in dynamic crack growth problems

occurs within the fracture process zone, where crack propagation velocity governs the

localized strain rates at the crack tip. Thus, it is instructive to discuss the relationship between

crack opening/sliding displacement rates, local strain rate ̅ at the crack tip within the

Embedded Process Zone (EPZ), which is depicted in Fig. 2b, and is called EPZ strain rate

hereinafter; and macroscale strain rate  which occurs outside the EPZ zone, e.g. in adjacent

plies. The macroscopic strain rate  , which is governed by continuum deformation

mechanisms in the medium, may remain finite outside the EPZ; although, the localized EPZ

strain rate ̅ may exceed 1000 s-1 due to the speed of crack advance (Glennie 1971, Glennie

1971a). The crack propagation may reach velocities in the order of 50 to 300 m.s-1 in the case

of “quasi-static” crack growth. The crack opening/sliding displacement can be correlated to

the EPZ strain rate via dilatational u2 and shear u; stress wave speeds in the medium

(Glennie 1971, Glennie 1971a). One possible correlation between normal # and sliding

8 displacement rates and the EPZ strain rate is as follows

2 2
|̅| ≅ ³´ j Gk + p1 − r j1 − | | k j Jk
| {|  |{| |{| 
{\µ¶·· u1 {\µ¶·· {1 u2
(21)

where ³ is a constant to be obtained experimentally. Eq. (21) implies that |̅| depends on the

dynamic energy density content of the loading. In the case of shear-dominated loading, i.e.

|{| → 0, the shear wave dominates the deformation mechanism in the interface and the shear

stress wave governs EPZ strain rate. For highly compressive stresses |€| → 1, and →1
|€| |€|
. €…†‡ˆˆ

for tensile stresses. Thus, the dilatational waves and normal crack opening rates affect the

EPZ strain rate. In an isotropic medium, the longitudinal and shear wave speeds can be

estimated through relations between elastic modulus, Poisson’s ratio and density of the

system (Glennie 1971, Glennie 1971a). Bridging the decohesion mechanisms in microscale

EPZ and macroscopic constitutive behavior of laminated composites is the subject of a

current investigation and the results will be reported elsewhere.

31
One may note that only delamination damage is considered in this work; while matrix

cracking, other types of damage and/or plastic deformation of plies may have a profound

effect on integrity of the system. For a full-scale health monitoring of the laminated FRP, the

present delamination model needs to be coupled with proper descriptions of plasticity and

damage of the plies. The plastic and damage processes in plies influence the delamination

damage mechanisms, and different damage distributions may be expected compared to that

shown in (15). This is the subject of the forthcoming paper by the authors where the matrix

cracking, fiber/matrix debonding and matrix plasticity are also considered in addition to

delamination damage to monitor the integrity of the laminated FRP under dynamic loading

conditions.

8. Conclusion

Dynamic delamination in FRPs is investigated in this work within the CDM framework.

Mechanics of decohesion is incorporated to provide a physically consistent CDM

delamination model for FRPs. Particularly the effects of the mixed-mode fractures and

dynamic energy density on the delamination responses are considered. The strain rate effects

on the strength and stiffness of the interlaminar layers are also studied in which empirical

relations are established for dynamic stiffness and strength of the interfaces. Comparing to the

existing cohesive models in the literature, the framework is developed in such a way that most

of the material parameters are directly obtained from the experimental testing that minimizes

the numerical curve fitting tasks. The developed CDM framework may be considered as an

alternative to the classical cohesive element in Finite Element Analysis (FEA) problems. The

fracture path in the developed CDM approach evolves naturally and conventional mesh types

can be utilized to mesh the FEA model; whilst in the case of discrete cohesive elements the

fracture paths need to be predefined to use specific types of cohesive element in that region.

In other words, in the case of discrete cohesive elements the fracture paths are always

predefined by the shape of two mating surfaces in which a single cohesive element is placed

between those mating surfaces. Consequently, the fracture path is always the same as the

32
shape of two bonding surfaces. In the case of CDM, multiple elements through the thickness

are allowed that result in natural fracture path evolution based on dissipative energy release

rates. Thus, the delamination fracture path in CDM models is not anymore dependent on the

shape of the two adjacent layup surfaces. One may note that still material properties of the

layups and interlaminar layers are assigned separately. Introduction of the developed scheme

into a commercial FEA package, through user-defined subroutines, provide a powerful design

tool to analyze the delamination in FRPs. The performance of the proposed scheme is shown

to compare well with existing experimental data in the literature, which demonstrates that the

proposed CDM model is capable of capturing a wide range of delamination scenarios.

Acknowledgment

This investigation was partially supported by a Cooperative Agreement NNX11AM17A

between NASA and the Louisiana Board of Regents under contract NASA/LEQSF (2011-

14)-Phase3-05. This study was also partially supported by the NSF, under grant number

CMMI0900064, and Army Research Office, under grant number W911NF-13-1-0145.

REFERENCES

Ahmed, A. and L. J. Sluys (2014). "A phantom node formulation for modeling
coupled adiabatic–isothermal cracking in FRP composites." Computer Methods in
Applied Mechanics and Engineering 278(0): 291-313.

Alfano, G. and M. A. Crisfield (2001). "Finite element interface models for the
delamination analysis of laminated composites: mechanical and computational
issues." International Journal for Numerical Methods in Engineering 50(7): 1701-
1736.

Allix, O. and P. Ladevèze (1992). "Interlaminar interface modelling for the prediction
of delamination." Composite Structures 22(4): 235-242.

Amini, M. R., J. Simon and S. Nemat-Nasser (2010). "Numerical modeling of effect


of polyurea on response of steel plates to impulsive loads in direct pressure-pulse
experiments." Mechanics of Materials 42(6): 615-627.

Azizi, R., C. F. Niordson and B. N. Legarth (2011). "Size-effects on yield surfaces for
micro reinforced composites." International Journal of Plasticity 27(11): 1817-1832.

Bao, Y. and T. Wierzbicki (2004). "On fracture locus in the equivalent strain and
stress triaxiality space." International Journal of Mechanical Sciences 46(1): 81-98.

33
Bao, Y. and T. Wierzbicki (2005). "On the cut-off value of negative triaxiality for
fracture." Engineering Fracture Mechanics 72(7): 1049-1069.

Barenblatt, G. I. (1959). "The formation of equilibrium cracks during brittle fracture:


general ideas and hypothesis, axially symmetric cracks." Applied Mathematics and
Mechanics 23: 622–636.

Benzeggagh, M. L. and M. Kenane (1996). "Measurement of mixed-mode


delamination fracture toughness of unidirectional glass/epoxy composites with mixed-
mode bending apparatus." Composites Science and Technology 56(4): 439-449.

Bouvard, J. L., J. L. Chaboche, F. Feyel and F. Gallerneau (2009). "A cohesive zone
model for fatigue and creep–fatigue crack growth in single crystal superalloys."
International Journal of Fatigue 31(5): 868-879.

Brassart, L., L. Stainier, I. Doghri and L. Delannay (2012). "Homogenization of


elasto-(visco) plastic composites based on an incremental variational principle."
International Journal of Plasticity 36(0): 86-112.

Brunner, A. J. (2000). "Experimental aspects of Mode I and Mode II fracture


toughness testing of fibre-reinforced polymer-matrix composites." Computer Methods
in Applied Mechanics and Engineering 185(2–4): 161-172.

Camanho, P. P., C. G. Davila and M. F. de Moura (2003). "Numerical Simulation of


Mixed-Mode Progressive Delamination in Composite Materials." Journal of
Composite Materials 37(16): 1415-1438.

Caner, F. C. and Z. P. Bažant (2009). "Size effect on strength of laminate-foam


sandwich plates: Finite element analysis with interface fracture." Composites Part B:
Engineering 40(5): 337-348.

Chaboche, J. L., S. Kruch, J. F. Maire and T. Pottier (2001). "Towards a


micromechanics based inelastic and damage modeling of composites." International
Journal of Plasticity 17(4): 411-439.

Chen, M., A. J. Crisfield, E. P. Kinloch, F. L. Busso, Y. Matthews and J. Qiu (1999).


"Predicting Progressive Delamination of Composite Material Specimens via Interface
Elements." Mechanics of Composite Materials and Structures 6(4): 301-317.

Chen, Y. and S. Ghosh (2012). "Micromechanical analysis of strain rate-dependent


deformation and failure in composite microstructures under dynamic loading
conditions." International Journal of Plasticity 32–33(0): 218-247.

Coker, D., A. J. Rosakis and A. Needleman (2003). "Dynamic crack growth along a
polymer composite–Homalite interface." Journal of the Mechanics and Physics of
Solids 51(3): 425-460.

34
Corigliano, A. (1993). "Formulation, identification and use of interface models in the
numerical analysis of composite delamination." International Journal of Solids and
Structures 30(20): 2779-2811.

Corigliano, A. and O. Allix (2000). "Some aspects of interlaminar degradation in


composites." Computer Methods in Applied Mechanics and Engineering 185(2–4):
203-224.

Crowell, M. W., T. A. Schaedler, B. H. Hazel, D. G. Konitzer, R. M. McMeeking and


A. G. Evans (2012). "Experiments and numerical simulations of single particle
foreign object damage-like impacts of thermal barrier coatings." International Journal
of Impact Engineering 48(0): 116-124.

de Borst, R., J. Pamin and M. G. D. Geers (1999). "On coupled gradient-dependent


plasticity and damage theories with a view to localization analysis." European Journal
of Mechanics - A/Solids 18(6): 939-962.

de Borst, R. and L. J. Sluys (1991). "Localisation in a Cosserat continuum under static


and dynamic loading conditions." Computer Methods in Applied Mechanics and
Engineering 90(1-3): 805-827.

de Morais, A. B., M. F. de Moura, A. T. Marques and P. T. de Castro (2002). "Mode-I


interlaminar fracture of carbon/epoxy cross-ply composites." Composites Science and
Technology 62(5): 679-686.

de Morais, A. B., A. B. Pereira, M. F. S. F. de Moura and A. G. Magalhães (2009).


"Mode III interlaminar fracture of carbon/epoxy laminates using the edge crack
torsion (ECT) test." Composites Science and Technology 69(5): 670-676.

De Moura, M. F. S. F., J. P. M. Gonçalves, A. T. Marques and P. M. S. T. De Castro


(1997). "Modeling Compression Failure after Low Velocity Impact on Laminated
Composites Using Interface Elements." Journal of Composite Materials 31(15): 1462-
1479.

Deng, H. and S. Nemat-Nasser (1992). "Dynamic damage evolution in brittle solids."


Mechanics of Materials 14(2): 83-103.

Dey, S., T. Børvik, X. Teng, T. Wierzbicki and O. S. Hopperstad (2007). "On the
ballistic resistance of double-layered steel plates: An experimental and numerical
investigation." International Journal of Solids and Structures 44(20): 6701-6723.

Doghri, I., L. Brassart, L. Adam and J. S. Gérard (2011). "A second-moment


incremental formulation for the mean-field homogenization of elasto-plastic
composites." International Journal of Plasticity 27(3): 352-371.

Eftis, J., C. Carrasco and R. A. Osegueda (2003). "A constitutive-microdamage model


to simulate hypervelocity projectile-target impact, material damage and fracture."
International Journal of Plasticity 19(9): 1321-1354.

35
Eftis, J. and J. A. Nemes (1991). "Evolution equation for the void volume growth rate
in a viscoplastic-damage constitutive model." International Journal of Plasticity 7(4):
275-293.

Eftis, J. and J. A. Nemes (1996). "On the propagation of elastic-viscoplastic waves in


damage-softened polycrystalline materials." International Journal of Plasticity 12(8):
1005-1022.

Espinosa, H. D., S. Dwivedi and H. C. Lu (2000). "Modeling impact induced


delamination of woven fiber reinforced composites with contact/cohesive laws."
Computer Methods in Applied Mechanics and Engineering 183(3–4): 259-290.

Fish, J. (2013). Practical Multiscaling. United Kingdom, John Wiley & Sons.

Freund, L. B. and J. W. Hutchinson (1985). "High strain-rate crack growth in rate-


dependent plastic solids." Journal of the Mechanics and Physics of Solids 33(2): 169-
191.

Freund, L. B. and S. Suresh (2003). Thin Film Materials Cambridge

Glennie, E. B. (1971). "A strain-rate dependent crack model." Journal of the


Mechanics and Physics of Solids 19(5): 255-272.

Glennie, E. B. (1971a). "The unsteady motion of a rate-dependent crack model."


Journal of the Mechanics and Physics of Solids 19(6): 329-338.

Grégoire, D., H. Maigre, J. Réthoré and A. Combescure (2007). "Dynamic crack


propagation under mixed-mode loading – Comparison between experiments and X-
FEM simulations." International Journal of Solids and Structures 44(20): 6517-6534.

Haddag, B., F. Abed-Meraim and T. Balan (2009). "Strain localization analysis using
a large deformation anisotropic elastic–plastic model coupled with damage."
International Journal of Plasticity 25(10): 1970-1996.

Haddag, B., T. Balan and F. Abed-Meraim (2007). "Investigation of advanced strain-


path dependent material models for sheet metal forming simulations." International
Journal of Plasticity 23(6): 951-979.

Hashagen, F. and R. de Borst (2000). "Numerical assessment of delamination in fibre


metal laminates." Computer Methods in Applied Mechanics and Engineering 185(2–
4): 141-159.

Hooputra, H., H. Gese, H. Dell and H. Werner (2004). "A comprehensive failure
model for crashworthiness simulation of aluminium extrusions." International Journal
of Crashworthiness 9(5): 449-464.

Horstemeyer, M. F. and D. J. Bammann (2010). "Historical review of internal state


variable theory for inelasticity." International Journal of Plasticity 26(9): 1310-1334.

36
Huang, Y., D. Meyer and S. Nemat-Nasser (2009). "Damage detection with spatially
distributed 2D Continuous Wavelet Transform." Mechanics of Materials 41(10):
1096-1107.

Hutchinson, J. W. (1982). CRACK-TIP SINGULARITY FIELDS IN NONLINEAR


FRACTURE MECHANICS: A SURVEY OF CURRENT STATUS. Advances in
Fracture Research. D. Francois, Pergamon: 2669-2684.

Hutchinson, J. W. and A. G. Evans (2000). "Mechanics of materials: top-down


approaches to fracture." Acta Materialia 48(1): 125-135.

Jain, J. R. and S. Ghosh (2008). "Homogenization Based 3D Continuum Damage


Mechanics Model for Composites Undergoing Microstructural Debonding." Journal
of Applied Mechanics 75(3): 031011-031011.

Ji, G., Z. Ouyang and G. Li (2011). "Effects of bondline thickness on Mode-II


interfacial laws of bonded laminated composite plate." International Journal of
Fracture 168(2): 197-207.

Ji, G., Z. Ouyang and G. Li (2012a). "Local Interface Shear Fracture of Bonded Steel
Joints with Various Bondline Thicknesses." Experimental Mechanics 52(5): 481-491.

Ji, G., Z. Ouyang and G. Li (2012b). "On the interfacial constitutive laws of mixed
mode fracture with various adhesive thicknesses." Mechanics of Materials 47(0): 24-
32.

Ji, G., Z. Ouyang and G. Li (2013). "Effects of bondline thickness on Mode-I


nonlinear interfacial fracture of laminated composites: An experimental study."
Composites Part B: Engineering 47(0): 1-7.

Ji, G., Z. Ouyang, G. Li, S. I. Ibekwe and S. S. Pang (2010). "Effects of Adhesive
Thickness on Global & Local Mode-I Interfacial Fracture of Bonded Joints."
International Journal of Solids and Structures 47: 2445-2458.

Johnson, G. R. and W. H. Cook (1985). "Fracture characteristics of three metals


subjected to various strains, strain rates, temperatures and pressures." Engineering
Fracture Mechanics 21(1): 31-48.

Kachanov, L. M. (1958). "Rupture time under creep conditions." Izvestija Academii


Nauk SSSR 8: 26-31 (Reprinted in International Journal of Fracture, 97, 11-18).

Kenane, M. and M. L. Benzeggagh (1997). "Mixed-mode delamination fracture


toughness of unidirectional glass/epoxy composites under fatigue loading."
Composites Science and Technology 57(5): 597-605.

Khan, A. S. and B. Farrokh (2006). "Thermo-mechanical response of nylon 101 under


uniaxial and multi-axial loadings: Part I, Experimental results over wide ranges of
temperatures and strain rates." International Journal of Plasticity 22(8): 1506-1529.

37
Khan, S. M. A. and M. K. Khraisheh (2004). "A new criterion for mixed mode
fracture initiation based on the crack tip plastic core region." International Journal of
Plasticity 20(1): 55-84.

Kruch, S. and J.-L. Chaboche (2011). "Multi-scale analysis in elasto-viscoplasticity


coupled with damage." International Journal of Plasticity 27(12): 2026-2039.

Krueger, R., P. J. Minguet and T. K. O’Brien (1999). A Method for Cal-culating


Strain Energy Release Rates in Preliminary Design of Compos-ite Skin/Stringer
Debonding Under Multi-Axial Loading. NASA TM-1999-209365.

Kubair, D. V. and P. H. Geubelle (2003). "Comparative analysis of extrinsic and


intrinsic cohesive models of dynamic fracture." International Journal of Solids and
Structures 40(15): 3853-3868.

Landis, C. M., T. Pardoen and J. W. Hutchinson (2000). "Crack velocity dependent


toughness in rate dependent materials." Mechanics of Materials 32(11): 663-678.

Le Quang, H. and Q. C. He (2008). "Effective pressure-sensitive elastoplastic


behavior of particle-reinforced composites and porous media under isotropic loading."
International Journal of Plasticity 24(2): 343-370.

Lemaitre, J. and J. Dufailly (1987). "Damage measurements." Engineering Fracture


Mechanics 28(5-6): 643-661.

Li, G. and A. Shojaei (2012). "A viscoplastic theory of shape memory polymer fibres
with application to self-healing materials." Proceedings of the Royal Society A:
Mathematical, Physical and Engineering Sciences 468(2144): 2319-2346.

Li, H. and N. Chandra (2003). "Analysis of crack growth and crack-tip plasticity in
ductile materials using cohesive zone models." International Journal of Plasticity
19(6): 849-882.

Li, S., S. R. Reid and P. D. Soden (1998). "A continuum damage model for transverse
matrix cracking in laminated fibre–reinforced composites." Philosophical
Transactions of the Royal Society of London. Series A: Mathematical, Physical and
Engineering Sciences 356(1746): 2379-2412.

Li, S., S. R. Reid and Z. Zou (2006). "Modelling damage of multiple delaminations
and transverse matrix cracking in laminated composites due to low velocity lateral
impact." Composites Science and Technology 66(6): 827-836.

Liu, J., J. Li and B. Wu (2013). "The Cohesive Zone Model for Fatigue Crack
Growth." Advances in Mechanical Engineering 2013: 16.

Maimi, P., J. A. Mayugo and P. P. Camanho (2008). "A Three-dimensional Damage


Model for Transversely Isotropic Composite Laminates." Journal of Composite
Materials 42(25): 2717-2745.

38
Mariani, S. and A. Corigliano (2005). "Impact induced composite delamination: state
and parameter identification via joint and dual extended Kalman filters." Computer
Methods in Applied Mechanics and Engineering 194(50–52): 5242-5272.

Murakami, S. (1988). "Mechanical Modeling of Material Damage." Journal of


Applied Mechanics 55(2): 280-286.

Needleman, A. (1987a). "A Continuum Model for Void Nucleation by Inclusion


Debonding." Journal of Applied Mechanics 54(3): 525-531.

Needleman, A. and V. Tvergaard (1987b). "An analysis of ductile rupture modes at a


crack tip." Journal of the Mechanics and Physics of Solids 35(2): 151-183.

Obrezanova, O., A. B. Movchan and J. R. Willis (2002). "Dynamic stability of a


propagating crack." Journal of the Mechanics and Physics of Solids 50(12): 2637-
2668.

Obrezanova, O., A. B. Movchan and J. R. Willis (2002). "Stability of an advancing


crack to small perturbation of its path." Journal of the Mechanics and Physics of
Solids 50(1): 57-80.

Ouyang, Z., G. Ji and G. Li (2011). "On Approximately Realizing and Characterizing


Pure Mode-I Interface Fracture between Bonded Dissimilar Materials." ASME
Journal of Applied Mechanics 78(3): paper number 031020.

Ouyang, Z. and G. Li (2009b). "Local Damage Evolution of DCB Specimens during


Crack Initiation Process: A Natural Boundary Condition Based Method." ASME
Journal of Applied Mechanics 76(5): paper number 051003.

Ouyang, Z. and G. Li (2009c). "Cohesive Zone Model Based Analytical Solutions for
Adhesively Bonded Pipe Joints under Torsional Loading." International Journal of
Solids and Structures 46(5): 1205-1217.

Park, K. and G. H. Paulino (2013). "Cohesive Zone Models: A Critical Review of


Traction-Separation Relationships Across Fracture Surfaces." Applied Mechanics
Reviews 64(6): 060802-060802.

Pereira, A. B., A. B. de Morais, A. T. Marques and P. T. de Castro (2004). "Mode II


interlaminar fracture of carbon/epoxy multidirectional laminates." Composites
Science and Technology 64(10–11): 1653-1659.

Reedy, E. D., F. J. Mello and T. R. Guess (1996). Modeling the Initiation and Growth
of Delaminations in Composite Structures SAND95-3070 New Mexico, CA, USA,
Sandia National Laboratories.

Remmers, J. J. C., R. de Borst and A. Needleman (2008). "The simulation of dynamic


crack propagation using the cohesive segments method." Journal of the Mechanics
and Physics of Solids 56(1): 70-92.

39
Rubio-Gonzalez, C. and J. J. Mason (2000). "Dynamic stress intensity factors at the
tip of a uniformly loaded semi-infinite crack in an orthotropic material." Journal of
the Mechanics and Physics of Solids 48(5): 899-925.

Sayers, C. M. and M. Kachanov (1991). "A simple technique for finding effective
elastic constants of cracked solids for arbitrary crack orientation statistics."
International Journal of Solids and Structures 27(6): 671-680.

Schellekens, J. C. J. and R. De Borst (1993). "A non-linear finite element approach


for the analysis of mode-I free edge delamination in composites." International
Journal of Solids and Structures 30(9): 1239-1253.

Shojaei, A., A. Dahi Taleghani and G. Li (2014). "A continuum damage failure model
for hydraulic fracturing of porous rocks." International Journal of Plasticity 59(0):
199-212.

Shojaei, A. and J. Fish (2015). Assessment of Reduced Order Homogenization for


Damage Tolerant Design Principles (DTDP) of Advanced Composite Aircraft
Structures. 56th AIAA/ASCE/AHS/ASC Structures, Structural Dynamics, and
Materials Conference, American Institute of Aeronautics and Astronautics.

Shojaei, A. and G. Li (2013). "Viscoplasticity analysis of semicrystalline polymers: A


multiscale approach within micromechanics framework." International Journal of
Plasticity 42: 31-49.

Shojaei, A., G. Li, J. Fish and P. J. Tan (2014). "Multi-scale Constitutive Modeling of
Ceramic Matrix Composites by Continuum Damage Mechanics." International
Journal of Solids and Structures 51(23-24): 4068-4081.

Shojaei, A., G. Li and G. Z. Voyiadjis (2012). "Cyclic Viscoplastic-Viscodamage


Analysis of Shape Memory Polymers Fibers With Application to Self-Healing Smart
Materials." Journal of Applied Mechanics 80(1): 1-15.

Shojaei, A., G. Z. Voyiadjis and P. J. Tan (2013). "Viscoplastic Constitutive Theory


for Brittle to Ductile Damage in Polycrystalline Materials under Dynamic Loading."
International Journal of Plasticity 48: 125–151.

Shojaei, A., G. Z. Voyiadjis and P. J. Tan (2013). "Viscoplastic Model of Ductile to


Brittle Damage for Polycrystalline Materials under Dynamic Loading." International
Journal of Plasticity.

Slepyan, L. I. (2010). "Dynamic crack growth under Rayleigh wave." Journal of the
Mechanics and Physics of Solids 58(5): 636-655.

Sluys, L. J. and R. de Borst (1994). "Dispersive properties of gradient-dependent and


rate-dependent media." Mechanics of Materials 18(2): 131-149.

Sprenger, W., F. Gruttmann and W. Wagner (2000). "Delamination growth analysis in


laminated structures with continuum-based 3D-shell elements and a viscoplastic

40
softening model." Computer Methods in Applied Mechanics and Engineering 185(2–
4): 123-139.

Steinberg, D. J. (1996). Equation of State and Strength Properties of selected


Materials. Livermore, CA, Tech.Rep.UCRL-MA-106439, Lawrance Livermore
National Labratory.

Stringfellow, R. G. and L. B. Freund (1993). "The effect of interfacial friction on the


buckle-driven spontaneous delamination of a compressed thin film." International
Journal of Solids and Structures 30(10): 1379-1395.

Tvergaard, V. and J. W. Hutchinson (1992). "The relation between crack growth


resistance and fracture process parameters in elastic-plastic solids." Journal of the
Mechanics and Physics of Solids 40(6): 1377-1397.

Tvergaard, V. and J. W. Hutchinson (1996). "Effect of strain-dependent cohesive


zone model on predictions of crack growth resistance." International Journal of Solids
and Structures 33(20–22): 3297-3308.

Tzaferopoulos, M. A. and P. D. Panagiotopoulos (1993). "Delamination of


composites as a substationarity problem: Numerical approximation and algorithms."
Computer Methods in Applied Mechanics and Engineering 110(1–2): 63-85.

Voyiadjis, G. Z., A. Shojaei and G. Li (2011). "A thermodynamic consistent damage


and healing model for self healing materials." International Journal of Plasticity 27(7):
1025-1044.

Voyiadjis, G. Z., A. Shojaei and G. Li (2012b). "A Generalized Coupled Viscoplastic-


Viscodamage- Viscohealing Theory for Glassy Polymers." International Journal of
Plasticity 28(1): 21-45.

Voyiadjis, G. Z., A. Shojaei, G. Li and P. Kattan (2012c). "Continuum Damage-


Healing Mechanics with Introduction to New Healing Variables." International
Journal of Damage Mechanics 21(3): 391-414.

Voyiadjis, G. Z., A. Shojaei, G. Li and P. I. Kattan (2012a). "A theory of anisotropic


healing and damage mechanics of materials." Proceedings of the Royal Society A:
Mathematical, Physical and Engineering Science 468(2137): 163-183.

Voyiadjis, G. Z., A. Shojaei and N. Mozaffari "Strain gradient plasticity for


amorphous and crystalline polymers with application to micro- and nano-scale
deformation analysis." Polymer(0).

Wei, Y. and J. W. Hutchinson (1997). "Steady-state crack growth and work of


fracture for solids characterized by strain gradient plasticity." Journal of the
Mechanics and Physics of Solids 45(8): 1253-1273.

Wisnom, M. R. (2012). "The role of delamination in failure of fibre-reinforced


composites." Philosophical Transactions of the Royal Society A: Mathematical,
Physical and Engineering Sciences 370(1965): 1850-1870.

41
Xu, Q. and Z. Lu (2013). "An elastic–plastic cohesive zone model for metal–ceramic
interfaces at finite deformations." International Journal of Plasticity 41(0): 147-164.

Zhang, Z. and G. H. Paulino (2005). "Cohesive zone modeling of dynamic failure in


homogeneous and functionally graded materials." International Journal of Plasticity
21(6): 1195-1254.

Zhu, X.-K. and K.-C. Hwang (2008). "Asymptotic crack-tip fields for dynamic crack
growth in compressive power-law hardening materials." International Journal of
Solids and Structures 45(13): 3644-3659.

Zi, G. and Z. P. Bažant (2003). "Eigenvalue method for computing size effect of
cohesive cracks with residual stress, with application to kink-bands in composites."
International Journal of Engineering Science 41(13–14): 1519-1534.

Zou, Z., S. R. Reid, S. Li and P. D. Soden (2002). "General expressions for energy–
release rates for delamination in composite laminates." Proceedings of the Royal
Society of London. Series A: Mathematical, Physical and Engineering Sciences
458(2019): 645-667.

Zuo, Q. H., D. Disilvestro and J. D. Richter (2010). "A crack-mechanics based model
for damage and plasticity of brittle materials under dynamic loading." International
Journal of Solids and Structures 47(20): 2790-2798.

Appendix A: CDM Computational Aspects

A.1. Stability Check

In most of the classical FEA simulations utilizing a stable material model is essential before

conducting a finite element analysis. Drucker stability criteria ensures that the material model

is stable


  > 0 (A.1)

It is worth noting that Drucker’s stability law is not a thermodynamic law, and many

materials show clear signs that they are not stable in the sense of Drucker’s stability check

including the traction-separation laws of delamination. However, it is well-understood now

that the violation of Drucker's stability postulation results in volumetric locking in FEA

element formulation, and failure of linear elasticity theory; unless an additional length– or

time–scale is specified in the constitutive relations (de Borst et al. 1999). The material

instability is a part of a progressive failure simulation and dealing with the material instability

42
is unavoidable in the case of delamination analysis. In this work a length scale is included into

the FEA simulation to remove the material instability as discussed in Section 6.

A.2. Continuum Damage Computation through Incremental Stiffness Updates

The damage mechanisms, i.e. microcracking and microvoiding, induce residual strains due to

mismatch between non-damaged and damaged stiffness (Bouvard et al. 2009). As depicted in

Fig. A1, there are two elastic strains that can be computed based upon either damaged E or

undamaged E elastic stiffness. In present work, the total strain in Eq. (1) is decomposed into

recoverable damaged elastic strain ϵº and inelastic damage strain ϵ² . One may note that the

name damaged elastic strain indicates that the recoverable elastic strain is computed based

upon the damaged stiffness E.

Figure A1. Typical stress straincurve in the case pure elastic and damage deformation
mechanisms (no plasticity)

Appendix B: Mesh Convergence Study

As discussed in Section 6, the mesh dependency is alleviated by calculating the displacement

at full damage from damage dissipation energy. The convergence study to support this

43
argument is provided in this appendix. Fig. B1 compares the traction separation response of

DCB models to the experimental data. The mesh convergence study is conducted by

systematically halving the mesh sizes and it reveals that fine meshes ensure the convergence

of the results. Only three mesh numbers are shown in Fig. B1 for DCB sample, and the

corresponding mesh sizes for each case are as follows: Mesh#1: 0.1 mm, Mesh#2: 0.025 mm,

and Mesh#3: 0.00625 mm.

Figure B1. Experimental and FEA results for DCB sample

44

You might also like