You are on page 1of 15

Engineering Fracture Mechanics 284 (2023) 109255

Contents lists available at ScienceDirect

Engineering Fracture Mechanics


journal homepage: www.elsevier.com/locate/engfracmech

Load-control vs. displacement-control strategy in fatigue threshold


analysis of adhesives: Effects of temperature
D. Santos a , A. Akhavan-Safar b ,∗, R.J.C. Carbas b , E.A.S. Marques b , S. Wenig c ,
L.F.M. da Silva a
a
Departamento de Engenharia Mecânica, Faculdade de Engenharia (FEUP), Universidade do Porto, Porto, Portugal
b
Instituto de Ciência e Inovação em Engenharia Mecânica e Engenharia Industrial (INEGI), Porto, Portugal
c
Sika Automotive AG, Kreuzlingerstrasse 35, 8590 Romanshorn, Switzerland

ARTICLE INFO ABSTRACT

Keywords: For the design of adhesive joints for service in fatigue, parameters such as the rate of
Fatigue fracture fatigue crack growth and the fatigue threshold energy release rate are of extreme importance.
Polyurethane These parameters can be obtained through fatigue fracture tests using different strategies,
Fatigue threshold
namely Constant load control and constant displacement control. Despite the extensive studies
Temperature
conducted on the fatigue performance of bonded joints, the role of testing methodology on
Mode I
the fatigue fracture behaviour of adhesives is less considered, especially in conjunction with
the effect of temperature in these tests. Accordingly, in this study, a structural polyurethane
adhesive suitable for automotive applications is characterized in terms of its fatigue behaviour
in pure mode I using DCB (double cantilever beam) specimens. Fatigue tests were performed
at different temperatures under both constant load control and constant displacement control
conditions. A compliance-based method was used to measure the crack size throughout tests.
The threshold energy release rate was calculated for different test temperatures and the two
testing strategies were compared. It was seen that a constant displacement control methodology
is superior to a constant load control methodology when testing for the threshold energy release
rate, due to its shorter duration and low sensitivity to initial testing conditions. Differences
observed between fatigue testing methodologies were exacerbated by the effect of temperature
on the adhesive.

1. Introduction

Adhesive bonding has been increasingly used in the last few decades for structural and non-structural components in the
automotive industry, and even more so now with the appearance and growth of electric vehicles on the market [1]. With the
use of adhesive bonding, it is possible to significantly reduce the weight of structures and increase the fuel efficiency of vehicles.
Due to the nature of adhesives, their use allows for the joining of different materials that otherwise could not be joined. They are
also useful for acoustic insulation and elimination of corrosion and vibration problems while presenting the great advantage of
providing a more uniform stress distribution when compared to traditional joining processes [2–5].
Fatigue life is supremely important in the design of adhesive joints, and especially in structural adhesive joints. For design,
two approaches are common: a safe-life design philosophy which is more conservative, based on the number of cycles required
for joint failure, and a fatigue crack growth approach (FCG) which allows for more versatile and precise design. It is possible to

∗ Corresponding author.
E-mail address: aAkhavan-Safar@inegi.up.pt (A. Akhavan-Safar).

https://doi.org/10.1016/j.engfracmech.2023.109255
Received 19 December 2022; Received in revised form 9 February 2023; Accepted 6 April 2023
Available online 12 April 2023
0013-7944/© 2023 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
D. Santos et al. Engineering Fracture Mechanics 284 (2023) 109255

design a joint so that the threshold energy release rate 𝐺𝑡ℎ is never surpassed, but this is at times limiting and the FCG approach
gives the designer more freedom to define a level of fatigue crack growth that may be acceptable for a given application, although
a full characterization of the fatigue behaviour is required [6]. In the past, several authors have studied the fatigue behaviour
of adhesives using different joint geometries and adhesive systems [7,8], although the focus has been mainly on epoxy adhesives
than, for example, polyurethanes. Mode I fatigue behaviour in particular has also previously been studied [9–11]. Using ADCB
(asymmetric double cantilever beam) joints, Azari et al. [12] showed that the fatigue threshold increases with an increase in load
or displacement ratio of the fatigue test, when measured in terms of 𝐺𝑚𝑎𝑥 or 𝛥𝐺. Azari et al. [13] found no difference between
the threshold obtained through load-control and displacement-control tests. The same authors [13] found no difference between
the fatigue threshold measured from a fatigue crack versus an uncracked fillet, provided a cohesive crack path, which opposed the
suggestion by Ashcroft and Shaw [14] that they would be different, although it was also found that a lower 𝐺 was measured from
a fatigue pre-crack if the crack path was instead interfacial. Both Mall and Johnson [15] and Mangalgiri et al. [16] found a higher
debond rate under displacement-control versus load-control. Additionally, Mangalgiri et al. [16] saw that this greater debond rate
increased with increasing adherend thickness. The bond-line thickness plays a major part in the fatigue resistance of a joint, which
tends to increase with thickness [11,17], although for large thicknesses diminishing returns are observed [2].
Much as fatigue life, the temperature is an equally important factor in the design of joints for the automotive industry, considering
the environments road vehicles are subjected to year-round and around the world. This effect has been studied before by several
researchers [7,18,19], and it is generally true that an increase in temperature weakens adhesive properties, such as tensile and shear
strength. Temperature also has a pronounced effect on fatigue behaviour as reported, for example, by Usman et al. [20], that an
increase in temperature from −55 ◦ C to 80 ◦ C for an FM94 epoxy adhesive led to an increase in FCG rate. Tan et al. [21] equally
showed that an increase in temperature, between −40 ◦ C and 80 ◦ C, led to a weakening of fatigue performance for a SikaFlex® -
265 polyurethane adhesive, especially around the measured Tg of −66 ◦ C; their work used Thick Adherend Shear Joints (TASJs),
and showed a strong change in the failure surface. Usman et al. [20] suggest that the main effect of temperature is a change in
the failure mechanism. Usually, the temperature does not come on its own in applications, and humidity is an equally important
environmental factor. The effect of ageing tends to be that of a reduction in the tensile properties of an adhesive, as reported by
several authors [22–24]. In particular, using an epoxy adhesive in Arcan joints, da Costa et al. [25] found that cyclic ageing has
an irreversible effect on fatigue performance, which can only be partially recovered after drying, with diminishing returns with
increasing cycles.
Despite the existence of standardized methods for fatigue crack growth and fatigue threshold measurement of metals and
composites, ASTM E647 and ASTM D6115 respectively, no standardized approach has been proposed for adhesives specially for
the threshold energy evaluation of bonded joints. Regarding the loading strategy, ASTM E 647 proposes a load control fatigue test
in which the load level decreases during the test until the threshold energy is reached. However, the crack closure phenomenon is a
major concern when using this standard [26–29]. On the other hand, ASTM D6115 proposes another loading strategy (displacement
control test) for mode I loading of composite materials. While fatigue methodology has been extensively studied in the past for both
metals and composites, not only the fatigue threshold analysis of adhesives is not standardized, but this parameter has also rarely
been addressed in the literature. Furthermore, to the best of the authors’ knowledge, the influence of temperature on fatigue crack
growth in combination with the test approach (load control vs. displacement control) for adhesive materials has not been analysed
so far. This work looks to do exactly this, by performing fatigue tests using different methodologies and temperatures. While the
whole regions in the Paris law curve are analysed, however, the work mainly focuses on the threshold energy measurement of
bonded joints. Accordingly, a high-cycle fatigue regime is contemplated, where service loads should be well below the threshold
range. Some real applications are assumed in which adhesives are used in the automotive industry, where the joint is exposed to a
wide range of temperatures in service e.g. in the batteries as discussed by which Zwicker et al. [1].
Firstly, quasi-static strength and fracture tests are performed at 23 ◦ C and 60 ◦ C to characterize the adhesive in terms of strength
and mode I fracture and to define fatigue testing parameters. Then, fatigue tests are performed in both displacement-control and
load-control, at 23 ◦ C and 60 ◦ C (frequency=10 Hz, load ratio=0.1 in constant load control tests and almost 0.1 for the constant
displacement control tests), in order to not only compare the effect of using different methodologies for testing but also what impact
temperature may have on the results given two different methodologies.

2. Experimental

2.1. Materials

The adhesive used in this study was a two-component polyurethane adhesive with the properties listed in Table 1. Physical
properties are taken directly from the manufacturer’s datasheet, while tensile properties were determined through tensile tests using
bulk specimens.
This adhesive was cured through a two-step process of 24 h curing at room temperature followed by 4 h post-curing at 80 ◦ C
for all specimens used.

2.2. Bulk specimen manufacturing

Bulk specimens were manufactured in order to determine the tensile strength, stiffness and ductility of the adhesive. This
manufacturing consisted of machining a ‘‘dogbone’’ profile using a CNC machine, based on the British Standard BS 2782 [30],
from strips cut out of a 2 mm thick cured adhesive plate. Right after cutting the samples into the final shape they are ready for the
tensile test. The scheme of these specimens is shown in Fig. 1.

2
D. Santos et al. Engineering Fracture Mechanics 284 (2023) 109255

Table 1
Physical and mechanical properties of the adhesive used.
Property Polyol Isocyanate Mixed
Specific gravity at 25 ◦ C (g/cm3 ) 1.57 1.22 –
Viscosity at 25 ◦ C (mPa s) 7000 20 1100
Glass transition temperature (Tg ) (◦ C) −5
Maximum tensile strengtha (MPa) 3.4 ± 0.09
Maximum tensile straina (%) 33.4 ± 1.34
Young’s Modulusa (MPa) 20.3 ± 1.23
a Mechanical properties at room temperature.

Fig. 1. Geometry of manufactured specimens: (a) bulk specimen; (b) standard DCB specimen; (c) short DCB specimen; a0 shows the pre-crack size, dimensions
in mm.

2.3. Double cantilever beam (DCB) specimen manufacturing

Mode I quasi-static tests, as well as fatigue tests, were performed using DCB specimens. These specimens used substrates made
(machined into the final shape) from a high-strength steel plate, DIN 40CrMnMo7, with geometry according to ASTM D3433-99 [31].
A scheme of these specimens is shown in Fig. 1. For fatigue tests at high temperature, a shortened version of these specimens was
used, with a substrate length of 192 mm, shown in Fig. 1. For either specimen type, the crack length is considered from the loading
pin to the crack tip introduced in manufacturing at the start of the adhesive layer. For the shorter specimens, the fourth, innermost
pin hole is used, and crack length is measured from here. The adhesive thickness used was 4 mm, and the initial crack length
value was 45 mm for standard specimens, and 20 mm for shortened specimens in order to leave enough space for fatigue crack
propagation in the joint.
As the adhesive used was liquid, a special mould was designed and constructed in order to manufacture joints in a repeatable and
simple way. The mould was manufactured through 3D printing using a polycarbonate filament in a Prusa i3 MK3S+ 3D printer. Each
originally designed part of the mould was broken into two fitting halves to be printed, which are divided into end parts, guaranteeing

3
D. Santos et al. Engineering Fracture Mechanics 284 (2023) 109255

Fig. 2. Mould filled with adhesive, only missing the top substrates.

the alignment of the substrates; middle parts, separating the adhesive joints; side parts, covering the sides of the outermost joints,
and which are merely one-sided versions of the middle parts. These parts were designed with the dimensions required to fit the 3D
printed mould inside a perimeter of metal pins mounted on a plate typically used for holding substrates in joint manufacturing.
With the use of this mould, it was possible to keep the adhesive stable inside the joints for the time required for the first stage of
curing, while controlling the thickness of the adhesive with the use of proper spacers. The mould also allowed for minimizing the
effect of the slight adhesive shrinkage which was observed in a trial manufacturing without a mould, through having extra adhesive
held at the sides of the joints.
Specimens were first sandblasted, a method chosen due to its speed and capability of achieving a homogeneous and repeatable
surface roughness [32]. Subsequently, the sandblasted surfaces were cleaned with acetone to remove any leftover dirt, and a thin
primer layer was applied using a foam applicator. This primer requires twenty minutes left out in the air before it is dry and the
manufacturing process can proceed. The use of this primer is important for improving adhesion between adhesive and substrate
beyond what simple mechanical surface treatments are capable of achieving, therefore reducing the likelihood of adhesive failure,
undesirable during the characterization of the adhesive. With the surface treatment complete for all substrates, half of these were
introduced in the mould with the spacers and mould parts coated in a release agent for better disassembly after the first curing
stage. The surfaces of the substrates were wet with adhesive to push out any air trapped in the surface roughness, and adhesive was
poured into the mould up to the required thickness (Fig. 2). After this, the top substrates were introduced in order to complete the
joint. A metal plate fitting the pins was placed on top of the joints and the mould was pressed with 80 kg of weights placed on top,
for a period of 24 h, at room temperature.
With the first stage of curing completed, the joints were removed from the mould and cleaned of any extra adhesive on their
sides. The joints then went through the post-curing phase, and spacers were carefully removed after this, leaving a short crack on
one end due to a shim included in one of the spacers. Spacers and 3D-printed parts could then be reused in the manufacturing of
new joints.

2.4. Testing conditions

The first two tensile strength tests made at room temperature used a clip-on extensometer (with the gauge length of 25 mm) in
order to measure the strain in the dogbone specimen. However, since the adhesive is very flexible, the specimen rotated significantly
due to the weight of the extensometer, leading to possible errors in the results. Due to this, subsequent tests were made using only
a digital image correlation (DIC) technique to measure the material strain, although less than 5% difference was found relative
to the extensometer results. Therefore, results for strength tests at room temperature presented in this work are a mix between
extensometer and DIC results. Tests at high temperature also used a DIC technique, exclusively. All tests used a displacement rate
of 1 mm/min in an Instron 3367 electro-mechanical testing machine with a load cell capacity of ± 30 kN.
For quasi-static fracture tests, there was a desire to use standard specimens, due to their larger size when compared with shortened
specimens, which would allow for a more clear strain energy release rate plateau in testing. These standard specimens did not fit the
environmental chamber available, an Instron 3119-606, and so a custom chamber was built in-house for high temperature testing.
This thermal chamber was made in a cost-effective manner, having been designed as a double wall of 304 aluminium plating,
assembled through riveting, inside which was inserted an insulation layer made of rock wool. Due to its size, this chamber could
only be used with the available servo-hydraulic testing machine, an Instron 8801 with a load cell rated to ± 100 kN, and not with
the electro-mechanical machine. This chamber also did not possess automatic temperature control systems, and so could only be
used for quasi-static testing. It should be noted that the load cell used in this study could measure the load with acceptable accuracy,
however, to increase the accuracy of the obtained load, a load cell with lower capacity should be used.

4
D. Santos et al. Engineering Fracture Mechanics 284 (2023) 109255

In the end, quasi-static fracture tests were performed using standard DCBs, with room temperature tests performed in the Instron
3367 electro-mechanical machine, and high temperature tests performed in the Intron 8801 servo-hydraulic machine. Fatigue
fracture tests were performed using Intron 8801 servo-hydraulic machine, with high temperature fatigue tests conducted using
the available Instron 3119–606 environmental chamber.
For fatigue fracture tests, two different control methodologies were used: a displacement-control and a load-control methodology.
When using a displacement-control methodology, a maximum and minimum displacement are set for each cycle. Due to the
joint being initially undamaged, these displacement values translate into high loads at the start of the test, which leads to a quick
opening of the joint, i.e. high crack growth rates, and high strain energy release rates. As the joint is damaged, the crack opens
considering the constant maximum displacement applied a progressive reduction in load is observed leading to lower strain energy
release rates and slower crack growth. This development entails that the test begins near the unstable crack growth region of the
Paris curve and progresses towards the threshold region of the curve, where crack growth slows down to very low levels. For this
methodology, the maximum displacement values were chosen corresponding to a quasi-static load equal to 70% of the maximum
load in quasi-static tests at room and high temperature although, since the goal is to obtain the threshold value, a rather wide range
of maximum displacement values is possible, as all will lead to the threshold region. Care should be taken mostly to guarantee
that the displacement is not so high that the joint breaks before reaching the threshold region, or so low that the crack does not
propagate initially, although it should be kept as high as possible in order to capture the maximum possible portion of the Paris
law curve. During the tests, the minimum displacement values were altered in order to keep the monitored load ratio at a value of
around 0.1. The frequency was kept constant at 12 Hz during the fatigue tests.
On the other hand, when using a load-control methodology, a set maximum and minimum loads are defined for each cycle. With
an initially undamaged joint, these loads lead to generally small displacements; the strain energy release rate is low and the fatigue
crack growth rate is equally small. The test, therefore, begins at the threshold region and, when the joint is sufficiently damaged,
crack growth begins. The crack grows in a stable manner for most of the test, together with the displacement, with increasing
strain energy release rates and crack growth rates. Eventually, the joint is damaged to a point where the crack grows unstably,
and the joint fails completely in a few cycles. Tests were performed in a one-shot manner, setting a maximum and minimum load
and leaving the joint in fatigue until the joint fracture. Tests were performed with different load levels, starting at around 70% of
the maximum quasi-static load, with subsequent tests performed at load levels below this. For each load level, only one test was
performed. This was done since fatigue fracture tests have little scatter, as was observed with displacement-control tests, performed
before load-control tests. Additionally, and more importantly, with load-control tests the trend in approaching 𝐺𝑡ℎ with reducing
load level is more important than, for example, having very secure results for a high load level but no trend. For this reason, several
load levels were tested only once to get an idea of the trend, and eventually, the value of the threshold was approached.
The temperatures picked for testing, 23 ◦ C and 60 ◦ C, are above the adhesive’s Tg of −5 ◦ C. Since this is an adhesive suitable for
automotive applications, and since vehicles experience a wide range of temperatures in the environment, these two temperatures
were picked within those that vehicle parts might experience, to see what effect they may have on this adhesive. For high temperature
tests, the temperature of the specimen was controlled using a thermocouple placed in the testing chamber. The test conditions
considered in this study are listed in Table 2.

3. Data reduction method

For all fracture tests knowledge of the crack length over a test is needed, although direct measurement of this value is not always
possible due to several reasons such as lack of vision, high loading frequency in fatigue, and the existence of a fracture process zone.
A method capable of solving these problems is the compliance-based beam method (CBBM) [33], which is based on linear elastic
fracture mechanics and only requires the joint’s compliance, calculated from load (𝑃 ) and load line displacement (𝛿) values during
the test. From this compliance, the method is capable of estimating the equivalent crack length and critical fracture energy. This
critical energy release rate is calculated for mode I through:
( 2 )
6𝑃 2 2𝑎𝑒𝑞 1
𝐺𝐼𝐶 = + (1)
𝐵 2 ℎ ℎ2 𝐸 𝑓 5𝐺

where 𝑃 is the load, 𝐵 is the width of the specimen, ℎ is the height (thickness) of the substrates, 𝑎𝑒𝑞 is the equivalent crack length, 𝐺
is the shear modulus of the substrate, and 𝐸𝑓 is the flexural modulus of the specimen, obtained from the reference compliance in the
test. This flexural modulus helps to account for the FPZ (fracture process zone), stress concentrations, and some beam deformation
and is given by:
( )
12(𝑎0 + |𝛥|) −1 8(𝑎0 + |𝛥|)3
𝐸 𝑓 = 𝐶0 − (2)
5𝐵ℎ𝐺 𝐵ℎ3
where 𝐶0 is the initial compliance and 𝛥 is a correction factor for the rotation of the specimen, given by:

( )( ( )2 )
1 𝐸 𝛤 1.18𝐸
𝛥=ℎ 3−2 , with 𝛤 = (3)
11 𝐺 1+𝛤 𝐺
where 𝐸 and 𝐺 are Young’s modulus and shear modulus of the adherend, respectively, and 𝑘 is the shear stress distribution constant,
which serves to correct for the deflection caused by shear forces, and is obtained as the x-intercept of a linear regression applied to
𝐶 1∕3 as a function of 𝑎 data.

5
D. Santos et al. Engineering Fracture Mechanics 284 (2023) 109255

Table 2
Listing of performed tests.
Temp. (◦ C) Disp. rate DIC N. of tests
(mm/min)
Tensile test 60 1 Yes 3
(bulk) 23 1 Partial 3
Temp. (◦ C) Disp. rate No. of tests
(mm/min)
Quasi-static 60 0.2 3
fracture tests 23 0.2 3
Temp. (◦ C) Freq. (Hz) Short DCB Max. Disp. N. of tests
(mm)
Fatigue 60 12 Yes 0.500 3
disp. control 23 12 No 0.925 3
Temp. (◦ C) Freq. (Hz) Short DCB Load level N. of tests
(N)
60 12 Yes 600 1
60 12 Yes 520 1
Fatigue 23 12 No 1080 1
load-control 23 12 No 810 1
23 12 No 630 1
23 12 No 400 1

Table 3
Quasi-static strength test results at different temperatures.
Temperature 23 ◦ C 60 ◦ C
Young’s modulus [MPa] 20.3 ± 1.23 20.5 ± 1.14
Tensile strength [MPa] 3.4 ± 0.09 2.4 ± 0.08
Elongation at break [%] 33.4 ± 1.34 15.9 ± 0.88

Since the compliance can be calculated from 𝑃 − 𝛿 data as:

𝛿 8𝑎3𝑒𝑞 12𝑎𝑒𝑞
𝐶= = + (4)
𝑃 𝐵ℎ3 𝐸𝑓 5𝐺𝐵ℎ
then 𝑎𝑒𝑞 , the equivalent crack length, can be calculated from this equation.
CBBM is capable to measure the crack size per load cycle. Using this compliance-based method the crack size for each loading
cycle was obtained from the load–displacement information provided by the fatigue test machine. By knowing the size of the crack at
each cycle the rate of crack growth per cycle was calculated. A filtering process was used to make the curves smoother by reducing
the noise, which is mainly due to the noise on displacement measured by the test machine. To achieve this, the Savitzky–Golay
approach was considered, a filtering technique that is applied to the digital data points to make them smooth by increasing the
accuracy of the data and removing the noise without distorting the data trend.

4. Experimental results and discussions

4.1. Quasi-static strength results

In Table 3 the quasi-static tensile test results are presented, for which the stress–strain curves are shown in Fig. 3.
Both temperatures tested are above the −5 ◦ C Tg of the adhesive. With the increase in temperature, there is a significant reduction
in maximum tensile strength and maximum strain. Studies with other adhesives show similar results. Banea and da Silva [34]
reported this decrease in strength and strain when using an AV118 epoxy adhesive. The same trend was found by both Houjou [19]
and Banea et al. [35], the latter when their epoxy adhesive was tested above its Tg of 155 ◦ C.

4.2. Quasi-static fracture test results

In Fig. 4(a) a load–displacement curve for each temperature tested is shown, and in Fig. 4(b) the corresponding resistance curves
are presented. From Fig. 4(a) it can be seen that an increase in temperature of 35 degrees has the effect of reducing the maximum
load and displacement attainable by the joint by 50% and 61%, respectively, although it should be noted that joint geometry will
have a strong effect on these values, and also on the stiffness. For this reason, load–displacement curves should not be used by
themselves to judge adhesive properties. Due to the very consistent geometry that was attained through a precise manufacturing
procedure, it is possible to observe some correlations with the stress–strain curves shown before (where the stiffness was maintained
with temperature while maximum stress and strain decreased). This was also expected as both tests were performed above the Tg .

6
D. Santos et al. Engineering Fracture Mechanics 284 (2023) 109255

Fig. 3. Tensile stress–strain curves at different temperatures.

Fig. 4. Representative load–displacement (a) and resistance curves (b) of mode I fracture tests at different temperatures where the horizontal lines show the
GIc obtained based on the plateau part of the R-curves.

Tan et al. [21] also found a reduction in maximum stress and displacement for a polyurethane adhesive tested above its Tg ,while
the joint stiffness remained nearly the same for high temperatures considered (with a difference of about 30 ◦ C as in the current
study).
From these resistance curves, it can be also found that 𝐺𝐼𝑐 is reduced by 82% with temperature increased from 23 ◦ C to 60 ◦ C.
This makes sense as the fracture toughness is related to energy absorption, and this energy absorption is related to stress and strain
values, which are equally reduced with temperature, as we have already seen, and a correlation has previously been found in the
literature [35]. These results are summarized in Table 4.
From looking at the resistance curves shown (Fig. 4(b)), a peak may be noted at the start of each one. Although an ideal resistance
curve should present the longest and clearest plateau possible, this was not striven for due to the reasons presented next. Due to
several factors dealing with crack propagation which will be discussed ahead, secondary cracks formed ahead of the crack tip upon
executing a pre-cracking procedure. Here ‘‘pre-cracking’’ means a process that consists of loading the DCB to its maximum strength
in an opening mode in order to create a real crack ahead of the artificial crack made during the manufacturing process. However,
this pre-cracking procedure results in a multi-site cracking phenomenon where multiple cracks initiate through the adhesive layer

7
D. Santos et al. Engineering Fracture Mechanics 284 (2023) 109255

Table 4
Critical strain energy release rate values for quasi-
static fracture tests at different temperatures.
Temperature 23 ◦ C 60 ◦ C
𝐺𝐼𝑐 (N/mm) 2.82 ± 0.05 0.51 ± 0.06

at the same time (see also Section 4.4) Although the use of CBBM does not require the constant monitoring of crack length, an
initial crack size value is needed, which was very difficult to account for with these secondary cracks, and for this reason, tests were
made straight away, with no pre-cracking, using only the crack introduced in manufacturing, therefore leading to this initial peak.
However, this was not seen as a problem as the plateaus obtained were still quite easily discernible and consistent over several tests.
According to Fig. 4(a) for presented curves, the range of crack length considered for the calculation of 𝐺𝐼𝑐 is (85–125 mm) and
(75–120 mm) for room and high temperatures, respectively.

4.3. Fatigue fracture test results

Fatigue test data in this study is presented using Paris law curves. These curves follow the Paris law relation:
𝑑𝑎
= 𝐶𝑓 (𝐺)𝑚 (5)
𝑑𝑁
𝑑𝑎
where 𝑑𝑁 is the crack growth rate observed, 𝑓 (𝐺) is a function of the strain energy release rate and 𝑚 is the slope of the stable crack
growth region of the Paris law curve. The equivalent crack length obtained using CBBM (discussed in Section 3) was considered in
this analysis. Several parameters are possible to be considered in the Paris law function, but for this study, 𝐺𝐼,𝑚𝑎𝑥 was used for its
simplicity. This parameter is calculated for every cycle by applying CBBM to the maximum load and displacement values at that
cycle.
Fig. 5 shows fatigue crack growth curves for constant load-control tests at several different load levels, for tests at room
temperature, as a function of maximum 𝐺𝐼,𝑚𝑎𝑥 . As mentioned in the experimental details section, several load levels were tested,
starting at around 70% of the maximum quasi-static strength and then decreasing the load with each subsequent test until reaching
a condition at which there is no significant crack growth. The threshold is considered as the strain energy release rate for which
the fatigue crack growth rate is sufficiently low to be almost imperceptible. This value is typically considered to be around 10−6
mm/cycle, or 1 mm crack growth per one million cycles, and was the one considered in this study.
It can be seen that, for each subsequently lower load level, a higher number of cycles is observed, and a stationary value for the
strain energy release rate is seen at ever lower values of 𝐺𝐼,𝑚𝑎𝑥 . This can be seen as a pseudo-threshold value, although the ‘‘real’’
threshold, as considered by definition, must fall below 10−6 mm/cycle. For higher load levels, this stationary value is limited to a
crack growth rate of 10−4 to 10−5 mm/cycle, and from here the test progresses to failure. Meanwhile, the 630 N load level almost
attains a slow enough fatigue crack growth rate for consideration of a threshold, and the 400 N load level attains it but does not
evolve from there with growing strain energy release rates. For this reason, the threshold for constant load-control tests can be said
to fall somewhere between 0.2 N/mm and 0.3 N/mm for room temperature, where a small enough fatigue crack growth rate would
be reached, while the test then progressed to eventual failure of the joint.
It can be seen that, for higher load levels, the maximum attained 𝐺𝐼 value surpasses 𝐺𝐼𝑐 . A possible reason for this is the increased
strain rate that is observed in fatigue tests. Polyurethanes are known to be affected by strain rate [36] and it is possible that the
adhesive is tougher under higher strain rates. In the fatigue tests performed, the frequency is 12 Hz. In a constant load-control test,
the constant load requires ever-increasing displacement over the course of the test, which coupled with the high frequency of the
test leads to greatly increasing strain rates. In a constant displacement-control test, on the other hand, the initial set displacement
coupled with this frequency would lead to a much higher strain rate than quasi-static tests at the start of the test.
In Fig. 6 the previous constant load-control test results closest to the threshold are compared to a constant displacement-control
test result. It can be seen that either methodology resulted in similar slopes for the Paris law curve. However, for the constant
displacement-control tests, an ever so slightly lower threshold value than that which would be considered given the constant
load-control results is attained. The reason for this is likely to be different crack conditions at the threshold point. For constant
displacement-control tests, the threshold is reached after significant crack propagation. This crack propagation occurred near the
interface of the joint, after the initial crack introduced in manufacturing kinked towards one of the interfaces following a few
fatigue cycles. Therefore the threshold obtained for constant displacement control is one where the crack is close to the interface.
For constant load-control tests, however, the threshold is achieved at the start of the test, where the crack has not yet propagated.
This crack is, therefore, at the middle of the adhesive layer, and it could be expected that the value of the threshold would be
different, in this case, higher, due to different fatigue resistances between a crack close to the interface versus the middle of the
adhesive. This is in agreement with Azari et al. [13], who found a lower threshold with an interfacial crack. After achieving the
threshold with the crack in the middle of the adhesive layer, however, for constant load-control tests, the crack again kinks to the
interface and propagates very close to it.
Constant load-control tests are typically executed by changing the load level mid-test, incrementally, such that the threshold level
is found. However, in this study, in order to compare simple constant displacement- versus constant load-controlled methodologies,
one-shot tests were made. For the fatigue crack propagation, constant load-controlled tests made in a one-shot manner were generally

8
D. Santos et al. Engineering Fracture Mechanics 284 (2023) 109255

Fig. 5. Representative curves of fatigue crack growth as a function of maximum 𝐺𝐼 , for constant load-control tests at different load levels, at room temperature.

faster than constant displacement-controlled tests, and would naturally be faster than other constant load-controlled tests where the
load was incrementally reduced and increased until crack propagation was observed, although here the one-shot variety of testing
will require more joints to test. The constant load-controlled tests performed did not always attain a low enough crack growth
rate, meaning they were not always useful in order to determine the threshold. If tests were made by incrementally changing
the load, the new problem that would arise is that of time, since the multiple adjustments of the load would lead to a test of a
length equivalent to the sum of several constant load-controlled tests as per this work, sometimes in the several millions of cycles.
Additionally, for automatic alteration of the loading level during the test one should know in advance at which loading cycle to
increment or decrement the load level, which is difficult to say. Several tests may be required, at constant stress levels, to calibrate
the loading blocks in the test. If one does not know in advance how exactly to program the constant load-control test, the alteration
will need to be manual. Depending on the equipment, this may require stopping the test and restarting it, which would introduce
errors due to phenomena such as stress relaxation, may require someone to be near the machine to monitor it, and would make this
long-term fatigue test harder to perform. In this regard, the constant displacement-control methodology is much superior for finding
the threshold, as each constant displacement-controlled test invariably reached the threshold in less than 1.5 million cycles for this
adhesive, in a more reliable way than constant load-controlled tests could. Using an infrared thermometer the local temperature of
the test specimen was monitored during the test for joints tested at room temperature. No considerable increase in temperature was
observed due to the cyclic loading.
To further compare both methodologies, Fig. 7 shows the evolution of crack size over normalized constant displacement- and
constant load-control tests. It can be seen that these tests progress in opposite manners, the constant displacement-control test
starting with faster crack growth and tapering off when it approaches the threshold, while the constant load-control test begins at
slow crack growth, and this rate grows until unstable crack growth is reached. While for a constant load-control test one must define
a load capable of attaining the threshold, with the correct fatigue crack growth rate, and change this load if the correct rate is not

9
D. Santos et al. Engineering Fracture Mechanics 284 (2023) 109255

Fig. 6. Representative curves of fatigue crack growth as a function of maximum 𝐺𝐼 , for constant load- and constant displacement-control tests, at room
temperature.

Fig. 7. Evolution of the crack size as a function of cycles for different test approaches.

reached, with a constant displacement control there exists almost an automatic guidance towards the threshold region, which makes
this kind of test much easier to define experimentally. It should also be noted that a properly made constant displacement-control
test does not lead to joint failure, and possibly a second test could be performed with the same joint, the crack being arrested at a
slightly larger size.

10
D. Santos et al. Engineering Fracture Mechanics 284 (2023) 109255

Fig. 8. Representative curves of fatigue crack growth as a function of maximum 𝐺𝐼 , for constant displacement- and constant load-control tests at room and
high temperature.

In Fig. 8 tests at 60 ◦ C are compared with the previous room temperature ones. Here, the development of constant displacement-
and constant load-control tests, when compared to room temperature results, is disparate. It would be expected, from quasi-static
strength and fracture results, that 𝐺𝐼,𝑡ℎ should be reduced with the increase in temperature, due to the weakening of properties with
temperature. This indeed happens with constant displacement-control tests, but does not happen with constant load-control tests,
for which 𝐺𝐼,𝑡ℎ slightly increases instead. This variation in threshold energy release rate value, including the ranges observed for
each temperature, can be seen in Fig. 9. In this regard, the constant displacement-control test gives a more predictable and precise
result, correlating with quasi-static tests, while the constant load-control test fails to correlate. With regard to the fatigue threshold,
the aforementioned effect of the crack condition would yet again be present, leading to different fatigue threshold values. While
the constant load-control test remains with the crack in the middle of the adhesive layer initially and measures the threshold there,
the constant displacement-control test, with its high initial load, instantly moves the crack close to the interface, and only then
measures the threshold.
The slope of the curves also seems to increase massively for constant load control at high temperature, while for constant
displacement control, it is similar to that observed for room temperature, although for each value of 𝐺𝐼,𝑚𝑎𝑥 the fatigue crack growth
rate is higher than it was for room temperature, which would make sense given the weakening of mechanical properties with
temperature. One possible reason for this high slope in the constant load control is that, after the crack starts propagating from the
threshold, requiring a higher load due to being achieved with a crack in the middle of the adhesive, the load is then too high for
the crack growing close to the interface, leading to much greater crack growth rates. In any case, the slope observed is very short
also as a consequence of this. The values of 𝐺𝐼 above 𝐺𝐼𝑐 observed for both constant load- and constant displacement-control tests
could again be related to strain rate effects. Another explanation is the following: the value of 𝐺𝐼𝑐 reported for quasi-static tests is
an average value from the stable part, the plateau, of the resistance curve, with conditions also differing from fatigue. It was seen
in Fig. 4(b) that the value of 𝐺𝐼 varies in this resistance curve and is a function of crack size, with the highest point at the start of
the test with a smaller crack. For constant displacement-control, this would explain the initially high 𝐺 values, while for constant
load-control it would partially explain the higher threshold, as we divide 𝐺𝑚𝑎𝑥 by a 𝐺𝐼𝑐 that is from the middle of the R-curve and
not from the start. This explanation for constant load control is aided by the fact that the shorter specimen with a shorter initial
crack size will show a stronger effect of this R-curve than a specimen with a longer initial crack.
Other tests were performed with lower load levels, not reported in this study, but none led to crack propagation even after 1
million cycles. The differing slope of the constant load-control tests at the high temperature could be a consequence of the high load
values required for initial crack propagation, but in fact, the slopes observed are unreliable, being made up of only a few points at
the end of the tests, with a somewhat flatter zone near the vertical drop of the threshold. In comparison with constant load-control
tests, constant displacement-control tests should be much more reliable and insensitive to initial conditions of the crack tip and
manufacturing problems when a 𝐺𝐼,𝑡ℎ value is desired, as they fly through the initial crack tip and retrieve the desired quantity
somewhere in the middle of the joint. However, further studies are required to analyse the effect of adhesive thickness specially

11
D. Santos et al. Engineering Fracture Mechanics 284 (2023) 109255

Fig. 9. Evolution of the value of 𝐺𝑡ℎ observed in constant displacement- and constant load-control tests with temperature. Ranges for each temperature are
shown.

Table 5
Paris law parameters for fatigue tests at different temperatures.
Temperature 23 ◦ C 60 ◦ C
Control method Disp. Load Disp. Load
𝐺𝐼,𝑡ℎ (N/mm) 0.1–0.2 0.2–0.3 0.02–0.04 0.5
𝑚 3.2 ± 0.5 3.7 ± 0.9 3.0 ± 0.3 11.8 ± 1.1

when the crack kinks toward the interface. Creep/relaxation effects on fatigue crack growth rate and fatigue threshold energy were
not analysed in this study. Understanding the impact of these phenomena requires a broader range of fatigue experiments and also
creep/relaxation testing to more accurately determine the role of each phenomenon on the fatigue threshold energy of the tested
adhesive. However, due to the very low level of stress (400 N) applied to the joint, no significant creep/relaxation is expected for the
constant load control tests to obtain the fatigue threshold energy. On the other hand, it should be noted that the fatigue threshold
energy is obtained in constant displacement control tests at the end of the experiments when the stress level is very low and the
relaxation cannot significantly change the 𝐺𝐼,𝑡ℎ obtained.
In Table 5 the parameters defining the Paris law for the different test conditions are shown.

4.4. Fracture surface analysis

The fracture surfaces from quasi-static tests were cohesive for both temperatures tested and the fracture mechanism was in every
way similar, as is visible in Fig. 10. Fracture began at the crack introduced in manufacturing and turned to one of the interfaces in
just a few millimetres, continuing to develop along the interface from then on, although always cohesively, in the adhesive layer,
as evidenced by the fracture surfaces, where no primer is directly visible. An example of the crack path evolution can be seen in
Fig. 11. The reason for the crack not propagating through the middle of the adhesive layer, where the initial crack was introduced,
is likely one of the stress concentrations that are typically observed at the interface for DCB specimens. As mentioned previously,
smaller secondary cracks are observed ahead of the main crack. These smaller cracks form at both interfaces, only sometimes across
the whole width of the specimen, and tend to grow as the test goes on, eventually coalescing into one larger crack or leading to
crack-bridging, a process where the main crack jumps from one interface to another where a secondary crack already exists. The
existence of these secondary cracks is likely due to two main factors: stress concentrations along the interfaces and the ductility of
the adhesive, which lends itself to the formation of a large fracture process zone. Banea et al. [2] found the same kind of secondary
cracks forming ahead of the main crack with their polyurethane adhesive, attributing this to the large fracture process zone present.
Meanwhile, Hasegawa et al. [37] found that the use of thinner, more flexible substrates could fix the crack-bridging they initially
observed in their polyurethane adhesive joints. The use of these more flexible substrates made the stress more localized around the
crack tip and reduced the size of the fracture process zone, thus avoiding a jump to the interfaces, as had happened with the stiffer
substrates.
Other than changing the thickness of adhesive present on the surfaces, the increase in temperature also seems to increase the
homogeneity of this leftover adhesive, making the fracture look more ductile. These two factors, the thickness leftover and the
homogeneity, are the reason for the slight change in tone between the fracture surfaces at the two temperatures.
In Fig. 12 the fracture surfaces of fatigue tests at room temperature are shown. It can be seen that, while for constant displacement
control, the crack eventually stops growing somewhere in the middle of the joint, for constant load control the joint breaks fully.
For the joint from constant displacement control, the point where the crack stops is curved. This is due to the existence of plane
stress conditions at the edge of the joint, which then becomes more fatigue resistant than the middle of the joint, where plane
strain conditions exist. It can also be noted that the fracture surface of the joint tested using constant displacement control is darker
than the one from the test using constant load control, which is due to the thickness of the adhesive left on the surface. This
difference was apparent for all load levels of constant load-controlled tests when compared with constant displacement-controlled

12
D. Santos et al. Engineering Fracture Mechanics 284 (2023) 109255

Fig. 10. Representative fracture surfaces of quasi-static fracture tests at different temperatures.

Fig. 11. Crack path observed in a quasi-static fracture test at room temperature.

Fig. 12. Fracture surfaces resulting from constant displacement- and constant load-control tests at room temperature.

tests. Still, the slight change in tone over the start and middle of the constant load-controlled joint and at the start of the constant
displacement-controlled joint might indicate an effect of strain rate, although stress distribution differences are also possible as a
cause.
The fracture surfaces for the tests at high temperature are shown in Fig. 13. These present similar features to the room
temperature fracture surfaces, although now the tone is much more similar between them, indicating a more similar depth in
the propagation of the fatigue crack. The same comments can be made for these surfaces as were made for surfaces from room
temperature tests, although now the differences are less pronounced.
As pointed out earlier, the crack tip generated during the manufacturing process is in general less sharp than the crack created
by fatigue loading. This means that in load control tests more energy is required for crack initiation. Moreover as shown in Fig. 12
and Fig. 13 the shape of the crack front in the constant displacement control test is curved (typical shape of a defect growing under
plane stress/strain conditions), while the crack tip for the threshold analysis in constant load control tests is flat due to the pre-crack
made artificially during the manufacturing procedure. This difference can be also a possible reason behind the difference in fatigue
threshold energy obtained by the two different test approaches under consideration.

13
D. Santos et al. Engineering Fracture Mechanics 284 (2023) 109255

Fig. 13. Fracture surfaces resulting from constant displacement- and constant load-control tests at 60 ◦ C.

5. Conclusions

The polyurethane adhesive studied in this work is very sensitive to temperature. An increase in temperature by 37 degrees from
23 ◦ C to 60 ◦ C led to a 30% decrease in strength and a 47% decrease in ductility for the adhesive in the study, which translated
into an 82% decrease in fracture toughness in mode I. The effect of increasing temperature in fatigue is an increase in crack growth
rate for a given strain energy release rate, as well as a reduction by around 80% of the fatigue threshold energy release rate.
Different crack conditions at the point of threshold in constant load- and constant displacement-control tests led to differing
values for the threshold energy release rate between testing methodologies, with constant load-control sporting the higher threshold,
measured with a crack at the middle of the adhesive layer, as opposed to close to the interface like for constant displacement control.
Given this, it would be expected for the threshold value to be the same if the crack condition for both methods were the same.
The size of the initial crack size likely affected constant load-control tests at high temperature, aided by changing adhesive
properties, but this was not the case for constant displacement control tests, which proved to be more resilient to changing initial
testing conditions. Additionally, constant displacement control tests proved to be much more reliable and easy to setup in order to
achieve a value for the threshold energy release rate, while also being faster in general. Therefore, a constant displacement control
methodology is given as the preferred way to measure a fatigue threshold.

CRediT authorship contribution statement

D. Santos: Writing – original draft, Investigation, Data curation. A. Akhavan-Safar: Writing – review & editing, Validation,
Supervision, Methodology, Conceptualization. R.J.C. Carbas: Supervision, Resources, Methodology. E.A.S. Marques: Writing –
review & editing, Supervision, Methodology. S. Wenig: Resources, Methodology. L.F.M. da Silva: Writing – review & editing,
Supervision, Project administration, Conceptualization.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared
to influence the work reported in this paper.

Data availability

All the data used is included in the paper.

Acknowledgement

The authors gratefully acknowledge the funding provided by Fundação para a Ciência e Tecnologia (FCT), Portugal, for the
Scientific Employment Grant No. 2020.04017.CEECIND.

14
D. Santos et al. Engineering Fracture Mechanics 284 (2023) 109255

References

[1] Zwicker MFR, Moghadam M, Zhang W, Nielsen CV. Automotive battery pack manufacturing – a review of battery to tab joining. J Adv Join Process
2020;1:100017.
[2] Banea MD, da Silva LFM, Campilho RDSG. The effect of adhesive thickness on the mechanical behavior of a structural polyurethane adhesive. J Adhes
2015;91(5):331–46.
[3] Boutar Yasmina, Naïmi Sami, Mezlini Salah, Carbas Ricardo JC, da Silva Lucas FM, Ben Sik Ali Moez. Cyclic fatigue testing: Assessment of polyurethane
adhesive joints’ durability for bus structures’ aluminium assembly. J Adv Join Process 2021;3:100053.
[4] Adams RD, Brearley T, Nehammer E, Rouse E, Vaughan D. Frictional damping in hollow beam structures joined by bolts, rivets, and adhesive. Proc Inst
Mech Eng L J Mater Des Appl 2021;235(6):1477–87.
[5] Antelo J, Akhavan-Safar A, Carbas RJC, Marques EAS, Goyal R, da Silva LFM. Fatigue life evaluation of adhesive joints in a real structural component.
Int J Fatigue 2021;153:106504.
[6] Passos AC, Arouche MM, Aguiar RAA, Costa HRM, de Barros S, Sampaio EM. Adhesion of epoxy and polyurethane adhesives in pultruded composite
material. J Adv Join Process 2021;3:100045.
[7] Tan Wei, Jingxin Na, Guangbin Wang, Chen Hongli, Meng Huan. Effect of temperature on the fatigue performance and failure mechanism of a flexible
adhesive butt joint. J Adhes 2021.
[8] Sousa F Castro, Akhavan-Safar Alireza, rakesh Goyal, da Silva LFM. Fatigue life estimation of adhesive joints at different mode mixities. J Adhes
2022;98(1):1–23.
[9] Jablonski DA. Fatigue crack growth in structural adhesives. J Adhes 1980;11(2):125–43.
[10] Xu XX, Crocombe AD, Smith PA. Fatigue crack growth rates in adhesive joints tested at different frequencies. J Adhes 1996;58(3–4):191–204.
[11] Abou-Hamda MM, Megahed MM, Hammouda MMI. Fatigue crack growth in double cantilever beam specimen with an adhesive layer. Eng Fract Mech
1998;60(5):605–14.
[12] Azari S, Jhin G, Papini M, Spelt JK. Fatigue threshold and crack growth rate of adhesively bonded joints as a function of load/displacement ratio.
Composites A 2014;57:59–66.
[13] Azari S, Papini M, Spelt JK. Effect of adhesive thickness on fatigue and fracture of toughened epoxy joints – Part I: Experiments. Eng Fract Mech
2011;78(1):153–62.
[14] Ashcroft IA, Shaw SJ. Mode I fracture of epoxy bonded composite joints 2. Fatigue loading. Int J Adhes Adhes 2002;22(2):151–67.
[15] Mall S. Characterization of mode 1 and mixed-mode failure of adhesive bonds between composite adherends. NASA, Langley Research Center; 1985.
[16] Mangalgiri PD, Johnson WS, Everett Jr RA. Effect of adherend thickness and mixed mode loading on debond growth in adhesively bonded composite
joints. J Adhes 1987;23(4):263–88.
[17] Sekiguchi Yu, Sato Chiaki. Effect of bond-line thickness on fatigue crack growth of structural acrylic adhesive joints. Materials 2021;14(7).
[18] Ragni M, Castagnetti D, Spaggiari A, Dragoni E, Milelli M, Girlando S, et al. Thermomechanical characterization of metal-polyurethane bonded joints:
effect of manufacturing parameters and working temperature. J Adhes 2021;1–21.
[19] Houjou K, Shimamoto K, Akiyama H, Sato C. Effect of test temperature on the shear and fatigue strengths of epoxy adhesive joints. J Adhes 2021;1–19.
[20] Usman M, Pascoe JA, Alderliesten RC, Benedictus R. The effect of temperature on fatigue crack growth in FM94 epoxy adhesive bonds investigated by
means of energy dissipation. Eng Fract Mech 2018;189:98–109.
[21] Tan Wei, Na Jingxin, Wang Guangbin, Xu Qianhui, Shen Hao, Mu Wenlong. The effects of service temperature on the fatigue behavior of a polyurethane
adhesive joint. Int J Adhes Adhes 2021;107:102819.
[22] Mu Wen-Long, Xu Qian-Hui, Na Jing-Xin, Wang Heng, Tan Wei, Li De-Feng. Influence of temperature and humidity on the fatigue behaviour of adhesively
bonded CFRP/aluminium alloy joints. J Adhes 2022;98(10):1358–76.
[23] da Costa JA, Akhavan-Safar A, Marques EAS, Carbas RJC, da Silva LFM. Effects of cyclic ageing on the tensile properties and diffusion coefficients of an
epoxy-based adhesive. Proc Inst Mech Eng L J Mater Des Appl 2021;235(6):1451–60.
[24] Wei Tan, Jingxin Na, Wenlong Mu, Guangbin Wang, Yao Feng. Effects of hygrothermal aging on the mechanical properties of aluminum alloy adhesive
joints for high-speed train applications. J Adhes 2022;98(3):227–56.
[25] da Costa JA, Akhavan-Safar A, Marques EAS, Carbas RJC, da Silva LFM. The influence of cyclic ageing on the fatigue performance of bonded joints. Int
J Fatigue 2022;161:106939.
[26] Forth Scott. The purpose of generating fatigue crack growth threshold data. In: ASTM committee meeting. 2006.
[27] Yamada Y, Newman Jr JC. Crack-closure behavior of 2324-T39 aluminum alloy near-threshold conditions for high load ratio and constant Kmax tests.
Int J Fatigue 2009;31(11–12):1780–7.
[28] Yamada Y, Newman Jr JC. Crack closure under high load ratio and Kmax test conditions. Procedia Eng 2010;2(1):71–82.
[29] Jones Rhys. Fatigue crack growth and damage tolerance. Fatigue Fract Eng Mater Struct 2014;37(5):463–83.
[30] da Silva Lucas FM, Adams RD. Measurement of the mechanical properties of structural adhesives in tension and shear over a wide range of temperatures.
J Adhes Sci Technol 2005;19(2):109–41.
[31] ASTM International. Standard test method for fracture strength in cleavage of adhesives in bonded metal joints. Standard ASTM D3433-99R20, West
Conshohocken, Pensylvania: ASTM International; 2020.
[32] Naat Nidhal, Boutar Yasmina, Naïmi Sami, Mezlini Salah, Silva Lucas Filipe Martins Da. Effect of surface texture on the mechanical performance of bonded
joints: a review. J Adhes 2021;1–93.
[33] De Moura MFSF, Campilho RDSG, Gonçalves JPM. Crack equivalent concept applied to the fracture characterization of bonded joints under pure mode I
loading. Compos Sci Technol 2008;68(10–11):2224–30.
[34] Banea MD, da Silva Lucas FM. The effect of temperature on the mechanical properties of adhesives for the automotive industry. Proc Inst Mech Eng L J
Mater Des Appl 2010;224(2):51–62.
[35] Banea MD, da Silva LFM, Campilho RDSG. Effect of temperature on tensile strength and mode I fracture toughness of a high temperature epoxy adhesive.
J Adhes Sci Technol 2012;26(7):939–53.
[36] Jia Zhemin, Yuan Guoqing, Hui David, Feng Xiaoping, Zou Yun. Effect of high loading rate and low temperature on mode I fracture toughness of ductile
polyurethane adhesive. J Adhes Sci Technol 2019;33(1):79–92.
[37] Hasegawa K, Crocombe AD, Coppuck F, Jewel D, Maher S. Characterising bonded joints with a thick and flexible adhesive layer–Part 1: Fracture testing
and behaviour. Int J Adhes Adhes 2015;63:124–31.

15

You might also like