Download as pdf or txt
Download as pdf or txt
You are on page 1of 68

Experimental methods to determine model

parameters for failure modes of CFRP


DANIEL SVENSSON
Department of Applied Mechanics
CHALMERS UNIVERSITY OF TECHNOLOGY
G
oteborg, Sweden 2013

THESIS FOR THE DEGREE OF LICENTIATE OF ENGINEERING IN SOLID AND


STRUCTURAL MECHANICS

Experimental methods to determine model


parameters for failure modes of CFRP
DANIEL SVENSSON

Department of Applied Mechanics


CHALMERS UNIVERSITY OF TECHNOLOGY
G
oteborg, Sweden 2013

Experimental methods to determine model


parameters for failure modes of CFRP
DANIEL SVENSSON
c DANIEL SVENSSON, 2013

Thesis for the degree of Licentiate of Engineering 2013:07
ISSN 1652-8565
Department of Applied Mechanics
Chalmers University of Technology
SE-412 96 G
oteborg
Sweden
Telephone: +46 (0)31-772 1000

Chalmers Reproservice
G
oteborg, Sweden 2013

Experimental methods to determine model


parameters for failure modes of CFRP
Thesis for the degree of Licentiate of Engineering in Solid and Structural Mechanics
DANIEL SVENSSON
Department of Applied Mechanics
Chalmers University of Technology

Abstract
The focus of this thesis is to develop methods to predict the damage response of Carbon
Fibre Reinforced Polymers (CFRP). In the pursuit of reducing the manufacturing cost
and weight of CFRP components, it is crucial to enable modelling of the non-linear
response associated with various failure modes. Two failure modes are considered in
this thesis: fibre compressive failure and interlaminar delamination. Multidirectional
laminated composites are commonly used when a low weight is desired due to their high
specific strength and stiffness. In a carbon/epoxy composite, almost exclusively the
fibres carry the load. However, along the fibre direction, the compressive strength is
considerably lower than the tensile strength. With the same reasoning, the transverse
strength is considerably lower than the in-plane strength. This makes delamination and
fibre compressive failure two of the major concerns in structural design. Moreover, the
presence of delaminations severely reduces the compressive strength of a laminate. This
can cause catastrophic failure of the structure.
In Paper A, we suggest a test method for determining fracture properties associated
with fibre compressive failure. A modified compact compression specimen is designed
for this purpose and compressive failure takes place in a region consisting exclusively of
fibres oriented parallel to the loading direction. The evaluation method is based on a
generalized J-integral and full field measurements of the strain field on the surface of the
specimen. Thus, the method is not restricted to small damage zones.
Paper B focuses on measuring cohesive laws for delamination in pure mode loading.
The cohesive laws in mode I and mode II are measured with the DCB- and ENF-specimen,
respectively. With a method based on the J-integral, the energy release rate associated
with the crack tip separation is measured directly. From this, the cohesive laws are derived.
It is concluded that the nonlinear response at the crack tip is crucial in the evaluation of
the mode II fracture energy.
Keywords: composite, delamination, compressive failure, energy release rate, cohesive
modelling

ii

Preface
This work has been carried out during the years 2010-2013 at the Mechanics of Materials research group at the University of Skovde. Funding from the Swedish National
Aeronautical Research Program (NFFP5) is gratefully acknowledged.
First of all, my deepest gratitude to my co-supervisor Associate Professor Svante Alfredsson
for the patient guidance, encouragement and advice he has provided me during this time.
I am also grateful to my supervisor Professor Ulf Stigh for the generous sharing of his
knowledge and giving me the opportunity to do research.
Thanks to my present and former colleagues for their helpful attitude and creating such
an enjoyable environment.
Last but not certainly not least, a big thank you to my Anna for her never ending support,
patience and love.
Daniel Svensson
Sk
ovde, March 2013

iii

iv

Thesis
This thesis consists of an extended summary and the following appended papers:

Paper A

D. Svensson, K.S. Alfredsson, U. Stigh and N.E Jansson. An experimental method to determine the fracture properties of compressive
fibre failure in unidirectional CFRP. To be submitted for international publication (2013)

Paper B

K.S. Alfredsson, D. Svensson, U. Stigh and A. Biel. Measurement


of cohesive laws for initiation of delamination of CFRP. Submitted
for international publication (2013)

The appended papers were prepared in collaboration with the co-authors. Paper A:
The author of the thesis was responsible for the major progress in planning the paper,
developing the theory, performing the simulations and evaluating the experiments. Took
part in designing the experimental setup and performing the experiments. Paper B: The
author took part in planning the paper and developing the theory. Responsible for the
simulations and the evaluation of the experiments.

vi

Contents
Abstract

Preface

iii

Thesis

Contents

vii

Extended Summary

1 Introduction

2 Fibre compressive failure

3 Delamination

4 Summary of appended papers


4.1 Paper A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2 Paper B . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

6
6
7

References

II

Appended Papers AB

11

vii

viii

Part I

Extended Summary
1

Introduction

Long fibre composites such as Carbon Fibre Reinforced Polymers (CFRP) are commonly
used in advanced structural applications, e.g. in the aerospace and marine industry.
This is mainly due to favourable mechanical properties such as high specific stiffness and
strength, low density and high resistance to corrosion. However, the wide range of possible
failure modes and limited understanding of the material behaviour requires conservative
dimensioning. This is further complicated by the fact that the material behaviour is
dependent on e.g. lay-up, loading direction, specimen size and environmental effects such
as temperature and moisture. Thus, phenomenological determination of criteria to predict
failure on a structural level from coupon tests requires numerous testing. Thousands of
coupon tests are not uncommon in the design of a safe advanced structure. Therefore, a
major challenge in the structural design is to remove excessive safety factors to e.g. reduce
the manufacturing costs and environmental impact. It is well-known from experimental
experience that assuming an ideal brittle behaviour is excessively conservative since the
redistribution of stresses at high strain regions are not considered. Stresses around stress
risers such as holes or cut-outs are relaxed by damage evolution and further load can
be applied prior to failure. Therefore, a certain loss of integrity has to be allowed. It
is therefore desirable to model, experimentally determine and simulate the non-linear
response that precedes failure of the laminate. Moreover, it has been observed that
specimen size effects play an important role on the strength of a laminate, cf. e.g. [1].
Thus, a fracture criterion should be associated with a length scale. This is introduced in
a fracture mechanics based approach.
The scientific community is devoting substantial effort to develop test methods for
determination of the fracture energy for various failure modes. It is also desirable to
obtain accurate failure criteria of a laminate from the mechanical properties of the
fibres, the matrix and the lay-up geometry. To successfully calibrate such models, an
important step is to determine the fracture energy for different failure modes in isolation.
Several methods for each failure mode have been proposed and in [2] Laffan et al. give a
summary. Standardized methods have been developed for interlaminar testing, e.g. [3],
and for translaminar testing [4]. Furthermore, methods have been proposed for studying
longitudinal [5] and transverse intralaminar matrix failure [6]. Moreover, test methods for
fibre tensile failure have been proposed in e.g. [7]. Testing methods have been proposed
for fibre compressive failure in e.g. [8], [9] and [2].
In this thesis, experimental methods are developed for determining governing fracture
properties associated with fibre compressive failure and interlaminar failure. These failure
modes are considered to be two major design limiting fracture processes in structural
design. Thus, to accurately predict these failure modes is crucial in the design of improved
composite structures.
1

Fibre compressive failure

Unidirectional (UD) CFRP is known for its high stiffness and strength along the fibre
direction. However, the compressive strength is only about 60-70 % of the tensile strength
[10]. The tensile strength is mainly governed by the tensile strength of the fibres whereas
the compressive strength, to a larger extent, depends on several material properties of
the laminate. Multiple failure modes are associated with fibre compressive failure, e.g.
elastic microbuckling, plastic microbuckling, longitudinal fibre splitting, fibre crushing
and shear-driven failure. The distinction of elastic and plastic microbuckling refers to the
amount of shear strain induced in the matrix. Experimental identification of the failure
mechanisms preceding compressive failure is difficult due to the unstable and catastrophic
nature of fibre compressive failure. However, plastic microbuckling is identified as the
dominating failure mode, cf. e.g [11], [12]. Plastic microbuckling is the failure mode
where compressive loading introduces bending of fibres. This may start at a material
discontinuity such as fibre waviness or at a free edge where the fibres lack lateral support,
cf. e.g. [13]. This places the matrix in a shearing mode. As the load increases, the
fibres rotate further and yielding of the matrix is induced. Ultimately, the fibres fail at
two points due to a combination of axial compression and bending. A localized band of
broken fibres denoted a kink-band is created and propagates into the intact region of
the specimen, cf. Fig 2.1. According to the classical strength models by Budiansky [14]
and Budiansky and Fleck [15], the governing parameters of the compressive strength of a
UD laminate are the fibre misalignment and the shear yield strengths of the composite
material.
In the formation of a kink-band, matrix/fibre splitting occurs in-between the rotated
fibres. This induces further rotation of the fibres prior to fibre failure. This mechanism is
explained by Fleck in [16]. It has also been experimentally observed in [17], where the
failure process at the micro scale has been observed while the specimen is kept under
load. Thus, in compression the material properties of the matrix are important since the
matrix provides lateral support to the fibres. Moreover, the bond between the matrix
and the fibres is important for the compressive strength. It should be noted that, even
though microbuckling leading to kink-band formation is recognized as the dominating
failure mode, a shear-driven fibre failure mode is observed in e.g. [17]. In [17], Gutkin et
al. highlights conditions for whether the fibre fails in a shear-driven mode or in a kinking
mode.
Unidirectional composites are rarely used in structural applications due to their poor
transverse performance. Instead, a laminated composite is built-up by successive plies
with varying orientations. In [18], it is concluded that compressive failure occurs in a
multidirectional centre cracked specimen due to microbuckling of the 0 -plies and the
fracture energy is increased when the portion of 0 -plies is increased. It is therefore
important to determine the fracture energy associated with compressive failure of the
0 -plies in isolation. As yet there is no standardized test method for this purpose. A recent
review of earlier reported fracture energies associated with compressive failure is given
in [8]. From experiments with center crack specimens with a T800/924C laminate with
(0,902 ,0)3s layup reported in [19], Pinho et al. derive the fracture energy for kink-band
2

Figure 2.1: Kink band found in an experiment presented in Paper A. The scale bar
indicates 500 m
formation to about 76 kJ/m2 . In [8], Pinho et al. use a compact compression (CC)
specimen with a T300/913 laminate with cross ply layup and the fracture energy 79.9
kJ/m2 associated with the 0 -plies is reported. In the presented method, a normalized
energy release rate is calculated from linear FE-simulations. The fracture energy is then
directly determined from the maximum load. Later in [9], Catalanotti et al. use the same
specimen and layup geometry with a different material system (Hexcel IM7-8552) and
the fracture energy 47.5 kJ/m2 associated with the 0 -plies is reported. The authors use
digital image correlation to calculate the fracture energy from the actual strain field on
the lateral surface of the specimen, i.e. the assumption of a small scaled damage zone is
avoided. However, the use of cross ply laminates necessitates partitioning of the fracture
energy since the energy dissipated in the 90 -plies needs to be deducted. Furthermore,
interaction effects between the alternating 0 -plies and 90 -plies are neglected. Therefore
it is desired to determine the fracture energy by the use of a UD-layup. A test method
using a four point bend specimen with a UD-layup is presented by Laffan et al. in [2]. The
material system is the same as in [9]. Also here, linear elastic FE-simulations are used to
determine how the energy release rate relates to the applied load. A lower fracture energy
of 25.7 kJ/m2 is reported. The reason may be that the experiments are interrupted when
damage initiation is first detected. The fracture energy is then calculated from the load at
onset of damage. It is reported that damage is initiated by a shear crack initiation at the
notch. The shear crack is propagating a small distance with the direction approximately
45 to the mode I direction and then transforms into a kink band as the load increases.
This transition phase is described in [17] by Gutkin et al.
In Paper A of this thesis, a modified CC-specimen is used to determine the fracture
energy associated with longitudinal compressive failure. Compressive failure takes place
in a region with a UD-layup. Thus, partitioning of the fracture energy is not necessary
and interaction effects between the 0 -plies and off-axis plies are avoided. The evaluation
3

method is based on the concept of equilibrium of configurational forces [20] and full field
measurement of the strain field. This method is therefore applicable also in the case of a
large damage zone.

Delamination

In general, the interlaminar strength of laminated composites is substantially lower than


the in-plane strength of a multidirectional composite. The foremost reason for this is that
no fibres are oriented in the transverse direction. Moreover, the resin rich regions between
the lamina are zones of weakness. Thus, due to the comparatively poor performance in
transverse loading, interlaminar delamination is a major concern in the design of structural
components. In Fig 3.1, delamination in a cross ply laminate is shown. Delamination may
start at a stress concentration arising from an initial defect or damages occurring in the
use of the component. Moreover, a structural component often includes curved laminates
or laminates with a varying thickness. If the curved laminate is subjected to out-of-plane
bending, delamination may initiate due to the transverse stresses. Furthermore, ply
drops are often used to progressively reduce the out of plane thickness along a composite
laminate. The axial load has to be transferred to the thinner section by interlaminar
shear stresses. Thus, at the ply-dropping region, delamination can initiate in a shearing
mode. Studies of the influence of ply-drops and out-of-plane curvatures can be found
in e.g. [22] and [23]. Moreover, delamination may also initiate and propagate due to
low-velocity impact [24]. Delaminations are difficult to detect and it can severely decrease
the structural compressive strength. Thus, the low compression strength after impact
is a limiting design parameter in the industrial design of aero-structures. Since the
structural strength is severely decreased by the presence of delaminations, a conservative
prediction of the strength can be carried out by including delaminations in the model to
e.g. determine the critical buckling load [25]. Two commonly used models for simulating
delamination are, the Virtual-Crack-Closing-Technique (VCCT) based on linear elastic
fracture mechanics and Cohesive Zone Modelling (CZM) that is not limited to small
process zones.
Fracture mechanics based methods have been successfully used for modelling onset and
propagation of delamination in laminated composites. In a fracture mechanics approach,
propagation of delamination is assumed to initiate when the total energy release rate,
G = GI + GII , is equal to the fracture energy, Gc , associated with the current mode
mix. Here, GI is the energy release rate in mode I (opening), GII is the energy release
rate in mode II (shearing) and the mode mix is given by e.g. the fraction GII /G. The
fracture energy is determined experimentally for various mode-mixes. Commonly used
test methods for measuring the fracture energies in mode I and mode II are, the Double
Cantilever Beam (DCB)-test for mode I and the End Notch Flexure (ENF)-test for mode
II. The most commonly used mixed-mode test is the Mixed Mode Bending (MMB)-test,
cf. [26]. With this setup, the mode mix can be varied by adjusting the load point position.
By determining Gc for a range of mode-mixes, a relation between Gc and the mode mix
can be obtained by a best curve fit. The energy release rate from these tests methods has
been calculated with different levels of sophistication. In the simplest case, the energy
4

Figure 3.1: 2. Delamination, tensile fibre fracture and tensile/shear matrix cracks in a
cross-ply laminate. Picture taken from [21]

release rate is calculated by assuming the Euler-Bernoulli beam theory and a rigid crack
tip. The methods have later been extended to consider the flexibility ahead of the crack
tip by adding a crack length correction to account for the influence of the anisotropic
composite material. In [27], Juntti et al. give a comprehensive review of the developed
evaluation methods. For the case of orthotropy, Bao et al. [28] present crack length
corrections based on an ingenious rescaling technique.
As a step toward a more complete model of the processes involved at the crack tip,
the cohesive zone model is a strong candidate. A brief historical review on cohesive
modelling is given in Paper B. In a cohesive zone model, a planar damage zone is assumed
where the behaviour is governed by a traction separation law. The onset of damage is
modelled by a stress criterion and the traction is assumed to decrease as the cohesive
separation increases. At a large enough separation, the traction acting on the cohesive
surface drops to zero and crack propagation is initiated. For a given load history, the
area beneath the cohesive laws in mode I and mode II corresponds to the total fracture
energy. It is noted that not only the fracture energy needs to be accurately determined.
The complete shape of the cohesive law has to be accurately modelled to predict critical
loads of structures suffering from delamination. For example, at regions subjected to high
interlaminar stresses such as ply-drop regions, the critical interlaminar stresses have a
great influence on the structural behaviour. Therefore, it is desirable to experimentally
measure the cohesive law with high accuracy to model the interlaminar behaviour of a
laminated composite.
Cohesive models of delamination are rather widely used in the scientific community.
Several methodologies have been used for determining the cohesive law. One common
method is to use a best fit approach. A function by which the cohesive traction varies with
the cohesive separation is pre-selected. Then, the cohesive parameters that best reproduce
the force-displacement curves are chosen. However, for many of the test specimen
5

geometries, the shape of the cohesive law has a minor influence of the force-displacement
relation. Thus, the procedure is not sensitive to the parameter that it aims at measure.
Another methodology is to experimentally measure the cohesive law with a method based
on evaluation of the path-independent J-integral, cf. [31]. Several test methods are
designed so that the energy release rate, J, associated with the crack tip separation can be
directly measured from external loads and displacements, cf. [29] and [30]. Subsequently,
the cohesive laws in mode I and mode II are determined by differentiation of J with
respect to the opening separation and the shearing separation, respectively. This method
does not require any assumption of the material behaviour. Furthermore, the method is
valid for large damage zones. However, if the complete cohesive law for all mode-mixes is
to be determined, it requires the existence of an associated potential. If a potential does
not exist, then in mixed-mode loading, the cohesive law is path dependent. Thus, the
cohesive law can only be determined for the specific load histories that were applied in
the experiments. However, this is not an issue when the measurement of the cohesive
laws is restricted to the pure mode cases, cf. Paper B.
For delamination, it is noted that two different fracture processes are involved. One is
associated with bridging of fibres in the wake of the growing crack and one is associated
with the fracture process at the crack tip. The bridging stress behind the crack tip is
small compared to the initiation stress at the crack tip. However, in the presence of fibre
bridging, the bridging fibres can contribute to substantially higher fracture energies. Thus,
these two mechanisms act on two very different length scales. In Paper B, we focus on
the fracture process associated with the crack tip until initiation of delamination. This is
the important process when onset of delamination growth has to be avoided.

4
4.1

Summary of appended papers


Paper A

In this work a modified CC-specimen is designed to study longitudinal compressive


failure. Normally when using a CC-specimen a cross ply laminate is used and an in-plane
notch is used to achieve a stress rising effect to nucleate compressive failure. Thus,
interaction effects between the 0 -plies and off-axis-plies are neglected and partitioning
of the fracture energy is needed to determine the fracture energy associated with the
0 -plies. Furthermore, the data reduction scheme is often based on linear elastic fracture
mechanics, i.e. the damage zone is assumed to be small.
In this work, localized high strains are achieved by decreasing the out-of-plane thickness towards the anticipated damage region that consists exclusively of 0 -plies. Thus,
compressive failure of 0 -plies is obtained in isolation. The data reduction scheme is
derived from Eshelbys concept of equilibrium of configurational forces [20]. The method
is similar to earlier work where the J-integral [31] is used to determine the energy release
rate associated with the damage zone. However, the J-integral is not applicable for the
present geometry due to the varying out of plane thickness. Thus, a generalized form of
the J-integral is used to determine the energy release rate associated with the damage
zone. Full field strain measurement with the DIC-system Aramis is used in the evaluation.
6

Thus, the assumption of a small damage zone is avoided. Numerical simulations are used
to verify the experimental results. The damage region is idealized as a cohesive zone
model, i.e. the material behaviour within the damage zone is governed by a cohesive law
relating the compressive stress and compressive separation. A cohesive law is proposed
and the simulated results agree well with the main features of the experimental results.

4.2

Paper B

In this paper, methods are presented for measurement of the cohesive laws in mode
I and mode II associated with interlaminar delamination. Cohesive laws for mode I
and mode II are measured with the DCB- and ENF-test, respectively. With a method
based on the path-independent J-integral, the energy release rate, J, associated with
the crack tip separation can be measured directly from the applied load, load point
rotations and the separation at the crack tip. By differentiation of J with respect to
the separation at the crack tip, the cohesive laws are determined. FE-simulations are
performed with the cohesive laws implemented and the simulations show good agreement
with experimental results. The results indicate that the fracture energy in mode II can
be severely underestimated if the inelastic behaviour at the crack tip is ignored.

References
[1] J. Lee and C. Soutis. Measuring the notched compressive strength of composite
laminates: Specimen size effects. Composite Science and Technology 68 (2008),
23592366.
[2] M. Laffan et al. Measurement of the fracture toughness associated with the longitudinal fibre mode of laminated composites. Composites Part A 43 (2012), 1930
1938.
[3] ASTM. D5528 Standard test method for mode I interlaminar fracture toughness of
unidirectional fiber-reinforced polymer matrix composites. 2007.
[4] ASTM. E1922-04 Standard test method for translaminar fracture toughness of
laminated polymer matrix composite materials. 2004.
[5] B. F. Sorensen and T. K. Jacobsen. Large-scale bridging in composites: R-curves
and bridging laws. Composites Part A: Applied Science and Manufacturing 29.11
(1998), 14431451.
[6] S. Pinho, P. Robinson, and L. Iannucci. Developing a four point bend specimen to
measure the mode I intralaminar fracture toughness of unidirectional laminated
composites. Composites Science and Technology 69 (2009), 13031309.
[7] M. Laffan et al. Measurement of the in situ ply fracture toughness associated with
mode I fibre tensile failure in FRP. Part I: Data reduction. Composites Science and
Technology 70 (2010), 606613.
[8] S. Pinho, P. Robinson, and L. Iannucci. Fracture toughness of tensile and compressive
fibre failure modes in laminated composites. Composites Science and Technology
66 (2006), 20692079.

[9] G. Catalanotti et al. Measurement of resistance curves in the longitudinal failure of


composites using digital image corroleation. Composites Science and Technology
70.13 (2010), 198693.
[10] C. Soutis. Compressive strength of unidirectional composites; measurement and
predictions. ASTM STP 1242 (1997), 168176.
[11] P. Berbinau, C. Soutis, and I. Guz. Compressive failure of 0 unidirectional carbonfibre-reinforced plastic (CFRP) laminates by fibre microbuckling. Composites Science
and Technology 59 (1999), 14511455.
[12] W. Slaughter, N. Fleck, and B Budiansky. Microbuckling of Fibre Composites:
The Roles of Multi-Axial Loading and Creep. J Eng. Mater. & Techn 115 (1993),
308313.
[13] A. Jumahat et al. Fracture mechanisms and failure analysis of carbon fibre/toughened
epoxy composites subjected to compressive loading. Composites Structures 92 (2010),
295305.
[14] B. Budiansky. Micromechanics. Computer and structures 16 (1983), 312.
[15] B. Budiansky and N. Fleck. Compressive failure of fibre composites. J. Mechanics
and Physics of Solids 41.1 (1993), 183211.
[16] N. Fleck. Compressive failure of fibre composites. Advances in applied mechanics
33 (1997), 43117.
[17] R. Gutkin et al. On the transition from shear-driven fibre compressive failure to fibre
kinking in notched CFRP laminates under longitudinal compression. Composites
Science and Technology 70 (2010), 12231231.
[18] C. Soutis, P. Curtis, and N. Fleck. Compressive failure of notched carbon fibre
composites. Proc R Soc 440.1909 (1993), 24156.
[19] C. Soutis and P. Curtis. A method for predicting the fracture toughness of CFRP
laminates failing by fibre microbuckling. Composites Part A 31 (2000), 733740.
[20] J. Eshelby. The force on an elastic singularity. Phil. Trans. R. Soc. London 244
(1951), 87112.

E. Characterization of impact damage in composite laminates. FFA TN


[21] M. Alvarez
1998-24. Tech. rep. The Aeron Res Inst of Sweden, Bromma., 1998.
[22] Z. Petrossian and M. R. Wisnom. Prediction of delamination initiation and growth
from discontinuous plies using interface elements. Composites Part A: Applied
Science and Manufacturing 29.56 (1998), 503 515.
[23] A. Weiss et al. Influence of ply-drop location on the fatigue behaviour of tapered
composites laminates. Procedia Engineering 2.1 (2010), 1105 1114.
[24] A. Turon et al. Accurate simulation of delamination growth under mixed-mode
loading using cohesive elements: Definition of interlaminar strength and elastic
stiffness. Composite Structures 92 (2010), 18571864.
[25] E. Barbero and J. Reddy. Modeling of delamination in composite laminates using a
layer-wise plate theory. International Journal of Solids and Structures 28.3 (1991),
373 388.
[26] J. Reeder and J. Crews Jr. Mixed-Mode Bending Method for Delamination Testing.
AIAA Journal 28 (1990), 12701276.

[27]

[28]
[29]
[30]

[31]

M. Juntti, L. Asp, and R. Olsson. Assessment of Evaluation Methods for the MixedMode Bending Test. Journal of Composites Technology and Research 21 (1999),
3748.
G. Bao et al. The role of material orthotropy in fracture specimens for composites.
International Journal of Solids and Structures 29 (1991), 11051116.
B. F. Sorensen and P. Kirkegaard. Determination of mixed mode cohesive laws.
Engineering Fracture Mechanics 73 (2008), 2006.
U. Stigh et al. Some aspects of cohesive models and modelling with special application
to strength of adhesive layers. International Journal of Fracture 165 (2010), 149
162.
J. Rice. A path independent integral and the approximative analysis of strain
concentration by notches and holes. Journal of applied mechanics 88 (1968), 379
386.

10

Part II

Appended Papers AB

11

12

Paper A

An experimental method to determine the fracture properties of compressive fibre failure in unidirectional CFRP

14

An experimental method to determine the fracture properties of


compressive fibre failure in unidirectional CFRP
D. Svenssona , K.S. Alfredssona , U. Stigha , N.E. Janssonb
a

University of Sk
ovde, S-541 28 Sk
ovde, Sweden
GKN Aerospace Engine Systems Sweden, S-461 81 Trollh
attan, Sweden

Abstract
A novel test specimen geometry for studies of longitudinal compressive failure of composites is proposed. Damage is localized without using a premade in-plane notch. Instead,
localized high strains are achieved by decreasing the out-of-plane thickness towards the
anticipated damage region which contains exclusively unidirectional lamina. Thus, an isolation of the compressive fibre failure mode is permitted. Experiments performed show
that the test geometry produces fracture in the form of kink-band formation progressing
along the section with the smallest out-of-plane thickness. Fibre kinking takes place in
the direction with the least support from the surrounding material, i.e. in the out-of plane
direction which is perpendicular to the direction of kink band progression. A method to
extract the fracture energy associated with initiation of fibre-dominated compressive failure is developed. The method is based on the concept of equilibrium of configurational
forces and full-field measurements of the strain field. Numerical simulations are used as a
tool for evaluating the local response in the most strained region of the test specimen. The
essential features of the response is captured by modelling the body of the test specimen as
a continuum and the damage region as a cohesive zone. An important finding is that, for
the present test configuration, a cohesive law should, prior to softening, include a region
where the stress remains constant or is increasing with a reduced tangential stiffness as the
compression increases.
1. Introduction
Longitudinal compressive failure in Carbon Fibre Reinforced Polymer (CFRP) laminates
has been a topic for intensive research during the last decades. These materials are in
general brittle. However, stresses around e.g. holes are relaxed by progressive damage
in a fracture process zone. If this process is ignored in an elastic stress analysis, the
load capacity is underestimated since the high stresses at the boundary of the holes are
redistributed. Thus, the process of progressive damage needs to be modelled. In [1],
Guynn and Bradley analyze the zone of compressive failure based on the Dugdale model.
Later Soutis et al. [2] analyses open hole compression specimens with a linear softening
cohesive zone model. The compressive strength and the size of the damage zone at failure
are successfully predicted for various lay-ups and hole sizes. In [3], Budiansky and Fleck
1

develop micromechanical models to predict the behaviour in the damage region. Pinho et
al. gives an overview of the processes involved in a recent review [4].
Modelling of the structural integrity requires the fracture energy as input parameter.
Standard tests to measure the fracture energy have been developed for certain failure modes
but no method has yet been successfully developed for measurement of fibre dominated
compressive failure. In [5] Pinho et al. use a compact compression (CC) specimen with a
cross ply laminate and a fracture energy of 79.8 kJ/m2 is reported. Later Catalanotti et al.
[6] use the same specimen geometry and layup with another material system and a fracture
energy of 45.7 kJ/m2 is reported. It is noted that partitioning of the fracture energy is
needed since the longitudinal fibre failure is accompanied by intralaminar cracking in the
90 -plies. The fracture energy of intralaminar cracking is however found to be of negligible
magnitude compared to the magnitude associated with fibre failure. However, possible
interaction effects are neglected. Thus, measurement of the fracture energy using a UDlayup is desirable. In [7], Laffan et al. use a 4-point bending specimen with a UD-layup
and a fracture energy of 25.9 kJ/m2 is reported. It is reported that failure initiates by a
mode II crack that propagates a small distance prior to transition to in-plane kink-band
propagation. This phenomenon is explained by Gutkin et al. in [8]. However, fibre splitting
sometimes occur prior to fibre failure when using UD-layups, cf. e.g. [9]. This is avoided
with cross ply laminates.
In this work, a modified CC-specimen is used to estimate the fracture energy associated
with longitudinal compressive failure and a data reduction scheme is derived by use of
Eshelbys concept of the equilibrium of configurational forces, [10]. How the concept of
equilibrium of configurational forces and cohesive modelling are related is presented in e.g.
[11]. In CC-tests, the compressive strain is usually localized by the use of a premade notch
that achieves a stress rising effect. Here, the damage is localized without using a premade
notch. Instead, the out of plane thickness is decreased towards the anticipated damage
region and longitudinal compressive failure is obtained within a small region consisting of
a UD-layup. FE-simulations are performed to validate the experimental results where a
cohesive zone models the damage region. Cohesive zone models fit well into the structure
of displacement based FE-programs and offers a convenient way to model the damage zone.
To this end, not only the associated critical energy release rate has to be determined; the
shape of the cohesive law must be defined. In this paper, governing parameters of the
cohesive law are estimated from experimental results. These are implemented in Abaqus
6.11 as a cohesive user material (UMAT). Simulations capture the main features of the
experimental behaviour.
The plan of the paper is as follows: First the method is presented in section 2. In
section 3 the experiments are presented. FE-simulations are presented in section 4. This
is followed by a discussion of the results. The paper ends with a summary of the main
conclusions.

110

24.7

L1

L2
65

15.6
9

8.7

0.8

14

90

Figure 1: Sketch of specimen geometry. All dimensions in mm


2. Method
2.1. Design of specimen and test setup
The geometry of the specimen is shown in Fig. 1. A multidirectional (MD) laminate is
used with nominal thickness, height and length of 4 mm, 65 mm and 110 mm, respectively.
The material is manufactured using Resin Transfer Moulding (RTM) with a UD fabric
reinforcement made from HTS carbon fibres and RTM6 resin. The nominal laminate layup
indicated by L1 in Fig. 1 is [0/+60/-60/0/0/+60/-60/0]s with the 0-direction parallel to
the loading direction. At the mid-section, the thickness is 0.8 mm and all fibres are oriented
parallel to the loading direction. To achieve this, the four inner 60-plies are dropped off
creating layup L2 Fig. 1. Closer to the mid-section, the outer plies including the 60s
are removed with a milling operation. The region at the mid-section with a height and
thickness of 1 mm and 0.8 mm, respectively, is denoted the waist, cf. Fig. 1.
The test-setup is visualized in Fig. 2. The specimen is mounted into a LLOYD loading
frame with a 10 kN load cell by inserting the free arms between two steel plates which ensures that the load is applied symmetrically and also prevents global buckling deformation.
Solid cylinders with a slightly smaller diameter are inserted into the holes in the plates
and the specimen. Frictional forces on the contact surfaces of the specimen are reduced
by lightly grinding them down. The lower cylinder is held fixed and the upper cylinder is
subjected to quasi-static loading by applying a downwards prescribed displacement of 50
m/s.
To capture relevant displacements, the specimen is prepared for digital image correlation (DIC) using an Aramis system by spraying a black-on-white speckle pattern on one
of the lateral surfaces of the specimen. A sufficiently large region is monitored with the
Aramis system so that displacements and rotations of the loading points are measured
3

Figure 2: Experimental setup


simultaneously with the strains along the notch of the specimen. Throughout the experiments, images are taken every five seconds by the Aramis system. For the present
experimental series about 300 pictures are taken in each experiment.
2.2. Theoretical framework
In fracture mechanics, fracture energies are often evaluated by using a curve integral designated the J-integral [12],
Z
Z
J=
(W dy Ti ui,1 ) dC =
(W n1 Ti ui,1 )dC
(1)
C

where W is the strain energy density and Ti = ij nj is the traction vector acting on
the outside of the path C with the outward normal vector ni . Index notation is used
where i = 1, 2 indicate components along the x and ydirection; summation is indicated
by repeated indices and partial differentiation by a comma. The J-integral is zero for
any closed path if the enclosed region does not include any singularity. With the notion
singularity, we here mean any object or feature of the body that, when moved in the elastic
field, changes the potential energy of the body.
For a test geometry with a constant out-of-plane thickness, evaluation of the J-integral
along the path indicated in Fig. 3 would yield an energetic balance in terms of energy
per unit surface. However, the present specimen geometry has a varying out-of-plane
thickness. Thus, a generalized form of the J-integral has to be introduced in terms of the
surface integral,
ZZ
ZZ
P =
(W 1j ij ui,1 ) nj dS =
(W n1 Ti ui,1 )dS
(2)
S

B
r
y

H
E

G
b

G
F

waist

x
s

F
r
D

Figure 3: Integration path (left) and detail of the notch (right)


where ni is the outward normal vector on the surface S. In the following, indices i = 1,2,3
indicate components along the x, y and z direction, respectively. From the P-integral,
an energetic balance can be formulated in terms of energy per unit length by exploiting
the fact that P is zero for a closed surface if the enclosed volume does not include any
singularity. The P-integral can also be interpreted as the sum of all configurational forces
that act on the volume with outer boundary S. For the present geometry, the P-integral
is evaluated over the closed surface that consists of the lateral surfaces and the surface
created when the dashed path indicated in Fig. 3 cuts through the thickness.
The resulting configurational force from each sub-surface is studied for the present
geometry. Both terms in Eq. (2) are zero for the lateral surfaces and the surfaces that
includes the paths A-B, C-D, E-F and G-H, since both n1 = 0 and T = 0 on these
surfaces. Moreover, in the design of the specimen, special precaution is taken so that
surfaces that include the vertical boundaries A-H, E-D and B-C are virtually undeformed,
i.e. W = 0 and T = 0. That is, these surfaces are not associated with any configurational
forces. Thus, the only surfaces that provide non-zero configurational forces are the surfaces
along the notch G-F and along the holes where the loads are applied. The magnitudes of the
corresponding configurational forces are denoted Pnotch and Pload , respectively. Equilibrium
of the configurational forces may then be formulated as
Pload = Pnotch

(3)

The left hand side in Eq. (3) is dominated by the second term in Eq. (2) and can be
formulated as
Pload = F (1 + 2 )

(4)

where F, 1 and 2 are the applied load and the rotations at the loading points, respectively
[11]. No traction is acting on the surface that includes the boundary F-G, whereby the
configurational force associated with this surface can be formulated as

Pnotch =

ZZ

W dydz

(5)

Snotch

This configuational force can be separated into three terms corresponding to the configurational forces associated with the sub-surfaces that includes the boundaries indicated by
r, b and waist, respectively, cf. Fig. 3. Thus we write
Pnotch = Pb + Pr + Pwaist

(6)

Since the fracture process is assumed to be concentrated to the waist, the main focus is to
determine Pwaist . From Eqs. (3) and (6) we derive
Pwaist = Pload Pr Pb

(7)

In the experimental evaluation, the strains measured along the boundaries r and b are used
to calculate Pb and Pr . Since the strains are only measured on one of the lateral surfaces,
assumptions have to be made regarding the variation of stress and strain components
through the thickness. Since the tangential stress is the dominating stress component
along the boundary, it is assumed that only this stress component contributes to the strain
energy. Furthermore, we assume that the state of strain is constant through the thickness
and that the deformation along b and r in Fig. 3 is governed by linear elasticity. Thus,
the strain energy density is calculated as if each element along the boundary is subjected
to uniaxial stress, i.e.
1
W = Et 2t
(8)
2
where t is the tangential strain along the boundary and Et is the direct elastic modulus
in the same direction. It is noted that Et varies along the notch due to the continuously
varying laminate layup and due to the change in direction along the radii.
Since the specimen has a constant out-of-plane thickness at the waist, its configurational
force can be interpreted as the mean value of the energy release rate consumed in the waist
multiplied with its out-of plane width, b. A nominal energy release rate associated with
the waist is therefore given by
Jwaist = Pwaist /b

(9)

The maximum value of Jwaist is identified as a measure of the fracture energy of the unidirectional material in the waist. It is denoted Jwaist,c .
2.3. Data reduction scheme
The configurational force Pload is determined with high precision by measuring the load
and the load point rotations, cf. Eq. (4). To determine Pr and Pb , the tangential strain,
t , along the boundary G-F is measured with the Aramis system. Strains are recorded in
points located every 0.15 mm along the boundary.
6

Given the strains along the boundary b, the strain energy density W is obtained from Eq.
(8), where the elastic modulus, Et , is determined with laminate theory. Note that Et varies
along the boundary b due to the reduction of thickness by dropping of the inner 60 -plies
and machining of the outer plies, as described earlier. Finally, Pb is formed by integration
similarily to Eq. (5).
The resolution of the strain measurements is not high enough to accurately measure
the strain along the radii and calculate Pr directly by integration. It is here assumed that
Pr can be established based on measurement of the tangential strains at the points F and
G, i.e. Fxx and G
xx , cf. Fig 3. The relation between Pr,sim and the square of the tangential
strain at points F or G is determined from a linear elastic FE-simulation. A coefficient
is first determined from the FE-simulation according to
=

Pr,sim
Pr,sim
= G
(Fxx,sim )2
(xx,sim )2

(10)

where it is noted that Fxx,sim = G


xx,sim due to symmetry. However, in the experiments
symmetry cannot be guaranteed, i.e. Fxx,exp and G
xx,exp are generally not equal. Thus,
the configurational force associated with the radii is evaluated from the experimental data
through
2
2 i
h F
Pr,exp =
xx,exp + G

(11)
xx,exp
2
where a correction factor is introduced to account for difficulties in measuring the tangential strains exactly at the positions F and G, at the start of the radii. To set the
value of for a certain experiment it is assumed that, in the initial stage of the experiment, Pr,exp is related to Pb,exp with the same ratio as in a linear elastic FE-simulation,
i.e. Pr,exp = Pb,exp where = Pr,sim /Pb,sim . For the present test specimen 0.46 and
the correction factor varies between 1.4-1.7 for the experiments presented in the next
section. This suggests that the tangential strains, Fxx,exp and G
xx,exp , are underestimated
by 15-23 %, cf. Eq. (11). Estimating the distance between the measuring points and
points F/G based on the resolution in the DIC-measurement; this error seems reasonable
considering the strain variation near the radii obtained with a FE-simulation. This type
of measurement error is considered neglectible in the evaluation of Pb . In FE-simulations,
the strain gradient component normal to the loading direction is noticed to be small along
the boundary b.
3. Experiments
Two experimental series, denoted series 1 and series 2, are carried out. Series 1 and 2 consist
of five and four experiments, respectively. In series 2, measurement problems preclude
evaluation according to the method discussed in the previous section. However, the loads
at failure are considered accurate and are presented below. Therefore, only experiments
from series 1 (denoted experiment 1-5 in the following) are fully evaluated in this paper.
Experiments 1-4 are loaded beyond compressive failure whereas experiment 5 is unloaded
7

prior to failure and a microscopy study is performed in an attempt to detect damage


mechanisms that eventually lead to compressive failure. These are difficult to identify in a
failed specimen since compressive failure occurs in an unstable and catastrophic manner.
Composite specimens subjected to compressive loading are sensitive for variations of
geometry and imperfections such as voids and fibre waviness. Therefore, the quality of the
specimens are studied. Specimen 2 is, after failure, cut and analysed in a microscopy study.
It is observed that the present specimen is manufactured as intended with all lamellas
present with the intended fibre orientation and ply drops are accurately positioned. Also,
three tensile tests are performed on coupon specimens cut from tested specimens by cutting
ten mm to the left, parallel with the boundary BC, cf. Fig 3. A FE-model of the tensile
test is analysed with Abaqus 6.11 using shell elements and the composite layup application
to model the material. The elastic stiffnesses measured in the coupon tests are in good
agreement with the FE-simulations. A certain amount of fibre waviness is observed in
the UD-material at the waists of the specimens which probably reduces the compressive
strength, cf. e.g. [13].
The specimens fail due to compressive failure in the waist. In the microscopy study
of specimen 2, a kink-band is observed in the waist and the kink-band tip is found approximately 47 mm behind the start of the waist. Magnified images of the kink-band tip
are shown in Fig. 4. A kink-band height of about 200 m is observed. This corresponds
to about 2-3 times larger than earlier reported kink-band heights, cf. e.g. [8] and [13].
Possible reasons are briefly discussed in section 5.

Figure 4: Kink-band observed in the waist after failure. Two different magnifications
are shown and the scale bar in the left and right image indicates 1000 m and 500 m,
respectively
In Fig. 5, load-displacement curves are shown; results from series 1 are presented in Fig. 5a
and results from series 2 are presented in Fig. 5b. In series 1, the load point displacement,
, is measured with high accuracy using the Aramis system. Experiments presented in Fig.
5a are denoted experiment 1-5 from left to right. However, the displacements presented in
Fig. 5b do not correspond to the load point displacements. Due to measurement problems,
8

4
F (kN)

F (kN)

0
0

0.2

0.4

0.6
0.8
(mm)

1.2

0
0

1.4

0.2

0.4

0.6

0.8
1
(mm)

1.2

1.4

1.6

Figure 5: Experimental results a) Load-displacement curves from experiment 1-5. Experiment 5, unloaded at F =4.4 kN, is indicated with +-signs b) Load versus displacement of
the loading frame actuator from series 2
the applied load is plotted versus the position of the loading frame actuator denoted .
At the critical load, all experiments fail in an unstable manner. The load drops rapidly.
Afterwards, the load decreases with increasing displacement. The critical load varies with
about 20% which seems reasonable since failure is governed by an instability and small
variations of the local geometry may lead to large variation of the stability load. The events
at the unstable failure are not captured since the displacements are only recorded every
five seconds and a jump of load and displacement is recorded at failure. The vertical drop
shown in Fig. 5b is recorded since the displacement of the load actuator is prescribed at
a rate of 50 m/s. All experiments show in principal an elastic brittle behaviour with an
initial stiffness of about 20 kN/m. At higher loads, the behaviour is slightly nonlinear and
at 90 % of the critical load, the stiffness is reduced to about 85 % of the initial stiffness.
The applied load versus local compression in the waist (F-w ) is shown in Fig. 6 for
the five experiments in series 1. The curves are only recorded to the point of failure since
measurement of w is only possible prior to failure. Here, w is measured from the relative
displacement of two points located in the waist with an initial distance of 0.7 mm. Note
that values of w are presented with different scales in Fig. 6.
It is observed, that all experiments show similar initial behaviour. At higher loads
at about 2.5-4.0 kN, a substantial decrease of the slope is clearly visible. At this point
w 7 10 m, which corresponds to a nominal strain at the waist of 1.0-1.4 %. This
point is interpreted as damage initiation in the waist. As the load increases, the form of
the F-w -curves differs a lot and a general local response is difficult to identify. A large
variation of w at failure is observed. This type of variation is expected at the local level
since the behaviour is highly dependent on material imperfection and variations of the local
geometry. Experiment 5 was unloaded at F =4.4 kN and in Fig. 6 a noticable decreased
9

Experiment 1

Experiment 2

F (kN)

F (kN)

4
3
2

1
0
0
6

0
0

10 15 20 25 30 35 40
w (m)

Experiment 3

10

20

30
40
w (m)

50

60

25

30

Experiment 4

F (kN)

4
3

3
2

2
1

1
0
0

20

40

60
80
w (m)

4
F (kN)

F (kN)

100

0
0

120

10

15
20
w (m)

Experiment 5

3
2
1
0
0

10
15
w (m)

20

Figure 6: Applied load vs local compression at the waist

10

t /F (1/MN)

0
1
2
3
4
5
6
7

8
10
S (mm)

12

14

Figure 7: Strain along S. Solid line: strain at F =1.4 kN. Dashed line: strain at F =4.5 kN
Table 1: In-plane mechanical properties in the material principal directions.
E11 (GPa)
120

E22 (GPa) G12 (GPa)


10

3.5

12
0.25

slope is recorded prior to unloading. However, in the microscopy study, no obvious signs of
damage are apparent that can explain the reduced slope in Fig. 6. One may note that Fig.
6 indicates a remnant deformation after unloading of experiment 5 of about 2 m. This
corresponds to a strain of about 0.3 %. It is questionable if such a small variation in local
geometry is detectable in an optical microscopy study. Since the F-w -curves after initiation
of local nonlinearity has a large variation, more experiments are needed to determine the
local mechanisms that lead to compressive failure.
The tangential strain along the notch, i.e. the boundary indicated with b-waist-b in
Fig. 3, is shown in Fig. 7. The two curves correspond to two different stages in experiment
4. The solid line corresponds to a low value of the load, F = 1.4 kN, and the dashed line
corresponds to a higher load level F = 4.5 kN. Both curves are normalized with respect
to the currently applied load. It is observed that the strain is concentrated in the waist
at the higher load level while strains elsewhere along the notch are relaxed as compared
to the case for the low load level. That is, at points along the notch located remote from
the waist, strains do not increase linearly with the load. This indicates a redistribution of
stresses and strains when damage is progressing in the waist.
Mechanical stiffnesses used in the data reduction scheme are presented in Table 1. The
evaluation method discussed in the previous section is applied to experiments 1-4 and the
results are shown in Fig. 8. To the left, the evaluated configurational forces Pload , Pr and
Pb are shown versus the applied load. From these results, the energy release rate associated
11

70

Pload
Pb + Pr
Pb
Pr

60
50
P (N)

Jwaist (kN/m)

40
30

30

20
15
10

20

10
0

70

40
30

20
15

5
2

3
F (kN)

40
Pload
Pb + Pr
Pb
Pr

Jwaist (kN/m)

110
100
90
80
70
60
50
40
30
20
10
0

60
50

3
F (kN)

30
25
20
15
10
5

3
4
F (kN)

Pload
Pb + Pr
Pb
Pr

70

Experiment 3

35

25

80

25

10

2
3
F (kN)

Experiment 2

30

10
0

35

20

Jwaist (kN/m)

P (N)

50

P (N)

Pload
Pb + Pr
Pb
Pr

60

P (N)

2
3
F (kN)

Jwaist (kN/m)

Experiment 1

25

40
30
20

3
4
F (kN)

Experiment 4

20
15
10

505

10
0

3
F (kN)

3
F (kN)

Figure 8: Left: Configurational forces versus applied load. Right: Jwaist versus applied
load

12

with the waist, Jwaist , is calculated according to Eq. (7) and Eq. (9). To the right in Fig. 8,
Jwaist is shown versus the applied load. The fracture energy is determined as the maximum
value of Jwaist .
It is observed that Jwaist 0 in the early stages of the experiments. This is expected
prior to damage initiation in the waist. Near failure, the Jwaist -curves show varying behaviour, arising from the measured strain along the boundary b. It is resonable to assume
that the assumption of linear elasticity results in conservative measurement of the fracture
energy since Pb is overestimated. This effect is particularly revealed in experiment 2, where
Jwaist is decreasing due to a sudden increase of Pb . At this stage, it is observed in the DICmeasurement that the damage zone appears to expand into the boundary b and since linear
elasticity is assumed along b, it results in a substantial increase of Pb . Of course, thorough
conclusions are difficult to establish since only strains on the lateral surface are avaliable
for measurement. The fracture energy, Jwaist,c , associated with compressive failure in the
waist for experiment 1-4 are: 25.5, 30.5, 39.8, and 23.8 kN/m respectively. In the following
section, these values are compared with corresponding results from FE-simulations of the
present experimental series.
4. Finite element simulations
Numerical simulations are performed in order to verify the experimental results and to
increase the understanding of the fracture process. For the present geometry, a global
instability occurs at the critical load and the load drops instantaneously, cf. Fig. 5.
Therefore, implicit dynamic simulations with numerical damping are performed to enable
capturing of the fracture process. This is done by using the quasi-static application in
Abaqus.
A 2D-model of the specimen is created. The body of the test specimen is modelled
as a continuum and the fractured region as a cohesive zone. To be more specific, the
anticipated damage zone in the waist is modelled as a cohesive zone with the height 200
m corresponding to the measured height of the kink-band, cf. Fig. 4.
Linear elastic plane stress elements are used for the continuum part of the FE-model.
The mesh consists mainly of fully integrated 4-node elements, except near the radii where
triangular elements occur. The varying thickness is modelled by dividing the specimen
into several sections as shown in Fig. 9. The height of the sections in the area of reduced
thickness is typically 1 mm and the thickness of each section is determined as the mean
thickness of the corresponding section of the specimen. Each section is assigned orthotropic
properties calculated with laminate theory and material properties according to Table 1.
The cohesive zone is modelled with 4-node cohesive elements. The deformation of the
cohesive elements is governed by a cohesive law describing the normal compressive stress,
, as a function of the compression, w.
At this point it is important to note the difference
between w and w;
w is the compression measured as the relative displacement of two
points/nodes at the waist with an initial distance of 0.7 mm while w is the compression in
the cohesive element with height 200 m.

13

Figure 9: Geometry of FE-model with indicated sections


Along the horizontal lines where the continuum elements and the cohesive elements share
nodes, the element sides are about 50 m long. This element size has proven to be small
enough to capture the global instability without numerical problems. Simulations have
been performed with smaller elements, only to obtain the same results and longer computation time. A 3D-model has also been used to validate that the 2D-model captures the
global response satisfactorily. In the following, only results from the 2D-simulations are
presented and compared to the experimental results.
To model the loading, the boundary of each hole is subjected to a pin-joint constraint
with respect to the centre of the hole. The center of the hole is subjected to displacement
boundary conditions that corresponds to the experimental loading.
Compressive failure of CFRP has earlier been modelled using cohesive laws. In e.g. [2]
a cohesive zone model is used to model the fracture process in an open hole compression
specimen. The damage evolution is often formulated assuming that, after onset of damage,
the cohesive stress is decreasing linearly with the compression, cf. Fig. 10a. Such models
are also available for use in conjunction with continuum elements in commercial FE-codes,
e.g. Abaqus. However, simulations of the present geometry show that a linear softening
model incorrectly predicts that the load decreases at the onset of damage.
For the present experiments, the substantial decrease in the slope of the F-w -curves
in Fig. 6 is interpreted as initiation of damage. Beyond this stage the load increases
monotonically with increased local compression until failure of the waist. Therefore, a
cohesive law should in this case, prior to softening, include a region where the stress
remains constant or is increasing with a reduced tangential stiffness as the compression
increases. Thus, we propose a cohesive law according to Fig. 10b. In this cohesive law,
three values of the compressive stress 1 , 2 and 3 and the corresponding compressions
w1 , w2 and w3 are input parameters to be defined.
The first part of the cohesive law, i.e. stage 1 in Fig. 10, is assumed to be linearly
elastic. The corresponding initial stiffness is taken as the longitudinal stiffness of the
present UD-laminate, i.e. w1 is given by w1 = 1 hcz /E1 , where E1 is the longitudinal
14

2
1
3
w1

wc

3

4

w1

w2 w3

Figure 10: Left: Linear softening cohesive model Right: Proposed cohesive model
Table 2: Parameters derived from the simulations
Experiment 1 (MPa)
1
2
3
4

540
750
645
645

2 (MPa)

3 (MPa)

w1 (m)

737
750
790
925

0
0
0
0

0.90
1.25
1.07
1.07

w2 (m) w3 (m)
31
40
105
23

50
75
140
68

elastic stiffness of the unidirectional composite in the waist and hcz is the height of the
cohesive element. The stress 1 at the end of stage 1 is determined from a linear elastic
simulation by matching the load at damage initiation to the corresponding experimental
value. The tangent stiffness of stage 2 in the cohesive law is determined from nonlinear
simulations by matching the slopes in the experimental and simulated F-w -curves during
damage growth. The values of 2 and w2 are determined by matching the critical load
from the simulations to the experimental value.
Table 2 gives the numerical values of the parameters obtained for the four experiments.
The graphs in the left part of Fig. 11 show comparisons between the experimental F-w curves and the corresponding curves obtained from simulation of the first two stages of
the cohesive law, cf. Fig. 10. The experimental and simulated results are indicated with
solid lines and dashed lines, respectively. The numbered arrows indicate the corresponding
stage in the cohesive law in the cohesive element at the notch, i.e. the most strained
cohesive element. These stages are indicated by the encircled numbers in Fig. 10. The
initiation of stage 3 is shown in the simulated curves. At this point, the softening initiates
and the simulated compression exhibits a jump indicated by the horizontal ending of the
simulated F-w -curves. As described earlier, the experimental F-w -curves are shown until
failure, since the measuring points the are lost at this stage. It is believed that the essential
features of the cohesive law up to the maximum stress, 2 , are well captured, even though
the agreement is far from perfect.
As mentioned previously, unstable failure occurs when proceeding the simulations into
15

2
1

20
w(m)

30

40

4
F (kN)

3
2

F (kN)

10

1

0

10

20

30
40
w(m)

50

60

40

60 80
w(m)

1

0

0.2

100 120

1

0

w(m)

20

0.6 0.8
(mm)

0.2

1.2

1.2

3

4

0.4

0.6 0.8
(mm)

3
2

30

0.4

1
10

0.8

20

0.4
0.6
(mm)

1

0

1

0

0.2

F (kN)

F (kN)

F (kN)

F (kN)

F (kN)

F (kN)

0.2

4

0.4

0.6
0.8
(mm)

Figure 11: Comparisons of simulated and experimental results. Experimental results are
indicated with solid lines and simulated results with dashed lines

16

Table 3: Critical loads and fracture energies from experiments and simulations
Experiment

Fcexp (kN)

exp
Jwaist,c
(kN/m)

Fcsim (kN)

sim
Jwaist,c
(kN/m)

1
2
3
4

4.17
4.78
5.90
4.88

25.5
30.5
39.8
23.8

4.20
4.75
5.90
4.87

21.2
31.3
76.8
20.2

the softening part of the cohesive law, i.e. stage 3 in Fig. 10. Thus, the dynamic simulation
predicts a sudden drop of the load exactly as recorded in the experiments. During the
sudden load drop, the deformation of the most strained cohesive element reaches the end
of the softening part of the cohesive law. Thus, this part of the cohesive law cannot
be determined by comparison of the simulation with experimental results. Instead, the
negative slope in stage 3 of Fig. 10 is chosen small enough so that convergence is obtained
in the simulation increments immediately after the load drop. To determine the stress
level 3 in stage 4 of the cohesive law, the experimental load-displacement curves to the
right in Fig. 11 are studied. In the present case 3 = 0 gives the best agreement with the
experimental load-displacement curves after failure. It seems unreasonable that no stress is
acting between the fractured surfaces. However, it can be expected that significant damage,
which is not modelled here, is introduced at neighbouring areas where compressive failure
has occurred.
The graphs in the right part of Fig. 11 show comparisons of simulated and experimental
load-displacement (F ) curves. Note that experimental values of displacement are only
recorded every five seconds and therefore the experimental response during the sudden load
drop is not recorded. With this in mind, the comparison shows that the essential features
of the global behaviour are well captured.
The simulated and experimental values of the fracture energies, Jwaist,c , and critical
loads, Fc , are presented in Table 3 and the results are also visualized in Fig. 12. The
sim
presented values of Jwaist,c
are determined by using Eqs. (5), (8) and (9). Since failure
sim
occurs in the simulation when stage 3 of the cohesive law is reached, Jwaist,c
is given by
Z
Z w2
2 (hwaist hcz )
sim
Jwaist,c
=
W dy =
dw + 2
(12)
2Et
waist
0
where the first term is the area beneath the cohesive law until initiation of stage 3. The
second term is the energy release rate consumed by the linear elastic elements along the
waist at failure. The parenthesis in the second term equals 0.8 mm since the height of the
waist and cohesive zone is 1 mm and 0.2 mm, respectively.
The fracture energies in the FE-models agree well with the experimental results for
three experiments. However, in the simulation of experiment 3, the simulated fracture
energy is about twice the experimentally obtained fracture energy. This may have several
explainations but one that appears reasonable is that the large difference arise from the
17

Jwaist,c (kN/m)

80
70
60
50
40
30
20
10
0

Simulation
Experiment

3
4
Fc (kN)

Figure 12: Comparison of fracture energies from simulations and experiments


experimental measurement of w. It is noted that w at failure is about 110 m in experiment
3 while the corresponding values in experiment 1,2 and 4 are about 20-40 m. In the Fw -curve of experiment 3, three jumps of 10, 5 and 17 m, are noticed at F 3, 4.2
and 5.6 kN without visibly affecting the global behaviour. It is believed that some event
occured at the outer ply, where the compression is measured, which is not representative
for the material in the waist. Assuming that the specimen fails due to an unstable kinking
process, this magnitude of compression seems unlikely prior of failure. This could be an
explanation to the poor correlation in experiment 3 between the experimental fracture
exp
sim
energy Jwaist,c
and the fracture energy used in the simulation Jwaist,c
. Thus, the simulated
fracture energy appears precarious for experiment 3.
Clearly, numerical simulation of compressive failure is a complicated task. All aspects
of the complicated fracture process can not be expected to be captured by the idealized FEmodel used in the present paper. Moreover, uncertainties associated with the experimental
measurements cannot be overlooked. However, the comparison between experimental and
simulated results indicates a relevance of modelling compressive failure as a cohesive zone.
5. Discussion
In the present paper a novel test specimen geometry for studies of compressive failure of
composites is proposed. Localized high strains are achieved by decreasing the out-of-plane
thickness towards the anticipated damage region which contains exclusively unidirectional
lamina. Experiments performed show that the test geometry produces fracture in the form
of kink-band formation progressing along the waist, i.e. along the section with the smallest
out-of-plane thickness. Kinking takes place in the direction with the least support from the
surrounding material, i.e in the out-of plane direction, cf. Fig. 4, which is perpendicular
to the direction of kink band progression. In [14], Berbinau et al. concludes that laminates
having 0 outer layers tend to fail due to out-of-plane microbuckling, whereas outer off-axis
plies permits in-plane microbuckling failure.
The fact that the damage region contains exclusively unidirectional lamina has several
implications. On the one hand, the unidirectional plies do not have any support from
surrounding plies as is the case when compression failure is tested using a cross ply laminate.
18

This might give a different failure mechanism as compared to the cross ply situation. This
thought is supported by the fact that the height of the kink-band is 200 m in the present
study, which is about 2-3 times larger than kink-band heights reported in the literature,
cf. e.g. [8] and [13]. Part of this difference may also be attributed to differences in the
materials studied. For example, the ply thickness in the composite studied here is 0.250 mm
while the composite studied in [8] has a ply thickness of 0.125 mm. On the other hand,
the present approach with exclusively unidirectional lamina in the damage region does
not require the use of a special procedure to identify the contribution of the 0 -directed
lamina from the measured properties for the cross ply laminate. Indeed, the fracture
energy detected with the proposed method can be attributed to longitudinal compression.
The proposed method yields fracture energies for the specific composite of about 20-40
kN/m. In [7], the critical energy release rate is measured on a UD-specimen and a value
of 25.7 kN/m is reported. This value corresponds to the energy release rate at the onset
of damage by the formation of a shear crack initiating at the notch. It is reasonable to
assume that the fracture process is different in the present work due to the substantially
smaller thickness used in this work. This promotes out-of-plane microbuckling rather than
shear crack initiation due to longitudinal compression. Moreover, the failures observed
in this work does not initiate at a premade in-plane stress riser. Using cross ply compact
compression specimens, fracture energies of 47.5 kN/m and 79.9 kN/m are reported for two
different materials in [5] and [6]. In [15] a fracture energy of 38.8 kN/m is reported for a
(0/902 /0)3S laminate from which the fracture energy associated with kink-band formation
was calculated as 76 kN/m [5].
Experiments reported in section 3 show that failure takes place in an unstable manner, i.e. the load drops rapidly to about half the value prior to fracture as the kink-band
formation propagates rapidly. The fact that the sudden load drop takes place under prescribed displacements, indicates that we are dealing with a global instability inherent to the
geometry of the test specimen. That is, the instability is analogous to similar problems encountered for test specimens for delamination of composites, cf. e.g [16]. This conclusion
is supported by the fact that the FE-simulations display a similar force/displacementresponse, i.e. the load drops rapidly as soon as the most strained cohesive element enters
the softening part of the cohesive law. The obvious practical consequence of this is that
the softening part of the cohesive law cannot be captured with the present test geometry.
It is, however, believed that this can be circumvented in future developments of the test
geometry. By increasing the distance between the loading points and the most strained region, a geometry more favourable from stability aspects is achieved. This kind of geometry
changes also reduces the length of the damage zone along the waist, i.e. the redistribution
of stresses during an experiment is less pronounced. In an ideal situation, this would enable
extraction of the fracture energy using methods not relying on full field measurement of
strains.
The numerical simulations performed in the present paper are used as a tool for evaluating the local response in the most strained region. The objective is to capture the
essential features of the response with a model that is as simple as possible. Thus, in the
FE-model, the body of the test specimen is modelled as a linear elastic continuum and the
19

fractured region as a cohesive zone with and idealized cohesive law. With a more refined
model, better agreement with the experimental results can be obtained. For instance, a
non-linear elastic model for the continuum would be necessary to capture the nonlinear
response initiating at about 1 kN in the experiments, cf. Fig. 11. In order to capture the
continuous change of slope in the experimental F-w -curves, a more complex form of the
cohesive law is needed. However, it is believed that the present level of model complexity
is sufficient for the present purpose, i.e. to verify the experimental results and to increase
the understanding of the fracture process. In this context, an important finding is that
the cohesive law, prior to softening, includes a region where the stress remains constant or
is increasing with a reduced tangential stiffness as the compression increases. The present
evaluation method results in a maximum stress of the cohesive law, i.e. 2 in Fig. 10 in
the interval 740-925 MPa. These values are relatively low as compared to what might be
expected from previous results from the literature [17]. It is possible that the fibre waviness
that is observed in the waist of the specimens may introduce microbuckling at a fairly low
longitudinal stress, cf. e.g. [13]. Moreover, it is reasonable to assume that out-of-plane
buckling initiates at a fairly low compressive stress in the UD material since no off-axis
plies provide lateral support to prevent out-of-plane kinking, cf. e.g. [13].
6. Conclusions
A method is proposed for measurement of the fracture energy associated with longitudinal
compressive failure. Failure occurs in a region consisting of a unidirectional layup. Thus,
a compressive fibre failure process is isolated and reduction of dissipated energy associated
with non-0 -plies and interaction effects are not necessary. The fracture energy is evaluated
by using a generalized J-integral and full field measurement of the strains with a DICsystem. Thus, the assumption of a small damage zone is avoided. Experimental results
show large scatter and a fracture energy of 20-40 kN/m is reported. FE-simulations of the
reported experiments are performed. Comparing the experimental and simulated results
indicates a relevance of modelling compressive failure as a cohesive zone. The simulations
capture the main features of the experimental behaviour.
Acknowledgements
The authors would like to acknowledge funding from the Swedish National Aeronautical
Research Program (NFFP5) in support of this work through the joint project FIKOM
with GKN Aerospace Sweden. The authors are grateful to Dr. Anders Biel for his help in
connection with the experiments and to Dr. Fredrik Edgren at GKN Aerospace Sweden
for performing the microscopy study.

20

References
[1] E.G. Guynn and W.L. Bradley. Micromechanics of compression failures in open hole
composite laminates. Annual Progress Report, Apr. 1986-Auq. 1987 for NASA research grant NAG-1-659. 227p, 1987.
[2] C. Soutis, N.A. Fleck, and P.A. Smith. Failure prediction technique for compression
loaded carbon fiber-epoxy laminate with open holes. Journal of Composite Materials,
25:147698, 1991.
[3] B. Budiansky and N.A. Fleck. Compressive failure of fibre composites. J. Mechanics
and Physics of Solids, 41(1):183211, 1993.
[4] S.T. Pinho, R. Gutkin, S. Pimenta, N.V. De Carvalho, and P. Robinson. On longitudinal compressive failure of carbon-fibre-reinforced polymer: from unidirectional to
woven, and from virgin to recycled. Phil. Trans. R. Soc. A, 370(165):18711895, 2012.
[5] S.T. Pinho, P. Robinson, and L. Iannucci. Fracture toughness of tensile and compressive fibre failure modes in laminated composites. Composites Science and Technology,
66:20692079, 2006.
[6] G. Catalanotti, P.P. Camanho, J. Xavier, C.G. Davila, and A.T. Marques. Measurement of resistance curves in the longitudinal failure of composites using digital image
corroleation. Composites Science and Technology, 70(13):198693, 2010.
[7] M.J. Laffan, S.T Pinho, P. Robinson, L. Iannucci, and A.J. McMillan. Measurement
of the fracture toughness associated with the longitudinal fibre mode of laminated
composites. Composites Part A, 43:19301938, 2012.
[8] R. Gutkin, S.T. Pinho, P. Robinson, and P.T. Curtis. On the transition from sheardriven fibre compressive failure to fibre kinking in notched cfrp laminates under longitudinal compression. Composites Science and Technology, 70:12231231, 2010.
[9] C. Soutis, P.T. Curtis, and N.A. Fleck. Compressive failure of notched carbon fibre
composites. Proc R Soc, 440(1909):24156, 1993.
[10] J.D. Eshelby. The force on an elastic singularity. Phil. Trans. R. Soc. London, 244:87
112, 1951.
[11] U. Stigh and T. Andersson. An experimental method to determine the complete
stress-elongation relation for a structural adhesive layer loaded in peel. Fracture of
Polymers, Composites and Adhesives (Eds. Williams J.G. and Pavan A.), 27:297306,
2000.
[12] J.R. Rice. A path independent integral and the approximative analysis of strain
concentration by notches and holes. Journal of applied mechanics, 88:379386, 1968.
21

[13] C. Soutis. Compressive strength of unidirectional composites; measurement and predictions. ASTM STP, 1242:168176, 1997.
[14] P. Berbinau, C. Soutis, P. Goutas, and P.T. Curtis. Effect of off-axis ply orientation
on 0 -fibre microbuckling. Composites Part A, 30:11971207, 1999.
[15] C. Soutis and P.T. Curtis. A method for predicting the fracture toughness of cfrp
laminates failing by fibre microbuckling. Composites Part A, 31:733740, 2000.
[16] K.S. Alfredsson and U. Stigh. Stability of beam-like fracture mechanics specimens.
Engineering Fracture Mechanics, 89:98113, 2012.
[17] J. Lee and C. Soutis. A study on the compressive strength of thick carbon fibre-epoxy
laminates. Composites Science and Technology, 67(10):20152026, 2007.

22

Paper B

Measurement of cohesive laws for initiation of delamination of CFRP

38

Measurement of cohesive laws for initiation of


delamination of CFRP
K.S. Alfredsson, D. Svensson, U. Stigh, A. Biel
University of Sk
ovde, S-541 28 Sk
ovde, Sweden

Abstract
With a cohesive zone, delamination of laminated material can be analysed using non-linear
finite element codes. For carbon fibre reinforced composites, two different delamination
processes are identified; one at the crack-tip with a sub millimetre sized process zone
and one with a large-scale process zone associated with fibre bridging. In the present
paper we focus on the crack-tip process that is important when crack growth has to be
avoided. Methods to measure the cohesive relations associated with the fracture processes
are developed. These methods are used to measure the cohesive relations in mode I and
II, respectively. The material is a carbon fibre reinforced composite.
Keywords: CFRP, Cohesive modelling, Experimental, Cohesive law
1. Introduction
Delamination of Carbon Fibre Reinforced Polymer composites (CFRP) is one of the major
concerns in the design and use of advanced composite structures. Delamination may start
at unidentified defects originating from the production process or damages occurring in the
use of the component. Two different mechanisms and corresponding length scales can be
identified in the process of delamination. At the close proximity of a crack tip, a process
region can be identified. With epoxy resins, the associated fracture energy is in the range of
102 N/m and the yield strength is in the range of 101 MPa. A simple estimate predicts the
size of the process zone to about 101 mm. That is too small to indicate interaction with
outer boundaries of structures, but large enough to imply interaction with the fibres. From
an experimental point of view, this shows that the fracture properties should be measured
in the relevant composite and one cannot rely on bulk properties for the resin. At the
larger length scale and in the wake of a growing crack, crack bridging may occur. This
process often contributes significantly and increases the total fracture energy to about 103
N/m. The bridging stress is however small, in the range of 100 MPa. That is, the process
zone is very large, in the range of 101 mm, [1]. Thus, the two fracture processes are
associated with two very different length scales. In some applications, the enhancement of
the strength due to crack bridging can be considered. However, in the aeronautic industry,
no defects are allowed to grow during the use of a composite structure. Moreover, defects
from the production stage are likely to lack bridging fibres. Therefore, if no defects from
1

t
crack tip

T
w

y
x

v
t

Figure 1: Left: Cohesive zone heading a crack tip. Traction T holds the cohesive surfaces
together. Right: Traction and separation separated in orthogonal components relative the
middle surface of the cohesive surfaces.
the productions stage are allowed to grow during use, fibre bridging cannot be considered
in aeronautical applications.
In the present paper, we study delamination at the smaller length scale, i.e. without considering fibre bridging. The study is performed within the framework of cohesive
modelling. As compared to linear elastic fracture mechanics, this can be viewed as a step
toward a more complete model of the actual damage process. With cohesive modelling we
assume the existence of a planar process zone heading the crack tip. All inelastic material
processes in the real process zone are modelled by a cohesive law acting on the cohesive
surface. Figure 1 illustrates a cohesive model. The traction T is assumed to decrease
as the separation of the cohesive surfaces increases. At large enough separation, the
traction is zero indicating the formation of new crack surfaces. Historically, Barenblatt
introduced the cohesive model to increase the understanding of brittle fracture in his seminal paper 1962, [2]. Later, a number of researchers showed the usefulness of the concept
to model fracture in a large variety of applications: e.g. strength of structures of concrete,
[3], in-plane strength of composites, [4], creep crack growth, [5], and fracture of adhesives,
[6]. A major step forward was the realization that cohesive models fits well within the
structure of deformation based finite element analysis, [6], [7]. That is, strength analysis
of structures can be performed as non-linear stress analyses using conventional FE-codes.
Today, cohesive models are included in many commercially available FE-codes.
In the present paper, we develop methods to measure the cohesive laws in mode I
and II, respectively. The theory is developed in section 2 and the experiments and their
evaluation are reported in section 3. In section 4 one of the methods developed in section
2.2 is used to re-evaluate mode II experiments previously presented in [8]. The paper ends
with some conclusions.

2. Theory and methods


Methods to measure cohesive laws have been relatively sparingly reported. The embryo to
such methods can be traced to [9]. The J-integral gives the release of potential energy of
an elastic body per unit created crack area, associated with the propagation of the crack
front. It can be calculated from the integral
Z
J = (W dy Ti ui,x dC) .
(1)
C

where C is a counter-clockwise integration path, and W , Ti and ui are the strain energy
density, the traction vector and the displacement vector, respectively, cf. Fig. 1. Index notation is employed with index i = 1, 2 indicating components along the x and
ycoordinates, respectively; summation is indicated by repeated indices and partial differentiation by a comma. The crack is assumed to lie in a plane with y = constant. If
W does not contain any explicit dependence of the xcoordinate, the integration path C
starting at the lower crack surface and ending at the upper crack surface can be chosen
freely as long as C does not encircle objects that change the energy of the body if the
objects positions are changed. Thus, by choosing C close to the crack tip,1 we get
Z v
Z
Z w
d
v
(2)
dw +
J = W dy =
C

where , , w, and v are the cohesive normal stress, shear stress, opening and shear at
the crack tip, respectively, cf. Fig. 1.2 That is, if we are able to continuously measure J
from the external loads acting on a specimen during an experiment, and at the same time
measure v and w at the crack tip, we would be able to differentiate the measured J(v, w)
data to derive the cohesive laws (v, w) and (v, w), cf. Eq. (2). Two facts complicate
this idea. Firstly, the cohesive law is not likely to be elastic in nature. That is, it is not
likely that the strain energy density W exists. However, if the loading within the cohesive
zone can be regarded as monotonically increasing, a pseudo-potential A can replace W ,
[10]. In this case the mathematical difference between an elastic and inelastic material is
immaterial. It is only when un-loading from an inelastically deformed state occurs that the
difference between elasticity and inelasticity reveals itself. The second problem originates
from the fact that most expressions for J in terms of external loads implicitly or explicitly
depend on an assumption of the material behaviour. In [9], it is however shown that some
specimen geometries allow for a direct measurement of J from the applied load without
the need for a too restricted assumption of the behaviour of the material. This idea is

1
In the present paper, the crack tip is considered to be situated at the left end of the process zone,
cf. Fig. 1. It should be noted that the definition of the position of crack tip differs among authors. For
instance, in studies of crack bridging, the right end of the process zone is usually considered as the crack
tip and the process zone is referred to as a bridging zone.
2
Note that the cohesive shear stress is defined as = xy in order to get a positive shear stress near
the crack tip for the ENF-specimen, cf. Fig. 2.

P, D/2

P, D
y

H/2
H/2

P, D/2

H/2
H/2

a
L

aL

Figure 2: a) Double Cantilever Beam-specimen; b) End Notched Flexure-specimen. Both


loaded with prescribed load point displacements . Fibre orientation is indicated at the
right part of the specimens.
developed in [11]. In [12] a different path of derivation is taken. Starting from the basic
equations of Euler-Bernoulli beam theory, the authors show how the cohesive law can be
measured from the external loads. It was later shown that the same result can be derived
using the J-integral, [13] or by a direct application of the concept of energetic forces, [14].
These methods have previously been used to measure cohesive laws for adhesives and for
fibre bridging, cf. e.g. [13], [15].
In this section, two methods to measure the cohesive laws for delamination are presented. With a Double Cantilever Beam-specimen (DCB) the cohesive law in mode I is
measured; with an End Notched Flexure specimen (ENF) the cohesive law in mode II is
measured, cf. Fig. 2.
The basic assumption in the development of the experimental methods is the existence
of a cohesive law associated with a potential . Thus, we assume


= w
(3)

= v

As stressed above, if does not exist, it suffices that a pseudo-potential can replace in
the specific loading history. In the following sections, methods to measure cohesive laws in
mode I (DCB) and II (ENF) are presented.
2.1. Mode I
With an inner integration path, J is given from Eq. (2) as
Z w
dw
J=

(4)

Taking an outer integration path according to Fig. 3 yields contributions to J only from
the left boundaries where the forces P act. As shown in Eq. (1), contributions to the first
term only appear at vertical boundaries. At the left boundary, the material is stressed
and W is non-zero. However, by using a long enough specimen, the right boundary is
unstressed and W = 0. The second term in Eq. (1) is only non-zero where traction acts,
i.e. only at the left boundary.
4

q1
q2

P, D/2

y
x

H/2
H/2

P, D/2

P
upper beam, undeformed

Figure 3: Outer integration path for Double Cantilever Beam-specimen and rotations
associated with the Timoshenko beam theory.
Using the Timoshenko beam theory, we distinguish the rotation of the cross-section of the
beam, 1 , from the rotation of the neutral axis, 2 , cf. Fig. 3. These rotations are given
by the displacements, u1 and u2 , by

1
1 = u
y
(5)
u2
2 = x

From these, we identify the shear of the beam as = 2 R1 . The shear results in
strain energy which contributes to the first part of Eq. (1), W dy = P (1 2 )/2B,
Rwhere B is the out-of-plane width of the specimen. The second part of Eq. (1) gives
Ti ui,x dC = P 2 /B. Thus, for the upper beam of the DCB-specimen we collect the
contribution P (1 + 2 )/2B to J. By doing the same exercise for the lower beam, we arrive
at
2P m
(6)
B
where m (1 + 2 )/2 is the average of the cross-section and neutral axis rotations.3 By
equating the two expressions for J, Eq. (4) and Eq. (6), we get an expression from which
the cohesive stress can be calculated,
J=

2 (P m )
(7)
B w
As shown in [14], this expression is not limited to beam theory and it can be derived by
direct application of the concept of energetic forces for 3D-elasticity. It is also readily
extended to large deformations by replacing with sin , cf. [17].
=

3
As demonstrated in [16], the error is often small in using the cross-section rotation instead of the
average rotation in FE-analyses.

A
P

y
H/2
H/2
a

b
x

aL

Figure 4: Integration path in ENF-specimen. Paths A and C are taken just above and
below the cohesive zone. Path D is taken below the loading point.
2.2. Mode II
Figure 4 shows an outer integration path used to derive a relation between the external
load P and J. The contributions to J are denoted JA , JB , JC , and JD , respectively. If b is
large enough, it is safe to neglect JB and JD as compared to JA and JC , cf. [18] and [19].
In this case, the shear stress in the cohesive zone switches sign below the loading point.
With a too small b, the shear stress switches sign at a point further to the right. The
contributions JA and JC to J originate from the second term in Eq. (1), viz.
Z
Z
Z b
JA + JC =
Ti ui,x dx Ti ui,x dx =
v dx
(8)
A

where v uC
x ux and ( ) denotes ( )/x. A careful re-derivation of the theory in [18] and
extending to the Timoshenko beam theory and an asymmetrically loaded ENF-specimen
gives

E1 H
v +
(9)
16
where = 3P/2BH for 0 < x < b and = 3(1 )P/2BH for b < x < b + L is
identified as the shear stress according to Jouravskis theory in the plane y = 0; E1 denotes
the Youngs modulus in the fibre direction and H is the thickness of the ENF-specimen,
cf. Fig. 2. Multiplying both sides of Eq. (9) with v and integrating from x = 0 to x = b
gives
Z b
Z b
Z
E1 H b
v dx =
v dx
(10)
v v dx +
16 0
0
0
=

whereby we have constructed the negative of the right-hand-side of Eq. (8). The righthand-side of Eq. (10) can be integrated to form
Z b

E1 H  2
v dx =
v (b) v (0)2 + [v(b) v(0)]
(11)
32
0

Following [18], we now neglect v(b) as compared to v(0) and v (b) as compared to v (0) =
24P a/E1 BH 2 . The latter expresses the boundary conditions that the bending moments
in each leg at x = 0 equal P a/2 and the horizontal stress resultants equal zero at this
point. Collecting all terms in Eq. (1) yields
J

182 P 2 a2 3P v(0)
+
E1 B 2 H 3
2BH

(12)

where it should be noted that a is the initial crack length and v(0) is the shear deformation
at the initial position of the crack tip, cf. [18]. Hence, the experimental method to measure
J does not require monitoring of the position of the crack tip during crack propagation.
By comparing with the corresponding expression used in e.g. [18] and [19], we note
that the only difference appears due to the unsymmetrically positioned load.4 That is, the
extension to Timoshenkos beam theory does not influence J. The reason is that the elastic
energy due to shear deformation is inversely proportional to the height of a Timoshenko
beam. That is, this part of the energy is unaltered if the crack propagates infinitesimally
under prescribed loading; the shear energy per unit length at the crack tip before and
after crack propagation are the same. Thus, shear deformation of the beams does not
contribute in mode II. This is obviously not the case in mode I, where no energy is stored
in the region heading the crack tip, cf. Eq. (6). We also note that the first term in
Eq. (12) corresponds to the energy release rate in an unsymmetrically loaded and elastic
ENF-specimen where the influence of a cohesive zone is neglected, i.e. the expression to
be used when the conditions of LEFM are met.
With Eqs. (2) and (12) we arrive at


182 P 2 a2 3P v(0)
=
+
(13)
v E1 B 2 H 3
2BH
Thus, we measure the applied force P and the shear deformation v at the crack tip.

3. Material, specimens, experiments and evaluation


The material studied is a CFRP-laminate with all fibres in the longitudinal direction of the
test specimens. In order to minimize effects of fibre bridging, one of the laminas at the crack
is oriented slightly off-axis with a nominal direction 5 . The crack is formed by cutting one
of the lamina to a shorter length and replacing it by an equally thick Teflon film in the crack
area. In the ENF-experiments, the Teflon film remains between the crack surfaces in order
4

Note that the beam height is defined differently in [18] and [19].

Figure 5: Vertical DCB-specimen in test machine. The legs of the specimen are separated
with a prescribed rate during an experiment. Two LVDTs are fixed at the crack tip region
and measures the expansion of the specimen at this point.
to minimize friction. After manufacturing in an autoclave, the specimens are thoroughly
examined for defects by NDE-techniques. Before conducting the experiments, a wedge is
used to propagate the crack beyond the resin filled crack-tip-area formed during the curing
process. The new crack tip and crack length are carefully identified and measured.
Finite element models of the test specimens are analysed using Abaqus [20] in order to
verify the experimentally obtained cohesive laws. Since B < H, 2D plane stress FE-models
are used. In the FE-models, the composite material is modelled as a continuum using the
fully integrated element CPS4. The four-node cohesive element COH2D4 is used to model
a cohesive zone heading the crack tip.
3.1. Mode I
A total of four successful DCB-experiments are conducted. The nominal dimensions of the
specimens are H = 16 mm, B = 8.3 mm, a = 155 mm and L = 275 mm, cf. Fig. 2a. Each
specimen is measured individually and the data are used in the evaluation.
A custom-made test machine is used to conduct the experiments, cf. [21]. In the DCBexperiments, the loading points are separated, , with a prescribed rate, cf. Fig. 2a.
During an experiment, the reaction forces P , the rotation at the loading points and the
exp
separation, wext
, of the two points on the outside of the specimen at the position of the
crack tip are measured, cf. Fig. 5. A semi-inverse method, described below, is used to
derive the cohesive law (w) in pure mode I using Eq. (7).
In experiments, all data suffer from measurement errors. To limit the influence, we first
exp
adapt a Prony-series to the J(wext
)-data using a suitable number of terms, and afterwards
differentiate the series. The procedure is described in some detail in [21]. From Eq. (2)
8

1.2
1

0.8
J/Jcexp

exp
P/Pmax

0.6
0.4

0.6
0.4

0.2
0
0

0.8

0.2
0.5

1
1.5
exp
/
c

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
exp
w/w
ext,c

Figure 6: Graphs from DCB-experiments. Left: P graphs. As shown, the specimens


exp
remain virtually linearly elastic until the peak load is reached. Right: J wext
graphs.
we conclude that the fracture energy, JIc , is given by the maximum of J recorded in the
experiment. This maximum should correspond to the start of crack propagation. However,
a maximum is not always easy to identify. Therefore, a procedure is developed whereby a
preliminary cohesive law is first identified by a good guess of the fracture energy. A finite
element model is then used to simulate the experiment. By comparing the simulated P
curve to the experimental, iteratively better estimates of the fracture energy are derived,
cf. [22] for a more detailed description.
exp
Figure 6 shows the P and J wext
graphs. The P graphs indicate that LEFM
would work well with the current geometry. That is, the in-elastic zone at the crack tip is
exp
very small compared to the size of the specimen. As noted from the J wext
graphs, crack
propagation occurs with almost constant J. Note that Eq. (6) does not depend on a. That
is, we get the propagation value of J with high accuracy since we are not depending on
measuring the crack length during the conduction of the experiment. The fracture energy
varies within about 5% from the average value Jcexp .
Initial simulations of the specimen reveal that the specimens expand in the transverse
direction at the crack tip region. This expansion is an effect of the very limited stiffness
in the transverse direction and the small cohesive zone. Similar effects are not shown with
adhesive joints where the cohesive zone is substantially larger. At the moment of crack
propagation, i.e. when the cohesive stress is zero at the crack tip, the expansion right above
the crack tip is small. However, by moving less than a millimetre in the fibre direction,
the expansion is much larger. Moreover, the expansion right above the crack tip is larger
at the earlier stages of loading during which the cohesive zone expands at the crack tip.
exp
Thus, wext
does not equal w at the crack tip.
To back-out the cohesive law from the experiments, a simulation model is designed. In
the crack-tip region, the longitudinal length of the elements is 0.3375 mm. This corresponds
to about one seventh of a fully developed damage zone.5 In the transverse direction, the
5
The damage zone is here defined as the part where the cohesive stress is non-proportional to w. That
is, for w w0 , cf. Fig. 7.

s
s

s
s wc = 2JIc
JIc
w0

wc

w0

wexp
ext,c

wc

Figure 7: Cohesive model used to back-out cohesive law from DCB-experiments with
parameters
, w0 , wc . These are constrained by
= 2JIc /wc .
size of the continuum elements varies between 0.252 to 0.269 mm depending on which
experiment to simulate. Convergence studies are performed to secure that the results do
not depend on the element size.
A cohesive law with linear softening in Abaqus allows for a triangularly shaped (w)relation. Thus, three parameters govern the law: the peak stress,
, the corresponding
separation w0 and the critical separation wc , cf. Fig. 7. All of these can be varied
to achieve a good fit to the experiments. However, the fracture energy, i.e. the area
under the cohesive law, is considered to be measured with good accuracy; J in Eq. (6) is
independent of w. Thus, we constrain two of the parameters to give the measured fracture
energy. Hence, JIc =
wc /2, cf. Fig. 7. At this point, w0 and wc can be chosen freely while

= 2JIc /wc . As noted above, the transverse expansion and thus the difference between
wext and w, is virtually zero at the moment of crack propagation. This indicates that we
exp
can choose wc = wext,c
. However, careful examinations of the specimens show that the
LVDTs were not positioned exactly above the crack tip. The initial crack tip is more easily
identified after the crack has separated the specimen in two halves. Moreover, the LVDTs
leave marks on the surface of the specimen. Individual measurements on each specimen
give the initial crack length and the individual positions of the LVDTs. Typically, the
LVDTs were positioned less than one millimetre to the left of the crack tip, cf. Fig. 3.
Taking this into account, a suitable value of wc is chosen for each experiment. This value,
together with JIc set
= 2JIc /wc . At this point, only w0 remains to be determined. It
is noted that the specimen is linearly elastic until w0 is reached in the loading history.
Thus, J increases parabolically with w. By adapting w0 , a good agreement is reached in
exp
the linearly elastic regime. The procedure gives good agreement to the complete J wext
curve, cf. the left part of Fig. 8. The right part of Fig. 8 shows a comparison of simulated
and experimental versions of the F wext curve. Each experiment is analyzed individually
using this procedure, the resulting cohesive parameters are summarized in Table 1.

10

J/Jc

0.8
0.6

0.8
0.6

0.4

0.4

0.2

0.2

0
0

Simulation
Experiment

1
exp
P/Pmax

0.5

1
w/wc

1.5

0
0

0.5

1
w/wc

1.5

Figure 8: Comparison of DCB-experiment 1 and simulation. Left: Crosses show a typical


exp
J wext
graph. Dashes indicates an adapted J w relation that gives the simulated
sim
J wext relation indicated by the solid curve. Data are normalized in relation to critical
values. Right: Force-separation graph.

Table 1: Results from DCB-experiments. The results are normalized with respect to the
average values that are indicated with an overbar.
Experiment

w0 /w0

wc /wc

JIc /JIc

1
2
4
5

0.8
1.2
0.8
1.2

0.9
1.1
0.9
1.1

1.2
0.9
1.2
0.8

1.0
1.1
1.0
1.0

11

1.4
1.2

1.2

1
J/Jc

P/Pc

1
0.8
0.6

0.8
0.6

0.4

0.4

0.2

0.2

0
0

0.2 0.4 0.6 0.8


c
/

0
0

1.2 1.4

0.5

1
v/
vc

1.5

Figure 9: Graphs from ENF-experiments. Solid curves are related to experiments with
= 0.273 and dash-dotted curves to experiments with = 0.5. Left: P graphs. Right:
J v graphs.
3.2. Mode II
A total of five successful ENF-experiments are conducted. The nominal dimensions of the
ENF-specimens are H = 20 mm, B = 10.4 mm, a = 75 mm and L = 200 mm, cf. Fig. 2b.
Each specimen is measured individually and the data are used in the evaluation.
The experiments are performed using a servo-hydraulic testing machine (MTS322).
During the experiments the displacement of the loading point, , is increased with a
constant velocity. The force, P , the displacement, , and the shear deformation at the
crack tip, v, are measured continuously during the experiment. A digital image correlation
system, Aramis, is used to measure the shear deformation, v.
Figure 9 shows the P - and J-v graphs. In the first two experiments, the load is
applied centrally between the supports (=0.5). For this loading configuration, no clearly
visible maximum in the P - graph is observed. This indicates that the process zone ahead
of the crack tip extends towards the position of the load, which means that the conditions
used in the derivation of Eq. (12) might be violated when fracture is approached. As a
precaution, the load is applied further away from the crack tip in the last three experiments
( = 0.273). These experiments display a clearly visible load maximum. It should be noted
that, even though the P --curves indicate that LEFM would work well with the latter
geometry ( = 0.273), the second term in Eq. (12) comprises about 40 % in average to
the total fracture energy.
The individual ENF-experiments are evaluated by adapting a Prony-series to the J(v)data. A subsequent differentiation yields the form resembling a bilinear stress-deformation
relation, cf. Eq. (13). The resulting cohesive parameters are given in Table 2. Verifying finite element simulations are performed using the experimentally obtained stressdeformation relation. The comparison is good for both configurations used ( = 0.5 and
0.273), cf. Fig. 10.
Two extra experiments (not reported in Table 2) are carried out on non-precracked
ENF-specimens. For these specimens the fracture energy is about twice as large as for
the pre-cracked specimens. Unstable crack growth is observed for both non-precracked
12

Table 2: Results from ENF-experiments. The results are normalized with respect to the
average values that are indicated with an overbar.
Experiment
3
4
5
6
7

/ v0 /
v0
1.4
1.1
1.0
0.8
0.7

vc /
vc

JIIc /JIIc

0.8
0.8
1.2
0.9
1.2

1.2
1.0
1.2
0.8
0.9

0.500
0.500
0.273
0.273
0.273

1.2
1.3
0.6
0.9

Exp
Sim

1
0.8
P/Pc

P/Pc

0.8
0.6

0.6

0.4

0.4

0.2

0.2

0
0

Exp
Sim

0.2 0.4 0.6 0.8


v/vc

0
0

1.2 1.4

0.2 0.4 0.6 0.8


v/vc

1.2 1.4

Figure 10: Comparison of P v relations between ENF-experiments and simulations. Left:


Experiment 3 with = 0.5. Right: Experiment 5 with = 0.273.
specimens, indicating that the fracture energy decreases rapidly as the crack propagates
through the resin filled crack-tip-area formed during the curing process.
3.3. Cohesive properties
The focus in the present paper is the development of methods to determine the cohesive
relations for initiation of delamination of composites in mode I and mode II. The specific
composite material examined is of secondary importance. In order to emphasize this, the
experimentally determined cohesive properties presented in Tables 1 and 2 are given in
normalized form. However, some additional comments to complete the information are
necessary.
For the specific CFRP composite studied, the maximum cohesive stresses in modes I
and II are about the same size, i.e. /
1. On the other hand, the critical deformation
in mode II is twice the corresponding value for mode I, i.e. vc /wc 2. Thus, the fracture
energy in mode II is twice the fracture energy in mode I, i.e. JIIc /JIc 2. Regarding the
shapes of the cohesive relations, v0 /vc 0.2 and w0 /wc 0.5, indicating that mode II
exhibits a more pronounced softening behaviour than mode I.

13

4. Analysis of previously reported experimental data


In [8], experiments are reported with a material6 with very similar properties as the material
studied in the present paper. However, the evaluation method used in [8] is based on linear
elastic fracture mechanics (LEFM) and yields a fracture energy in mode II which is highly
dependent on the length of the starter crack; GIIc =1002 N/m for a=35 mm and GIIc =742
N/m for a=15 mm. In the experiments, the dimensions are: span length between the
supports 2L = 100 mm, width B = 20 mm and total beam height H = 3.1 mm. The
elastic properties of the laminate are E1 = 120 GPa, E2 = 10.5 GPa, 12 = 0.3 and G12
= 5.25 GPa. Directions 1 and 2 correspond to the fibre direction and transverse direction
in Fig. 2b, respectively. In order to demonstrate the strength of the method presented in
sections 2.2 and 3.2, the ENF-experiments reported in [8] are re-evaluated.
According to beam theory, and assuming the fracture energy to be independent of crack
growth, the crack length must exceed a certain critical length, acr , in order to achieve
stable crack growth under controlled displacement. For the ENF-specimen the critical
crack length is acr 0.693L= 34.6 mm, cf. [23].7 That is, the stability condition is
only fulfilled with the longer starter crack (a=35 mm). An important note is that in [8]
no further pre-cracking to sharpen the starter crack was made after manufacturing the
specimens. Thus, as discussed in section 3.2, the mode II fracture energy is most likely
large as compared to the value for a propagating crack. This presumption is supported by
the fact that unstable crack growth is reported for all mode II experiments in [8], indicating
that the fracture energy decreases rapidly as the crack propagates through the resin filled
crack-tip-area formed during the curing process.
In [8], a crack length correction is used to account for the flexibility at the crack tip.
With this correction the strain energy release rate for the ENF-specimen is evaluated as


2 a2
E1 H 2
GII = 2E9P
1 + 36KG
(14)
2 3
2
1B H
12 a
where K=5/6 is the cross-sectional shear factor. The first term equals the first term in
Eq. (12) with = 1/2 as used in these experiments.The second term is the correction for
the extra flexibility at the crack tip.
As mentioned above, the method to evaluate the ENF-experiments in [8] (cf. Eq. (14))
yields a fracture energy which is highly dependent on the length of the starter crack. The
reason for the discrepancy is that the evaluation method used in [8] is based on LEFM.
Since no measurements of the shear deformation at the crack tip are reported in [8], Eq.
(12), based on a large-scale process zone, cannot be used directly to obtain improved values
of the fracture energy. However, the two different values for GIIc reported for two different
crack lengths in [8] can be used to extract the critical value of the shear deformation, vc , in
6
In [8] specimens are produced from the toughened resin HTA/6376C carbon/epoxy prepreg from Ciba
Geigy. The layup is [012 //(5/04 )S ], where // indicates the plane of delamination.
7
The fact that the specimens in [8] are orthotropic and that a cohesive zone is present, will reduce the
critical crack length values by a few percent as compared to the values predicted by beam theory, cf. [23].

14

an indirect manner. Therefore, to set the material parameters in the cohesive law for mode
II (
, v0 and vc ), the procedure is as follows: The mode II fracture energy, JIIc , from the
ENF-tests in [8] is assumed to be a material parameter independent of the starter crack
length. The critical loads in the experiments in [8] are determined8 with Eq. (14) for both
the long and the short crack length. These are denoted Pc,l and Pc,s , where subscript l and
s denote a long and short starter crack, respectively. The assumption that the fracture
energy given by Eq. (12) is independent of the starter crack length yields (with =1/2),
2 2
2 2
9Pc,l
al
9Pc,s
as
3Pc,s vc
3Pc,l vc
+
=
+
(15)
2
3
2
3
2E1 B H
4BH
2E1 B H
4BH
from which the critical shear deformation vc =46 m is solved using the second equality.
The resulting fracture energy JIIc =1280 N/m then follows from the first equality in Eq.
(15). It is noted that this procedure does not require any presumptions regarding the shape
of the cohesive law. With a bi-linearly shaped cohesive law, the peak stress is given by
= 2JIIc /vc = 56 MPa. The initial stiffness, kt = exp /v|(w,v)=(0,0) , is taken from the
experiments reported in section 3.2 and the value of v0 is set accordingly, i.e v0 = /kt =
28 m.
Table 3 shows the results from the present evaluation and the original evaluation of
the experiments in [8]. The last column contains the critical forces, Pc,sim , obtained in
FE-simulations9 using the parameters of the cohesive law (
, v0 and vc ) determined above.
It is first noted that the critical forces from the FE-simulations agree very well with the
experimental values, Pc,exp , for both crack lengths. The maximum difference is about 3
%. Secondly, we note that the fracture energy, JIIc , is substantially larger than the GIIc values reported in [8]; 73 % for the short crack (a=15 mm) and 28 % for the long crack
(a=35 mm). The reason for the large discrepancy is that the evaluation of JIIc assumes the
existence of a process zone heading the crack tip while the method used in [8] is based on
LEFM. That is, with GIIc according to Eq. (14) the fracture process is assumed to occur
at a point at the crack tip. It can also be mentioned that the terms in Eq. (15) involving
the critical shear deformation, vc , comprises 44 % of the fracture energy for the short crack
and 22 % for the long crack. Hence, consideration of the local shear deformation at the
crack tip is crucial if the correct fracture energy is to be captured. Since Eq. (14) is based
on Timoshenkos beam theory without consideration of the flexibility of the cohesive zone
(cf. [8]), it consequently underestimates the fracture energy.

JIIc =

It is here assumed that the influence of gravity and forces from the dial guages can be neglected, cf.

[8].
9

The 2D FE-models used resemble the ones described in the beginning of section 3, but plane strain is
assumed due to the large width, B.

15

Table 3: Results from alternative evaluation of ENF-experiments reported in [8].


a0 (mm)

GIIc (N/m)

Pc,exp (N)

JIIc (N/m)

Pc,sim (N)

15
35

742
1002

1007
508

1280
1280

1040
512

5. Conclusions
Methods to accurately measure the cohesive laws associated with the crack-tip region are
developed. The cohesive laws are measured using the path-independent J-integral. The
measured cohesive law accurately represents the fracture processes at a delamination cracktip in mode I and II, respectively. However, the cohesive zone associated with the crack-tip
region is very small and the identification of the location of the crack-tip is crucial to achieve
the correct cohesive stress for the DCB-experiments. A semi-inverse method is developed to
accurately back-out the cohesive law for mode I. It is also shown that experiments in mode
II with apparent LEFM-behaviour can give large process zones. By ignoring the inelastic
zone and its contribution, the fracture energy in mode II can be severely underestimated.
Acknowledgement
The authors would like to acknowledge funding from the Swedish National Aeronautical
Research Program (NFFP4) in support of this work through the joint project KEKS with
Saab AB.
References
[1] Sorensen L, Botsis J, Gm
ur Th, Humbert L. Bridging tractions in modeI delamination:
Measurements and simulations. Composites Science and Technology 2008;68:23502358.
[2] G. Barenblatt. The mathematical theory of equilibrium cracks in brittle fracture.
Advances in Applied Mechanics 1962;7:55-129.
[3] Hillerborg A, Modeer M, Petersson PE. Analysis of crack formation in concrete by
means of fracture mechanics and finite elements. Cement and Concrete Research
1976;6:773-782.
[4] Backlund J. Fracture analysis of notched composites. Computers and Structures
1981;13:145-154.
[5] Janson J, Hult J. Fracture mechanics and damage mechanics - A combined approach.
Journal de Mecanique Appliquee 1977;1:69-84.
16

[6] Stigh U. Initiation and growth of an interface crack. In: Mechanical behaviour of
adhesive joints. Paris, France: Pluralis, 1987.
[7] Needleman A. A Continuum Model for Void Nucleation by Inclusion Debonding. Journal of Applied Mechanics 1987;54:525-531.
[8] Juntti M, Asp LE, Olsson R. Assesment of Evaluation Methods for the Mixed-Mode
Bending Test. Journal of Composite Technology and Research 1999;21:37-48.
[9] Rice JR. A path independent integral and the approximative analysis of strain concentration by notches and cracks. J Appl Mech 1968;88:379-386.
[10] Nilsson F. Fracture mechanics from theory to applications. Stockholm, Sweden: Department of Solid Mechanics, KTH, 2001.
[11] Suo Z, Bao G, Fan B. Delamination R-curve phenomena due to damage. J Mech Phys
Solids 1992;40:1-16.
[12] Olsson P, Stigh U. On the determination of the constitutive properties of thin interphase layers - An exact inverse solution, Int J Fract 1989;41:R71-R76.
[13] Andersson T, Stigh U. The stress-elongation relation for an adhesive layer loaded in
peel using equilibrium of energetic forces. Int J Solids Struct 2004;41:413-434.
[14] Stigh U, Andersson, T. (2000) An experimental method to determine the complete
stress-elongation relation for a structural adhesive layer loaded in peel. Proceedings of
the 2nd ESIS TC4 Conference on Polymers and Composites, Eds. J.G. Williams and
A. Pavan, pp. 297-306. ESIS Publication 27, Elsevier, Amsterdam.
[15] Srensen BF, Jacobsen TK. Large-scale bridging in composites: R-curves and bridging
laws. Composites Part A 1998;29A:1443-1451.
[16] Biel A, Stigh U. An analysis of the evaluation of the fracture energy using the dcbspecimen. Archives of Mechanics 2007;59:311-327.
[17] Nilsson F. Large displacement aspects on fracture testing with double cantilever beam
specimen. International Journal of Fracture 2006;139:305-311.
[18] Alfredsson KS. On the instantaneous energy release rate of the end-notch flexure
adhesive joint specimen. International Journal of Solids and Structures 2004;41:47874807.
[19] Leffler K, Alfredsson KS, Stigh U. Shear behaviour of adhesive layers. International
Journal of Solids and Structures 2007;44:530-545.
[20] Abaqus 6.11-2. Dassault Systemes Simulia Corp. Rising Sun Mills, 166 Valley Street,
Providence, RI 02909-2499, USA.
17

[21] Andersson T, Biel A. On the effective constitutive properties of a thin adhesive layer
loaded in peel. International Journal of Fracture 2006;141:227-246.
[22] Carlberger, T., A. Biel, Stigh U. Influence of temperature and strain rate on cohesive properties of a structural epoxy adhesive. International Journal of Fracture
2009;155:155-166.
[23] Alfredsson KS, Stigh U. Stability of beam-like fracture mechanics specimens. Engineering Fracture Mechanics 2012;89:98-113.

18

You might also like