You are on page 1of 19

Engineering Fracture Mechanics 69 (2002) 865–883

www.elsevier.com/locate/engfracmech

Experimental and numerical investigations of mixed mode


crack growth resistance of a ductile adhesive joint
K.S. Madhusudhana, R. Narasimhan *
Department of Mechanical Engineering, Indian Institute of Science, Bangalore 560 012, India
Received 19 February 2001; received in revised form 7 August 2001; accepted 20 August 2001

Abstract
Polymeric adhesive joints are extensively employed in various industrial and technological applications. It has been
observed that in ductile adhesive joints, interface fracture is a common mode of failure which may involve stable crack
propagation followed by catastrophic growth. The objectives of this paper are to investigate the effects of bondline
thickness and mode mixity on the steady state energy release rate Jss of such a joint. To this end, a combined experi-
mental and numerical investigation of interfacial crack growth is carried out using a modified compact tension shear
specimen involving two aluminium plates bonded by a thin ductile adhesive layer. A cohesive zone model along with a
simple traction versus separation law is employed in the finite element simulations of crack growth. It is observed that
Jss increases strongly as mode II loading is approached. Also, it enhances with bondline thickness in the above limit.
These trends are rationalized by examining the plastic zones obtained from the numerical simulations. The numerically
generated Jss values are found to agree well with the corresponding experimental results. Ó 2002 Elsevier Science Ltd.
All rights reserved.

Keywords: Interface fracture; Adhesive layer; Crack growth; Finite elements

1. Introduction

In recent years, adhesively bonded joints are used quite extensively in various industrial and techno-
logical applications including space technology, micro-electronic packaging, aerospace and automobile
industries. For example, epoxy and phenolic based adhesives are used to bond friction materials for au-
tomotive brake linings, clutch parts and transmission bands. Since a common mode of failure in adhesively
bonded components is interface fracture, it is important to understand the fracture behavior of a thin
ductile adhesive layer sandwiched between two elastic adherends.
It has been observed experimentally [1–3] that the fracture behavior of the above system is associated
with stable crack propagation followed by a catastrophic growth. Further, it has been found that under
mode I loading, the fracture energy shows little variation with the thickness of the layer. However, under

*
Corresponding author. Tel.: +91-80-309-2589; fax: +91-80-360-0648.
E-mail address: narasi@mecheng.iisc.ernet.in (R. Narasimhan).

0013-7944/02/$ - see front matter Ó 2002 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 1 3 - 7 9 4 4 ( 0 1 ) 0 0 1 1 0 - 2
866 K.S. Madhusudhana, R. Narasimhan / Engineering Fracture Mechanics 69 (2002) 865–883

mode II loading, the fracture energy increases considerably with the layer thickness [1] and is much higher
than the mode I fracture energy. Also, the results obtained in [1] show that in the limit as the adhesive
layer thickness h approaches zero, the fracture energy corresponding to modes I, II and III converge to a
single value C0 , which is the intrinsic work of separation. It has been proposed in Ref. [1] that the above
noted enhancement in fracture toughness with layer thickness is caused by plastic deformation in the
adhesive layer. This is supported by the observation of a large plastic zone, which, in some cases, extends
up to thousand times the layer thickness [2]. An estimate of the plastic zone dimension in interleaved
composites based on elastic stress fields was also obtained by Odzil and Carlsson [4]. They observed
that the trends in variation of fracture toughness with interleaf thickness, obtained from experiments,
were consistent with predicted plastic zone height and yield strength. It must be noted that the experi-
ments conducted in Refs. [1–3] are limited to the pure modes I, II or III. Swadener and Liechti [5] carried
out an experimental investigation of mixed mode interface crack propagation in an epoxy layer bonding a
glass and an aluminium block. They observed asymmetry in the variation of fracture toughness with
mode mixity and attributed this asymmetry to the pressure sensitive nature of yielding in the adhesive
layer.
Quasi-static crack propagation along the interface of thin constrained ductile layers has also been
simulated in many numerical studies. Tvergaard and Hutchinson [6] employed a cohesive zone model and
simulated interfacial crack growth in a sandwiched metallic layer system under remote mode I conditions.
They studied the role of layer thickness and material strain hardening parameter on the plasticity contri-
bution to toughness. In a subsequent work, Tvergaard and Hutchinson [7] analyzed the effect of residual
stress, elastic modulus mismatch, and various model parameters on the fracture toughness of the system.
Roy Chowdhury and Narasimhan [8] adopted the cohesive zone approach to simulate quasi-static crack
propagation under small scale yielding conditions in a pressure sensitive polymeric adhesive layer. They
examined the effect of layer thickness, mode mixity and pressure sensitivity on the steady state fracture
toughness. It must be noted here that in the above studies [6–8], direct comparison was not made with
experimental results. Thus, the efficacy of the cohesive zone approach in predicting the interface fracture
behavior observed in constrained ductile layers needs to be assessed. Also, such a direct comparison will
enable rationalization of the experimental trends in both qualitative and quantitative terms. The above
reasons provide the impetus for carrying out a combined experimental and finite element investigation of
interfacial crack growth in a ductile adhesive joint.
Hence, in this paper, interface crack growth experiments are first carried out using the modified
compact tension shear (CTS) specimen proposed recently by De and Narasimhan [9]. This specimen
comprises of two aluminium plates bonded by a thin adhesive layer. A ductile epoxy based structural
adhesive, LAPOX is employed in this work. The thickness of the layer is varied from 0.075 to 0.5 mm in
the experiments. It has been shown in Ref. [9] that the entire range of mixed mode loading from mode I to
II can be realized with the modified CTS specimen by changing the angle of loading with respect to the
crack plane. The load and crack extension are simultaneously measured in the experiments and analyzed
using a LEFM calibration of the specimen to obtain the energy release rate. Next, finite element simu-
lations of stable crack propagation along the interface in the fracture specimen are undertaken by em-
ploying a cohesive zone model with a simple traction–separation law. The cohesive zone parameters are
obtained by comparing the results of the simulations and the corresponding experiments for a few cases.
The adhesive layer is taken to obey the Drucker–Prager yield condition with power law hardening, while
the aluminium adherends are assumed to be linear elastic. The crack growth resistance curves are predicted
from the simulations for the entire range of layer thickness considered in the experiments as well as mode
mixity from mode I to II. The variations of the steady state energy release rate Jss (determined from the
simulated resistance curves) with layer thickness and mode mixity agree well with the corresponding ex-
perimental results. These trends are rationalized by examining the steady state plastic zones obtained from
the simulations.
K.S. Madhusudhana, R. Narasimhan / Engineering Fracture Mechanics 69 (2002) 865–883 867

2. Experimental investigation

2.1. Materials and specimen preparation

The interface fracture specimen proposed by De and Narasimhan [9] is used in this work. It is a modified
version of the CTS specimen suggested by Richards and Benitz [10] for mixed mode fracture testing of
homogeneous materials like engineering alloys. The modified CTS specimen is prepared by joining two
rectangular plates of aluminium alloy (having dimensions of 74  90 mm) by an adhesive layer of thickness
h, which is varied from 0.075 to 0.5 mm in the experiments conducted here. The thickness of the specimen is
taken as 10 mm. This specimen is employed along with a loading fixture (see Fig. 1) which enables several
combinations of normal (mode I) and shear (mode II) loading. The loading angle with respect to the crack
plane, a (Fig. 1), can be changed by choosing the appropriate hole on the loading fixture. De and Nara-
simhan [9] have shown that the entire range of loading from mode I to II can be obtained with this single
specimen geometry.
The adhesive employed in this work is LAPOX, a thermosetting epoxy based structural adhesive
manufactured and supplied by Cibatul Ltd., India. The adherends were machined from a plate of high
strength aluminium alloy. The specimen preparation involved milling the bonding surfaces of the adherends
to achieve flatness, and subsequently abrading against emery paper of 150 grit size to induce adequate
roughness. This was followed by ultrasonic cleaning with acetone for degreasing and removal of dust.
Spacers in the form of filament tapes of pre-determined thickness were introduced at the corners of the
bonding surfaces in order to control the bond line thickness. The adhesive was prepared by mixing the resin
and hardener in the specified proportion and was applied to the surfaces of the adherend plates. These are

Fig. 1. Schematic of the adhesively bonded, interface fracture specimen along with loading grips. The angle a of the line of the ap-
plication of the load can be varied by changing the loading holes. All dimensions are in mm.
868 K.S. Madhusudhana, R. Narasimhan / Engineering Fracture Mechanics 69 (2002) 865–883

then bonded together in a specially designed fixture and held together for the stipulated curing period. A
crack was introduced by placing a strip of thin release film, 10 lm thick, across the interface before bonding
to a length of 45 mm (so that the initial crack length a ¼ 0:5w). After curing, the bond line thickness was
measured with an optical microscope at several locations and an average value was taken. The variation in
the thickness along the bond line was usually very small. In order to measure the extent of stable crack
growth, a special crack propagation gauge (supplied by Measurements Group, USA) was installed across
the bond line. This gauge comprises of 20 thin conducting strands (with spacing of 0.25 mm) arranged in
parallel. It was installed in such a way that the first strand coincides approximately with the location of the
crack tip.

2.2. Experimental procedure

The experimental setup consists of the fracture specimen along with the loading grips mounted on an
INSTRON 8502 testing machine. The load and actuator signals from the INSTRON machine are logged
into a multi-channel data acquisition system. The crack propagation gauge is connected to an electrical
circuit which involves a 12 V DC supply and 50 X shunt resistance as shown in Fig. 2. The propagation of
the crack through the gauge pattern causes successive open circuiting of the strands which is reflected as
step increases in the output voltage from the electrical circuit shown in Fig. 2. This output is also logged
into the data acquisition system. Since the distance between each strand is 0.25 mm, one step in the output
voltage (i.e., the breaking of a strand) signifies crack extension of 0.25 mm. Since the crack propagation
gauge used here contains twenty such strands, it is capable of detecting upto 5 mm of stable crack growth.
The fracture tests were conducted in displacement controlled mode at a rate of 0.005 mm/s, for various
loading angles a (see Fig. 1), ranging from 0° to 90° and for different bond line thicknesses from 0.075 to 0.5
mm. The test pertaining to each case was performed a few times, till three repeatable results (with fracture
load varying within 5%) were obtained. In addition to the above, uniaxial compression tests were also
conducted on bulk adhesive specimens of length equal to 20 mm and diameter equal to 15 mm. The
compression specimen was placed between two plates with grease applied at the ends and loaded in dis-
placement controlled mode such that an engineering strain rate of _ ¼ 104 =s was maintained. The me-
chanical properties like Young’s modulus and yield strength in compression of the adhesive are determined
from the stress–strain curves obtained in these tests.

Fig. 2. Electrical circuit used for amplifying the crack gauge signal.
K.S. Madhusudhana, R. Narasimhan / Engineering Fracture Mechanics 69 (2002) 865–883 869

2.3. Linear elastic calibration of the fracture specimen

A linear elastic finite element analysis of the modified CTS specimen corresponding to the chosen ad-
hesive–adherend combination was carried out to obtain J integral solutions and mode mixity (see also Ref.
[9]). The material properties of the adhesive determined from uniaxial compression tests (see Section 2.4)
were used in these analyses. The domain representation proposed by Li et al. [11] was used to evaluate the J
integral. The interaction integral [12] was employed for extracting the complex stress intensity factor K
associated with the interface crack. The mode mixity angle w, which gives the relative amounts of in-plane
shear and normal tractions on the interface, was calculated from K as
 ie

1 ImðKL Þ
w ¼ tan ; ð1Þ
ReðKLie Þ
where e is a bimaterial constant [13] and L, a reference length scale which is taken as 100 lm in this work. A
range of a=w starting from a=w ¼ 0:5 was considered [9] so that stable crack growth under small scale
yielding conditions can be taken into account while using the calibration. These calibration results enabled
computation of the energy release rate by using the experimentally determined load and extent of crack
propagation.

2.4. Results

The Young’s modulus E and yield strength in compression rc0 for the adhesive LAPOX were determined
as 3.0 GPa and 51 MPa, respectively, from the uniaxial compression tests. Also, the hardening exponent n
n
was obtained by fitting a curve of the type =c0 ¼ ðr=rc0 Þ , where c0 ¼ rc0 =E, to the stress–strain curve
beyond the yield point. The value of n was found to be around 4.
The experimental results obtained from the fracture test corresponding to a ¼ 90° and bond line
thickness of h ¼ 0:5 mm are presented in Fig. 3. Here, the load versus time is shown in Fig. 3(a). In Fig. 3(b)
and (c), detailed views of the load and crack gauge output are presented from from t ¼ 115–140 s. It may be
seen from Fig. 3(a) that the load increases monotonically upto about t ¼ 140 s and then abruptly drops to
zero. On examining Fig. 3(c), a series of six steps in the crack gauge output can be noticed between t ¼ 133
and 138 s which implies stable crack growth of around 1.5 mm. During the above phase, the load remains
almost constant at 3.2 kN (see Fig. 3(b)). In some experiments, an unidirectional electrical resistance strain
gauge was installed 8 mm above the crack tip on one of the adherends along a line perpendicular to the
interface and its output was also logged into the data acquisition system. The strain gauge signal first
increased monotonically and then showed a gradual reduction during the time period where the steps
occurred in the crack gauge output. Thus, it is confirmed that the steps in the crack gauge output seen in
Fig. 3(c) correspond to slow, stable crack growth. However, there is some uncertainity about the exact
crack initiation time as signalled by the crack propagation gauge. This is attributed to the fact that unlike in
bulk materials, the crack displacements are small in adhesive joints. Hence, a small amount of crack growth
may occur before the first strand of the crack propagation gauge is sufficiently stretched and broken.
The results obtained for a ¼ 30° are qualitatively similar to a ¼ 90° and are not shown here. In this case,
the extent of stable crack growth is estimated as around 2.1 mm. The results for a ¼ 0° and h ¼ 0:5 mm are
presented in Fig. 4(a)–(c). Here, the load versus time is shown in Fig. 4(a). In Fig. 4(b) and (c), detailed
views of the load and crack gauge output from t ¼ 973–985 s are presented. It may be noticed from the Fig.
4(c) that the steps in the crack gauge output are not equally spaced and are of varying height unlike the case
a ¼ 90° (Fig. 3(c)). Since some of the steps may correspond to breakage of more than one strand, the exact
amount of stable crack growth is determined as 3.75 mm by correlating the step increases in the crack gauge
output and the number of strands broken by using manufacturer’s catalog. During the above phase (973–
985 s), the load increases by about 0.6 kN (see Fig. 4(b)).
870 K.S. Madhusudhana, R. Narasimhan / Engineering Fracture Mechanics 69 (2002) 865–883

Fig. 3. (a,b) Load (in kN) versus time and (c) crack gauge output (in volts) versus time corresponding to a ¼ 90° (mode I) and h ¼ 0:5
mm.

The fracture loads (corresponding to onset of unstable crack growth) for different layer thicknesses and
angles of loading are summarized in Table 1. These are average values obtained from three repeatable tests
performed for each case. From Table 1, it may be noticed that the fracture load increases strongly (by a
factor of 3.5) as the loading angle changes from 90° to 0°. However, the effect of layer thickness h on the
fracture load is negligible for a ¼ 90°. A decrease in the fracture load of about 10% is noticed for a ¼ 0° as
the bond line thickness is reduced from 0.5 to 0.075 mm. The experimental results for the energy release
rate, obtained from the LEFM calibration described in the Section 2.3, will be presented along with those
predicted by the numerical simulations using the cohesive zone model in Section 3.5.

3. Numerical simulations

In this section, numerical simulations of quasi-static crack growth along the adhesive–adherend interface
are carried out using cohesive zone model. In the following, the constitutive models and computational
aspects are briefly presented. The simulation results are discussed in Section 3.5.

3.1. Constitutive model

Experimental evidence suggests that yield behavior of polymeric materials is considerably different from
that of metals. In particular, the yield stress under uniaxial tension for a polymer is different from that in
K.S. Madhusudhana, R. Narasimhan / Engineering Fracture Mechanics 69 (2002) 865–883 871

Fig. 4. (a,b) Load (in kN) versus time and (c) crack gauge output (in volts) versus time corresponding to a ¼ 0° (mode II) and h ¼
0:5 mm.

Table 1
Values of fracture load corresponding to different angles of loading and layer thickness h for the adhesive LAPOX
Loading angle (a°) Fracture load (kN)
h ¼ 0:5 mm h ¼ 0:375 mm h ¼ 0:25 mm h ¼ 0:075 mm
90° 3.20 3.11 3.11 3.10
30° 5.4 5.25 5.10 4.9
15° 8.20 8.10 8.0 7.6
0° 11.2 10.8 10.5 10.1

uniaxial compression (see, for example, Bowden and Jukes [14]). Such a behavior can be explained by
assuming a pressure-dependent yield criterion. Hence, in this work, the Drucker–Prager model [15] is
employed to describe the constitutive response of the polymeric adhesive layer.
The Drucker–Prager yield function is given as [15]:
pffiffiffiffiffiffiffi  
Uðrij ; rc Þ ¼ 3J2 þ rm tan b  1  13 tan b rc ¼ 0; ð2Þ

where, J2 is the second invariant of the deviatoric part of the Cauchy stress rij , and rm ¼ rkk =3 is the
hydrostatic stress. Further, rc is the true yield stress in a uniaxial compression test and b is a pressure
sensitivity index which is usually less than 23° for polymers [14,16,17]. An associated flow rule is used so
that the plastic part of the total deformation rate is directed along the normal to the yield surface.
872 K.S. Madhusudhana, R. Narasimhan / Engineering Fracture Mechanics 69 (2002) 865–883

A complete description of the finite deformation elastic–plastic constitutive equations for the above model
is given in Ref. [18].
The true stress–strain response of the polymeric material in uniaxial compression is idealized here by a
piecewise power hardening law of the form:
 c c
c r =r0 rc 6 rc0 ;
¼ c n ð3Þ
c
0 ðr =r0 Þ rc > rc0 :
c

Here, rc0 is the initial yield stress, c0 ¼ rc0 =E is the initial yield strain under uniaxial compression and n is the
strain hardening exponent of the material. The initial yield stress under uniaxial tension, rt0 , is related to rc0
by,
 
rt0 1  13 tan b
¼  : ð4Þ
rc0 1 þ 13 tan b

3.2. Cohesive zone model

A simple traction versus separation law is employed to represent the mechanical response of the ad-
herend–adhesive interface as described below.
Let dn and dt be the normal and tangential components of the relative separation of two initially co-
incident material points across the interface in the vicinity of the crack tip. A non-dimensional effective
separation parameter k is defined as,
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 2  2
dn dt
k¼ c þ ; ð5Þ
dn dct

where, dcn and dct are critical values of dn and dt for the case of pure normal and tangential separation,
respectively. The normal and tangential components of the nominal traction vector across the interface are
given by [6],

rðkÞ dn rðkÞ dt dcn


Tn ¼ ; and Tt ¼ ; ð6Þ
k dcn k dct dct

where, r is an equivalent stress measure. The variation of rðkÞ is assumed as shown in Fig. 5. In this figure,
rf denotes the critical interfacial stress for pure normal separation (dt ¼ 0). The work of separation per unit
surface area of the interface is given by the area under the traction–separation curve shown in Fig. 5 as,

C0 ¼ 12rf dcn ð1  k1 þ k2 Þ; ð7Þ

irrespective of the combination of normal and tangential displacements across the interface. The viewpoint
adopted here is that the energy release rate during steady state crack growth, Jss , is the key parameter which
characterizes the fracture resistance of the adhesive joint. The plastic dissipation in the ductile layer, which
would depend on the mode mixity and layer thickness, contributes to enhancing Jss over C0 . In other words,
Jss would reduce to C0 , irrespective of the mode mixity, in the limit as h ! 0, which is supported by the
experimental results of Chai [1].
Tvergaard and Hutchinson [6] have shown that the shape parameters k1 and k2 have little influence on
the crack growth process. In this work, k1 and k2 are taken as 0.15 and 0.5, respectively. The other cohesive
zone parameters C0 , rf and dcn =dct are determined in Section 3.5.1.
K.S. Madhusudhana, R. Narasimhan / Engineering Fracture Mechanics 69 (2002) 865–883 873

Fig. 5. The prescribed traction–separation law for the interface.

3.3. Finite element procedures

An updated Lagrangian finite element procedure [19], in which the reference configuration coincides
instantaneously with the current configuration, is employed in this work. The cohesive zone model de-
scribed in Section 3.2 is used to represent the mechanical response of the interface ahead of the crack tip.
Since the typical crack speeds deduced from the crack length versus time records during the stable crack
growth phase, such as Figs. 3(c) and 4(c), are very small (of the order of 0.3 mm/s), quasi-static conditions
are assumed in the simulations. The virtual work rate due to the interface tractions is included in the
variational principle given in Ref. [19] (see Ref. [20] for details). The interface is discretized into cohesive
elements in which the relative separation (of initially coincident material particles) is interpolated using
linear shape functions. The stiffness matrix and force vector of these cohesive elements are derived from the
virtual work rate due to interface tractions [21] and are assembled along with those of the continuum ele-
ments. The cohesive elements lose their stress carrying capacity when k (see Eq. (5)) reaches a unit value.
This gives rise to a pair of new traction-free surfaces and the crack propagates by the length l of the element.
In order to control nodal instabilities that arise when a cohesive element loses its stress carrying capacity, a
special Ritz-finite element procedure [21] is employed. The novel feature of this procedure is that it enables
controlling directly the relative separation of initially coincident nodes across the interface.

3.4. Modelling aspects

An overall view of a typical finite element mesh for h ¼ 0:5 mm is presented in Fig. 6(a). The loading and
boundary conditions used in finite element simulations are indicated in this figure. The distribution of the
load among the three holes is described in Ref. [23]. The simulations are carried out by gradually increasing
the load and using the Ritz method [22] to control nodal instabilities as mentioned above. An enlarged view
of the region near the interface between the adhesive layer and the elastic adherends is displayed in Fig.
6(b). The crack which lies at the interface between the top adherend and adhesive layer is modelled as one
of zero thickness. In other words, the nodes along the upper and lower crack flanks coincide with one
another without being merged. Fig. 6(c) shows the details of the highly refined mesh in the region close to
the crack tip. A set of cartesian axes X1 , X2 , established with origin O at the crack tip in the undeformed
configuration is indicated in this figure. The uniform mesh region ahead of the crack tip shown in Fig. 6(c)
is used to model the crack growth. The dimension of the smallest elements l in Fig. 6(c) is about 0.02 mm
and has been chosen after conducting a mesh convergence study. Similar meshes with l ¼ 0:01, 0.005 and
874 K.S. Madhusudhana, R. Narasimhan / Engineering Fracture Mechanics 69 (2002) 865–883

Fig. 6. (a) Overall view of a typical mesh used for the finite element analysis. (b) View of the mesh near the interface region between the
adhesive layer and the elastic adherends. The thickness of the layer, h is indicated in the figure. (c) Details of the highly refined mesh
near the crack tip. The uniform mesh region ahead of the tip is used to model crack growth.

0.002 mm were employed in modelling adhesive layers with thickness h ¼ 0:25, 0.075 and 0.025 mm, re-
spectively. It must be mentioned that the above values of l are very small compared to the computed length
of the fracture process zone which is defined as the distance between the location where k ¼ k1 and the
current crack tip [24].
The material properties determined from the uniaxial compression tests are used in the analysis. Thus,
the Young’s modulus E, yield stress in compression rc0 and strain hardening exponent n are taken as 3 GPa,
51 MPa and 4, respectively. The value of Poisson’s ratio is assumed as 0.34. The pressure sensitivity index b
K.S. Madhusudhana, R. Narasimhan / Engineering Fracture Mechanics 69 (2002) 865–883 875

(refer Eq. (2)) is chosen as 15° which corresponds to a ratio, rc0 =rt0 of 1.2. This value is typical for most of
the polymers [17,25]. The aluminium adherends are taken to be linear elastic with Es ¼ 70 GPa and
ms ¼ 0:3.

3.5. Results and discussion

3.5.1. Determination of cohesive zone parameters


The intrinsic work of separation, C0 for the adhesive–adherend interface is first chosen tentatively as 0.2
N/mm based on the experimental results for a ¼ 90° (see Section 3.5.4). The peak normal stress rf and the
ratio of critical normal and tangential displacements (dcn =dct ) are determined by comparing the load versus
crack extension plots obtained from the numerical simulations pertaining to different choices of these
parameters with the corresponding experimental results. A similar procedure was adopted by Kolhe et al.
[26] for the extraction of cohesive zone parameters for nickel–alumina interface. To this end, two extreme
loading angles, a ¼ 90° and 0°, and a fixed layer thickness h ¼ 0:5 mm are considered, for which load P
versus crack extension Da variations are shown in Fig. 7(a) and (b). Since, there is some ambiguity in the
exact crack initiation load as mentioned in Section 2.4, the experimental variation of load versus crack
extension is marked as a continuous line starting from Da ¼ 0:25 mm in these figures.
It can be observed from Fig. 7(a) and (b) that the effect of dcn =dct on P versus Da is marginal for mode I
whereas it has a pronounced effect for mode II. On the other hand, C0 and rf have considerable effect on P
versus Da under both mode I as well as mode II. By trial and error, it is found that the combination of
cohesive zone parameters C0 ¼ 0:19 N/mm, dcn =dct ¼ 0:8 and rf ¼ 1:75rt0 represents the experimental results
closely (see Fig. 7(a) and (b)) for both modes I and II corresponding to the case h ¼ 0:5 mm. Hence, these
values are chosen in the subsequent simulations.

3.5.2. Contours of plastic strain


The contours of effective plastic strain ep in the adhesive layer around the crack tip when steady state
crack growth conditions are attained (see Section 3.5.3 for a precise definition of steady state) are presented
in Figs. 8 and 9 for a ¼ 0° and 30°, respectively. The current crack tip is indicated by a vertical arrow in all
the plots. Also, the thickness of the adhesive layer is marked as h at the right-hand side of each figure. The
same contour levels are presented for each thickness case in Figs. 8 and 9 to facilitate direct comparison.
On examining the plastic strain contours for a ¼ 0° (see Fig. 8(a)–(d)), it is evident that these contours
start a little behind the current crack tip at the top interface and spread predominantly in a direction
parallel to the interface. They extend to a long distance ahead of the current crack tip. It can be seen from
Fig. 8(a) and (b) that none of the contours shown fully occupy the layer thickness for h ¼ 0:5 and 0.25 mm.
Also, the contours pertaining to different levels in Fig. 8(a) and (b) are similar both in shape and size. Thus,
for example, the contour pertaining to ep ¼ 0:04 spreads to a distance of about 1.2 mm ahead of the current
crack tip and to a depth of about 0.1 mm in both the figures. Thus, the contours shown in Fig. 8(a) and (b)
are expected to be similar to those pertaining to crack growth at the interface between semi-infinite ductile
and brittle solids. On the other hand, for h ¼ 0:075 mm, the contour labelled as 1 is constrained by the
lower interface (see Fig. 8(c)), whereas for h ¼ 0:025 mm, contours marked as 1, 2 and 3 fully engulf the
adhesive layer (Fig. 8(d)).
The plastic strain contours for a ¼ 15° are qualitatively similar to a ¼ 0° and are not shown here. It can
be noted from Fig. 9(a)–(d), that for a ¼ 30°, the contours are more rounded in shape behind the current
crack tip as compared to a ¼ 0° indicating the presence of a strong mode I component. Also, the magnitude
of ep at comparable distances near the crack tip during steady state crack growth is much smaller than that
for a ¼ 0° (compare Figs. 8 and 9). This is attributed to much less shear deformation for a ¼ 30°. Further,
876 K.S. Madhusudhana, R. Narasimhan / Engineering Fracture Mechanics 69 (2002) 865–883

Fig. 7. Variation of load with crack extension for various choices of cohesive zone parameters corresponding to h ¼ 0:5 mm and
(a) a ¼ 90°, (b) a ¼ 0°.

similar to a ¼ 0°, the lower level contours fully engulf the layer thickness for h ¼ 0:075 and 0.025 mm (see
Fig. 9(c) and (d)), whereas, the contours shown in Fig. 9(a) and (b) are well contained within the adhesive
layer.
Following Roy Chowdhury and Narasimhan [18], the plastic zone size rp in the adhesive layer is defined
as the maximum distance of the effective plastic strain contour ep ¼ 0:001 from the current crack tip.
During the initial part of the crack growth, rp increases strongly with crack extension. However, as steady
state conditions are approached, the rate of increase of rp reduces. The plastic zone size during steady state
crack growth is denoted here by rpss . In Fig. 10, rpss is plotted against the layer thickness h for various loading
angles a. It can be observed from this figure that rpss increases initially with layer thickness and thereafter
remains almost constant (for h > 0:25 mm). This can be attributed to the fact that the plastic strain con-
tours are not constrained by the lower interface for large h (see Figs. 8 and 9). Hence, as h increases, rpss
K.S. Madhusudhana, R. Narasimhan / Engineering Fracture Mechanics 69 (2002) 865–883 877

Fig. 8. Contours of effective plastic strain, ep , in the adhesive layer around the crack tip, during steady state crack extension corre-
sponding to a ¼ 0° for (a) h ¼ 0:5 mm, (b) h ¼ 0:25 mm, (c) h ¼ 0:075 mm and (d) h ¼ 0:025 mm. All the dimensions marked on the
axes are in mm. The contour levels are same in all the figures.

attains the limiting value corresponding to steady quasi-static crack growth at the interface between semi-
infinite ductile and brittle solids. Similar variations of plastic zone size with layer thickness were observed
under remote mode II loading by Chai [2] and Chiang and Chai [25] for a toughened epoxy layer con-
strained between two aluminium adherends.
Since the increase in rpss with h for small layer thickness will result in increased energy dissipation in the
plastic zone, an enhancement in the steady state fracture toughness will occur. It may be noted from Fig. 10
that for a given h, rpss increases strongly as a approaches 0°. For example, corresponding to h ¼ 0:25 mm,
the values of rpss are 3.3, 9.2 and 23.2 mm for a ¼ 30°, 15° and 0°, respectively. This behavior is expected in
view of the larger plastic strain near the tip as a decreases, owing to larger shear deformation in the adhesive
layer. In fact, corresponding to the case a ¼ 0°, the plastic zone for some values of h extends up to 500 times
the layer thickness. The enhancement in rpss as a ! 0° is expected to reflect in the steady state fracture
toughness as will be seen in Section 3.5.4.
878 K.S. Madhusudhana, R. Narasimhan / Engineering Fracture Mechanics 69 (2002) 865–883

Fig. 9. Contours of effective plastic strain, ep , in the adhesive layer around the crack tip, during steady state crack extension corre-
sponding to a ¼ 30° for (a) h ¼ 0:5 mm, (b) h ¼ 0:25 mm, (c) h ¼ 0:075 mm and (d) h ¼ 0:025 mm. All the dimensions marked on the
axes are in mm. The contour levels are same in all the figures.

3.5.3. R-curves
In Fig. 11(a)–(d), the computed crack growth resistance curves (R-curves) for a ¼ 90°, 30°, 15° and 0°,
respectively, are shown. In these figures, the energy release rate J is plotted against crack extension Da for
different layer thicknesses. The variations of J with Da obtained from the experimental results for two
thicknesses, h ¼ 0:5 and 0.075 mm, are also indicated starting from Da ¼ 0:25 mm. In all these figures,
except Fig. 11(a), it can be observed that J increases sharply during the initial part of crack growth.
K.S. Madhusudhana, R. Narasimhan / Engineering Fracture Mechanics 69 (2002) 865–883 879

Fig. 10. Variation of plastic zone size during steady state crack growth, rpss , with the layer thickness for various loading angles.

Fig. 11. Crack growth resistance curves obtained from numerical simulations corresponding to different layer thicknesses for
(a) a ¼ 90°, (b) a ¼ 30°, (c) a ¼ 15° and (d) a ¼ 0°. The experimental results for h ¼ 0:5 and 0.075 mm are also shown.

However, the rate of increase in J gradually reduces with continued crack extension and J reaches a
plateau. For the purpose of definiteness, when the slope of the resistance curve falls below 2.5% of its
880 K.S. Madhusudhana, R. Narasimhan / Engineering Fracture Mechanics 69 (2002) 865–883

initial value, the corresponding magnitude of J is called as the steady state energy release rate and is
denoted as Jss . This stage is indicated by an ‘’ mark on each curve in Fig. 11(b)–(d). It should be
mentioned here that larger extents of stable crack growth were observed in the experiments (see Section
2.4) as compared to the simulations because they were carried out under displacement controlled con-
ditions.
It can be seen from Fig. 11(a) that there is no noticeable enhancement in J during crack growth for
a ¼ 90°. This is due to the small plastic zone size and low levels of plastic strain for a ¼ 90°, which indicates
lack of sufficient plastic dissipation in the adhesive layer. On examining Fig. 11(b)–(d), a strong en-
hancement in J (high tearing resistance) may be observed during crack growth as a decreases, which is
attributed to larger plastic zone and, hence, larger plastic dissipation in the adhesive layer as mode II
conditions are approached (see Section 3.5.2). The extent of stable crack growth (before attainment of
steady state condition) also increases as a decreases. For example, considering h ¼ 0:25 mm, this extent is
0.12, 0.25 and 0.31 mm for a ¼ 30°, 15° and 0°, respectively. In all the cases, the R-curves predicted by the
numerical simulations using the cohesive zone approach are in good agreement with the experimentally
measured J versus Da variations.
It can be seen from Fig. 11(b)–(d) that the initial part of the R-curves coincide for all values of thickness.
This is because, during early stages of crack growth, the plastic zone emanating from crack tip is small in
size and is not constrained by the lower interface. During later stages of crack growth, the R-curves per-
taining to different values of h deviate from each other. In this phase, the resistance to crack growth in-
creases with increase in h. For example, considering a ¼ 15° (see Fig. 11(c)), the J values at steady state for
h ¼ 0:025, 0.075, 0.25 and 0.5 mm are 0.32, 0.39, 0.47 and 0.54 N/mm, respectively. The corresponding
extents of stable crack growth are 0.09, 0.18, 0.25 and 0.3 mm, respectively. Another interesting observation
is that for a ¼ 30° (see Fig. 11(b)), the R-curve pertaining to any given h value saturates (i.e., steady state is
attained) soon after deviating from the other curves. However, for low a (see Fig. 11(c) and (d)), J per-
taining to any given h continues to increase for a considerable extent even after deviating from the other
curves.

3.5.4. Steady state fracture toughness


3.5.4.1. Effect of layer thickness. Fig. 12 shows the variation of steady state energy release rate Jss predicted
by numerical simulations with adhesive layer thickness h. Also, the corresponding J determined from the
experimental data is plotted against layer thickness in this figure to facilitate direct comparison. It can be
seen from Fig. 12 that the enhancement in Jss with h is negligible for mode I (a ¼ 90°). Similar observations
were made by Roy Chowdhury and Narasimhan [18] and Chai [1]. As a decreases, Jss shows more elevation
with h which becomes very pronounced for a ¼ 0°. However, it should be noted that the slope of Jss versus h
curve for all angles of loading decreases with increase in h. Thus, considering a ¼ 0°, Jss increases from 0.52
to 0.68 N/mm as h changes from 0.025 to 0.075 mm, whereas, it increases from 0.76 to 0.82 N/mm for
change in h from 0.25 to 0.5 mm. Thus, not much enhancement in fracture energy is expected for h greater
than 0.25 mm. Further, the qualitative similarity between the Jss versus h curves for different a values and
the corresponding variations of rpss with h shown in Fig. 10 should be noted. This clearly establishes the fact
that the elevation in Jss with h for loadings close to mode II is due to enhancement in the plastic zone size
and, hence, the plastic dissipation in the adhesive layer.
It can be seen from Fig. 12 that the agreement between the numerically predicted Jss versus h curves and
the corresponding experimental variations is close for the entire range of loading from a ¼ 0–90°. It should
be recalled that only the experimental results for two cases, viz., a ¼ 90° and 0° pertaining to h ¼ 0:5 mm,
are used to compare with the numerical results to determine the cohesive zone parameters. The close
agreement obtained in Fig. 12 for other a and h values validates the predictive capability of the approach
employed in this work for simulating interface crack growth using the cohesive zone method.
K.S. Madhusudhana, R. Narasimhan / Engineering Fracture Mechanics 69 (2002) 865–883 881

Fig. 12. Variation of the computed steady state energy release rate, Jss , with adhesive layer thickness, h, for different loading angles a.
The corresponding experimental results are also shown.

Fig. 13. Variation of the computed steady state energy release rate, Jss , with the mode mixity angle, w, for various values of layer
thickness.

3.5.4.2. Effect of mode mixity. In Fig. 13, the computed steady state energy release rate Jss is plotted
against mode mixity angle w (see Section 2.3) for several layer thicknesses. It can be noticed from Fig. 13
882 K.S. Madhusudhana, R. Narasimhan / Engineering Fracture Mechanics 69 (2002) 865–883

that Jss increases as w changes from 0° (mode I) to 90° (mode II). This elevation in Jss is quite steep for
w P 60°. Thus, for h ¼ 0:5 mm, Jss as w approaches 90° is more than three times that pertaining to w ¼ 0°.
Two factors cause this enhancement. Firstly, at a given load, the stresses ahead of the crack tip decrease
with increase in w. Hence, a higher load is required to meet the critical stress level in the cohesive zone
ahead of the tip to drive the crack as w increases. Secondly, and more importantly, the size of the plastic
zone, and thus, the contribution from plastic dissipation increases with increase in w. This can be clearly
seen from Fig. 10 which shows that the size of plastic zone during steady state crack growth is almost
seven times larger for a ¼ 0° as compared to a ¼ 30°. It should be mentioned here that the general trend
exhibited by the variation of Jss with w is similar to that obtained by Swadener and Liechti [5] from
experiments conducted on an epoxy layer bonding a glass and an aluminium block. However, unlike
in the present simulations, stable crack extension (with increasing J versus Da) was not modelled in their
work.

4. Conclusions

In this paper, a combined experimental and numerical investigation of stable crack growth in a ductile
adhesive joint was carried out. The modified CTS specimen was employed in the study with different layer
thicknesses and angle of loading (from mode I to II). The numerical simulations were carried out using the
cohesive zone model with the relevant parameters determined by comparing with the experimental results
for two representative cases. The following are the important conclusions of this work.
1. From the results of the mixed mode fracture tests, it is found that both the fracture load, as well as, the
extent of stable crack growth, increase as mode II conditions are approached.
2. The simulations show that the adhesively bonded system displays R-curve behavior (increasing J with
crack extension) except for loadings close to mode I. The R-curves corresponding to different layer
thicknesses initially coincide, but deviate from each other with continued crack extension. The slope of the
R-curves gradually decreases as steady state conditions are approached.
3. The steady state energy release rate, Jss , increases with layer thickness particularly for mode II predom-
inant loading. But it shows a tendency to saturate beyond h ¼ 0:25 mm. The above observation is ratio-
nalized by examining the plastic strain contours in the adhesive layer during steady state crack growth. The
plastic strain contours fully engulf the adhesive layer thickness for small h. However, for h > 0:25 mm, these
contours do not fully occupy the layer thickness and are qualitatively similar to those pertaining to crack
growth at the interface between semi-infinite ductile and brittle solids. Thus, not much enhancement in
fracture toughness caused by plastic dissipation in the adhesive layer occurs for h > 0:25 mm. In fact, there is
qualitative similarity between the Jss versus h variations and steady state plastic zone size (rpss ) versus h curves.
4. It is observed that for a given h, Jss increases as mode mixity angle w changes from 0° to 90°. This el-
evation is very steep for w P 60° which is attributed to the growth in rpss , and hence, the contribution from
plastic dissipation. Thus, for example, corresponding to h ¼ 0:25 mm, rpss for mode II predominant loading
(w close to 90°) is about seven times that for mode I predominant loading.
5. The numerically predicted Jss versus h variations are found to compare well with the corresponding
experimental results for all angles of loading. This good agreement between the simulations and the ex-
periments demonstrates the predictive capability of the cohesive zone approach used in this work to model
interfacial crack growth.

Acknowledgements

The authors would like to gratefully acknowledge the Indian Space Research Organization for financial
support through the sponsored project ISTC/ME/RN/90.
K.S. Madhusudhana, R. Narasimhan / Engineering Fracture Mechanics 69 (2002) 865–883 883

References

[1] Chai H. Shear fracture. Int J Fract 1988;37:137–59.


[2] Chai H. Micromechanics of shear deformations in cracked bonded joints. Int J Fract 1992;58:223–39.
[3] Chai H, Chiang MYM. A crack propagation criterion based on local shear strain in adhesive bonds subjected to shear. J Mech
Phys Solids 1996;44:1669–89.
[4] Odzil F, Carlsson L. Plastic zone estimates in mode I interlaminar fracture of interleaved composites. Engng Fract Mech
1992;41:645–58.
[5] Swadener JG, Liechti KM. Asymmetrical shielding mechanism in the mixed-mode fracture of a glass/epoxy interface. ASME J
Appl Mech 1998;65:25–9.
[6] Tvergaard V, Hutchinson JW. Toughness of an interface along a thin ductile layer joining elastic solids. Phil Mag A 1994;70:641–
56.
[7] Tvergaard V, Hutchinson JW. On the toughness of ductile adhesive joints. J Mech Phys Solids 1996;44:789–800.
[8] Roy Chowdhury S, Narasimhan R. A finite element analysis of quasistatic crack growth in a pressure sensitive constrained ductile
layer. Engng Fract Mech 2000;66:551–71.
[9] Deepankar D, Narasimhan R. Analysis of an interface fracture specimen for adhesively bonded joints. Int J Fract 1998;92:L35–40.
[10] Richard HA, Benitz K. Loading device for the creation of mixed mode in fracture mechanics. Int J Fract 1983;22:R55–8.
[11] Li FZ, Shih CF, Needleman A. A comparison of methods for calculating energy release rates. Engng Fract Mech 1985;21:405–21.
[12] Shih CF, Asaro RJ. Elastic plastic analysis of cracks on bimaterial interfaces. Part I: small-scale yielding. J Appl Mech
1988;55:299–316.
[13] Shih CF. Cracks on bimaterial interfaces: elasticity and plasticity aspects. Mater Sci Engng 1991;A143:77–90.
[14] Bowden PB, Jukes JA. The plastic flow of isotropic polymers. J Mater Sci 1972;7:52–63.
[15] Chen WF, Han DJ. Plasticity for structural engineers. Berlin: Springer; 1988.
[16] Quinson R, Perez J, Rink M, Pavan A. Yield criteria for amorphous glassy polymers. J Mater Sci 1997;32:1371–9.
[17] Brown N. Creep, stress relaxation and yielding. Engineered materials handbook, vol. 2, Engineering plastics, ASM International;
1987.
[18] Roy Chowdhury S, Narasimhan R. A finite element analysis of stationary crack tip fields in a pressure sensitive constrained ductile
layer. Int J Solids Struct 2000;37:3079–100.
[19] McMeeking RM, Rice JR. Finite element formulation for problems of large elastic–plastic deformation. Int J Solids Struct
1975;11:601–16.
[20] Needleman A. A continuum model for void nucleation by inclusion debonding. ASME J Appl Mech 1987;54:525–31.
[21] Roy Chowdhury S, Narasimhan R. A cohesive finite element formulation for modelling fracture and delamination in solids.
Sadhan a, Proc Ind Acad Sci 2000;25:561–87.
[22] Needleman A. Finite elements for finite strain plasticity problems. In: Lee EH, Mallet RL, editors. Plasticity of metals at finite
strain: theory, experiment and computation. Troy, NY 12181: RPI, 1982. p. 387–437.
[23] Aoki S, Kishimoto K, Yoshida T, Sakata M, Richards HA. Elastic–plastic fracture behavior of an aluminium alloy under mixed
mode loading. J Mech Phys Solids 1990;38:195–213.
[24] Tvergaard V, Hutchinson JW. The relation between crack growth resistance and fracture process parameters in elastic–plastic
solids. J Mech Phys Solids 1992;40:1377–97.
[25] Chiang MYM, Chai H. Plastic deformation analysis of cracked adhesive bonds loaded in shear. Int J Solids Struct 1994;31:2477–
90.
[26] Kolhe R, Tang S, Hui CY, Zehnder AT. Cohesive properties of nickel–alumina interfaces determined via simulation of ductile
bridging experiments. Int J Solids Struct 1999;36:5573–95.

You might also like