Download as pdf or txt
Download as pdf or txt
You are on page 1of 36

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

net/publication/357023983

Analysis of unbalanced laminate composites with imperfect interphase:


Effective properties via asymptotic homogenization method

Article in Proceedings of the Institution of Mechanical Engineers Part L Journal of Materials Design and Applications · December 2021
DOI: 10.1177/14644207211060004

CITATIONS READS

4 77

3 authors:

Bruno Christoff Humberto Brito Santana


Technische Universität Dresden University of São Paulo
21 PUBLICATIONS 83 CITATIONS 22 PUBLICATIONS 155 CITATIONS

SEE PROFILE SEE PROFILE

Volnei Tita
University of São Paulo
308 PUBLICATIONS 2,734 CITATIONS

SEE PROFILE

All content following this page was uploaded by Volnei Tita on 03 January 2022.

The user has requested enhancement of the downloaded file.


Journal name

Comparative study of adhesive fatigue in aeronautical


bonded-joints: a numerical approach in the frequency-
domain

Journal: Part L: Journal of Materials: Design and Applications

Manuscript ID JMDA-21-0289.R1

Manuscript Type: Special issue: Composite Materials

Date Submitted by the


n/a
Author:

Complete List of Authors: Marques, Denys; University of Sao Paulo, Department of Aeronautical
Engineering
Fo
Ribeiro, Marcelo; University of Sao Paulo, Aeronautical Engineering
Tita , Volnei ; Universidade de Sao Paulo Escola de Engenharia de Sao
Carlos,
rP

Keywords: Fatigue, Random Loads, Bonded Joints, Cohesive Failure, GLARE

The use of adhesively bonded structures has increased over the years,
ee

together with the development of composite materials. This work


investigates a procedure for fatigue life prediction of an aeronautical
bonded joint under random loads, in particular, the cohesive failure of
Abstract: the adhesive layer in a skin-to-stiffener bonded joint. The use of two
rR

different adhesives is investigated, and Dirlik’s method is employed to


predict the stress response in the adhesive layer, from which the fatigue
life is obtained. The effect of damping is also investigated, and it is
shown that increases in damping result in higher fatigue life estimations.
ev
iew

http://mc.manuscriptcentral.com/(site)
Page 1 of 34 Journal name

1
2
3
4 COMPARATIVE STUDY OF ADHESIVE FATIGUE IN AERONAUTICAL BONDED JOINTS:
5
6 A NUMERICAL APPROACH IN THE FREQUENCY-DOMAIN
7
8
Denys Eduardo Teixeira Marquesa, Marcelo Leite Ribeiroa, Volnei Titaa
9
10
a University
of São Paulo, São Carlos School of Engineering, Department of Aeronautical Engineering
11
12 Av. João Dagnone, 1100, Jardim Santa Angelina, São Carlos – SP, 13563-120, Brazil
13
14
15
16 Abstract. The use of adhesively bonded structures has increased over the years, together with the
17 development of composite materials. This work investigates a procedure for fatigue life prediction of an
18 aeronautical bonded joint under random loads, in particular, the cohesive failure of the adhesive layer in
19 a skin-to-stiffener bonded joint. The use of two different adhesives is investigated, and Dirlik’s method is
Fo
20 employed to predict the stress response in the adhesive layer, from which the fatigue life is obtained. The
21
22
effect of damping is also investigated, and it is shown that increases in damping result in higher fatigue
23 life estimations.
rP

24
25 Keywords: Fatigue, Random Loads, Bonded Joints.
26
ee

27
28
29
1. INTRODUCTION
rR

30
31
32 The aircraft industry is always looking for new technologies to improve aircraft performance and
ev

33 safety, while meeting several regulatory requirements. Also, accurate information on the fatigue life of
34
aircraft structures is increasingly required to improve aircraft safety and dispatchability. The last decades
35
36 have also seen an increase in the use of lightweight components made of composite materials for
iew

37 secondary and even primary structures. Included in these recent investigations are the Fiber-Metal-
38 Laminates (FMLs) [1], [2].
39
40
41 A Fiber-Metal-Laminate is a hybrid structural material composed of laminated layers of aluminum
42 and a fiber reinforced polymer. In particular, the second generation of this material, made of aluminum
43 layers and glass fiber composite layers and denominated GLARE (GLAss fiber REinforced aluminum
44
45 laminates), has been successfully applied in structural aircraft components [3]. This type of material offers
46 long fatigue life, high impact resistance and low weight density [4].
47
48 Bonding of composite materials requires special attention. Adhesive joining has attractive intrinsic
49
50 advantages for composite components compared to the classical joining techniques such as riveting
51 fasteners, for two reasons. Rivets require that holes be drilled in the component, which is highly
52 disadvantageous for continuous fiber-reinforced structure. In addition, rivets are very sensitive to fatigue,
53
54 as they introduce stress concentrations that may lead to complete failure of an assembled component.
55 Furthermore, structures in the aerospace and automotive industries are often subjected to random dynamic
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Journal name Page 2 of 34

1
2
3 loads which can lead to excessive vibrations and fatigue damage [5]. Thus, it becomes important to
4
5 characterize the fatigue life of the bonded joints under this type of load.
6
7 Over the past decades, most of the studies in fatigue of adhesively bonded skin/stiffener have been
8 concentrated on composite joints. Hoyt et al. [6] proposed a non-linear 2-D finite element analysis for
9
10 composite bonded joints capable of predicting the failure modes and tracking damage growth due to static
11 or fatigue loading. The authors presented the application of their methodology to a skin reinforced with a
12 T-stiffener and for the case of a single lap joint. Krueger et al. [7] proposed a fatigue life methodology for
13
14 skin/stringer bonded composite. They used experimental fatigue data to obtain a P-N fatigue curve that
15 was used to predict matrix damage onset, and a second fatigue curve based on the mixed-mode strain
16 energy release rate, G-N, that was used to predict delamination in the composite. Combined, the approach
17
18 allowed the prediction of the cumulative fatigue life of bonded composites with good agreement with
19 experimental tests, according to the authors.
Fo
20
21 Composite skin/stiffener debonding under monotonic tension load was investigated by Hosseini-
22
23
Toudeshky et al. [8]. In their experimental tests, which were also modelled by finite element analysis, a
rP

24 tension load was applied to the composite skin. The authors observed that damage was initiated as matrix
25 crack in the composite skin, near the stiffener, which would lead to crack propagation inside the layers of
26
the skin.
ee

27
28
29 Freitas and Sinke [9] studied the fatigue behavior of an inverted T-shape composite stiffener in a
rR

30 Fiber Metal Laminate (FML) skin and observed the composite stiffener as the weakest link, with
31
predominant inter and intra-laminar failure of the composite. It was also observed a large scatter in the
32
fatigue life for crack initiation, which was attributed to manufacturing defects of the stiffener noodle due
ev

33
34 to the complex geometry in the region of curvature between the web and the foot of the stiffener.
35
36
iew

Despite the effort in recent years to fully characterize the fatigue properties of structures such as
37
38 the skin-to-stiffener joint, there is still a lack of knowledge on the behavior of this structure under fully
39 random loads, such as the ones that arise from wind spectra. In this respect, Thawre et al. [10]
40
experimentally investigated a composite T-joint under cyclic loads. They used constant amplitude fatigue
41
42 tests to build load-life (P-N) curves of the joint and used a constant life diagram to estimate the fatigue
43 life of the joint under aeronautical spectrum loads (mini-FALSTAFF), which was found to be in good
44 agreement with experiments.
45
46
47 Regarding the failure of the adhesive layer, Antoniou et al. [11] investigated crack initiation of
48 thick adhesive bond lines under variable amplitude loading. They took into account the residual stresses
49 due to the manufacturing curing process of the adhesive which was superimposed to the expected loads
50
51 caused by wind spectra. The modified Goodman diagram and linear damage accumulation were used to
52 predict the cohesive failure in the adhesive.
53
54 When it comes to the adhesive itself, adhesively bonded structures are subjected to two types of
55
56 failure scenarios: cohesive failure or adhesive failure. Cohesive failure is the one that occurs in the bulk
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Page 3 of 34 Journal name

1
2
3 layer of the adhesive, while adhesive failure happens at the interface between the adhesive and the
4
5 adherent. It should be noted that cohesive failure is usually the preferred failure mode. In this work, the
6 fatigue life of a skin-to-stiffener bonded joint under spectrum loads will be evaluated regarding the
7 cohesive failure of the adhesive.
8
9
10 Due to the random nature of the spectrum loads, the stress history can only be described in a
11 statistical sense, thus, the traditional approach to fatigue analysis in the time-domain cannot be performed.
12 To overcome this problem, frequency-domain methods can be used. These methods usually rely on the
13
14 properties of the random processes to estimate the Probability Density Function (PDF) of the stress
15 response at a given point of the structure. For stationary, Gaussian, and wideband processes, Dirlik’s
16 method is usually the preferred choice [12] and it has been shown to perform extremely well for metallic
17
18 [13], [14] and also (more recently) for composite structures [15]. However, to the best of the authors'
19 knowledge, it has not yet been applied to fatigue estimation of adhesively bonded joints.
Fo
20
21 2. FATIGUE ANALYSIS FOR RANDOM LOADS
22
23
rP

24
Random loads can be characterized in the frequency-domain by their Power Spectrum Density
25 (PSD). Due to their random nature, these loads are better described by a Probability Density Function
26 (PDF), rather than by a time history. The PDF gives the probability of occurrence of a certain range of
ee

27
28
stress amplitudes. For wideband, Gaussian processes, Dirlik [16] proposed a semi-empirical method for
29 the estimation of the stress PDF based on its PSD. The method is based on the spectral moments of the
rR

30 PSD, which, for a one-sided PSD, 𝑃𝑆𝐷𝑋(𝑓), are defined by:


31
32 ∞
ev

33
34
𝑚𝑖 = ∫𝑓 × 𝑃𝑆𝐷 (𝑓)𝑑𝑓
𝑖
𝑋 (1)
35 0
36
iew

37 where 𝑚𝑖 is the spectral moment of i-th order, and 𝑓 denotes frequency (in Hertz). The PDF of the stress
38
39 amplitudes, 𝑃𝐷𝐹(𝑆), can then be calculated as:
40
41 𝑍2 2

[ ]
𝑍 𝑍
1 𝐺1 ― 𝑄 𝐺2𝑍 ― 2𝑇2 ―
42 𝑃𝐷𝐹(𝑆) = 𝑒 + 2𝑒 + 𝐺3𝑍𝑒 2 (2)
43 𝑚0 𝑄 𝑇
44
45
46 where 𝑍 is the normalized amplitude, and 𝑥𝑚 is the mean frequency, which are defined as:
47
48 𝑆
49 𝑍= (3)
50 𝑚0
51
52 and
53
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Journal name Page 4 of 34

1
2
3 1
4 𝑚1 𝑚2
5
6
𝑥𝑚 =
𝑚0 𝑚4( ) 2
(4)

7
8 Another important parameter is the spectral width, which takes the general form of:
9
10
𝑚𝑖
11 𝛼𝑖 = (5)
12 𝑚0𝑚2𝑖
13
14
15
The parameters 𝐺1, 𝐺2, 𝐺3, 𝑇 and 𝑄 are defined as:
16
17 2(𝑥𝑚 ― 𝛼22) 1 ― 𝛼2 ― 𝐺1 + 𝐺21
18 𝐺1 = , 𝐺2 = , 𝐺3 = 1 ― 𝐺1 ― 𝐺2 (6)
19 1 + 𝛼22 1―𝑇
Fo
20 𝛼2 ― 𝑥𝑚 ― 𝐺21 1.25(𝛼2 ― 𝐺3 ― 𝐺2𝑇)
21 𝑇= , 𝑄 = (7)
22 1 ― 𝛼2 ― 𝐺1 + 𝐺21 𝐺1
23
rP

24 Knowing the PDF of the stress, and assuming a linear combination of damage (Palmgren-Miner’s
25
26 rule), the expected damage per second, 𝐷, can be calculated as:
ee

27
28 ∞
𝑓𝑟𝑒𝑞
29 𝐷= ∫𝑆 𝑏
× 𝑃𝐷𝐹(𝑆)𝑑𝑆 (8)
rR

30
𝐶
0
31
32
where 𝐶 and 𝑏 are the material coefficient and the Basquin exponent of the S-N curve of the material, as
ev

33
34 given by:
35 𝑁 = 𝐶 × 𝑆 ―𝑏 (9)
36
iew

37
38 and 𝑓𝑟𝑒𝑞 is the expected peak occurrence per second, which can be calculated as:
39
40 𝑚4
41 𝑓𝑟𝑒𝑞 = (10)
𝑚2
42
43
44 3. AERONAUTICAL BONDED JOINT
45
46 The component used in this work consists of an aluminum stiffener bonded to a Fiber Metal
47
48 Laminate (FML) skin, which is commonly found in aircraft structures such as in fuselage panels. The
49 geometry chosen is the one used in Stiffener Pull-Off Tests (SPOT), which are used to characterize the
50 performance and failure mode of this type of joint. The geometry and material properties used in this work
51
52 were taken from Freitas and Sinke [17], who used SPOT to study the behavior of two different types of
53 adhesives under monotonic loads.
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Page 5 of 34 Journal name

1
2
3 The stiffener is an extruded inverted T-shape stiffener, 1.5 mm thick, made of the aluminum alloy
4
5 2024-T3. The skin is made of Glare 5-3/2-0.3, an FML which consists of three 2024-T3 aluminum alloy
6 layers 0.3 mm thick bonded together with glass fiber reinforced plastic (GFRP) with the layup
7 [0º/90º/90º/0º]. The final layup of the skin is then [Al/[0º/90º/90º/0º]/Al/[0º/90º/90º/0º]/Al], with a total
8
9 thickness of 3.134 mm. It should be noted that this configuration results in a metal-to-metal joint since the
10 outer layers of the FML are made of aluminum. The adhesive layer is assumed to be 0.35 mm thick, and
11 perfect adhesion is assumed between the skin and the stiffener. The skin has a width of 100 mm and a
12
13 span of 300 mm, as shown in Figure 1, which also shows the stiffener’s dimensions.
14
15
16 4. NUMERICAL MODEL
17
18
The whole structure shown in Figure 1 is modeled in AbaqusTM using second order solid elements
19
with reduced integration (C3D20R). The structure is held in place by four steel blocks (shown in Figure
Fo
20
21 1) at which a fixed boundary condition is applied (i.e. zero displacements and rotations), and perfect
22 adhesion is assumed between the steel blocks and the skin. The aluminum stiffener is modeled with 23800
23
rP

24 elements, while the structure containing the FML skin and the steel blocks is modeled with 6996 elements.
25 The mechanical properties of the skin and stiffener are shown in Table 1.
26
ee

27 Table 1: Mechanical properties of the materials used in the simulation [17].


28
29
rR

30 GFRP Al 2024-T3
31 Property Value Property Value
32 𝐸11 48.9 GPa 𝐸 72.4 GPa
ev

33
34 𝐸22, 𝐸33 5.5 GPa 𝑣 0.33
35 𝐺13, 𝐺23, 𝐺12 5.5 GPa 𝜌 2.78 kg/m^3
36
iew

37
𝑣12, 𝑣23, 𝑣13 0.33 - -
38 𝜌 2.48 kg/m^3 - -
39
40 Since the fatigue analysis demands the accurate knowledge of the stress state, a finer mesh is used
41
42 in the adhesive layer, which is modeled with five elements along its thickness and with a total of 500000
43 elements. A mesh convergence study was performed to make sure the stress values had already converged.
44 The adhesive layer was placed between the GLARE skin and the stiffener using the “tie constraint” in
45
46 AbaqusTM. Two different types of adhesive were investigated, and their mechanical properties are given
47 in Table 2 (further details about the adhesives and their fatigue properties are given in Section 5).
48
49 Table 2: Mechanical properties of the adhesives used in the simulation [19], [20], [21].
50
51
52 Adhesive I Adhesive II
53 Property Value Property Value
54 𝐸 1.6 GPa 𝐸 2.80 GPa
55
56 𝑣 0.4 𝑣 0.42
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Journal name Page 6 of 34

1
2
3 𝜌 1.20 kg/m^3 𝜌 1.17 kg/m^3
4
5
6 The finite element model is then used to obtain the Frequency Response Function (FRF) of the
7 stress response in the most critical point of the adhesive layer. The FRF is obtained from a unity impact
8 load that is applied at the top corner of the stiffener, in a frequency range from 100 to 160 Hz. The first
9
10 (bending) mode of the structure was previously identified to be at around 134 Hz while the second mode
11 lies at around 230 Hz, thus the spectrum load applied is capable of exciting only the first vibrational mode
12 of the joint. The stress response is registered for a point located at the bottom corner of the stiffener,
13
14 exactly in the middle of the adhesive layer. Both the impact force and the critical point in the adhesive
15 layer are shown in Figure 2.
16
17 The FRF obtained from the numerical analysis is then used to obtain the PSD of the stress response,
18
19 𝑃𝑆𝐷𝑆, to a given force input, 𝑃𝑆𝐷𝑖, according to Equation (11):
Fo
20
21 𝑃𝑆𝐷𝑆 = |𝐹𝑅𝐹|2𝑃𝑆𝐷𝑖 (11)
22
23
rP

24
25 5. FATIGUE LIFE ESTIMATION
26
ee

27 In the present work, the use of two different adhesives is investigated regarding their fatigue
28 properties. The analysis is restricted to damage nucleation inside the adhesive layer, and the fatigue
29
rR

30 behavior of the joint is analyzed based on the fatigue properties of the bulk adhesives found in the
31 literature. As pointed out by Zuo and Vassilopoulos [18], it is still a matter of debate whether or not bulk
32 adhesive properties are representative of the properties in thin film adhesive bonds. It is difficult to make
ev

33
34 a fair comparison because bulk adhesives are often tested under simpler stress states, while in a real joint
35 a complex, non-uniform, tri-axial stress state is expected. Nevertheless, the authors pointed out that similar
36
iew

material properties have been found between bulk adhesives and thin adhesive bonds.
37
38
39 The first adhesive analyzed (Dow® Betamate 1496 v, from hereafter identified as Adhesive I)
40 corresponds to a one-component epoxy-based adhesive whose fatigue behavior was studied by Beber et
41 al. [19], who reported its fatigue properties under different temperatures.
42
43
44 The second adhesive (Araldite-F and Piperidine from Sigma-Aldrich, from hereafter identified as
45 Adhesive II) comes from a study by Wang et al. [20], who investigated the fatigue behavior of polymer
46 nanocomposites. They observed that the addition of hard particles (silica) could improve the fatigue life
47
48
when compared to that of the pure epoxy used in their study, while the addition of soft nano rubber
49 particles reduced the fatigue performance. For the sake of comparison, this study is based on the fatigue
50 properties from a mix of epoxy with 6% in weight of nano rubber particles and 6% in weight of nano silica
51
52
particles. The experimental S-N curves for both adhesives are shown in Figure 3.
53
54 It should be noted that the experimental S-N curve of Adhesive I, shown in Figure 2(a), was
55 obtained from tests with a positive stress ratio 𝑅 = 0.1, while Adhesive II was tested under fully reversed
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Page 7 of 34 Journal name

1
2
3 cycles, i.e., 𝑅 = ―1. Since the stress cycles in the spectrum load are assumed to be always fully reversed,
4
5 the Smith-Watson-Topper equation was used to convert the S-N curve of Adhesive I to an equivalent S-
6 N curve at 𝑅 = ―1.
7
8 Damping has the potential to have a huge impact on the FRF of the structure and, consequently,
9
10 may play an important role in fatigue estimation in frequency-domain. Unfortunately, it is extremely
11 difficult to model damping in a structure based solely on the material properties of its components, and
12 thus damping values are usually obtained from experimental investigations. To better evaluate the impact
13
14 of damping in the fatigue life prediction, two different and arbitrary values of damping are used in this
15 study. Thus, a choice was made for a modal damping of 1.3% and 1.5%, which is applied in the first
16 vibrational mode of the numerical model.
17
18
19 A schematic illustration of the procedure to obtain the fatigue life using a frequency-domain
Fo
20 approach is shown in Figure 4. As shown, the fatigue life depends on a structure (geometry and material)
21 and a load case. The Finite Element Method (FEM) is used to obtain the FRF of the stress response in the
22
23 most critical point of the structure. The PSD of the stress response (‘Output PSD’ in Figure 4) is obtained
rP

24 by combining the load PSD with the structure’s FRF through Equation (11). Dirlik’s method is then used
25 to obtain the PDF of the stress amplitudes, and the Palmgren-Miner’s rule is used to obtain the expected
26
ee

27
damage per second from Equation (8). Finally, the inverse of the expected damage per second gives the
28 fatigue life of the structure.
29
rR

30 For this work, a load case with a constant amplitude spectrum of 𝑃𝑆𝐷𝑖 = 15.5 𝑁2/𝐻𝑧 in the
31
32 frequency range from 100 to 160 Hz will be considered, applied in the top corner of the stiffener, as shown
by Figure 2.
ev

33
34
35 It should be noted that due to their random nature, the spectrum loads can only be described in a
36
iew

37 statistical sense, and the history information cannot be retrieved. Thus, in the procedure outlined in this
38 paper, effects such as load interactions are not accounted for in the fatigue analysis.
39
40 6. RESULTS
41
42
43 The numerical analysis revealed that the first natural frequency of the structure corresponds to a
44 bending mode that happens at 133.9 Hz. The PSD of the stress response was obtained from the numerical
45 FRF and a constant force 𝑃𝑆𝐷𝑖 = 15.5 𝑁2/𝐻𝑧, calculated according to Equation (11). The stress PSD of
46
47 both adhesives for a modal damping of 1.5% is shown in Figure 5.
48
49 It is possible to observe from Figure 5 that Adhesive II, which is stiffer, has a higher stress response
50 in all the frequency range analyzed. In fact, the root mean square of the stress process, 𝑆𝑟𝑚𝑠, is equal to
51
52 4.6 𝑀𝑃𝑎 for Adhesive I and has a value of 5.9 𝑀𝑃𝑎 for adhesive II, both at a modal damping of 1.5%.
53
54 Applying Equation (2), the PDF of the stress amplitudes was obtained for all the study cases, as
55
shown in Figure 6.
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Journal name Page 8 of 34

1
2
3 Comparing the PDF for Adhesive I and II, it is possible to observe that the PDF curve is shifted to
4
5 the left, showing a higher probability of occurrence of stress levels of lower amplitude for Adhesive I, and
6 also decreasing the maximum expected stress amplitude. When looking at the effect of damping alone,
7 comparing Figures 6(a) and 6(b) reveals only a small difference in the amplitudes of the stress. However,
8
9 closer inspection, as shown in Figure 7, demonstrates that increasing the damping shift the PDF curve to
10 the left, which then shows a high probability of occurrence of lower amplitude stress cycles and lower
11 probabilities at higher amplitudes. To quantify the impact of damping on the fatigue life of the structure,
12
13 Equation (8) was used. The results for fatigue life are shown in Table 3.
14
15 As it can be seen, damping has a huge impact on the fatigue life of the structure. As damping
16 increases from 1.3% to 1.5%, the fatigue life increases by almost three times for adhesive I. Similarly, in
17
18 Adhesive II the same change in damping increases the fatigue life by a factor of almost 1.5.
19
Table 3: Fatigue life for different damping levels.
Fo
20
21 Adhesive I Adhesive II
22
23 1.3% damping 1.7e+05 cycles 1.1e+06 cycles
rP

24 1.5% damping 4.5e+05 cycles 1.7e+06 cycles


25
26 Comparing the performance of both materials, Adhesive II has shown better fatigue life regardless
ee

27
28 of the damping level. At 1.3% damping, Adhesive II improves the crack initiation life by almost one order
29 of magnitude. In the prevalence of lower amplitudes of stresses (i.e. at 1.5% damping), the fatigue life of
rR

30 Adhesive II is about 3.8 times higher.


31
32
7. CONCLUSIONS
ev

33
34
35 Crack nucleation under spectrum loads is investigated using the S-N approach and the probability
36
iew

density function of the stresses is calculated via Dirlik’s method. The method is applied to predict the
37
38 fatigue life of a skin-to-stiffener bonded joint due to cohesive failure of the adhesive.
39
40 The study compares the performance of two epoxy-based adhesives whose fatigue properties had
41 been previously characterized in the literature and it provides a simple methodology for engineers to easily
42
43
access the fatigue performance of a bonded joint under random loads using a simple finite element solution
44 of the frequency response function of the structure. Since the method is implemented in the frequency-
45 domain, there is no need of time-domain stress history, helping saving time and computational costs in
46
47
the fatigue analysis.
48
49 Numerical results showed the strong influence of damping in the structure’s response and its effect
50 on the fatigue life, in which an increase in the modal damping of the structure also leads to an increase in
51
its life. This happens due to the dissipation effect caused by damping, which decreases the stress level in
52
53 the adhesive layer. This phenomenon highlights the importance of correct assessment of damping values
54 under the risk of non-conservative life estimates in the case of overestimation of damping in the structure.
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Page 9 of 34 Journal name

1
2
3 Future works will investigate the experimental validation of the numerical approach outlined in
4
5 this paper and help assess the potentialities and limitations of the methodology. It should be noted that
6 while the approach shown in this paper concerns the case of fatigue life for crack nucleation, a similar
7 approach can be used for crack propagation in the context of Linear Elastic Fracture Mechanics, as
8
9 demonstrated by the authors for the case of metallic beams [23].
10
11 ACKNOWLEDGEMENTS
12
13 This study was financed in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível
14
15 Superior - Brasil (CAPES) - Finance Code 001.
16
17 REFERENCES
18
19 [1] Sinmazçelik T, Avcu E, Bora MO, Çoban O. A review: Fibre metal laminates, background, bonding
Fo
20
21 types and applied test methods. Materials and Design 2011; 32: 3671-3685.
22
23 [2] Bikakis GSE, Kalfountzos CD, Theotokoglou EE. Elastic buckling response of rectangular GLARE
rP

24 fiber-metal laminates subjected to shearing stresses. Aerosp Sci Technol 2019; 87: 110-118.
25
26
ee

27
[3] Zhu S, Chai GB. Low-velocity impact response of fiber-metal laminates – A theoretical approach.
28 Proc IMechE, Part L: J Materials: Design and Applications 2014; 228: 301-311.
29
rR

30 [4] Tsamasphyros GJ, Bikakis GS. Analytical modeling to predict the low velocity impact response of
31
32
circular GLARE fiber-metal laminates. Aerosp Sci Technol 2013; 29: 28-36.
ev

33
34 [5] Bachoo R, Bridge J. Random vibration analysis and modal energy characteristics of fiber-reinforced
35 composite beams. Proc IMechE, Part L: J Materials: Design and Applications 2020.
36
iew

37
[6] Hoyt DM, Ward SH, Minguet PJ. Strength and fatigue life modeling of bonded joints in composite
38
39 structure. J Compos Tech Res 2002; 24: 190-210.
40
41 [7] Krueger R, Paris IL, O’Brien K, Minguet PJ. Fatigue life methodology for bonded composite
42
skin/stringer configurations. J Compos Tech Res 2002; 24: 56-79.
43
44
45 [8] Hosseini-Toudeshky H, Mohammadi B, HHamidi B, Ovesy HR. Analysis of composite skin/stiffener
46 debounding and failure under uniaxial loading. Compos Struct 2006; 75: 428-436.
47
48
[9] Freitas ST, Sinke J. Failure analysis of adhesively-bonded metal-skin-to-composite-stiffener: effect
49
50 of temperature and cyclic loading. Compos Struct 2017; 166: 27-37.
51
52 [10] Thawre MM, Pandey KN, Dubey A, Verma KK, Peshwe DR, Paretkar RK, Jagannathan N,
53 Manjunatha CM. Fatigue life of a carbon fiber composite T-joint under a standard fighter aircraft
54
55 spectrum load sequence. Compos Struct 2015; 127: 260-266.
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Journal name Page 10 of 34

1
2
3 [11] Antoniou AE, Verpermann MM, Sayer F, Krimmer A. Life prediction analysis of thick adhesive bond
4
5 lines under variable amplitude fatigue loading. ECCM18 – 18th European Conference on
6 Composite Materials. Athens 2018.
7
8 [12] Mrsnik M, Slavic J, Boltezar M. Frequency-domain methods for a vibration-fatigue-life estimation –
9
10 application to real data. Int J Fatigue 2013; 47: 8-17.
11
12 [13] Kong YS, Abdullha S, Schramm D, Omar MZ, Haris SM. Vibration fatigue analysis of carbon steel
13 coil spring under various road excitations. Metals 2018; 8: 617.
14
15
16 [14] Yeter B, Garbatov Y, Soares CG. Evaluation of fatigue damage model predictions for fixed offshore
17 wind turbine support structures. Int J Fatigue 2016; 87: 71-80.
18
19 [15] Sun Y, Zhang Y, Yang C, Liu Y, Chen X, Yao L, Gao W. Prediction on fatigue properties of the plain
Fo
20
21 weave composite under broadband random loading. Fatigue Fract Eng Mater Struct 2021; 44:
22 1515-1532.
23
rP

24 [16] Dirlik T. Application of computers in fatigue analysis. PhD Thesis. University of Warwick. England,
25
26
1985.
ee

27
28 [17] Freitas ST, Sinke J. Failure analysis of adhesively-bonded skin-to-stiffener joints: metal-metal vs.
29 composite-metal. Eng Fail Anal 2015; 56: 2-13.
rR

30
31
32
[18] Zuo P, Vassilopoulos AP. Review of fatigue of bulk structural adhesives and thick adhesive joints.
Int Mater Rev 2020; 1-26.
ev

33
34
35 [19] V. C. Beber, B. Schneider, M. Brede. Influence of temperature on the fatigue behaviour of a
36
iew

toughened epoxy adhesive. The Journal of Adhesives 2016; 92: 778-794.


37
38
39 [20] Wang GT, Liu HY, Saintier N, Mai YW. Cyclic fatigue of polymer nanocomposites. Eng Fail Anal
40 2009; 16: 2635-2645.
41
42
[21] Zhang J, Deng S, Wang Y, Ye L. Role of rigid nanoparticles and CTBN rubber in the toughening of
43
44 epoxies with different cross-linking densities. Compos Part A 2016; 80: 82-94.
45
46 [22] Berhanu G. Vibration durability testing and design validation based on narrow frequency band.
47
Master Thesis, Blekinge Institute of Technology, Sweden, 2011.
48
49
50 [23] Marques D ET, Vandepitte D, Tita V. Damage detection and fatigue life estimation under random
51 loads: A new structural health monitoring methodology in the frequency domain. Fatigue Fract
52 Eng Mater Struct 2021; 46: 1622-1636.
53
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Page 11 of 34 Journal name

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17 Figure 1: Model used in the simulation (left) and detail of the stiffener (right).
18
19 3626x1119mm (79 x 79 DPI)
Fo
20
21
22
23
rP

24
25
26
ee

27
28
29
rR

30
31
32
ev

33
34
35
36
iew

37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Journal name Page 12 of 34

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
Figure 2: (a) Impact force (green arrow) and point from which the stress response is recorded (red dot); (b)
Fo
20
Mesh details of the adhesive layer.
21
22 2733x1077mm (59 x 59 DPI)
23
rP

24
25
26
ee

27
28
29
rR

30
31
32
ev

33
34
35
36
iew

37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Page 13 of 34 Journal name

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
Fo
20
21 Figure 2: (a) Impact force (green arrow) and point from which the stress response is recorded (red dot); (b)
Mesh details of the adhesive layer.
22
23
rP

2484x1063mm (59 x 59 DPI)


24
25
26
ee

27
28
29
rR

30
31
32
ev

33
34
35
36
iew

37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Journal name Page 14 of 34

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
Figure 3: Relationship between stress amplitude and fatigue life (S-N curve) for: (a) Adhesive I [19]; and
Fo
20
(b) Adhesive II [20].
21
22 771x299mm (118 x 118 DPI)
23
rP

24
25
26
ee

27
28
29
rR

30
31
32
ev

33
34
35
36
iew

37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Page 15 of 34 Journal name

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
Fo
20
21
22
23
rP

24
25
26
ee

27
28
29
Figure 4: Schematic overview of the fatigue analysis in the frequency-domain (adapted from [22]).
rR

30
31 509x358mm (59 x 59 DPI)
32
ev

33
34
35
36
iew

37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Journal name Page 16 of 34

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
Fo
20
21
22
23
rP

24
25
26
ee

27
28
29
rR

30
31 Figure 5: PSD of the stress response for Adhesives I and II at 1.5% modal damping.
32
374x280mm (38 x 38 DPI)
ev

33
34
35
36
iew

37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Page 17 of 34 Journal name

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
Fo
20 Figure 6: PDF of the stress response for (a) 1.3% modal damping and (b) 1.5% modal damping.
21
753x300mm (118 x 118 DPI)
22
23
rP

24
25
26
ee

27
28
29
rR

30
31
32
ev

33
34
35
36
iew

37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Journal name Page 18 of 34

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
Fo
20
21
22
23
rP

24
25
26
ee

27
28
29
rR

30
31 Figure 7: Effect of damping on the PDF curve for Adhesive II.
32
374x280mm (38 x 38 DPI)
ev

33
34
35
36
iew

37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Page 19 of 34 Journal name

1
2
3
4 COMPARATIVE STUDY OF ADHESIVE FATIGUE IN AERONAUTICAL BONDED JOINTS:
5
6 A NUMERICAL APPROACH IN THE FREQUENCY-DOMAIN
7
8
Denys Eduardo Teixeira Marquesa, Marcelo Leite Ribeiroa, Volnei Titaa
9
10
a University
of São Paulo, São Carlos School of Engineering, Department of Aeronautical Engineering
11
12 Av. João Dagnone, 1100, Jardim Santa Angelina, São Carlos – SP, 13563-120, Brazil
13
14
15
16 Abstract. The use of adhesively bonded structures has increased over the years, together with the
17 development of composite materials. This work investigates a procedure for fatigue life prediction of an
18 aeronautical bonded joint this type of structures under random loads, in particular, the cohesive failure
19 of the adhesive layer in a skin-to-stiffener bonded joint. The use of two different adhesives is investigated,
Fo
20 and Dirlik’s method is employed to predict the stress response in the adhesive layer, from which the fatigue
21
22
life is obtained. The effect of damping is also investigated, and it is shown that increases in damping result
23 in higher fatigue life estimations.
rP

24
25 Keywords: Fatigue, Random Loads, Bonded Joints.
26
ee

27
28
29
1. INTRODUCTION
rR

30
31
32 The aircraft industry is always looking for new technologies to improve aircraft performance and
ev

33 safety, while meeting several regulatory requirements. Also, accurate information on the fatigue life of
34
aircraft structures is increasingly required to improve aircraft safety and dispatchability. The last decades
35
36 have also seen an increase in the use of lightweight components made of composite materials for
iew

37 secondary and even primary structures. Included in these recent investigations are the Fiber-Metal-
38 Laminates (FMLs) [1], [2].
39
40
41 A Fiber-Metal-Laminate is a hybrid structural material composed of laminated layers of aluminum
42 and a fiber reinforced polymer. In particular, the second generation of this material, made of aluminum
43 layers and glass fiber composite layers and denominated GLARE (GLAss fiber REinforced aluminum
44
45 laminates), has been successfully applied in structural aircraft components [1][3]. This type of material
46 offers long fatigue life, high impact resistance and low weight density [2][4].
47
48 Bonding of composite materials requires special attention. Adhesive joining has attractive intrinsic
49
50 advantages for composite components compared to the classical joining techniques such as riveting
51 fasteners, for two reasons. Rivets require that holes be drilled in the component, which is highly
52 disadvantageous for continuous fiber-reinforced structure. In addition, rivets are very sensitive to fatigue,
53
54 as they introduce stress concentrations that may lead to complete failure of an assembled component.
55 Furthermore, structures in the aerospace and automotive industries are often subjected to random dynamic
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Journal name Page 20 of 34

1
2
3 loads which can lead to excessive vibrations and fatigue damage [3][5]. Thus, it becomes important to
4
5 characterize the fatigue life of the bonded joints under this type of load.
6
7 Over the past decades, most of the studies in fatigue of adhesively bonded skin/stiffener have been
8 concentrated on composite joints. Hoyt et al. [4][6] proposed a non-linear 2-D finite element analysis for
9
10 composite bonded joints capable of predicting the failure modes and tracking damage growth due to static
11 or fatigue loading. The authors presented the application of their methodology to a skin reinforced with a
12 T-stiffener and for the case of a single lap joint. Krueger et al. [5][7] proposed a fatigue life methodology
13
14 for skin/stringer bonded composite. They used experimental fatigue data to obtain a P-N fatigue curve that
15 was used to predict matrix damage onset, and a second fatigue curve based on the mixed-mode strain
16 energy release rate, G-N, that was used to predict delamination in the composite. Combined, the approach
17
18 allowed the prediction of the cumulative fatigue life of bonded composites with good agreement with
19 experimental tests, according to the authors.
Fo
20
21 Composite skin/stiffener debonding under monotonic tension load was investigated by Hosseini-
22
23
Toudeshky et al. [6][8]. In their experimental tests, which were also modelled by finite element analysis,
rP

24 a tension load was applied to the composite skin. The authors observed that damage was initiated as matrix
25 crack in the composite skin, near the stiffener, which would lead to crack propagation inside the layers of
26
the skin.
ee

27
28
29 Freitas and Sinke [7][9] studied the fatigue behavior of an inverted T-shape composite stiffener in
rR

30 a Fiber Metal Laminate (FML) skin and observed the composite stiffener as the weakest link, with
31
predominant inter and intra-laminar failure of the composite. It was also observed a large scatter in the
32
fatigue life for crack initiation, which was attributed to manufacturing defects of the stiffener noodle due
ev

33
34 to the complex geometry in the region of curvature between the web and the foot of the stiffener.due to
35
its complex geometry.
36
iew

37
38 Despite the effort in recent years to fully characterize the fatigue properties of structures such as
39 the skin-to-stiffener joint, there is still a lack of knowledge on the behavior of this structure under fully
40
random loads, such as the ones that arise from wind spectra. In this respect, Thawre et al. [8][10]
41
42 experimentally investigated a composite T-joint under cyclic loads. They used constant amplitude fatigue
43 tests to build load-life (P-N) curves of the joint and used a constant life diagram to estimate the fatigue
44 life of the joint under aeronautical spectrum loads (mini-FALSTAFF), which was found to be in good
45
46 agreement with experiments.
47
48 Regarding the failure of the adhesive layer, Antoniou et al. [9][11] investigated crack initiation of
49 thick adhesive bond lines under variable amplitude loading. They took into account the residual stresses
50
51 due to the manufacturing curing process of the adhesive which was superimposed to the expected loads
52 caused by wind spectra. The modified Goodman diagram and linear damage accumulation were used to
53 predict the cohesive failure in the adhesive.
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Page 21 of 34 Journal name

1
2
3 When it comes to the adhesive itself, adhesively bonded structures are subjected to two types of
4
5 failure scenarios: cohesive failure or adhesive failure. Cohesive failure is the one that occurs in the bulk
6 layer of the adhesive, while adhesive failure happens at the interface between the adhesive and the
7 adherent. It should be noted that cohesive failure is usually the preferred failure mode. In this work, the
8
9 fatigue life of a skin-to-stiffener bonded joint under spectrum loads will be evaluated regarding the
10 cohesive failure of the adhesive.
11
12 Due to the random nature of the spectrum loads, the stress history can only be described in a
13
14 statistical sense, thus, the traditional approach to fatigue analysis in the time-domain cannot be performed.
15 To overcome this problem, frequency-domain methods can be used. These methods usually rely on the
16 properties of the random processes to estimate the Probability Density Function (PDF) of the stress
17
18 response at a given point of the structure. For stationary, Gaussian, and wideband processes, Dirlik’s
19 method is usually the preferred choice [10][12] and it has been shown to perform extremely well for
Fo
20 metallic [11], [12],[13], [14] and also (more recently) for composite structures [13][15]. However, to the
21
22
best of the authors' knowledge, it has not yet been applied to fatigue estimation of adhesively bonded
23 joints.
rP

24
25 2. FATIGUE ANALYSIS FOR RANDOM LOADS
26
ee

27
28
Random loads can be characterized in the frequency-domain by their Power Spectrum Density
29 (PSD). Due to their random nature, these loads are better described by a Probability Density Function
rR

30 (PDF), rather than by a time history. The PDF gives the probability of occurrence of a certain range of
31
stress amplitudes. For wideband, Gaussian processes, Dirlik [14][16] proposed a semi-empirical method
32
for the estimation of the stress PDF based on its PSD. The method is based on the spectral moments of
ev

33
34 the PSD, which, for a one-sided PSD, 𝑃𝑆𝐷𝑋(𝑓), are defined by:
35
36
iew


37
38
𝑚𝑖 = ∫𝑓 × 𝑃𝑆𝐷 (𝑓)𝑑𝑓
𝑖
𝑋 (1)
39 0
40
41 where 𝑚𝑖 is the spectral moment of i-th order, and 𝑓 denotes frequency (in Hertz). The PDF of the stress
42
43 amplitudes, 𝑃𝐷𝐹(𝑆), can then be calculated as:
44
45 𝑍2 2

[ ]
𝑍 𝑍
1 𝐺1 ― 𝑄 𝐺2𝑍 ― 2𝑇2 ―
46 𝑃𝐷𝐹(𝑆) = 𝑒 + 2𝑒 + 𝐺3𝑍𝑒 2 (2)
47 𝑚0 𝑄 𝑇
48
49
50 where 𝑍 is the normalized amplitude, and 𝑥𝑚 is the mean frequency, which are defined as:
51
52 𝑆
53 𝑍= (3)
54 𝑚0
55
56 and
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Journal name Page 22 of 34

1
2
3 1
4 𝑚1 𝑚2
5
6
𝑥𝑚 =
𝑚0 𝑚4( ) 2
(4)

7
8 Another important parameter is the spectral width, which takes the general form of:
9
10
𝑚𝑖
11 𝛼𝑖 = (5)
12 𝑚0𝑚2𝑖
13
14
15
The parameters 𝐺1, 𝐺2, 𝐺3, 𝑇 and 𝑄 are defined as:
16
17 2(𝑥𝑚 ― 𝛼22) 1 ― 𝛼2 ― 𝐺1 + 𝐺21
18 𝐺1 = , 𝐺2 = , 𝐺3 = 1 ― 𝐺1 ― 𝐺2 (6)
19 1 + 𝛼22 1―𝑇
Fo
20 𝛼2 ― 𝑥𝑚 ― 𝐺21 1.25(𝛼2 ― 𝐺3 ― 𝐺2𝑇)
21 𝑇= , 𝑄 = (7)
22 1 ― 𝛼2 ― 𝐺1 + 𝐺21 𝐺1
23
rP

24 Knowing the PDF of the stress, and assuming a linear combination of damage (Palmgren-Miner’s
25
26 rule), the expected damage per second, 𝐷, can be calculated as:
ee

27
28 ∞
𝑓𝑟𝑒𝑞
29 𝐷= ∫𝑆 𝑏
× 𝑃𝐷𝐹(𝑆)𝑑𝑆 (8)
rR

30
𝐶
0
31
32
where 𝐶 and 𝑏 are the material coefficient and the Basquin exponent of the S-N curve of the material, as
ev

33
34 given by:
35 𝑁 = 𝐶 × 𝑆 ―𝑏 (9)
36
iew

37
38 and 𝑓𝑟𝑒𝑞 is the expected peak occurrence per second, which can be calculated as:
39
40 𝑚4
41 𝑓𝑟𝑒𝑞 = (10)
𝑚2
42
43
44 3. AERONAUTICAL BONDED JOINT GEOMETRY AND NUMERICAL MODEL
45
46 The component used in this work consists of an aluminum stiffener bonded to a Fiber Metal
47
48 Laminate (FML) skin, which is commonly found in aircraft structures such as in fuselage panels. The
49 geometry chosen is the one used in Stiffener Pull-Off Tests (SPOT), which are used to characterize the
50 performance and failure mode of this type of joint. The geometry and material properties used in this work
51
52 were taken from Freitas and Sinke [15][17], who used SPOT to study the behavior of two different types
53 of adhesives under monotonic loads.
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Page 23 of 34 Journal name

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16 Figure 1: Model used in the simulation (left) and detail of the stiffener (right).
17
18
19
The stiffener is an extruded inverted T-shape stiffener, 1.5 mm thick, made of the aluminum alloy
2024-T3. The skin is made of Glare 5-3/2-0.3, an FML which consists of three 2024-T3 aluminum alloy
Fo
20
21 layers 0.3 mm thick bonded together with glass fiber reinforced plastic (GFRP) with the layup
22
23
[0º/90º/90º/0º]. The final layup of the skin is then [Al/[0º/90º/90º/0º]/Al/[0º/90º/90º/0º]/Al], with a total
rP

24 thickness of 3.134 mm. It should be noted that this configuration results in a metal-to-metal joint since the
25 outer layers of the FML are made of aluminum. The adhesive layer is assumed to be 0.35 mm thick, and
26
perfect adhesion is assumed between the skin and the stiffener. The skin has a width of 100 mm and a
ee

27
28 span of 300 mm, as shown in Figure 1, which also shows the stiffener’s dimensions.
29
rR

30 The structure is held in place by four steel blocks at which a fixed boundary condition is applied
31
(i.e. zero displacements and rotations), and perfect adhesion is assumed between the steel blocks and the
32
skin. The skin has a width of 100 mm and a span of 300 mm, as shown in Figure 1, which also shows the
ev

33
34 stiffener’s dimensions.
35
36
iew

37
38 4. NUMERICAL MODEL
39
40 The whole structure is modeled in AbaqusTM with 83076 second-order solid elements (C3D20)
41 and 92 second-order wedge elements (C3D15). Table 1 shows the mechanical properties of the FML skin
42
43
and aluminum stiffener.
44
45 The whole structure shown in Figure 1 is modeled in AbaqusTM using second order solid elements
46 with reduced integration (C3D20R). The structure is held in place by four steel blocks (shown in Figure
47
1) at which a fixed boundary condition is applied (i.e. zero displacements and rotations), and perfect
48
49 adhesion is assumed between the steel blocks and the skin. The aluminum stiffener is modeled with 23800
50 elements, while the structure containing the FML skin and the steel blocks is modeled with 6996 elements.
51
The mechanical properties of the skin and stiffener are shown in Table 1.
52
53
54 Table 1: Mechanical properties of the materials used in the simulation [15][17].
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Journal name Page 24 of 34

1
2
3 GFRP Al 2024-T3
4
5 Property Value Property Value
6 𝐸11 48.9 GPa 𝐸 72.4 GPa
7
8
𝐸22, 𝐸33 5.5 GPa 𝑣 0.33
9 𝐺13, 𝐺23, 𝐺12 5.5 GPa 𝜌 2.78 kg/m^3
10 𝑣12, 𝑣23, 𝑣13 0.33 - -
11
12 𝜌 2.48 kg/m^3 - -
13
14 Since the fatigue analysis demands the accurate knowledge of the stress state, a finer mesh is used
15 in the adhesive layer, which is modeled with five elements along its thickness and with a total of 500000
16
17 elements. A mesh convergence study was performed to make sure the stress values had already converged.
18 The adhesive layer was placed between the GLARE skin and the stiffener using the “tie constraint” in
19 AbaqusTM. Two different types of adhesive were investigated, and their mechanical properties are given
Fo
20
21 in Table 2 (further details about the adhesives and their fatigue properties are given in Section 5).
22
23 Table 2: Mechanical properties of the adhesives used in the simulation [17], [18], [19][19], [20], [21].
rP

24
25 Adhesive I Adhesive II
26
ee

27 Property Value Property Value


28 𝐸 1.6 GPa 𝐸 2.80 GPa
29 𝑣 0.4 𝑣 0.42
rR

30
31 𝜌 1.20 kg/m^3 𝜌 1.17 kg/m^3
32
ev

33 The finite element model is then used to obtain the Frequency Response Function (FRF) of the
34
stress response in the most critical point of the adhesive layer. The FRF is obtained from a unity impact
35
36 load that is applied at the top corner of the stiffener, in a frequency range from 100 to 160 Hz. The first
iew

37 (bending) mode of the structure was previously identified to be at around 134 Hz while the second mode
38 lies at around 230 Hz, thus the spectrum load applied is capable of exciting only the first vibrational mode
39
40 of the joint. The stress response is registered for a point located at the bottom corner of the stiffener,
41 exactly in the middle of the adhesive layer. Both the impact force and the critical point in the adhesive
42 layer are shown in Figure 2.
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Page 25 of 34 Journal name

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
Fo
20
21
22
23
rP

24
25
26
ee

27
28
29
rR

30
31
32
ev

33
34
35
36
iew

37
38
39
40
41
42 Figure 2: (a) Impact force (green arrow) and point from which the stress response is recorded (red dot);
43 (b) Mesh details of the adhesive layer.
44
45 The FRF obtained from the numerical analysis is then used to obtain the PSD of the stress response,
46
47 𝑃𝑆𝐷𝑆, to a given force input, 𝑃𝑆𝐷𝑖, according to Equation (11):
48
49 𝑃𝑆𝐷𝑆 = |𝐹𝑅𝐹|2𝑃𝑆𝐷𝑖 (11)
50
51
52
53
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Journal name Page 26 of 34

1
2
3 5. FATIGUE LIFE ESTIMATION
4
5
6
In the present work, the use of two different adhesives is investigated regarding their fatigue
7 properties. The analysis is restricted to damage nucleation inside the adhesive layer, and the fatigue
8 behavior of the joint is analyzed based on the fatigue properties of the bulk adhesives found in the
9
literature. As pointed out by Zuo and Vassilopoulos [16][18], it is still a matter of debate whether or not
10
11 bulk adhesive properties are representative of the properties in thin film adhesive bonds. It is difficult to
12 make a fair comparison because bulk adhesives are often tested under simpler stress states, while in a real
13
joint a complex, non-uniform, tri-axial stress state is expected. Nevertheless, the authors pointed out that
14
15 similar material properties have been found between bulk adhesives and thin adhesive bonds.
16
17 The first adhesive analyzed (Dow® Betamate 1496 v, from hereafter identified as Adhesive I)
18
corresponds to a one-component epoxy-based adhesive whose fatigue behavior was studied by Beber et
19
al. [17][19], who reported its fatigue properties under different temperatures.
Fo
20
21
22 The second adhesive (Araldite-F and Piperidine from Sigma-Aldrich, from hereafter identified as
23
rP

Adhesive II) comes from a study by Wang et al. [18][20], who investigated the fatigue behavior of polymer
24
25 nanocomposites. They observed that the addition of hard particles (silica) could improve the fatigue life
26 when compared to that of the pure epoxy used in their study, while the addition of soft nano rubber
ee

27 particles reduced the fatigue performance. For the sake of comparison, this study is based on the fatigue
28
29 properties from a mix of epoxy with 6% in weight of nano rubber particles and 6% in weight of nano silica
rR

30 particles. The experimental S-N curves for both adhesives are shown in Figure 2 Figure 3.
31
32
ev

33
34
35
36
iew

37
38
39
40
41
42
43
44
45
46
47
48
49
50 Figure 2Figure 3: Relationship between stress amplitude and fatigue life (S-N curve) for: (a) Adhesive I
51 [17][19]; and (b) Adhesive II [18][20].
52
53
54
It should be noted that the experimental S-N curve of Adhesive I, shown in Figure 2(a), was
55 obtained from tests with a positive stress ratio 𝑅 = 0.1, while Adhesive II was tested under fully reversed
56 cycles, i.e., 𝑅 = ―1. Since the stress cycles in the spectrum load are assumed to be always fully reversed,
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Page 27 of 34 Journal name

1
2
3 the Smith-Watson-Topper equation was used to convert the S-N curve of Adhesive I to an equivalent S-
4
5 N curve at 𝑅 = ―1. Table 2 the mechanical properties of the adhesives used in the simulation.
6
7 Table 2: Mechanical properties of the adhesives used in the simulation [17], [18], [19].
8
9 Adhesive I Adhesive II
10
11 Property Value Property Value
12 𝐸 1.6 GPa 𝐸 2.80 GPa
13
𝑣 0.4 𝑣 0.42
14
15 𝜌 1.20 kg/m^3 𝜌 1.17 kg/m^3
16
17 To obtain the Frequency Response Function (FRF) of the structure, a unity impact load is applied
18
19
at the top corner of the stiffener, in a frequency range from 100 to 160 Hz. The first (bending) mode of
the structure was previously identified to be at 133 Hz, thus the spectrum load applied is capable of
Fo
20
21 exciting this first vibrational mode of the joint.
22
23
rP

Damping has the potential to have a huge impact on the FRF of the structure and, consequently,
24
25 may play an important role in fatigue estimation in frequency-domain. Unfortunately, it is extremely
26 difficult to model damping in a structure based solely on the material properties of its components, and
ee

27
thus damping values are usually obtained from experimental investigations. To better evaluate the impact
28
29 of damping in the fatigue life prediction, two different and arbitrary values of damping are be used in this
rR

30 study. Thus, a choice was made for a modal damping of 1.3% and 1.5%, which is applied in the first
31
vibrational mode of the numerical model. The stress response is registered for a point located at the bottom
32
corner of the stiffener, exactly in the middle of the adhesive layer. Both points are shown in Figure 3.
ev

33
34
35 A schematic illustration of the procedure to obtain the fatigue life using a frequency-domain
36
iew

approach is shown in Figure 4. As shown, the fatigue life depends on a structure (geometry and material)
37
38 and a load case. The Finite Element Method (FEM) is used to obtain the FRF of the stress response in the
39 most critical point of the structure. The PSD of the stress response (‘Output PSD’ in Figure 4) is obtained
40 by combining the load PSD with the structure’s FRF through Equation 11. Dirlik’s method is then used
41
42 to obtain the PDF of the stress amplitudes, and the Palmgren-Miner’s rule is used to obtain the expected
43 damage per second (from Equation 8). Finally, the inverse of the expected damage per second gives the
44 fatigue life of the structure.
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Journal name Page 28 of 34

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
Fo
20
21
22
23
rP

24
25
26
ee

27
28
29
rR

30
31
32
ev

33
34 Figure 4: Schematic overview of the fatigue analysis in the frequency-domain (adapted from [22]).
35
36
iew

For this work, a load case with a constant amplitude spectrum of 𝑃𝑆𝐷𝑖 = 15.5 𝑁2/𝐻𝑧 in the
37
38 frequency range from 100 to 160 Hz will be considered, applied in the top corner of the stiffener, as shown
39 by Figure 2.
40
41
It should be noted that due to their random nature, the spectrum loads can only be described in a
42
43 statistical sense, and the history information cannot be retrieved. Thus, in the procedure outlined in this
44 paper, effects such as load interactions are not accounted for in the fatigue analysis.
45
46
The FRF obtained from the numerical analysis is then used to obtain the PSD of the stress
47
48 response, 𝑃𝑆𝐷 𝑆, to a given force input, 𝑃𝑆𝐷 𝑖, as shown:
49
50 𝑃𝑆𝐷 𝑆 = |𝐹𝑅𝐹| 2 𝑃𝑆𝐷 𝑖 (11)
51
52
53
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Page 29 of 34 Journal name

1
2
3 6. RESULTS
4
5
6 The numerical analysis revealed that the first natural frequency of the structure corresponds to a
7 bending mode that happens at 133.7 Hz. The PSD of the stress response was obtained for the case of a
8 constant force 𝑃𝑆𝐷 𝑖 = 5.0 𝑁 2 /𝐻𝑧 applied at the top of the stiffener, as shown in Figure 3, in the frequency
9
10 range from 100 to 160 Hz, calculated according to Equation (11). The stress PSD of both adhesives for a
11 modal damping of 1.5% is shown in Figure 4.
12
13
The numerical analysis revealed that the first natural frequency of the structure corresponds to a
14
15 bending mode that happens at 133.9 Hz. The PSD of the stress response was obtained from the numerical
16 FRF and a constant force 𝑃𝑆𝐷𝑖 = 15.5 𝑁2/𝐻𝑧, calculated according to Equation (11). The stress PSD of
17
18
both adhesives for a modal damping of 1.5% is shown in Figure 5.
19
Fo
20
21
22
23
rP

24
25
26
ee

27
28
29
rR

30
31
32
ev

33
34
35
36
iew

37
38
39
40
41
42
43
44
45
46
47 Figure 4Figure 5: PSD of the stress response for Adhesives I and II at 1.5% modal damping.
48
49
50 It is possible to observe from Figure 4Figure 5 that Adhesive II, which is stiffer, has a higher stress
51 response in all the frequency range analyzed. In fact, the root mean square of the stress process, 𝑆𝑟𝑚𝑠, is
52 equal to 4.6 𝑀𝑃𝑎 𝑆 𝑟𝑚𝑠 = 4.8 𝑀𝑃𝑎 for Adhesive I and has a value of 5.9 𝑀𝑃𝑎 𝑆 𝑟𝑚𝑠 = 6.3 𝑀𝑃𝑎 for
53
54 adhesive II, both at a modal damping of 1.5%.
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Journal name Page 30 of 34

1
2
3 Applying Equation (2), the PDF of the stress amplitudes was obtained for all the study cases, as
4
5 shown in Figure 5 Figure 6.
6
7
8
9
10
11
12
13
14
15
16
17
18
19
Fo
20
21
22
23
rP

24
25 Figure 5Figure 6: PDF of the stress response for (a) 1.3% modal damping and (b) 1.5% modal damping.
26
ee

27 Comparing the PDF for Adhesive I and II, it is possible to observe that the PDF curve is shifted to
28
29 the left, showing a higher probability of occurrence of stress levels of lower amplitude for Adhesive I, and
rR

30 also decreasing the maximum expected stress amplitude expected. When looking at the effect of damping
31 alone, comparing figuresFigures 6(a) and 6(b) reveals only a small difference in the amplitudes of the
32
stress. However, closer inspection, as shown in Figure 6 Figure 7, demonstrates that increasing the
ev

33
34 damping shift the PDF curve to the left, which then shows a high probability of occurrence of lower
35 amplitude stress cycles and lower probabilities at higher amplitudes. To quantify the impact of damping
36
iew

37
on the fatigue life of the structure, Equation (8) was used. The results for fatigue life are shown in Table
38 3.
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Page 31 of 34 Journal name

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
Fo
20
21
22
23
rP

24
25
26
ee

27
28 Figure 6Figure 7: Effect of damping on the PDF curve for Adhesive II.
29
rR

30
As it can be seen, damping has a huge impact on the fatigue life of the structure. As damping
31
32 increases from 1.3% to 1.5%, the fatigue life increases by almost three times for adhesive I. Similarly, in
ev

33 Adhesive II the same change in damping increases the fatigue life by a factor of almost 1.5 two.
34
35
Table 3: Fatigue life for different damping levels.
36
iew

37 Adhesive I Adhesive II
38 1.3% damping 1.7e+05 cycles 1.1e+06 cycles
39
40
1.5% damping 4.5e+05 cycles 1.7e+06 cycles
41
42 Comparing the performance of both materials, Adhesive II has shown better fatigue life regardless
43 of the damping level. At 1.3% damping, Adhesive II improves the crack initiation life by almost one order
44
45
of magnitude. In the prevalence of lower amplitudes of stresses (i.e. at 1.5% damping), the fatigue life of
46 Adhesive II is about 4.5 3.8 times higher.
47
48 7. CONCLUSIONS
49
50
Crack nucleation under spectrum loads is investigated using the S-N approach and the probability
51
52 density function of the stresses is calculated via Dirlik’s method. The method is applied to predict the
53 fatigue life of a skin-to-stiffener bonded joint due to cohesive failure of the adhesive.
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Journal name Page 32 of 34

1
2
3 The study compares the performance of two epoxy-based adhesives whose fatigue properties had
4
5 been previously characterized in the literature and it provides a simple methodology for engineers to easily
6 access the fatigue performance of a bonded joint under random loads using a simple finite element solution
7 of the frequency response function of the structure. Since the method is implemented in the frequency-
8
9 domain, there is no need of time-domain stress history, helping saving time and computational costs in
10 the fatigue analysis.
11
12 Numerical results showed the strong influence of damping in the structure’s response and its effect
13
14 on the fatigue life, in which an increase in the modal damping of the structure also leads to an increase in
15 its life. This happens due to the dissipation effect caused by damping, which decreases the stress level in
16 the adhesive layer. This phenomenon highlights the importance of correct assessment of damping values
17
18 under the risk of non-conservative life estimates in the case of overestimation of damping in the structure.
19
Future works will investigate the experimental validation of the numerical approach outlined in
Fo
20
21 this paper and help assess the potentialities and limitations of the methodology. It should be noted that
22
23
while the approach shown in this paper concerns the case of fatigue life for crack nucleation, a similar
rP

24 approach can be used for crack propagation in the context of Linear Elastic Fracture Mechanics, as
25 demonstrated by the authors for the case of metallic beams [23].
26
ee

27
28
ACKNOWLEDGEMENTS
29
rR

30 This study was financed in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível
31 Superior - Brasil (CAPES) - Finance Code 001.
32
ev

33
REFERENCES
34
35
36 [1] Sinmazçelik T, Avcu E, Bora MO, Çoban O. A review: Fibre metal laminates, background, bonding
iew

37 types and applied test methods. Materials and Design 2011; 32: 3671-3685.
38
39
[2] Bikakis GSE, Kalfountzos CD, Theotokoglou EE. Elastic buckling response of rectangular GLARE
40
41 fiber-metal laminates subjected to shearing stresses. Aerosp Sci Technol 2019; 87: 110-118.
42
43 [3] Zhu S, Chai GB. Low-velocity impact response of fiber-metal laminates – A theoretical approach.
44
Proc IMechE, Part L: J Materials: Design and Applications 2014; 228: 301-311.
45
46
47 [4] Tsamasphyros GJ, Bikakis GS. Analytical modeling to predict the low velocity impact response of
48 circular GLARE fiber-metal laminates. Aerosp Sci Technol 2013; 29: 28-36.
49
50 [5] Bachoo R, Bridge J. Random vibration analysis and modal energy characteristics of fiber-reinforced
51
52 composite beams. Proc IMechE, Part L: J Materials: Design and Applications 2020.
53
54 [6] Hoyt DM, Ward SH, Minguet PJ. Strength and fatigue life modeling of bonded joints in composite
55 structure. J Compos Tech Res 2002; 24: 190-210.
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Page 33 of 34 Journal name

1
2
3 [7] Krueger R, Paris IL, O’Brien K, Minguet PJ. Fatigue life methodology for bonded composite
4
5 skin/stringer configurations. J Compos Tech Res 2002; 24: 56-79.
6
7 [8] Hosseini-Toudeshky H, Mohammadi B, HHamidi B, Ovesy HR. Analysis of composite skin/stiffener
8 debounding and failure under uniaxial loading. Compos Struct 2006; 75: 428-436.
9
10
11 [9] Freitas ST, Sinke J. Failure analysis of adhesively-bonded metal-skin-to-composite-stiffener: effect
12 of temperature and cyclic loading. Compos Struct 2017; 166: 27-37.
13
14 [10] Thawre MM, Pandey KN, Dubey A, Verma KK, Peshwe DR, Paretkar RK, Jagannathan N,
15
16 Manjunatha CM. Fatigue life of a carbon fiber composite T-joint under a standard fighter aircraft
17 spectrum load sequence. Compos Struct 2015; 127: 260-266.
18
19 [11] Antoniou AE, Verpermann MM, Sayer F, Krimmer A. Life prediction analysis of thick adhesive bond
Fo
20
21 lines under variable amplitude fatigue loading. ECCM18 – 18th European Conference on
22 Composite Materials. Athens 2018.
23
rP

24 [12] Mrsnik M, Slavic J, Boltezar M. Frequency-domain methods for a vibration-fatigue-life estimation –


25
26
application to real data. Int J Fatigue 2013; 47: 8-17.
ee

27
28 [13] Kong YS, Abdullha S, Schramm D, Omar MZ, Haris SM. Vibration fatigue analysis of carbon steel
29 coil spring under various road excitations. Metals 2018; 8: 617.
rR

30
31
32
[14] Yeter B, Garbatov Y, Soares CG. Evaluation of fatigue damage model predictions for fixed offshore
wind turbine support structures. Int J Fatigue 2016; 87: 71-80.
ev

33
34
35 [15] Sun Y, Zhang Y, Yang C, Liu Y, Chen X, Yao L, Gao W. Prediction on fatigue properties of the plain
36
iew

weave composite under broadband random loading. Fatigue Fract Eng Mater Struct 2021; 44:
37
38 1515-1532.
39
40 [16] Dirlik T. Application of computers in fatigue analysis. PhD Thesis. University of Warwick. England,
41
1985.
42
43
44 [17] Freitas ST, Sinke J. Failure analysis of adhesively-bonded skin-to-stiffener joints: metal-metal vs.
45 composite-metal. Eng Fail Anal 2015; 56: 2-13.
46
47
[18] Zuo P, Vassilopoulos AP. Review of fatigue of bulk structural adhesives and thick adhesive joints.
48
49 Int Mater Rev 2020; 1-26.
50
51 [19] V. C. Beber, B. Schneider, M. Brede. Influence of temperature on the fatigue behaviour of a
52 toughened epoxy adhesive. The Journal of Adhesives 2016; 92: 778-794.
53
54
55 [20] Wang GT, Liu HY, Saintier N, Mai YW. Cyclic fatigue of polymer nanocomposites. Eng Fail Anal
56 2009; 16: 2635-2645.
57
58
59
60 http://mc.manuscriptcentral.com/(site)
Journal name Page 34 of 34

1
2
3 [21] Zhang J, Deng S, Wang Y, Ye L. Role of rigid nanoparticles and CTBN rubber in the toughening of
4
5 epoxies with different cross-linking densities. Compos Part A 2016; 80: 82-94.
6
7 [22] Berhanu G. Vibration durability testing and design validation based on narrow frequency band.
8 Master Thesis, Blekinge Institute of Technology, Sweden, 2011.
9
10
11 [23] Marques D ET, Vandepitte D, Tita V. Damage detection and fatigue life estimation under random
12 loads: A new structural health monitoring methodology in the frequency domain. Fatigue Fract
13 Eng Mater Struct 2021; 46: 1622-1636.
14
15
16
17
18
19
Fo
20
21
22
23
rP

24
25
26
ee

27
28
29
rR

30
31
32
ev

33
34
35
36
iew

37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 http://mc.manuscriptcentral.com/(site)

View publication stats

You might also like