You are on page 1of 28

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

net/publication/321496888

The effects of curing temperature on fracture energy and cohesive parameters


for the adhesive Araldite 2015

Article  in  Journal of Adhesion Science and Technology · December 2017


DOI: 10.1080/01694243.2017.1410080

CITATIONS READS

5 111

2 authors, including:

Engin Erbayrak
Yildiz Technical University
14 PUBLICATIONS   7 CITATIONS   

SEE PROFILE

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

Kendi-kendini Onaran Akıllı Yapıştırıcı Kırılma Tokluğunun Nümerik ve Deneysel Olarak Belirlenmesi View project

All content following this page was uploaded by Engin Erbayrak on 07 September 2020.

The user has requested enhancement of the downloaded file.


Journal of Adhesion Science and Technology

ISSN: 0169-4243 (Print) 1568-5616 (Online) Journal homepage: https://www.tandfonline.com/loi/tast20

The effects of curing temperature on fracture


energy and cohesive parameters for the adhesive
Araldite 2015

Halil Özer & Engin Erbayrak

To cite this article: Halil Özer & Engin Erbayrak (2018) The effects of curing temperature on
fracture energy and cohesive parameters for the adhesive Araldite 2015, Journal of Adhesion
Science and Technology, 32:12, 1287-1312, DOI: 10.1080/01694243.2017.1410080

To link to this article: https://doi.org/10.1080/01694243.2017.1410080

Published online: 03 Dec 2017.

Submit your article to this journal

Article views: 381

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=tast20
Journal of Adhesion Science and Technology, 2018
VOL. 32, NO. 12, 1287–1312
https://doi.org/10.1080/01694243.2017.1410080

The effects of curing temperature on fracture energy and


cohesive parameters for the adhesive Araldite 2015
Halil Özer and Engin Erbayrak
Mechanical Engineering Department, Faculty of Mechanical Engineering, Yıldız Technical University, Istanbul,
Turkey

ABSTRACT ARTICLE HISTORY


This paper attempts to investigate the effects of curing temperature Received 22 March 2017
on the fracture energy, the glass transition temperature (Tg) and Revised 22 November 2017
cohesive parameters for the adhesive Araldite 2015. Relationship Accepted 23 November 2017
between curing temperature and the glass transition temperature KEYWORDS
was taken into account. Tensile tests were performed on the dogbone- Cohesive zone modeling;
shaped bulk specimens to evaluate the effect of curing temperature curing temperature; glass
on the mechanical properties of the adhesive. DCB test results were transition temperature;
used to obtain the cohesive laws of the adhesive Araldite 2015. The fracture properties; Araldite
exponential and PPR cohesive zone models were used to obtain some 2015
of the fracture properties. Inverse analyses were also performed, if the
experimental softening curves are incompatible with the numerical
ones. It was seen that softening behavior of the adhesive can be easily
controlled by the shape parameters available in the PPR cohesive zone
model. It is seen from the DCB test results that curing the adhesive
about the temperature at which the Tg∞ is obtained caused the
adhesive to have more ductility, higher load-carrying capacity and
higher fracture energy than curing it below or above the temperature
at which Tg∞ is attained. Here, the Tg∞ is the Tg of the fully cured network.
Experimental and numerical R curves were obtained to account for
deviations between experiments and simulations. A good agreement
between the numerical and experimental load-displacement curves
was achieved showing the adequacy of the cohesive model used.

1. Introduction
Lap joints bonded with structural adhesives have been used extensively due to their high
strength-to-weight ratio in comparison with the traditional methods such as fastening and
riveting. Cohesive zone modeling (CZM) and associated traction separation laws have been
used for the strength and damage predictions of adhesive joints. CZM has been applied
to model fracture using various traction-separation laws, such as triangular, exponential,
trapezoidal ones [1–4].
If one takes a general look at CZM models related to the adhesive Araldite 2015, now
being used in this study, only a limited number of papers have been found in the open

CONTACT  Halil Özer  hozer@yildiz.edu.tr


© 2017 Informa UK Limited, trading as Taylor & Francis Group
1288   H. ÖZER AND E. ERBAYRAK

literature and most of them are related to the strength predictions. Campilho and others
[5] presented a trapezoidal mixed-mode cohesive damage model appropriate for ductile
adhesives in adhesively bonded joints. Double Cantilever Beam (DCB) and End Notched
Flexure (ENF) tests were performed to obtain the cohesive laws of the adhesive Araldite
2015. An inverse method was used to define the cohesive parameters of the trapezoidal
laws. They concluded that the proposed CBBM (Compliance-based beam theory) provides
accurate results on the critical fracture energies. Campilho et al. [6] studied the influence
of the CZM shape (triangular, exponential or trapezoidal) used to model a thin adhesive
layer in single-lap adhesive joints, for an estimation of its influence on the strength predic-
tion under different material conditions. In their study, two epoxy adhesives were consid-
ered, Araldite AV 138M which is brittle and Araldite 2015 which allows large plastic flow
prior to failure. Their results showed that joints bonded with ductile adhesives are highly
influenced by the CZM shape, and that the trapezoidal shape fits best the experimental
data. They reported that the smaller is the overlap length, the greater is the influence of
the CZM shape. They also observed that the influence of the CZM shape can be neglected
when using brittle adhesives, without compromising too much the accuracy of the strength
predictions. However, it should be noted here that, in the study, CZM-shape investigations
are only related to SLJs, but not DCB or ENF. Moura and others [7] studied the adequacy
of cohesive damage models for the strength prediction of bonded joints. A ductile epoxy
adhesive of Araldite 2015 was used in their study. A mixed-mode cohesive damage model
was presented and used to simulate the behavior of laminated composite single lap-shear
adhesive joints. They stated that the trapezoidal shape in the constitutive law is more valid
when ductile adhesive is used. Moura et al. [8] presented a new data reduction scheme
based on the crack equivalent concept and applied to the DCB test to obtain the respective
critical fracture energy. They compared the model with classical reduction schemes. Some
of the cohesive properties of the trapezoidal law were determined using an inverse method,
allowing the complete fracture characterization of the adhesive under mode I loading. As
in their previous study [7], Araldite 2015 adhesive was used in this study too. Campilho
et al. [9] performed the experimental and numerical study of the buckling behaviour under
pure compression of carbon–epoxy adhesively-bonded scarf repairs. The adhesive layer
including Araldite 2015 was simulated using cohesive elements with trapezoidal traction–
separation laws. They proposed different traction–separation laws using an inverse method,
which consisted in obtaining the fracture toughness with DCB and ENF tests. Tsouvalis
and Anyfantis [10] developed of a robust flexible cohesive law that can be used to simulate
the constitutive behavior of an adhesive layer under mixed mode loading. Araldite 2015, a
relatively stiff epoxy adhesive, was used in their study. Curing procedure consisted of heating
the specimens in 60 °C for 4 h. They used exponential cohesive law with linear softening for
SLJ and DSJ. They indicated the validity of the exponential type cohesive law for modelling
the elastoplastic behaviour of the Araldite 2015 adhesive.
A new potential-based cohesive model has recently come to the fore among the important
remedies to overcome some deficiencies arising in CZM models. The model is named the
PPR model, after the first initials of the authors’ last names, i.e. Park–Paulino–Roesler. It
is known that, in most CZMs, it is difficult to get softening behavior properly. In the PPR
model, softening can be easily controlled by the shape parameters. The group members, Park
and collaborators [11], discussed this constitutive model for cohesive fracture to characterize
different fracture energies and cohesive strengths. They showed that the potential-based
JOURNAL OF ADHESION SCIENCE AND TECHNOLOGY   1289

model is applicable to various material softening responses, i.e. plateau-type (e.g. ductile),
brittle and quasi-brittle, due to controllable softening given by the shape parameters. In
addition, Park and Paulino [12] implemented the PPR model in commercial software,
e.g. ABAQUS, as a user-defined element subroutine. Nguyen and Waas [13] presented an
overview of various mixed-mode cohesive laws including the PPR model. Tsouvalis and
Anyfantis [14], in another study, described the traction increasing part of the law by an
exponential function, whereas the softening part was described by a linear decrease for
the adhesive Araldite 2015. Comparisons were provided from finite element analyses with
the already known trapezoidal law and with the PPR model. Alfanoa et al. [15] developed
a local approach for the identification of the PPR model, considering only mode I fracture
condition. They also proposed a new inverse procedure based on the use of a genetic algo-
rithm. Encouraging preliminary results were obtained on a model DCB specimen.
It is known that curing temperature, curing process and post-curing affect the mechanical
and fracture properties of adhesives. Curing treatment is therefore of particular importance
in adhesive applications. In Gillham’s founding work [16], the formation and properties of
high Tg polymeric glasses were studied, in which Tg is the glass transition temperature. It
is known that the area is of particular importance in the making of composites, coatings,
and adhesives by thermosetting ‘cure’ reactions. In his study, an attempt was made to pro-
duce a generalized model for the formation and properties of thermosetting of systems.
Carbas et al. [17] studied the influence of the curing temperature on the physical and
mechanical properties of three structural adhesives, i.e. Araldite 2011, Araldite AV 138M
and Sikadur-30 LP. They investigated the relationship of the effect of curing temperature in
the glass transition temperature, Tg, and stiffness of epoxy adhesives. They concluded that,
when cured below the cure temperature, Tcure, at which the Tg of the fully cured network,
Tg∞, was achieved, the strength and stiffness of the adhesive increase as the cure temperature
increases and the Tg was higher than the cure temperature. They observed also that, when
cured above the Tcure at which the Tg∞ is achieved, the strength and stiffness decrease as
the cure temperature increases and the Tg is higher than the cure temperature. Carbas et al.
[18] studied the effects of post-curing and cure temperature on the Tg and the mechanical
properties of two epoxy adhesives, i.e. Araldite 2011 and Loctite Hysol 3422. They showed
that, when post-cured at a temperature above Tg∞, the mechanical and physical properties
tend to have a constant value for any cure temperature. Zhang et al. [19] studied the changes
of the adhesive Tg with different cure cycles and thermal histories. They concluded that
raising the temperature after a low temperature cure (say 20 °C) would be an ideal option
to increase the Tg of ‘cold curing’ adhesives. Pinto et al. [20] studied experimentally the
influence of temperature on the tensile strength of single and double-strap repairs. The
Araldite 2015 adhesive, characterized by a large ductility in tension and shear, was selected
for the study. Accounting for the temperature effects, tests were carried out at 23, 50 and
80 °C. Their results showed that high temperatures gradually decreased the repairs stiffness
and strength due to the degradation of the adhesive, and this degradation is slightly higher
at room temperature, reducing at higher temperatures.
The effect of curing temperature on fracture properties of the Araldite 2015 adhesive has
been studied in a limited number of papers in the literature [10,20]. To our knowledge, for
the Araldite 2015 adhesive, the effects of curing temperature on the fracture energy and
cohesive parameters are not available in the open literature. It is the main motivation to study
these effects for the adhesive. Tensile tests were performed on the dogbone-shaped bulk
1290   H. ÖZER AND E. ERBAYRAK

specimens for evaluating the mechanical properties of the adhesives cured with ­different
temperatures. DCB tests were performed to obtain the cohesive laws of the adhesive Araldite
2015. Relationship between curing temperature and the glass transition temperature is also
taken into account. In numerical modeling, the exponential and PPR cohesive zone mod-
els are considered, in which the PPR model applied for the adhesive firstly in this study.
DCB test results are used to obtain the cohesive laws of the adhesive in mode I. Inverse
analyses were also performed, if the experimental softening curves are incompatible with
the numerical ones.

2.  Experimental studies: methods and results


In this study, subjects studied and their results are given and then discussed in their related
sections and subsections in order to make it easier to follow the paper, due to the paper
including many substeps. Therefore, results and discussion sections will not appear as sep-
arate sections in the paper.

2.1.  Glass transition temperature measurements


The adhesive Araldite 2015 (Huntsman, Belgium) is a DGEBA based epoxy with an amine
hardener of the reaction between bisphenol A and epichlorohydrin with a basic catalyst
[21,22]. Mechanical properties of the adhesive Araldite 2015 were tabulated in Table 1.
Curing temperatures (Tcure) and times were specified as RT (Room Temperature)/6 h,
40 °C/2 h, 60 °C/35 min, 100 °C/7 min. A week was left after curing before experimen-
tal tests were undertaken. Waiting a week after the curing process can be regarded as a
post-curing treatment. Glass transition temperatures (Tgs) of the adhesive Araldite 2015
were determined using the 10 mg little adhesive samples cured at RT, 40, 60 and 100 °C,
respectively. In literature, many different methods have been used to determine the glass
transition temperature such as the Differential Scanning Calorimetry (DSC), Dynamic
Mechanical Analysis and Thermo Mechanical Analysis. In this study, the DSC method was
used to determine the glass transition temperature of the adhesive.
In the DSC application, cured little specimens were placed into a small pot on the DSC
measurement device (Perkin Elmer Pyris). DSC tests were performed in two cycles. In both
cycles, the adhesive samples were heated up to 150 °C and then cooled to RT (see Figure 1).
The temperature was increased every two minutes by 10 °C (10 °C/2 min) until 150 °C
and then decreased every two minutes by 10 °C until RT. As seen in Figure 1, the glass tran-
sition temperature was determined by using a line existing over the curve obtained from
the second cycle. By using this line, the device identified the glass transition temperature
of the adhesive specimen. Detailed steps defining the Tg of the specimen cured at 60 °C
were shown in Figure 1. (The heat flow curves related to other curing temperatures were

Table 1. Properties of the adhesive Araldite 2015 [23–26].


Elasticity modulus MPa 1850
Poisson’s ratio 0.33
Tensile strength MPa 21.6
Tensile yield strength MPa 12.6
Tensile failure strain % 4.77
JOURNAL OF ADHESION SCIENCE AND TECHNOLOGY   1291

Figure 1. Heat flow curves of the Araldite 2015 specimen cured at 60 °C.

Table 2. Variation of glass transition temperature as a function of curing temperature.


Curing temp (°C) Curing time Post-curing Tg (°C) ΔT (Tg – Tcure)
Tcure1 RT 6 h One week 66 43
Tcure2 40 2 h One week 68 28
Tcure3 50 35 min One week 67 17
Tcure4 60 35 min One week 74 14
Tcure5 70 35 min One week 68 −2
Tcure6 100 7 min One week 71 −29
SD = 2.83

assessed, but not reported here for saving space.) It was seen that the highest Tg value of
74 °C was attained at the curing temperature of 60 °C. The highest Tg value can be assumed
to be almost the glass transition temperature of the fully cured system (Tg∞). However, two
new curing temperatures and times of 50 °C/35 min and 70 °C/35 min were also taken into
account to more clarify the Tg∞ value. It is to be noted here that the first one is 10 °C below
and the second one 10 °C above the curing temperature of 60 °C at which the Tg∞ is approx-
imately achieved. Curing time of 35 min were held constant for the curing temperatures of
50 °C and 70 °C. The above procedures for defining the Tgs were also repeated for the last
two temperatures. The Tg values obtained from the DSC tests and temperature differences
(ΔT) were then presented in Table 2. (In addition, the sample standard deviation (SD) was
calculated and given in Table 2 [27].)
As seen in Table 2, curing temperatures affect the Tgs of the Araldite 2015 adhesive. The Tg
values are increasing with increasing the curing temperatures until it becomes peak at 60 °C.
Therefore, the maximum glass transition temperature of 74 °C, the highest Tg value, can be
assessed as the glass transition temperature of the fully cured system (Tg∞). It is seen from
Table 2 that all the curing temperatures are below their own glass transition temperatures,
except the curing temperatures of 70 and 100 °C. Meanwhile, the Tg value of the adhesive
cured at RT was reported as 67 °C by DSC method in manufacturer’ data sheet [28]. It is
1292   H. ÖZER AND E. ERBAYRAK

also interesting that Tgs for the specimens cured at 40 and 70 °C have equal values (see Table
2). In addition, temperature difference is the lowest one at the curing temperature of 70 °C.
However, to further clarify the Tg∞ value and elaborate the effect of curing temperature
on the mechanical properties of the adhesives cured with different temperatures, it was
decided to carry out tensile tests on the dogbone-shaped bulk specimens. Steps in prepar-
ing the bulk specimens were briefly discussed below. Three specimens were prepared for
each curing temperature. Therefore, a total of 18 dogbone samples were produced. Using a
vacuum device designed by authors, the adhesive were prepared in a vacuum medium to
prevent air bubbles and micro voids. Adhesives were manufactured in a special mold as a
thin plate. PID-controlled oven was used for curing operations. The thin plates were then cut
into the required dimensions by using KMT waterjet machine. During the cutting process,
pressure water (3800 bar) enriched with the abrasive-waterjet-additive sand (80 μ) was used.
Bulk specimens were prepared, in accordance with BS 2782 standard. Tensile tests were
performed on the test specimens with Instron 5982 at 1 mm/ min tensile rate speed. The
longitudinal strains were recorded during the tests using the video extensometer.
Stress–strain curves were obtained for the dogbone specimens cured with the six differ-
ent temperatures (see Figure 2). As seen in Figure 2, the tensile strength is increasing with
increasing the curing temperatures until it becomes peak at 60 °C, and above the 60 °C the
strengths tend to decrease. (It must be stated that strength values for the specimens cured
at the 40 °C and 50 °C are approximately equal ones, probably due to the distinct differ-
ence in curing times. Similar behaviors were also reported in the Ref. [17] for the adhesive
Araldite 2011.) As stated above, tensile strengths tend to decrease above Tg∞. Carbas et al.
[17] stated that, above Tg∞, as the temperature increases the adhesive performance decreases.
This decrease of the adhesive performance is probably due to the thermal degradation or
oxidative crosslinking. Above Tg∞, the network degradation can occur and with it comes
a change of the properties.

Figure 2. Stress-strain curves of the dogbone specimens.


JOURNAL OF ADHESION SCIENCE AND TECHNOLOGY   1293

Finally, it is assumed that the highest Tg value of 74 °C, attained at the curing temperature
of 60 °C, can be assessed as the glass transition temperature of the fully cured system (Tg∞).
A time–temperature–transition (TTT) isothermal cure diagram can help us in better
understanding the relationship between curing and glass transition temperatures. S-shaped
vitrification curve and the gelation curve divide the TTT plot into four distinct states: liquid,
gelled rubber, ungelled glass and gelled glass. Three important temperatures are marked
on the temperature axis of the TTT cure diagram: Tg0 is the glass transition temperature of
the unreacted resin mixture, Tg∞ is the glass transition temperature of the fully cured resin,
and Tgel is the temperature where gelation and vitrification occur simultaneously as well
as the point where the vitrification and gelation curves intersect. At temperatures below
Tg0 reaction takes place in the glassy state and is therefore slow to occur [29,p.441–442].
According to the TTT cure diagram, it is clear that, depending on the cure temperature
Tcure, there are three types of cure behavior. If Tg0 < Tcure < Tgel, gelation cannot occur, only
vitrification can be observed, the system would convert directly from liquid into glass. This
is the first type of cure behavior. If Tgel < Tcure < Tg∞, both gelation and vitrification can be
observed, and the system would convert from liquid into gelled rubber and then into gelled
glass. This is the second type of cure behavior. If Tg∞ < Tcure < Tdecomposition, only gelation can
be observed, and the system would change into gelled rubbery state. This is the third type
of cure behavior [30]. In the light the considerations above, when it comes to Table 2, it is
seen that Tcure1, Tcure2, Tcure3 and Tcure4 are below their own glass transition temperatures
and also the glass transition temperature of the fully cured network (Tcure < Tg, Tcure < Tg∞).
As stated the above, two curing temperatures of 50 and 70 °C were chosen only to more
clarify the Tg∞ value. Curing temperatures of RT, 40, 60, 100 °C are therefore taken into
account in the rest of the paper.

2.2.  Fracture energy estimation of DCB specimens


Our aim to perform DCB tests are to obtain load–displacement (P–Δ) curves and then
evaluate the critical energy release rates (GIC) of the adhesives cured with four different
temperatures. DCB tests were carried out in accordance with ASTM 3433-99.
The geometric properties of the DCB specimen are as follows: the adherend length
L = 230 mm, adherend thickness h = 12 mm, adherend width b = 25 mm and adhesive
thickness t = 0.2 mm. An initial pre-crack of 40 mm was introduced to the DCB speci-
mens (Appendix 1). DCB adherends were fabricated using mild steel (St37) and the epoxy
adhesive of Araldite 2015 was introduced to adherend surfaces. Mechanical properties of
the steel adherend were determined in the scope of this study and tabulated in Table 3.
Adherends were cut into the required dimensions and then adherend surfaces were
polished by using acetone. Sand blasting was implemented on the adherend surfaces.
Aliminium oxide sand was used during the sand-blasting treatment. After the sand-blast-
ing treatment, adherends were washed into the ultrasonic cleaner machine.

Table 3. Mechanical properties of St37 adherend.


Elasticity modulus (GPa) Poisson’s ratio Tensile yield strength (MPa) Tensile strength (MPa)
200 0.3 483 546
1294   H. ÖZER AND E. ERBAYRAK

The Araldite 2015 was applied as a layer of 0.2 mm between two adherends. A wire of
0.2 mm was used to adjust the adhesive thickness. Four DCB specimens were prepared for
each curing temperature. Therefore, a total of 16 DCB samples were produced. Then, all
four groups of the DCB specimens were cured at four different temperatures. (As stated
above, curing temperature and times were specified as RT/6 h, 40 °C/2 h, 60 °C/35 min,
100 °C/7 min, according to the adhesive data sheet [28]). PID-controlled oven was used for
curing operations and digital thermocouple [VERTH; ± (0.2% reading + 1 °C) accuracy;
0.1 °C resolution] for recording the internal temperature of the oven. As stated above, a week
was left after curing before experimental tests were undertaken, and waiting a week after
curing was regarded as the post-curing treatment. According to [18], post-cure conditions
affect the mechanical and physical properties of adhesives. However, all the specimens
used in this study were under the same post-cure condition. After completing the curing
operations, DCB tests were repeated for each curing temperature.
DCB tests were done by using Mecmesin universal tensile machine at room temperature
[31]. Loading rate was chosen as 1 mm/min. In the experimental studies, it was seen that
cohesive failure occurred inside the adhesive bondlines. Load–displacement (P–Δ) curves
were then obtained and presented in Figure 3.
As seen in Figure 3, curing temperature affects on the peak load and the deformation
corresponding to the peak load and the final deformation of the adhesive. The peak load is
increasing with increasing the curing temperature until it becomes peak at 60 °C, at which
it was assumed that the Tg∞ was achieved (see also Table 2). Therefore, it can be seen from
Figure 3 that curing the adhesive about the temperature at which the Tg∞ is achieved caused
the adhesive to be more ductile than curing it below or above the temperature at which
the Tg∞ is achieved. However, as expected, at the curing temperature of 100 °C, i.e. above
the temperature at which the Tg∞ was attained, the peak load decreases and has the lowest
value. As seen from Figure 3, sample cured at 60 °C has the most deformation capacity

Figure 3. Experimental P vs. Δ curves of the DCB specimens as a function of cure temperature.
JOURNAL OF ADHESION SCIENCE AND TECHNOLOGY   1295

among all the other samples. (It has also the highest fracture energy, when considering the
areas under the curves.)
After performing the above steps, using the resulting P–Δ curves, the critical energy
release rates of the adhesives were evaluated with three different methods. The methods
are the modified beam theory (MBT), the compliance calibration method (CCM) and the
area method (AM). As a first estimate, by using the MBT, the critical energy release rates
were evaluated through the relation as follow [32]:
3PΔ
GIC = (1)
2ba
where GIC is the Mode I strain energy release rate (Mode I fracture energy), P the applied
load, Δ the displacement at DCB ends. a is the equivalent crack length obtained from sum-
ming the current crack length and initial crack length a0.
(Crack length was not measured during the tests and it is calculated from the relation
given in Appendix 2.) Secondly, the CCM was used in computing the fracture energy
through the relation below [32, Appendix 3]:
nPΔ
GIC = (2)
2ba
where n is the slop of the log C vs. log a curve, here C the compliance of the adhesive. Last
method for evaluating the energy is the AM, and the following relationship was used to
calculate the fracture energy [33, Appendix 4]:
1
GIC = (P Δ −P Δ )
2b(a2 −a1 ) 1 2 2 1 (3)

The areas under the curves were calculated using the data obtained from P–Δ curves. In
addition, software built in Mecmesin universal tensile machine can automatically calculate
the areas under the curves. This gave also the opportunity to check our results. As a result,
the fracture energies were evaluated using the three methods and tabulated in Table 4.
As seen from Table 4, the fracture energies predicted by MBT and CCM methods are
very close to each other. The fracture energy is increasing with increasing the curing tem-
perature, and the GIC has the highest value at 60 °C This can be attributed to the fact that,
at the curing temperature of 60 °C, the specimen was cured at the temperature at which
the Tg∞ was achieved (see Table 2). The energy was then decreasing at the highest curing
temperature of 100 °C.

Table 4. Estimations of the strain energy release rates with respect to curing temperatures.
Experimental GIC (N/mm)
Curing temp (°C) MBT CCM AM SD
RT 0.500 0.510 0.543 0.022
40 °C 0.705 0.710 0.767 0.034
60 °C 0.789 0.795 0.863 0.041
100 °C 0.544 0.554 0.635 0.049
SD 0.135 0.133 0.141
1296   H. ÖZER AND E. ERBAYRAK

3.  Numerical studies: CZM modelings and results


3.1.  Exponential cohesive zone model
Cohesive zone model can describe debonding process under the effect of the cohesive trac-
tion and material separation. For pure mode I fracture problem, the potential parameters
mainly involve the normal traction (Tn) and the normal separation distance (Δn). Since
there is no shear separation considered (i.e. Δt  =  0), the work of separation (potential)
can be expressed using the exponential Xu and Needleman cohesive zone law through the
relation below [4, Appendix 5]:
[ ( ) ( )]
Δ Δ
𝜙(Δn ) = 𝜙n 1 − 1 + n exp − n (4)
𝛿nc 𝛿nc

where ϕn is the normal work of separation (i.e. the mode I fracture energy), Δn the normal
separation distance and 𝛿nc the critical normal separation distance. Figure 4 schematically
shows an illustrative representation of the exponential CZM shape.
In Figure 4, 𝛿n represents the normal conjugate final crack opening width. Derivative
of the potential with respect to the normal separation distance gives the normal traction.
Differentiating Equation (4) with respect to Δn, we have the normal traction as :
( ) ( )
𝜕𝜙 𝜙 Δn Δn
Tn = = n exp − (5)
𝜕Δn 𝛿nc 𝛿nc 𝛿nc

when Δn = δnc, the normal traction reaches a maximum value and reads:

(6)
[ ]
Tn,max = 𝜙n ∕ 𝛿nc exp(1)

Figure 4. Illustration of the exponential traction-separation law concepts for pure Mode I [34].
JOURNAL OF ADHESION SCIENCE AND TECHNOLOGY   1297

where Tn,max represents the cohesive strength of the interface. Using the exponential cohesive
zone option available in the ANSYS software, finite element analyses (FEA) were performed.
A state of plane strain was considered. Adherends were meshed with four-node structural
solid elements (Plane 182). Interface area (cohesive layer) was composed from cohesive
zone (interface) elements. Adhesive was then meshed with four-node cohesive interface
elements (INTER202) generated using CZMESH command. Different meshing schemes
were investigated to optimize number and size of the elements. The model was discretized
into 4955 elements including 155 cohesive elements. Non-linear analyses were performed
under 159 iterations. Detail of the finite element result was given in Figure 5 (Figure 5 shows
the reproduction results obtained from the original Ansys plot).
Cohesive parameters introduced into numerical modeling are the cohesive strength
(Tn,max) and the final crack opening width (𝛿n). It must be stated that the final deformations
should be taken from the DCB test results, which are in range between 3.5 and 5 mm in this
study (see Figure 3). Campilho and collaborators stated that the adhesive bulk properties are
not adequate to characterize these laws [5]. Therefore, during the curve fitting process, the
cohesive strengths (Tn,max) and final opening widths (δn) were changed to ensure coinciding
the critical point (𝛿nc, Pmax) on the numerical P–Δ curve with those on the experimental one.
The numerical curves were also checked whether they are also providing convergence to
other points on the experimental P–Δ curve. Range of cohesive strengths and final opening
widths were given in Table 5.
It was seen that changing the strength and final opening widths in the range above was
not a significant effect on the P–Δ curves. (In Ref. of [8], the cohesive stresses used in the
trapezoidal cohesive law were reported as 18, 23 and 28 MPa for the same adhesive. It was
also reported the similar results for changing the strength values.) In this study, the average
ones representative for each curing temperature, given in the last column of Table 5, were
used in the numerical modelings [35,36,37]. Fracture energies obtained using the AM
method were taken into account to evaluate the potentials in cohesive modelings (see Table
4). Cohesive parameters used in the exponential CZM model were presented in Table 6.

Figure 5. Deformed shape of the DCB specimen cured at RT.


1298   H. ÖZER AND E. ERBAYRAK

Table 5. Range of cohesive strengths and final opening widths as a function of cure temperature.
Curing temp (°C) Test 1 Test 2 Test 3 Test 4 Test 5 Average SD
RT Tn,max 18 19 20 21 22 20 1.581
δn 0.225 0.215 0.205 0.185 0.175 0.201 0.020
40 °C Tn,max 20 21 22 23 24 22 1.581
δn 0.245 0.230 0.220 0.210 0.205 0.222 0.016
60 °C Tn,max 22 23 24 25 26 24 1.581
δn 0.300 0.295 0.290 0.285 0.280 0.290 0.007
100 °C Tn,max 15 16 17 18 19 17 1.581
δn 0.330 0.325 0.320 0.310 0.300 0.317 0.012

Table 6. Fracture parameters of the adhesive Araldite 2015 for the exponential CZM.
Curing temp (°C) ϕn (N/mm) Tn,max(MPa) δnc (mm) δn(mm)
RT 0.543 20 0.009987 0.201
40 °C 0.767 22 0.012825 0.220
60 °C 0.863 24 0.013228 0.290
100 °C 0.635 17 0.013741 0.317
SD 0.141 2.98 0.00168 0.055

Figure 6. Comparison between the experimental and numerical (Exponential CZM) P vs. Δ curves for
curing temperature of 60 °C.

Using the exponential CZM modeling, load-displacement curves were firstly obtained
for each curing temperatures and then resulting curves were compared with the experi-
mental ones. Load-displacement curve was presented for the curing temperature of 60 °C
in Figure 6.
It is seen from Figure6 that softening behavior of the adhesive aren’t properly modelled
by using the exponential CZM model, even if the two checkpoints (𝛿nc, Pmax) are exactly
JOURNAL OF ADHESION SCIENCE AND TECHNOLOGY   1299

satisfied. The curves related to other curing temperatures, not reported herein for the sake
of brevity, showed the similar softening incompatibility.

3.2.  PPR potential-based cohesive zone model


As presented above, the exponential cohesive law was firstly implemented to account for
the load-displacement curves. However, from the numerical simulations, it was seen that
the softening behaviors of the adhesive were not properly correlated with the experimental
softening ones. To overcome this difficulty, a potential-based cohesive zone model, the
Park–Paulino–Roesler cohesive model (PPR) was secondly used in modeling the soften-
ing behaviors [11,38]. However, the PPR model isn’t available in the ANSYS library and it
is applied using the MATLAB software. Before applying the PPR model, the method was
briefly discussed below. In PPR model, the PPR potential is given as [11]

Δ 𝛼 m Δn m
( ) ( )
𝜙 Δn = 𝜙n + Γn 1− n (7)
( )
+
𝛿n 𝛼 𝛿n

where α is the shape parameter and m is the non-dimensional exponent. Γn is the energy
constant and reads.
( )m
𝛼
Γn = −𝜙n (8)
m
Derivative of the PPR potential with respect to the normal separation distance Δn gives
the normal traction. Differentiating Equation (7) with respect to Δn, one can obtain the
normal traction as:

Δn 𝛼−1 m Δn m−1
( ) ( )
𝜕𝜙 𝜙n ( 𝛼 )m Δ
Tn = = 1− + (𝛼 + m) n (9)
𝜕Δn 𝛿n m 𝛿n 𝛼 𝛿n 𝛿n

The non-dimensional / exponent m is evaluated from the boundary conditions of the critical
separation, i.e. 𝜕Tn 𝜕Δn = 0 at Δn = δnc, as below:

𝛼(𝛼 − 1)𝜆2n
m= ( (10)
1 − 𝛼𝜆2n
)

where the initial slope indicator 𝜆n is defined as the ratio of the critical crack opening width
to the final crack opening width as follows:

𝛿nc
𝜆n = (11)
𝛿n

where 𝛿nis, as stated above, the final separation distance and reads [11]

𝜙n ( )𝛼−1 ( 𝛼 )(
𝛼
)m−1
𝛿n = 𝛼𝜆n 1 − 𝜆n +1 𝜆n + 1 (12)
Tn,max m m
1300   H. ÖZER AND E. ERBAYRAK

Figure 7 shows the effect of the shape parameter α on the softening behavior.
As seen in Figure 7, the shape parameter α defines material softening responses, e.g. brittle
(α ≈ 2), plateau (1 < α ≪ 2) and quasi-brittle (α ≫ 2). For α = 0, a triangular CZM law is
attained. In the study, the parameters of the cohesive zone model were tuned with comparing
the plots of the numerical load-displacement curves with those of the experimental ones.
The cohesive parameters, giving minimum deviations between numerical and experimental
P–Δ curves, were obtained by setting α = 12 and m = 1.5 for all the specimens cured with
different temperatures. Moreover, the initial slope indicator 𝜆n has also a constant value of
0.1, due to the fact that two parameters (α, m) were held constant above (refer to Equation
10). Cohesive parameters used in the PPR models were tabulated in Table 7.
It is seen from Table 7 that cohesive parameters {ϕn, δnc, δn} are minimum at curing
temperature of RT. As stated above, α = 12 was found to be the optimum value for the
numerical predictions that is also in accordance with quasi-brittle behavior (α ≫ 2) of the
adhesive used [11,12]. Resulting load-displacement curve obtained from the PPR modeling
was presented for the curing temperature of 60 °C in Figure 8.
It is seen in Figure 8 that the numerical P–Δ curve is still not matching the softening
parts of the experimental curve, although the PPR approach gives more appropriate form
of the softening part when compared to the softening part obtained from the exponential
CZM modeling (see Figure 6). The curves related to other curing temperatures, not reported
herein, showed the similar softening incompatibility. Therefore, it was decided to apply the
inverse analysis.

3.3.  Inverse analysis


Steps of the inverse analysis can be briefly summarized below and procedures of the inverse
analysis are given in Figure 9. Firstly, for each curing temperature, PPR softening curve is
plotted by using Equation 9. In plotting the softening curve, it is important to get maximum
stress corresponding to the requirement of Δn 𝛿n = 0.1. Secondly, three points on the PPR
/

softening curve were selected (Figure 9). Then, the values of the cohesive stresses and sep-
aration distances corresponding to these 3 points were recorded. Using these stresses and
corresponding separation distances, finite element analyses were performed on the updated

Figure 7. The effect of the shape parameter α on the softening behavior.


JOURNAL OF ADHESION SCIENCE AND TECHNOLOGY   1301

Table 7. Fracture parameters of the adhesive Araldite 2015 for the potential based PPR modelings.
Curing temp (°C) ϕn (N/mm) Tn,max(MPa) δnc (mm) δn (mm) α Γn
RT 0.543 20 0.01325 0.1325 12 −12.988
40 °C 0.767 22 0.01590 0.1590 12 −17.174
60 °C 0.863 24 0.01680 0.1680 12 −20.093
100 °C 0.635 17 0.01770 0.1770 12 −14.956
SD 0.141 2.986 0.00192 0.0192 – 3.051

Figure 8. Comparison between the experimental and numerical (PPR CZM) P vs. Δ curves for the curing
temperature of 60 °C.

models with ANSYS and load-displacements graphs were replotted with the new data.
Thirdly, transferring the current data from the numerical and experimental load-displace-
ment curves to the polynomial-basic-fitting toolbox built in MATLAB software, polynomial
regression analyses were performed.
Lastly, as a result of the regression analyses, if the resulting P–Δ curves are incompatible
with experimental ones, the above procedure should be repeated by selecting more new
points different from the previously selected three points. (It is to be noted here that the
softening part of the curve can also be obtained without applying the inverse analysis, but
it is very difficult to have the best fit between the numerical critical-point (𝛿nc, Pmax) and the
experimental critical-one.) Finally, load-displacement curves obtained after applying the
inverse analyses were presented in Figure 10 for the specimens cured at the temperatures
of RT, 40, 60 and 100 °C.
It is seen from Figure 10 that the inverse analyses performed over the PPR models facil-
itated modeling the softening behaviors of the adhesive, without missing the critical point
(𝛿nc, Pmax). A group of eighth-order equations defining the softening parts shown in Figure
10 was presented in Table 8.
1302   H. ÖZER AND E. ERBAYRAK

Figure 9. Steps of the inverse analysis.

3.4. R-curves
In addition, experimental and numerical R curves were also obtained to account for devi-
ation between experiments and simulations. It is seen from Table 4 that there were no
significant differences among the fracture energies predicted by the methods. The experi-
mental load-displacement curves were used to obtain the mode I R-curves using the CCM
method. Figures 11 and 12 present the experimental and numerical R-curves as a function of
crack length, respectively. (The method followed for calculating the crack length in plotting
R-curves was presented in Appendix 2.)
It can be easily seen that the R curves are approximately consistent curves. Ideally, the
R-curves are horizontal lines, although experimentally fluctuations may occur due to issues
such as poor adhesive mixture, adhesion problems, defects and unstable crack growth [39].
The maximum error between the experimental and numerical fracture energies is 2.74 %
for RT, 1.97 % for 40 °C, 1.88 % for 60 °C and 3.42 % for 100 °C. It is therefore verified that
the numerical R-curves approximately agreed with the experimental ones and it can be said
that the cohesive model used accurately reproduced R curves.
The statistical method proposed by Budzik et al. [40] was also applied to our problem.
Steps followed and results obtained are summarized below. As seen from Figure 3, there is
a sudden drop after reaching the peak load. Moreover, after that point, crack propagating
takes place at random intervals including both continuous (stable) and stick-slip crack
growths, which is called as a hybrid crack growth. This can be also seen from the stepped
JOURNAL OF ADHESION SCIENCE AND TECHNOLOGY   1303

Figure 10. Comparison between the experimental and numerical P vs. Δ curves: (a) RT, (b) 40 °C, (c) 60 °C
and (d) 100 °C.

structures of the force vs. displacement plot (see Figure 3). In our study, crack growth is
therefore an unstable one. Firstly, Equation (1) in Budzik’s paper was rearranged for the
DCB problem, because the Equation (1) was related to SCB problem. The fracture energy
expression was then rewritten as follows:
3 ( 4 )−1∕3 4∕3 2∕3
GIC = bE P Δ (13)
h
If GIC is considered constant, Equation (13) can be arranged as below:

P = A� Δ−1∕2 (14)
where A′ is constant. It can be seen from Equation (14) that the fracture energy can be
described using the coordinates of P vs. Δ–1/2. P vs. Δ–1/2plots for all the curing temperatures
were obtained and presented in Figure 13.
Using the standard normal histogram, the statistical distribution of fracture energy var-
iation during propagation was evaluated via the Z-score function. The Z-score function
can be written as follows [27]:
1304 
 H. ÖZER AND E. ERBAYRAK

Table 8. Eighth-order equations defining the numerical softening parts shown in Figure 14.
P = AΔ8 + BΔ7 + CΔ6 + DΔ5 + EΔ4 + FΔ3 + GΔ2 + HΔ+ I
Curing temp (°C) A B C D E F G H I
RT −35.55 627.66 −4563.75 17,386.92 −36751.32 41,909.40 −23444.64 6128.46 −232.50
40 °C −43.77 811.77 −6218.19 25,306.57 −58530.07 76,359.48 −53093.40 18,473.55 −2203.77
60 °C −22.43 482.58 −4310.29 26,610.81 −56711.61 89,936.99 −78790.65 35,664.80 −6112.82
100 °C −1.70 73.94 −898.19 4962.77 −14077.35 20,408.07 −13820.11 4218.84 −152.38
JOURNAL OF ADHESION SCIENCE AND TECHNOLOGY   1305

Figure 11. Experimental R curves on the DCB specimens cured at different temperatures.

Figure 12. Numerical R curves on the DCB specimens cured at different temperatures.

GIC −𝜇
Z= (15)
SD
where μ is the population mean and SD is the standard deviation. These are defined as
below [27]:

GIC
𝜇= (16)
n
1306   H. ÖZER AND E. ERBAYRAK

Figure 13. P vs. Δ–1/2plots together with regression lines.

Figure 14. Statistical distribution of fracture energy variation for 60 °C.


∑� �2
GIC − 𝜇 (17)
SD =
n

in which n is the population number. Relative frequency vs. Z plots (for plotting steps,
see [27]) are presented for the curing temperatures of 60 and 100 °C in Figures 14 and 15,
respectively.
It was seen from the calculations that the standard deviations are relatively high for our
problem. This can be also seen from the Z values of the two plots above. It was concluded
that the statistical method is not valid for our problem, due to the unstable crack propaga-
tion, high standard deviation and viscoelastic properties of the adhesive used.
JOURNAL OF ADHESION SCIENCE AND TECHNOLOGY   1307

Figure 15. Statistical distribution of fracture energy variation for 100 °C.

4.  Summary and concluding remarks


In this study, the effects of curing temperature on the mode I fracture energy and cohesive
parameters were investigated experimentally and numerically. Relationship between curing
temperature and the glass transition temperature was also taken into account. In numerical
modeling, the Exponential and PPR cohesive zone models were considered. DCB tests were
performed to obtain the cohesive laws of the adhesive. Some of the fracture properties were
determined using the inverse method. The main conclusions are as follows:
(1) It was observed that the curing temperature should be examined together with
the glass transition temperature. Interpretations made only considering the curing
temperature are misleading and therefore not adequate.
(2) It was seen that the Tg values are increasing with increasing the curing tempera-
tures until the glass transition temperature of the fully cured system (Tg∞). It was
observed from the DCB test results that curing the adhesive about the temperature,
at which the Tg∞ is obtained, caused the adhesive to have more ductility, higher
load-carrying capacity and higher fracture energy than curing it below or above
the temperature at which Tg∞ is attained.
(3) Tensile tests were performed on the dogbone-shaped bulk specimens for evaluating
the mechanical properties of the adhesives cured with different temperatures. It was
seen that the tensile strength is increasing with increasing the curing temperatures
until it becomes peak at 60 °C, and above the 60 °C the strength tend to decrease.
(4) The influence of curing temperatures on the cohesive properties (i.e. cohesive
strength, fracture energy, shape parameter, separation distances, etc.) was stud-
ied. It was seen that cohesive stresses and separation distances in the tested range
do not a significant effect on the P–Δ curves (see Table 5). It was decided that the
most appropriate way is to select an appropriate group of cohesive strength and
final opening width giving the best fit between the numerical and experimental
simulations, without considering the adhesive bulk properties.
1308   H. ÖZER AND E. ERBAYRAK

(5) It is known that post-curing affects the glass transition temperature, mechanical
and fracture properties of adhesives. In this study, a week was left after curing
before experimental tests were under taken. Waiting a week after curing process
can be regarded as a post-cure treatment. It is seen that all the specimens used in
this study were under the same post-cure condition. Post-curing effect was not
measured in this study, but the authors are aware that the results given in the paper
were affected by the post-curing effect.
(6) It is known that it is difficult to model softening behavior properly in most CZMs.
It was seen that softening behaviors of the adhesive were not properly correlated
with the experimental softening ones by using the exponential model. However, the
PPR approach gives more appropriate form of the softening part when compared
to the softening part obtained from the exponential CZM modeling. Additionally,
softening was more easily controlled by performing the inverse analyses over the
PPR models.
(7) Experimental and numerical R curves were also obtained to account for deviation
between experiments and simulations. It was verified that the numerical R-curves
approximately agreed with the experimental ones and the cohesive model used
accurately reproduced R curves.
(8) To our knowledge, the effects of curing temperature on cohesive parameters and the
fracture energy have not been studied for the adhesive Araldite 2015 in the open
literature. It is expected that this study can make a contribution to the literature
on the adhesive.

Acknowledgements
This work was funded by Yıldız Technical University under the research project no of 2015-06-
01-DOP01. The authors gratefully acknowledge Dr. Özkan Öz and Dr. Fatih Çakar for giving the
opportunity to use some laboratory facilities.

Disclosure statement
No potential conflict of interest was reported by the authors.

References
  [1] Song SH, Paulino GH, Buttlar WG. A bilinear cohesive zone model tailored for fracture of
asphalt concrete considering viscoelastic bulk material. Eng Fract Mech. 2006;73:2829–2848.
  [2] Campilho RDSG, de Moura MFSF, Ramantani DA, et al. Tensile behaviour of three-dimensional
carbon-epoxy adhesively bonded single- and double-strap repairs. Int J Adhes Adhes.
2009;29:678–686.
  [3] Gilormini P, Diani J. Some features of the PPR cohesive-zone model combined with a linear
unloading/reloading relationship. Eng Fract Mech. 2017;173:32–40.
  [4] van den Bosch MJ, Schreurs PJG, Geers MGD. An improved description of the exponential Xu
and Needleman cohesive zone law for mixed-mode decohesion. Eng Fract Mech. 2006;73:1220–
1234.
  [5] Campilho RDSG, de Moura MFSF, Ramantani DA, et al. Obtaining the cohesive laws of a
trapezoidal mixed-mode damage model using an inverse method. Cienc Tecnol Mater.
2008;20:81–86.
JOURNAL OF ADHESION SCIENCE AND TECHNOLOGY   1309

  [6] Campilho RDSG, Banea MD, Neto JABP, et al. Modelling adhesive joints with cohesive zone
models: effect of the cohesive law shape of the adhesive layer. Int J Adhes Adhes. 2013;44:48–56.
  [7] de Moura MFSF, Campilho RDSG, Gonçalves JPM. Mixed-mode cohesive damage model
applied to the simulation of the mechanical behaviour of laminated composite adhesive joints.
J Adhes Sci Technol. 2009;23:1477–1491.
  [8] de Moura MFSF, Campilho RDSG, Goncalves JPM. Crack equivalent concept applied to the
fracture characterization of bonded joints under pure mode I loading. Compos Sci Technol.
2008;68:2224–2230.
  [9] Campilho RDSG, de Moura MFSF, Ramantani DA, et al. Buckling behaviour of carbon–epoxy
adhesively-bonded scarf repairs. J Adhes Sci Technol. 2009;23:1493–1513.
[10] Tsouvalis NG, Anyfantis KN. Numerical prediction of the response of metal-to-metal adhesive
joints with ductile adhesives. Appl Mech Mater. 2010;24–25:189–194.
[11] Park K, Paulino GH, Roesler JR. A unified potential-based cohesive model of mixed-mode
fracture. J Mech Phys Solids. 2009;57:891–908.
[12] Park K, Paulino GH. Computational implementation of the PPR potential-based cohesive model
in ABAQUS: educational perspective. Eng Fract Mech. 2012;93:239–262.
[13] Nguyen N, Waas AM. A novel mixed-mode cohesive formulation for crack growth analysis.
Compos Struct. 2016;156:253–262.
[14] Anyfantis KN, Tsouvalis NG. A novel traction–separation law for the prediction of the mixed
mode response of ductile adhesive joints. Int J Solids Struct. 2012;49:213–226.
[15] Alfano M, Lubineau G, Paulino H. On the inverse identification of cohesive zone models.
In: Vari A, editor. Proceedings of 41° Convegno Nazionale; 2012 Sept 5–8. Padova: Padova
University Press; 2012.
[16] Gillham JK. Formation and properties of thermosetting and high Tg polymeric materials. Polym
Eng Sci. 1986;26:1429–1433.
[17] Carbas RJC, Marques EAS, da Silva LFM, et al. Effect of cure temperature on the glass transition
temperature and mechanical properties of epoxy adhesives. J Adhes. 2014;90:104–119.
[18] Carbas RJC, da Silva LFM, Marques EAS, et al. Effect of post-cure on the glass transition
temperature and mechanical properties of epoxy adhesives. J Adhes Sci Technol. 2013;27:2542–
2557.
[19] Zhang Y, Adams RD, da Silva LFM. Effects of curing cycle and thermal history on the glass
transition temperature of adhesives. J Adhes. 2014;90:327–345.
[20] Pinto AMG, Campilho RDSG, Durão LMP, et al. Effect of plug-filling, testing velocity and
temperature on the tensile strength of strap repairs on aluminium structures. J Adhes Sci
Technol. 2012;26:1481–1496.
[21] Bordes M, Davies P, Cognard JY, et al. Prediction of long term strength of adhesively bonded
steel/epoxy joints in sea water. Int J Adhes Adhes. 2009;29:595–608.
[22] Souza JPB, Reis JML. Thermal behavior of DGEBA (Diglycidyl Ether of Bisphenol A) adhesives
and its influence on the strength of joints. Appl Adhes Sci. 2013;1:1–10.
[23] Azevedo JCS, Campilho RDSG, da Silva FJG, et al. Cohesive law estimation of adhesive joints
in mode II condition. Theor Appl Fract Mec. 2011;80:143–154.
[24] Pesquet G, da Silva LFM, Chiaki S. The use of thermally expandable microcapsules for increasing
the toughness and heal structural adhesives. F&IS. 2011;16:18–27.
[25] Hua P, Han X, Li WD, et al. Research on the static strength performance of adhesive single
lap joints subjected to extreme temperature environment for automotive industry. Int J Adhes
Adhes. 2013;41:119–126.
[26] Öz Ö, Özer H. An experimental investigation on the failure loads of the mono and bi-adhesive
joints. J Adhes Sci Technol. 2016;31:2251–2270.
[27] Newbold P, Carlson WL, Thorne BM. Statistics for business and economics. 4th ed. New Jersey
(NJ): Prentice Hall; 1994.
[28] Apps.huntsmanservice.com [Internet]. Switzerland: HUNSTMAN; [cited 2013 Jan]. Available
from: https://apps.huntsmanservice.com/WebFolder/2015_eur_e.pdf
1310   H. ÖZER AND E. ERBAYRAK

[29] Chartoff RP, Menczel JD, Dillman SH. Dynamic mechanical analysis (DMA). In: Menczel JD,
Prime RB, editors. Thermal analysis of polymers fundamentals and applications. New Jersey:
Wiley; 2009. p. 387–495.
[30] Jiawua G, Kuia S, Mong GZ. The cure behavior of tetraglycidyl diaminodiphenyl methane with
diaminodiphenyl sulfone. Thermochim Acta. 2000;352:153–158.
[31] Ozer H, Erbayrak E. Adhesives – applications and properties. 1st ed. Rijeka (Croatia): INTECH
Press; 2016. Chapter 6, Experimental investigation on the self-healing efficiency of araldite 2011
adhesive reinforced with thermoplastic microparticles; p. 169–185.
[32] Khoshravan M, Mehrabadi FA. Fracture analysis in adhesive composite material/aluminum
joints under mode-I loading; experimental and numerical approaches. Int J Adhes Adhes.
2012;39:8–14.
[33] Bardis JD, Kedward KT. Surface preparation effects on mode I testing of adhesively bonded
composite joints. J Compos Technol Res. 2002;24:30–37.
[34] Salehi S, Nygaard R. Full fluid–solid cohesive finite-element model to simulate near wellbore
fractures. J Energ Resour Asme. 2014;137:1–9.
[35] Monteiro JPR, Campilho RDSG, Marques EAS, et al. Experimental estimation of the mechanical
and fracture properties of a new epoxy adhesive. Appl Adhes Sci. 2015;25:1–17.
[36] Shokrieh MM, Rajabpour-Shirazi H, Heidari-Rarani M, et al. Simulation of mode I delamination
propagation in multidirectional composites with R-curve effects using VCCT method. Comp
Mater Sci. 2012;65:66–73.
[37] Fernandes RL, de Moura MFSF, Moreira RDF. Effect of temperature on pure modes I and II
fracture behavior of composite bonded joints. Compos Part B. 2016;96:35–44.
[38] Kwon SH, Zhao Z, Shah SP. Effect of specimen size on fracture energy and softening curve of
concrete: Part II. Inverse analysis and softening curve. Cem Concr Res. 2008;38:1061–1069.
[39] Monteiro JPR, Campilho RDSG, Marques EAS, et al. Experimental estimation of the mechanical
and fracture properties of a new epoxy adhesive. Appl Adhes Sci. 2015;3:1-17.
[40] Budzik MK, Jumel J, Shanahan MER. Experimental investigation of mesoscale crack front triple
line. Appl Phys A. 2014;114:495–501.
[41] Franklin VA, Christopher T. Fracture energy estimation of DCB specimens made of Glass/
Epoxy: An experimental study. Adv Mater Sci Eng. 2013;ID:412601:1–7.
[42] O’Brien TK. Composite materials: fatigue and fracture. 3th ed. Philadelphia (PA): ASTMSTP1110;
1991.
[43] Korjakin A, Rikards R. Comparative study of interlaminar fracture toughness of GFRP with
different fiber surface treatments. Polym Compos. 1998;19:793–806.

Appendix 1. The criteria for evaluating the adherend thickness and initial
crack length
For DCB tests, the following criteria is advised to use in evaluating the adherend thickness and initial
pre-crack length (a0), respectively [41,42]:
)1
E(2h)3 2
(
a0 ≤ 0.042 (A1)
GIC

) 13
GIC a02
(
2h ≥ 8.48 (A2)
E
JOURNAL OF ADHESION SCIENCE AND TECHNOLOGY   1311

Appendix 2. The method followed for calculating the crack length [8,41]
Using the strength of materials approach, the deflection of a DCB can be written as follows:
( 3)
Pa
Δ=2 (B1)
3EI
where E is the elasticity modulus of the adherend. I is the second moment of the area for the adher-
end and defined as I = bh3 12. Substituting Equation (B1) and I into the compliance expression
/
(C = Δ∕ P) and rearranging yield.
( ) 13
CEbh3
a= (B2)
8
The compliance is to be determined from the P-∆ curves. Crack tip rotation and deflection are
also considered for calculating the crack length.

Appendix 3. The Compliance Calibration Method (CCM)


The critical fracture energy can be written in terms of compliance as [43]:
𝜕Π 𝜕U P 2 𝜕C
GIC = − = = (C1)
b𝜕a b𝜕a 2b 𝜕a
where U is the strain energy and Π is the total energy. P is the critical load at which crack growth is
observed. C is the compliance of the adhesive and reads.
C = kan (C2)
where n is the slop of the log C vs. log a curve and k is a constant. Differentiating Equation (C2) with
respect to a and substituting it into Equation (C1) give
P 2 nkan
GIC = (C3)
2ba
The compliance can also be written as C = ∆/P. Then, substituting it into Equation (C3) and
simplifying the relation obtained, we get

nPΔ
GIC = (C4)
2ba

Appendix 4. The Area Method (AM)


Using the Irwin’s formula, the released energy due to crack extension can be evaluated for the critical
fracture energy as follows [43]:
𝜕Π ΔU
GIC = − = (D1)
b𝜕a bΔa
The change in strain energy can be written in terms of P and δ as:
1(
(D2)
)
ΔU = P 𝛿 −P 𝛿
2 1 2 2 1
1312   H. ÖZER AND E. ERBAYRAK

Appendix 5. The work of separation for the exponential cohesive zone model of the Xu
and Needleman law is given as [4]
� �
r−q
�⎧� � ⎡ Δn ⎥ � Δ2 �⎪
� �� ⎤ ⎫
Δn ⎪ Δn 1 − q ⎢ r−1 t
𝜙(Δn , Δt ) = 𝜙n + 𝜙n exp − 1 − r+ − q+ exp − 2 ⎬
𝛿n ⎨
⎪ 𝛿n r − 1 ⎢⎢ 𝛿n ⎥⎥ 𝛿t ⎪
⎩ ⎣ ⎦ ⎭
(E1)
Rearranging the formula with respect to Mode I conditions (i.e. q = 0, r = 0 and ∆t = 0) gives
[ ( ) ( )]
Δn Δn
(E2)
( )
𝜙 Δn = 𝜙n 1 − 1 + exp −
𝛿nc 𝛿nc

View publication stats

You might also like