JMR&T-2023 Final Paper
JMR&T-2023 Final Paper
Article history: This study investigates the microstructural, mechanical, high-cycle fatigue, and fracture
Received 13 April 2023 behaviour of a dissimilar 2017A-T451/7075-T651 Al alloy joint produced by single-pass
Accepted 11 June 2023 friction stir welding (FSW) without post-processing. Residual stress analysis revealed low
Available online 14 June 2023 tensile residual stresses in the longitudinal direction on the crown side and compressive
residual stress in the transverse direction and on the root side of the weld, indicating an
Keywords: almost zero-stress state. High-cycle fatigue tests showed the superior fatigue endurance of
Friction stir welding (FSW) the FSWed sample due to the synergistic effect of expressive grain refinement and
Dissimilar Al alloys compressive residual stress in the transverse direction. Fractographic observations
High-cycle fatigue enhancement confirmed a brittle type II failure with low fracture toughness in the base material 7075-
Cleaner manufacturing technology T651 alloy. In contrast, both the FSWed samples and the base 2017A-T451 alloy dis-
Sustainable engineering played a ductile transition with type I failure, resulting in higher average fracture tough-
ness values. The direction and propagation of the fatigue crack in the FSWed sample were
influenced by the microstructure, with the crack completely bypassing the stir region
consisting of larger crystal grains that acted as the crack arrester. This study highlights the
high potential of single-pass FSW technology for producing thick dissimilar Al alloy joints
with enhanced high-cycle fatigue life without requiring post-treatment. The findings of
this study can benefit industries that require high-performance lightweight structures with
improved fatigue life, such as aerospace and transportation.
© 2023 The Author(s). Published by Elsevier B.V. This is an open access article under the CC
BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
* Corresponding author.
** Corresponding author.
E-mail addresses: [email protected] (U. Trdan), [email protected] (Z. Bergant).
https://doi.org/10.1016/j.jmrt.2023.06.096
2238-7854/© 2023 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://
creativecommons.org/licenses/by-nc-nd/4.0/).
2334 j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 2 3 3 3 e2 3 5 2
catastrophic consequences. While for engineered compo- However, compared to the base materials, FSW joints
nents, most failures are caused by corrosion (29%), followed by usually still have lower fatigue strength [16]. Additionally, the
fatigue (25%) and only 3% by wear/abrasion/erosion, the trend dynamic response of an FSW joint can vary significantly in the
is reversed for aircraft. There most failures are caused by fa- cross-section and the in-plane rolling and/or welding direc-
tigue (55%), followed by corrosion (16%) and wear-related tion [17]. Moreover, the thickness reduction, flashes, and
failures with only 6% in last place. The high incidence of fa- keyholes that can form in FSW joints can significantly affect
tigue failure is related to the sudden destruction caused by the the mechanical properties of the joint [18,19]. To promote
development and propagation of the fatigue crack, while wider use of FSW in manufacturing industries, it is necessary
corrosion is a much slower process and more visible to the to address three inherent challenges: back support, weld
naked eye [1]. thinning, and keyhole defects. This is crucial for ensuring the
Although fibre-reinforced plastics (CFRP) are increasingly structural integrity, safety, and service life of the products
replacing traditional materials, metallic materials still ac- being manufactured [18]. For example, flashes on the crown
count for almost 50% of the total. For instance, Airbus' flagship side of the FSW joint may cause significant problems with
A350 long-range aircraft consists of 50% composites, 20% Al surface finish, corrosion resistance, and stress concentrations
alloys (AA), 15% titanium alloys, 10% steel and 5% other ma- [19]. In view of this, Zeng et al. [19] proposed an innovative
terials. However, joining different materials and alloys is still concept of combining friction stir welding and end milling to
mainly done using “heavy” mechanical connections with fabricate welds and remove flashes simultaneously. However,
screws and rivets. Friction Stir Welding (FSW), invented by the due to the high temperature generated during flash removal,
Welding Institute (TWI) in 1991, is an innovative, clean solid- the implementation of additional cooling with compressed air
state process for successfully joining light alloys, compos- is necessary to achieve proper mechanical properties with a
ites, and dissimilar materials [2,3]. AA7075 is one of the very low feed demand. On the other hand, Friction plug
strongest Al alloys in today's industrial applications. It is a welding (FPW) is a widely reported method for repairing local
precipitation-hardenable alloy with Zn as the main alloying volume defects, such as keyholes [18].
element and is widely used for highly stressed structural ap- Mimouni et al. [20] investigated the microstructure and
plications in the aerospace industry [4]. Similarly, the age- mechanical behaviour of the FSW 2017-T451 Al alloy and
hardenable AA2017 contains Cu as the primary alloying confirmed a significant decrease in fatigue life of the welded
element, whereas the addition of Mg results in the formation samples. Even when the surfaces of the welds were treated by
of Al2CuMg particles that provide a good combination of shot peening (SP) or laser shock peening (LSP), the fatigue life
strength, fracture toughness, and fatigue resistance. There- was still lower compared to the base materials [16,21]. This
fore, it is often used in aviation (fuselage, wings) and in the indicates the significant influence of welding parameters. If
automotive industry, where it can potentially be exposed to these parameters are not fully optimised, the mechanical
dynamic impact loads in car accidents or bird strikes in properties of the weld cannot reach the desired level, even
aircraft [5]. However, the presence of numerous intermetallic with sophisticated surface treatment processes. Components
phases such as q phase (Al2Cu), S phase (Al2CuMg), Mg2Si, in aircraft structures are constantly exposed to vibrations that
Al3Fe, Al7Cu2Fe, T phase (Al20Cu2Mn3), Al4Cu2Mg8Si7, AlCu- stress the material with small amplitude and high frequency
FeMnSi and others) [6e8], which have different melting tem- [22]. Cyclic loading in aerospace and transport structures
peratures and can form eutectic and fragile intermetallic likely exceeds one million cycles [4]. This makes the perfor-
phases in the fusion zone [9], limits the joining of dissimilar mance of structural materials, specifically high cycle (HCF)
2xxx and 7xxx Al alloys. As a result, heavy-weight riveting is and very high cycle fatigue (VHCF), of great practical
predominantly used for joining such materials, which con- importance.
tributes to higher weight, increased fuel consumption, and The magnitude and sign of residual stresses are closely
subsequently higher pollutant emissions [10]. On the other related to fatigue properties. Tensile residual stresses caused
hand, FSW represents a promising technology for joining by welding promote crack initiation and propagation, while
various 2xxx and 7xxx series high-strength aerospace Al compressive residual stresses increases fatigue strength. FSW
alloys. parameters such as tool material, geometry, welding speed
FSW uses a rotating and traversing tool to generate fric- and turning speed influence the residual stresses distribution
tional heat, causing plastic deformation and mechanical and, consequently, the fatigue performance. Jafari et al. [23]
mixing of semi-solid base materials below the melting tem- investigated the residual stresses distribution of dissimilar
perature, which are brought into contact as butt, lap or clad friction stir welded 7075-T6 and Al6051-T6 with and without
joints. FSW avoids the disadvantages that occur with fusion SiO2 nanoparticles. The results confirmed that without SiO2
welding processes, such as hydrogen embrittlement, strong nanoparticles the tensile residual stresses can be reduced for
shrinkage-induced porosity and equiaxed microstructure 63 MPa at high turning speed and for MPa at low turning
[11,12], and excellent bonding can be obtained [13]. Due to the speed. Furthermore, samples with lower tensile residual
formation of finer recrystallized areas in the stirred zone and stresses obtained a higher ultimate strength. Therefore, opti-
the minimization of the geometric notch effect, higher frac- mizing the FSW parameters to minimize residual stresses is
ture toughness and longer fatigue life than with fusion- crucial for achieving a high fatigue strength of the welded
welded joints can also be expected. Successful FSW applica- joint. FSW joints consist of four different microstructural
tions and several reports have shown that FSW welds have zones, namely the stir zone (SZ) or nugget zone (NZ), heat
higher tensile strength and several times longer fatigue life in affected zone (HAZ), and thermo-mechanically affected zone
the low and high cycle range [14,15]. (TMAZ) [24]. The heterogeneity of FSW joints, similar to fusion
j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 2 3 3 3 e2 3 5 2 2335
welding, requires the identification of detrimental micro- fatigue behaviour compared to the base material. Fatigue tests
structural formations. It is well-known that in fusion welding, were performed on an electromagnetically driven testing de-
equiaxed structures near the fusion line are the weakest link vice at an average starting resonant test frequency of 100 1 Hz
against fatigue, where cracks often occur under dynamic and a stress ratio of R ¼ 0.1 using Charpy V-notched samples.
loading [12]. Such fine equiaxed zone (FQZ) can also be formed In the final part of the study, the microstructure and fractog-
in the FSW process, which does not necessarily represent the raphy of both the base materials (BMs) and FSW samples were
material area with a higher fatigue crack growth rate. How- investigated and evaluated.
ever, according to the current state of the art (SOTA), there are
many contradictory arguments. For example, Vysotskiy et al.
[25] reported higher fatigue strength of FSWed heat treatable 2. Experimental methods
AA6061 than the base material. However, to retain mechanical
properties after FSW, the welds were artificially aged at 160 C 2.1. Materials and manufacturing process
for 8 h. The joints failed mostly in the base material in the low
cyclic regime and in HAZ in the high cyclic regime due to The BMs used in this study are two dissimilar precipitation
coarser microstructure in this region, but did not fail in the hardened Al alloys, artificially and naturally aged, namely
nugget zone. The authors stated that the increased fatigue 2017A-T451 and 7075-T651 Al alloys, respectively. Both BMs
resistance of the stir zone material can be related to the grain were supplied by Amari Austria GmbH in the form of 10 mm
refinement that occurred in this microstructural region. thick rolled plates. The chemical compositions of both alloys
In contrast, Akhtar et al. [26] confirmed a 2e3 times higher are depicted in Table 1.
fatigue crack growth rate in the NZ of the FSWed AleLi alloy Prior to friction stir welding (FSW), the samples were
compared to the base material. However, the weakest link sectioned using a water jet process along the longitudinal (L)
with the occurrence of delamination was confirmed within and long-transverse (T) rolling directions into rectangular
the TMAZ region. He et al. [4] also confirmed that the AA7075- shapes with dimensions of 500 70 10 mm3 (w l t). Just
T651 base material had 65% higher fatigue strength than before the FSW process, all the samples were thoroughly
FSWed without post-weld ageing in a VHCF regime. However, degreased with acetone and ethanol. Optimised FSW, based
both the HAZ and TMAZ on the retreating side (RS) achieved on preliminary tests, visual appearance, bending and tensile
higher fatigue strength than the NZ weld region. It has been results, and microstructural examination (results not pre-
reported that the Mg2Si precipitates found in the fine equiaxed sented herein), was performed in the longitudinal direction of
grains (2e7 mm) of the NZ act as stress raisers for the devel- the rolling direction of the BMs. To create the welds, the
opment of fatigue cracks. Nevertheless, the TMAZ had a lower samples were hydraulically clamped in a butt-weld configu-
fatigue strength on the advancing side (AS) than the NZ ration against a backing anvil plate (see Fig. 1a) and the
despite a similar horizontal plateau in the VHCF range as the welding was done in a force-control mode using an axial
TMAZ-RS. The different fatigue behaviour between TMAZ-AS downward force of 38 kN. The welding tool comprised a
and TMAZ-RS was attributed to microstructural differences, shoulder with a 25 mm diameter and a cylindrical pin with a
with a sharp boundary between NZ and TMAZ on the diameter of 9 mm and a length of 9.8 mm. The tool offset was
advancing side and an unclear boundary on the retreating positioned 1 mm from the joint centre line towards the RS
side. Post-artificial ageing could enhance yield, tensile, and (7075-T651), with a tilt angle of 1 , while the rotation per feed
fatigue properties since the hardening particles dissolved was kept constant at 20 rpcm, and the tool rotation was
during welding are precipitated again [27]. counterclockwise [17]. A photograph of the as-welded joint is
Despite considerable progress in this area, little attention shown in Fig. 1b.
has been paid to investigating the fatigue properties of dis-
similar FSWed alloys (2xxx - 7xxx Al), which are widely used in 2.2. Surface topography and residual stress
the aerospace industry. Moreover, there is no literature on the characterization
microstructural properties and mechanical behaviour of
2017e7075 All alloy butt joints. Therefore, in the present Since surface topography affects the origin of stress concen-
study, 10 mm thick dissimilar FSWed 2017e7075 Al butt joints, trators and can tremendously affect the fatigue life of
obtained by a single-pass welding, were analysed with a dynamically stressed machine parts [28,29], topographical
special emphasis on the fatigue and fracture behaviour since analyses of BMs and FSW welds along the centreline were
fatigue is the main cause of failure in welded structures. The determined in both the L and T directions (see Fig. 1b). Surface
main objective was to investigate how the change in surface roughness and waviness were obtained using unified high-
topography, microstructural features, residual stresses, and pass and low-pass Gaussian cut-off filtration (lc ¼ 800 mm),
mechanical properties through the FSWed process affects the respectively, with a Surtronic 3þ contact stylus profilometer
2336 j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 2 3 3 3 e2 3 5 2
Fig. 2 e Schematic presentation of the position of the tensile and fatigue notched samples on the first third of the FSWed
joint.
criterion to 95 Hz. The clamping detail and position of the V- T651 and 8 samples for 2017-T451 in the long-transversal (LT)
notched samples during fatigue testing are shown in Fig. 3. direction perpendicular to rolling. A total of 18 fatigue samples
The Cracktronic testing device consists of a basic module were cut along the weld, 6 from 1/3 of the weld length (FSW1),
with separate static and dynamic drives. The static load is 6 from 2/3 (FSW2), and 6 from the last third (FSW3) of the 2017-
generated by a ball screw spindle driven by a DC motor and T651/7075-T651 dissimilar FSW joint (see Fig. 2). Afterwards,
couples over the torsion rod into the oscillating system. The 2 mm deep V-notches with a tip radius of 0.2 mm and a notch
dynamic load is generated by an electromagnetically driven face angle of 45 were made using EDM in the T-S direction.
resonator with a maximum swing angle of 2 . The electro- The fracture toughness, Kc, was measured on the same
magnet is integrated into a closed control loop and excites the sample after fatigue crack growth tests using a universal
vibrating system at its natural resonant frequency, with the testing machine BETA 50 Messphysik with a piezoelectric
operating point at the peak of the resonance curve, with force sensor with a measuring capacity of 50 kN and with a
typical frequencies between 50 and 240 Hz, depending on the resolution of 1 N. The displacement was measured using an
stiffness of the sample, which is part of the mass-vibration optical laser speckle extensometer with an accuracy of within
system [33]. The dynamic and static parts are controlled by 2%. The initial fatigue crack length was measured under the
separate circuits connected to the Credo control unit and optical microscope from the line of load to the crack front,
computer software, allowing any combination of stress ratios where the measurement error is ±0.01 mm. The fracture
with high accuracy. toughness was measured after increasing the fatigue pre-
The alternating torque is measured using a Rumul load crack, made during the fatigue test.
transducer with a maximum moment of 70 Nm (±35 Nm) with Afterwards, the cross-sectional crack morphology (T-S
a resolution of 0.01 Nm. The Cracktronic high-frequency pul- plane) was analysed using a Keyence VHX-600 digital micro-
sator is designed for dynamic in-plane bending of CVN sam- scope after dynamic fatigue and static fracture toughness
ples. A total of 8 samples were used for the base material 7075- tests. In addition, fractographic analysis of the fracture
Fig. 4 e Profile line scans across the FSWed dissimilar 2017A/7075 joint in the transverse direction, with a schematic
presentation of the measuring lines for roughness and waviness analyses shown in blue colour.
surfaces (L-S plane) was conducted using the Thermo Scien- the T direction was 24.27 ± 3.98 mm, whereas in the opposite,
tific ESEM FEG Quattro S scanning electron microscope (SEM). i.e. L direction of the welding, it was 5.06 ± 0.97 mm.
The SEM analyses were performed at preselected locations on Nonetheless, three separated measurements of topog-
the fatigued surfaces and at the transition zone from fatigue raphy parameters on a FSWed region, in the specific direction
to static fracture surface using an accelerating voltage of have confirmed good surface uniformity with small standard
20 kV. deviation values. For a comparison, the lowest roughness was
obtained with the 7075-T651 BM sample in the parallel direc-
tion, with a roughness Ra value of 0.29 ± 0.02 mm and waviness
3. Results and discussion Wa of 0.68 ± 0.21 mm. As expected, in the opposite direction
slightly higher values were obtained as consequence of the
3.1. Surface topography and residual stresses preliminary rolling of the plates.
Fig. 5a shows the distribution of residual stresses along the
Fig. 4 displays the topographical line scan profiles taken in the FSW joint on the top/crown side in its “as-is” condition, prior to
transverse direction across the weld at the first, second, and the cutting of the sample by WaterJet (WJ). As expected, a
third parts of the FSWed dissimilar 2017A/7075 plate. The line higher magnitude of residual stresses with a tensile nature is
scan profiles depicted in Fig. 4 show similar topography in all prevailing within the longitudinal stress component (sL) [34].
specific regions, indicating the appropriate selection of opti- Although at the beginning and end of the weld, the residual
mised FSW parameters for manufacturing dissimilar joints. stresses is almost in a zero-stress state, with low tensile re-
Furthermore, since the similar topography was achieved, the sidual stresses values of 17 ± 4 and 8 ± 6 MPa, respectively, its
surface roughness and waviness of the BMs and FSW joint, in magnitude gradually increases with increasing distance along
the L and T directions, as shown in Table 2, represent the the weld. At a distance of 100 mm from the beginning of the
average values based on three separate measurements. weld, the residual stresses reaches 69 ± 4 and exceeds 80 MPa
As can be seen from the topography profiles in Fig. 4, the at 210 mm from the beginning of the weld. A stable region
furrows introduced by FSW significantly increase surface (from 250 to 350 mm) with the residual stresses of about
roughness and waviness in both directions. This effect may 60 MPa is present, followed by a sharp gradient decrease to-
contribute to stress concentration during fatigue testing. wards the end of the weld.
Additionally, as shown in Table 2 despite similar Ra and Rz In contrast to the longitudinal residual stresses, the
roughness values in the L and T directions (4.95 ± 0.18 vs. transverse component (sT) shows beneficial compressive re-
4.68 ± 0.15 and 24.77 ± 0.53 vs. 24.92 ± 0.66 mm), the results sidual stresses values ranging from 10 ± 6 to 16 ± 7 MPa
depict significantly higher waviness in the parallel/transverse over a larger portion of the FSW joint from 100 to 350 mm.
direction as a collateral effect of tool rotation and movement Although the results presented in this study are consistent
during welding. For example, the maximum waviness Wa in with other studies of within FSW joints [25,28,34e37], much
Table 2 e Surface roughness and waviness of the BMs and FSWed dissimilar joint in the specific direction.
Region Longitudinal (L) direction Transverse (T) direction
Ra [mm] Rz [mm] Wa [mm] Wz [mm] Ra [mm] Rz [mm] Wa [mm] Wz [mm]
2017A-T451 0.57 ± 0.09 3.51 ± 0.60 0.96 ± 0.21 1.46 ± 0.29 0.91 ± 0.06 5.23 ± 0.46 1.22 ± 0.20 1.97 ± 0.17
7075-T651 0.29 ± 0.02 1.80 ± 0.09 0.67 ± 0.19 0.91 ± 0.18 0.36 ± 0.06 2.62 ± 0.26 0.68 ± 0.21 1.15 ± 0.39
FSW 4.95 ± 0.18 24.77 ± 0.53 5.06 ± 0.97 9.95 ± 1.01 4.68 ± 0.15 24.92 ± 0.66 24.27 ± 3.98 20.87 ± 1.53
j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 2 3 3 3 e2 3 5 2 2339
Fig. 5 e Longitudinal (sL) and transverse (sT) residual stresses; (a) along dissimilar FSW joint on the top/crown side (Inset:
Residual stresses profile before and after WaterJet sectioning of the samples), (b) on the crown and root side on the specific
distance along the weld after WaterJet sectioning and (c) on the top and bottom side of the BMs samples.
lower tensile longitudinal and transverse residual stresses intercept grain size analyses according to the ASTM E112-10
were achieved, which, together with good surface uniformity, [38] and standard revealed a much larger average grain size in
confirms appropriate FSW manufacturing parameters. More- 2017A-T451 compared to the 7075-T651 Al alloy. The grain size
over, the inset diagram within Fig. 5a depicts that, after
sectioning the samples from the FSW joint using the WJ pro-
cess, the compressive state of transverse stresses slightly in-
creases due to the relaxation effect during cutting. Since
compressive residual stresses is beneficial for fatigue and
fracture behaviour, it should be emphasized that such stress
state within the FSWed joint is indeed desirable if the machine
part or structure is exposed to cyclic loading in the T direction.
In this way, transverse residual stresses has a predominant
effect on crack closure and mitigates crack propagation.
To gain a better understanding of the material state prior to
subjecting it to mechanical loading, residual stresses were
measured in the L and T directions on both the top and bottom
sides of the WJ-sectioned FSWed (Fig. 5b) and BM's (Fig. 5c)
samples. The results of the FSW WJ-sectioned samples in
Fig. 5b confirm a different residual stresses state at specific
sample locations. The initial first third (FSW1) of the weld has,
on average, the lowest compressive residual stresses, which
could contribute to lower fatigue strength compared to the
samples sectioned from the second (FSW2) and final third
(FSW3) of the weld.
Fig. 7 e Cross-sectional microstructures of the FSWed dissimilar 2017A-7075 joint; (a) macrograph at lower magnification
showing specific regions within the weld; (b) TMAZ/NZ interface on RS, (c) NZ towards AS, (d) NZ/TMAZ interface on AS, (e)
TMAZ-RS (Inset: dark-field micrograph demonstrating precipitate distribution), (f) NZ towards RS and (g) (e) TMAZ-AS (Inset:
dark-field micrograph demonstrating precipitate distribution).
was 55% larger in the T direction (129.7 vs. 83.6 mm) and as general view at lower magnification (Fig. 7a) shows the exis-
much as 200% larger in the S (i.e. short transverse e see Fig. 2) tence of seven specific regions within the weld. On the far left,
direction (43.4 vs. 14.3 mm), respectively. on the retreating side (RS) toward the centre, is the unaffected
Comparing the grain size morphology of the BMs in the 7075-T651, followed by the heat-affected zone (HAZ - RS), the
specific direction, a much higher anisotropic effect can be thermo-mechanically affected zone (TMAZ - RS), and the
seen in the 7075-T651 alloy, where the grains differentiate by nugget zone (NZ). Further on the advancing side (AS) are the
as much as a factor of 5.85 (83.6 vs. 14.3 mm), whereas with the TMAZ - AS, the HAZ - AS, and the unaffected Al alloy 2017A-
Al alloy 2017A-T451, the difference of 2.99 is almost half of T651 on the far right section. As shown in the macrograph in
that in the 7075-T651 (129.7 vs. 43.4 mm). Such results indicate Fig. 7a, there is no difference in the appearance of the NZ on
a lower degree of static recrystallization (SRX) effect for the the RS, whereas within the NZ region on the AS side, an
7075-T651 alloy, which occurred during annealing after rolling expressive banded/onion-like structure is present. Here,
[32]. Hence, due to the lower restoration of the microstructure alternating lighter and darker layers, consisting of different
with higher dislocation density, higher strength can be grain morphologies and sizes, coexist. Higher magnification
anticipated [39]. Nevertheless, both micrographs confirm that (Fig. 7c) confirms oriented, larger grains in the range of
complete precipitation took place in the microstructure of 10e20 mm within the lighter layer (denoted as Stir region “A")
both BMs during the artificial T651 or the natural T451 peak and small equiaxed grains of 2e7 mm within the darker layers
aging process. of the NZ (denoted as Stir region “B"). It is noteworthy that
The cross-section of the dissimilar 2017A/7075 FSW joint, despite sufficient energy and deformation strain enable the
divided into representative regions, is shown in Fig. 7. The occurrence of full dynamic recrystallization (DRX), the
j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 2 3 3 3 e2 3 5 2 2341
asymmetrical banded microstructure develops within the NZ and centre of the weld [29]. The first measurement line was
due to the asymmetrical plastic flow during the FSW process performed at 0.8 mm below the top/crown side, the second at
[40,41]. the middle of the cross-section, 5 mm below the surface, and
In addition, the TMAZ regions where the material is “trap- the third at 0.8 mm above the bottom/root side of the weld,
ped” between and flows around the rotating tool pin during the respectively. The hardness tests were carried out approxi-
FSW process exhibit significant differences on the RS (Fig. 7b) mately one month after the FSW process, without any addi-
and AS sides (Fig. 7d). As shown in Fig. 7d, a clear boundary tional heat treatment. As depicted in Fig. 8, softening occurred
between the TMAZ and NZ regions is observed on the AS side within the HAZ regions of the dissimilar 2017e7075 FSW joint,
due to a higher gradient of microstructure variation. On the with the lowest hardness value of 93 HV0.3 measured on the
other hand, the boundary between TMAZ and NZ on the advancing side, in the bottom region at approximately 5.5 mm
opposite, RS side (Fig. 7b) is not clear as the microstructural from the weld centre. On the other hand, the RS side had the
gradient is much lower than on the advancing side, where a lowest hardness of 116 HV0.3, measured in the upper region at
higher maximum centrifugal force and energy for recrystalli- about 10 mm from the weld centre. In addition to this soft-
zation growth is present [41]. As a result, microstructure with ening effect, a good hardness distribution can be seen along
equiaxed and smaller crystal grains than on the AS side the FSW joint. Within the TMAZ region, on the retreating side,
(Fig. 7c) is present in the NZ on the RS side (Fig. 7f). In addition hardness reached a value of 189 HV0.3, exceeding the hard-
to the expressive DRX, the precipitation dissolution and re- ness of the unaffected 7075-T651, with an average value of
precipitation also occurred. This is further confirmed by the 176 ± 8 HV0.3. In addition, despite complete recrystallization
dark-field inset micrographs within RS and AS TMAZ regions in the NZ region, the average microhardness is still higher
(Fig. 7e and g), where dense precipitation formation within the compared to the unaffected 2017A-T451 alloy (149 ± 12 vs.
crystal grains and on the grain boundaries can be seen, even 133 ± 6 HV0.3).
though DRX did not occur due to insufficient deformation Similar microhardness results along the FSW joint were
strain. It should be noted that the dissolution, reprecipitation also reported by He et al. [4] and Jafari et al. [23] for similar
kinetics its volume fraction and size modification of 7075-T6 and dissimilar 7075-T6/6061-T6 FSW Al alloy joints,
precipitation-hardened Al alloys during/after the FSW signif- respectively. Nonetheless, most researchers have reported a
icantly affect the mechanical properties [18]. significant reduction in microhardness within FSW joints
Since a previous study [4] reported that small equiaxed [25,26,36]. As discussed earlier [36,42], the microhardness of
grains in the NZ FSW region lead to detrimental fatigue heat treatable Al alloys in FSWed joints depends on several
behaviour with higher crack propagation than BM, this phe- features, such as the formation and distribution of pre-
nomenon will be thoroughly investigated in the following cipitates and intermetallic compounds, grain size, etc. Ac-
sections. Finally, there is an evident band of larger grains in cording to the Hall-Petch relationship, material hardness is
the centre of the 2017A-T451 cross-section from the macro- inversely proportional to the grain size [43]. Therefore, the fine
graph in Fig. 7a. Despite the fact that this alloy generally has a equiaxed grains formed within the NZ should result in high
lower texture and anisotropy effect in the specific direction hardness values. However, the heat generated during welding
than the Al alloy 7075-T651 (Fig. 6), such material's through- softens the material by simultaneously dissolving the
thickness texture can affect the repeatability of the fatigue strengthening precipitates, which causes a reduction in
behaviour and should also be investigated. strength and hardness. Hence, the optimum heat generation
is required for mixing the BMs while preventing complete
3.3. Microhardness and static tensile testing results dissolution of the precipitate within the weld [42].
The results of the static tensile testing for the 7075-T651
Fig. 8 shows the Vickers microhardness results along the FSW and 2017A-T451 base materials and the dissimilar FSW joint
joint from RS (left side) to AS (right side), at the top, bottom, are shown in Table 3, including the average measured values
of tensile strength, yield stress, and elongation. The
Table 4 e Fatigue testing data at R ¼ 0.1 for 2017A-T451 and 7075-T651 Al alloy samples.
# Мm [Nm] Ma [Nm] smax;nom [MPa] smax [MPa] Ds [MPa] N ðcyc:Þ f0 [Нz] t [h] Failed
1 2017-T451 11.73 9.60 200 380 342 45 708 100.2 0.1 Yes
2 8.80 7.20 150 285 256 142 281 99.7 0.4 Yes
3 7.04 5.76 120 228 205 280 297 100.7 0.8 Yes
4 5.87 4.80 100 190 171 585 277 100.8 1.7 Yes
5 5.28 4.32 90 171 154 977 699 100.4 2.7 Yes
6 4.69 3.84 80 152 137 1 712 043 100.4 3.9 Yes
7 4.41 3.61 75 142 128 2 559 614 101.1 5.8 Yes
8 4.11 3.36 70 133 120 15 602 591 101.0 43.3 No
1 7075-T651 17.60 14.40 300 570 513 12 594 100.21 0.03 Yes
2 11.73 9.60 200 380 342 18 821 100.77 0.05 Yes
3 8.80 7.20 150 285 256 39 198 100.48 0.11 Yes
4 7.04 5.76 120 228 205 84 051 99.91 0.2 Yes
5 6.45 5.28 110 209 188 125 253 101.19 0.2 Yes
6 6.16 5.04 105 200 180 149 601 100.32 0.4 Yes
7 5.57 4.46 95 181 163 279 695 101.03 0.8 Yes
8 4.41 3.61 80 152 137 680 160 101.22 1.5 Yes
Another crucial factor that affects fatigue behaviour is the For the plane strain fracture toughness KIC mode I test, the
state of residual stress in the material. As shown in Fig. 5a, thickness should be equal or higher than:
residual stresses in the T direction to welding (and fatigue 2
loading direction) is compressive nature, which has a benefi- Kc
Bq (17)
cial effect on the crack closure and inhibition of crack propa- Rp0:2
gation. Additionally, residual stresses in the longitudinal and where B is the thickness of the sample for plane-strain frac-
T direction on the upper and the lower side of the WJ ture toughness test, and q is the plane-strain/plane stress
sectioned samples (Fig. 5b and c) confirms a perfect correla- transition empirical constant set to q ¼ 2.5 and KC ¼ KQ . The
tion with the dynamic fatigue behaviour. Thus, the initial Charpy-V notch bend sample is a single edge notched sample
third of the weld (FSW1), which showed the lowest compres- SEN(B), loaded in a three-point bending configuration with a
sive transverse residual stresses, failed at a lower maximum support span of S ¼ 40 mm, which is nominally equal to four
stress compared to the samples further along the weld (Table times the width, W ¼ 10 mm. Normally, the SEN(B) has
6). The same argument applies to the BMs, where the 2017A- B ¼ 2W. The Charpy bend sample has a width of B ¼ 10 mm
T451 Al alloy, with lower tensile residual stresses in the T di- and W ¼ B, which also fulfills standard ASTM E399 [49] rec-
rection, exhibited higher fatigue strength. ommendations for an alternative sample (1 W=B 4).
Samples were loaded in a test rig with cylindrical supports
3.5. Fracture toughness of 10 mm in diameter (see Fig. 13). The spindle speed was set to
0.5 mm/min, which corresponds to a stress intensity increase
Fracture toughness tests were performed on bending samples pffiffiffiffiffi
of 0 :58 MPa m= s. This value is at the lower bound of stan-
after the fatigue pre-crack to evaluate the effects of FSW
dard ASTM E399 [49] interval between 0.55 and 2.7 MPa√m/s
welding compared to the base materials, according to stan-
and was chosen to avoid overestimation of fracture toughness
dard tests ASTM E1820 [48] and ASTM E399 [49]. The stress
at high strain rates. The initial tests of the base material were
intensity factor K for the SEN(B) sample at force Pi was
performed directly after the fatigue tests, which terminated at
calculated as:
95 Hz. Upon fracturing the first set of samples, the a/W ratio
Pi S a was determined to be of around 0.4, which was slightly below
i
KðiÞ ¼ $f (15)
ðBBN Þ
1
2 W
3
2 W the required ratio. Therefore, three samples in each testing
group were subsequently fatigued to 92 Hz to achieve a valid a/
where B represents the thickness, BN is net thickness (BN ¼ B, W ratio within the interval between 0.45 and 0.55.
if no side grooves are present), W is the depth (width), ai is the However, in the case of 2017A-T451, the linear elastic re-
current crack length, S is the span. At force PQ, which is the quirements are violated. For example, the Pmax/PQ is 1.41 and
force where the 95% secant line intersects the P-v graph, exceeds 1.1. The material closest to plane strain condition is
KðiÞ ¼ KQ . The geometry factor f(ai/W) is defined as: 7075-T651 Al alloy, with an average Pmax/PQ ratio of 1.08, while
1
a 3ðai =WÞ2 1:99 ðai =WÞð1 ai =WÞ$ 2:15 3:93ai =W þ 2:7ai 2 W2
i
f ¼ 3
(16)
W 2ð1 þ 2ðai =WÞð1 ai =WÞ2
j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 2 3 3 3 e2 3 5 2 2345
Table 5 e Fatigue testing data at R ¼ 0.1 for 2017e7075 Al alloy FSWed samples.
# Мm [Nm] Ma [Nm] smax;nom [MPa] smax [MPa] Ds [MPa] N ðcyc:Þ f0 [Нz] t [h] Failed
1 Region 1 11.73 9.60 200 380 342 14 363 101.38 0.04 Yes
2 8.80 7.20 150 285 256 42 707 101.17 0.14 Yes
3 7.04 5.76 120 228 205 173 535 101.21 0.48 Yes
4 5.87 4.80 100 190 171 591 897 101.40 1.64 Yes
5 5.28 4.32 90 171 154 17 102 153 101.33 47.51 No
6 4.69 3.84 80 152 137 16 647 463 101.47 46.24 No
1 Region 2 11.73 9.60 200 380 342 15 909 101.02 0.04 Yes
2 8.80 7.20 150 285 256 46 530 101.59 0.13 Yes
3 7.04 5.76 120 228 205 130 665 101.06 0.36 Yes
4 6.45 5.28 110 209 188 193 102 100.99 0.54 Yes
5 6.16 5.04 105 200 180 334 984 101.22 0.93 Yes
6 5.87 4.80 100 190 171 16 796 342 100.83 46.66 No
1 Region 3 10.27 8.40 175 332 299 26 740 101.31 0.07 Yes
2 8.80 7.20 150 285 256 44 416 100.52 0.12 Yes
3 7.04 5.76 120 228 205 151 899 101.01 0.42 Yes
4 6.45 5.28 110 209 188 253 369 101.32 0.70 Yes
5 6.16 5.04 105 200 180 417 388 101.23 1.16 Yes
6 5.87 4.80 100 190 171 16 253 907 101.31 45.15 No
Fig. 12 e SeN curves for the 2017-T451 and 7075-T651 base Al alloys and FSW dissimilar samples.
2346 j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 2 3 3 3 e2 3 5 2
J
J ¼ Je=0 þ p=0
a 0:5 Da (21)
1þ
a þ 0:5 b0
Table 7 e Results of initial crack length a, a/W ratio, Pmax, PQ, Pmax/PQ ratio, linear-elastic fracture toughness results KC, and
J-integral for 2017A-T451 and 7075-T651 base materials and dissimilar 2017A/7075 FSWed samples, respectively.
pffiffiffiffiffi
# Material a ½mm a=W Pmax [N] PQ [N] Pmax =PQ KC [MPa m] J [kJ/m2]
1 2017A-T451 4.69 0.47 4041 2779 1.45 26.9 136.9
2 2017A-T451 4.73 0.47 3849 2762 1.39 26.8 122.9
3 2017A-T451 4.63 0.46 3958 2858 1.39 26.9 151.6
1 7075-T651 4.8 0.48 4232 3595 1.12 35.9 20.1
2 7075-T651 4.68 0.47 4244 3836 1.06 37.2 20.4
3 7075-T651 4.8 0.48 4132 3804 1.09 38.1 21.3
1 FSW1 4.54 0.45 4302 3691 1.15 34.2 102.0
2 FSW1 4.68 0.47 4320 3643 1.20 35.1 89.8
3 FSW1 4.72 0.47 3956 2912 1.36 28.4 78.0
4 FSW1 4.72 0.47 3961 3404 1.16 33.2 83.5
5 FSW1 4.82 0.48 3909 3492 1.12 35.1 86.5
6 FSW1 4.86 0.48 3891 3375 1.15 34.0 83.1
7 FSW2 4.86 0.49 3821 3420 1.15 34.9 87.3
8 FSW2 4.85 0.48 3916 3440 1.12 34.9 96.7
9 FSW2 4.89 0.49 3772 3391 1.14 34.8 100.6
10 FSW2 4.87 0.49 3895 3445 1.11 35.2 99.9
11 FSW2 4.90 0.49 3842 3418 1.13 35.2 101.7
12 FSW2 4.95 0.49 3689 3281 1.12 34.4 96.7
13 FSW3 4.86 0.49 3902 3210 1.22 32.7 97.4
14 FSW3 4.78 0.48 3889 3521 1.10 35.0 87.7
15 FSW3 4.97 0.50 3693 3205 1.15 33.7 92.80
16 FSW3 4.85 0.49 3844 3345 1.15 34.0 97.37
17 FSW3 4.85 0.49 3848 3355 1.15 34.1 92.17
18 FSW3 4.86 0.49 3782 3372 1.12 34.4 97.41
j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 2 3 3 3 e2 3 5 2 2347
Fig. 14 e Force-displacement curves (a) and fracture toughness energy J-integral (b).
obtained. The bell-shaped curve reached its maximum at a crack surface in Al alloy 7075-T651 and 2017A-T451, and to
value about 40% above the PQ without cracking (red line), some extent in the first third of the FSW plate, appears more
while the average Pmax /PQ is 1.45. From the histogram of J- or less flat. In the latter case, the slant fracture mode has not
integral values (Fig. 14b), it can be observed that the 2017A- developed, despite the fractures occurring in plane stress
T451 alloy achieved the highest fracture toughness with an conditions. However, the most fractures of the FSW samples
average of 137, while the Al alloy 7075-T651 has the lowest (FSW2 and FSW3) consist of flat and slant fracture due to the
fracture toughness with an average of only 20.6 kJ/m2. The shear stresses in the presence of plane stress conditions.
energy required for fracture depends on the area under the P-v Nonetheless, there is more plastic deformation in 2017A-T451
curve (Eq. (20)). Therefore, the 2017A-T451 alloy and FSW with Al alloy samples, which can be seen from the final dimensions
type I failure and the bell-shaped P-v curve require a larger of the cracked surface at the bottom edge, designated as final
amount of energy to fracture than the brittle type II failure and bottom width, BF. The average measured widths of the
with a small area under the P-v curve of 7075-T651 Al alloy. cracked surface at the bottom edge of the sample were found
FSW1, FSW2 and FSW3 have very similar fracture toughness to be in the following order: 10.23 (AA7075-T651) < 10.43
energy values of 87.1, 97.2 and 93.9 kJ/m2 respectively, which (FSWed 2017A/7075) < 10.68 mm (AA2017A-T451). These re-
are between the fracture toughness values of the BMs but sults are in good agreement with the results of the fracture
slightly closer to the 2017A-T451 alloy. toughness tests (Fig. 14 and Table 7).
As depicted in the cross-sectional macrograph of the
3.6. Fractographic analyses Charpy fractured sample (Fig. 15f) and the magnified view of
the EDM notch to fatigue crack (Fig. 15g) and from fatigue to
The fracture surfaces of the SEN(B) samples of 2017A-T451 and static fracture (Fig. 15h), the crack propagated through the
7075-T651 Al alloys, and FSWed dissimilar joints are depicted banded/stirred regions “A” and “B” in the NZ of the FSWed
in Fig. 15. The fatigue crack initiates at the vicinity of the notch joint. However, no obvious effect of larger and/or smaller
with several vertical marks on the notch front, which disap- grains on the crack propagation could be observed in this
pear after approximately 1 mm of crack extension (Fig. 15a material direction (i.e. T-S direction). To investigate the frac-
and cee). This phenomenon is evident in all samples except tured surfaces in more detail, additional SEM analyses were
for 2017A-T451 (Fig. 15b). carried out in the L-S material direction at the boundary be-
The likely reason for the difference in 2017A-T451 Al alloy tween the EDM notch and the surface with the fatigue cracked
are comparatively larger grains of the 2xx Al grade. The vertical surface (region 1 - Fig. 16) and at the boundary between the
lines indicate the uneven fatigue surface at the beginning of surface with the fatigue cracked surface and the static frac-
the test, where the crack length was microstructurally small, ture surface (region 2 - Fig. 17).
and both crack surfaces were still in contact. About 1 mm Fig. 16 shows the SEM fracture surfaces in the L-S direction
below the crack origin, the crack surface became more ho- of the upper layer, i.e. at the boundary between the EDM notch
mogeneous as the crack opened under tensile-tensile loading and the fatigue cracked surface. The base Al alloys, i.e. 7075-
with a positive load ratio R ¼ 0.1. An obvious curvature of the T651 (Figs. 16a) and 2017A-T451 (Fig. 16b), depict similar frac-
crack front is visible due to the different residual stress state at ture features with the presence of multiple fatigue dimples,
the edges and core of the samples. The presence of compres- tear ridges, and cracks. Nonetheless, the fatigue fractured
sive stresses at the edges and/or tensile stresses in the core surface of Al alloy 7075-T651 indicates a more brittle fracture
leads to faster crack growth in the core of the sample [21]. nature compared to the base Al alloy 2017A-T451, consisting of
In our case, the curvature originates from the compressive expressive grooves due to the higher microstructural texture
transverse residual stresses at the surface, as seen from the X- with the preferential grain orientation (Fig. 6). The presence of
ray residual stresses measurements in in Fig. 5. The static highly oriented grains, intergranular constituents, and
2348 j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 2 3 3 3 e2 3 5 2
Fig. 15 e Charpy SEN(B) samples after fracture in the L-S (aee) and T-S (feh) directions, respectively. The marked yellow
squares in a), b) and d) indicate the regions of SEM analyses at the boundary between the EDM notch and fatigue-cracked
surface and at the boundary between fatigue-cracked surface and static fracture surface (presented in Figs. 16 and 17,
respectively).
j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 2 3 3 3 e2 3 5 2 2349
Fig. 16 e SEM fractographs at the boundary between the EDM notch and the fatigue-cracked surface; (a) 7075-T651; (b)
2017A-T451 and (c) FSWed sample with a magnified view of the marked region.
Fig. 17 e SEM fractographs at the boundary between the fatigue cracked surface and static fracture surface (a) 7075-T651; (b)
2017A-T451 and (c) FSW at 100£ (d) FSW sample at 500£ magnification.
2350 j o u r n a l o f m a t e r i a l s r e s e a r c h a n d t e c h n o l o g y 2 0 2 3 ; 2 5 : 2 3 3 3 e2 3 5 2
microcracks significantly induces stress concentration, thereby The resonant fatigue test frequency analyses confirmed
increasing local plastic deformation, and causing inter- the higher durability of the FSWed dissimilar samples, with
granular fracture (IGF) [12]. The effect of the IGF along the a longer fatigue crack initiation phase than the BM 7075-
elongated grains is clearly evident from the magnified view of T651 and 2017A-T451 Al alloys.
the marked region in Fig. 16a. In contrast to the BMs, the FSWed HCF tests confirmed the superior endurance of the FSWed
SEM fractography (Fig. 16c) indicates a smoother surface with sample, with a fatigue limit of 198 MPa. In contrast, both
non-dendritic equiaxed grains, consisting of larger (stir region BMs showed no fatigue limit.
“A”) and smaller grains (stir region “B”). FSWed samples also showed significantly better fatigue
More importantly, magnified SEM/BE images of the FSWed performance in the stress range of 120e125 MPa due to the
fractured surfaces in the L-S direction (marked region on synergistic effect of expressive grain refinement and
Fig. 16c; right image) clearly confirm the effect of the grain size compressive residual stress in the T direction.
within the NZ on the direction of crack propagation. It is The BM 7075-T651 exhibited a brittle type II failure with low
noteworthy and should be emphasized that within the NZ re- average fracture toughness. Conversely, both the FSWed
gion, the crack has turned, changed its direction and bypassed dissimilar samples and the BM 2017A-T451 alloy displayed
the stir region “A” consisting of larger crystal grains. Although a ductile transition with type I failure, resulting in higher
the results of this study confirm the superior high-cycle fatigue average fracture toughness values.
resistance of the dissimilar 2017e7075 FSWed sample SEM fractographic analysis confirmed the more brittle
compared to the BMs, some accordance with the previous nature of BM 7075-T651 compared to 2017A-T451.
studies [4,51] exists, since the crack propagates more easily Fractography of the FSWed sample revealed a
along the smaller grains within the “banded” NZ region. smoother surface with non-dendritic equiaxed grain
Nonetheless, similar to the BMs, the predominant fracture morphology.
mechanism within the FSW sample is the IGF, where the cracks The fractographic analyses confirm that the microstruc-
propagate along the grain boundary between two adjacent ture strongly influences the direction and propagation of
crystal grains. the fatigue crack in the FSWed sample. Specifically, the
To further investigate the fracture mechanism at the crack propagated through region “B” with smaller crystal
boundary between the fatigue cracked and static fractured grains and completely bypassed the stir region “A”, which
surface, additional SEM analyses in the L-S direction were consists of larger crystal grains.
performed (Fig. 17). An obvious difference can be observed SEM analyses of the static fracture surfaces again revealed
when comparing the fractographs of the BMs 7075-T651 and a ductile fracture mechanism in 2017A-T451 and the
2017A-T451 with the FSWed samples. As seen from the fracture FSWed sample and a brittle fracture in BM 7075-T651 with
toughness results (Fig. 14), the most inferior condition can be profound racks and tear ridges.
observed with BM 7075-T651, which displayed brittle fracture,
evident cracks, and tear ridges (Fig. 17a). On the other hand, the
BM 2017A-T451 (Fig. 17b) and FSWed samples (Fig. 17c and d)
exhibited a typical ductile fracture mechanism with dimple Data availability statement
rupture due to micro-void coalescence [21]. Additionally, the
FSWed sample had smaller dimples and larger intermetallic The data that support the findings of this study are available
phases (Fig. 17d), which were attributed to the dynamic from the corresponding authors upon reasonable request.
recrystallization, secondary precipitation, and redistribution of
intermetallic phases during/after the FSW process [17,51].
CRediT authorship contribution statement
and technology (IBCAST). Islamabad, Pakistan: IEEE; 2015. [39] Sla mova M, Oc
ena
sek V, Vander Voort G. Polarized light
p. 6e13. https://doi.org/10.1109/IBCAST.2015.7058471. microscopy: utilization in the investigation of the
[27] Malopheyev S, Vysotskiy I, Kulitskiy V, Mironov S, recrystallization of aluminum alloys. Mater Char
Kaibyshev R. Optimization of processing-microstructure- 2004;52:165e77. https://doi.org/10.1016/
properties relationship in friction-stir welded 6061-T6 j.matchar.2003.10.010.
aluminum alloy. Materials Science and Engineering: A. [40] Schmidt HNB, Dickerson TL, Hattel JH. Material flow in butt
2016;662:136e43. https://doi.org/10.1016/j.msea.2016.03.063. friction stir welds in AA2024-T3. Acta Mater
[28] Sano Y, Masaki K, Gushi T, Sano T. Improvement in fatigue 2006;54:1199e209. https://doi.org/10.1016/
performance of friction stir welded A6061-T6 j.actamat.2005.10.052.
aluminum alloy by laser peening without coating. 1980-2015 [41] Chen Y, Zhang R, He C, Liu F, Yang K, Wang C, Wang Q, Liu Y.
Mater Des 2012;36:809e14. https://doi.org/10.1016/ Effect of texture and banded structure on the crack initiation
j.matdes.2011.10.053. mechanism of a friction stir welded magnesium alloy joint in
[29] Abdulstaar MA, Al-Fadhalah KJ, Wagner L. Microstructural very high cycle fatigue regime. Int J Fatig 2020;136:105617.
variation through weld thickness and mechanical properties https://doi.org/10.1016/j.ijfatigue.2020.105617.
of peened friction stir welded 6061 aluminum alloy joints. [42] Khan NZ, Siddiquee AN, Khan ZA, Mukhopadhyay AK.
Mater Char 2017;126:64e73. https://doi.org/10.1016/ Mechanical and microstructural behavior of friction stir
j.matchar.2017.02.011. welded similar and dissimilar sheets of AA2219 and AA7475
[30] Trdan U, Skarba M, Porro JA, Ocan ~ a JL, Grum J. Application of aluminium alloys. J Alloys Compd 2017;695:2902e8. https://
massive laser shock processing for improvement doi.org/10.1016/j.jallcom.2016.11.389.
of mechanical and tribological properties. Surf Coating [43] Peyre P, Berthe L, Vignal V, Popa I, Baudin T. Analysis of laser
Technol 2018;342:1e11. https://doi.org/10.1016/ shock waves and resulting surface deformations in an
j.surfcoat.2018.02.084. AleCueLi aluminum alloy. J Phys D Appl Phys
[31]
Ravnikar D, Trdan U, Nagode A, Sturm R. Energy density 2012;45:335304. https://doi.org/10.1088/0022-3727/45/33/
effect of laser alloyed TiB2/TiC/Al composite coatings on 335304.
LMZ/HAZ, mechanical and corrosion properties. Metals [44] Peterson RE. Notch sensitivity. In: Sines G, Waisman JL,
2020;10:411. https://doi.org/10.3390/met10030411. editors. Metal fatigue. New York: MacGraw Hill; 1959.
[32] Trdan U, Skarba M, Grum J. Laser shock peening effect on the [45] Wang Z-J, Li Q-J, Cui Y-N, Liu Z-L, Ma E, Li J, Sun J, Zhuang Z,
dislocation transitions and grain refinement of AleMgeSi Dao M, Shan Z-W, Suresh S. Cyclic deformation leads to
alloy. Mater Char 2014;97:57e68. https://doi.org/10.1016/ defect healing and strengthening of small-volume metal
j.matchar.2014.08.020. crystals. Proc Natl Acad Sci USA 2015;112:13502e7. https://
[33] Klubberg F, Klopfer I, Broeckmann C, Berchtold R, Beiss P. In: doi.org/10.1073/pnas.1518200112.
Fatigue testing of materials and components under mean [46] Li J, Shan Z, Ma E. Elastic strain engineering for
load conditions; 2011. p. 419e24. unprecedented materials properties. MRS Bull
[34] Kumar N, Mishra RS, Baumann JA. Residual stresses in 2014;39:108e14. https://doi.org/10.1557/mrs.2014.3.
friction stir welding. Amsterdam: Elsevier/Butterworth- [47] Leonetti D, Maljaars J, Bert HH(, Snijder. Fitting fatigue test
Heinemann; 2014. data with a novel S-N curve using frequentist and Bayesian
[35] Kawashima T, Sano T, Hirose A, Tsutsumi S, Masaki K, inference. Int J Fatig 2017;105:128e43. https://doi.org/10.1016/
Arakawa K, Hori H. Femtosecond laser peening of friction stir j.ijfatigue.2017.08.024.
welded 7075-T73 aluminum alloys. J Mater Process Technol [48] ASTM E1820-20b. Standard test method for measurement of
2018;262:111e22. https://doi.org/10.1016/ fracture toughness. West Conshohocken, PA: ASTM
j.jmatprotec.2018.06.022. International; 2020. https://doi.org/10.1520/E1820-20B.
[36] Threadgill PL, Leonard AJ, Shercliff HR, Withers PJ. Friction [49] ASTM E399-08. Test method for linear-elastic plane-strain
stir welding of aluminium alloys. Int Mater Rev fracture toughness KIc of metallic materials. West
2009;54:49e93. https://doi.org/10.1179/174328009X411136. Conshohocken, PA: ASTM International; 2008. https://
[37] Campanelli S, Casalino G, Casavola C, Moramarco V. doi.org/10.1520/E0399-12E03.
Analysis and comparison of friction stir welding and laser _
[50] Iriç S, Ayhan AO. Dependence of fracture toughness on
assisted friction stir welding of aluminum alloy. Materials rolling direction in aluminium 7075 alloys. Acta Phys Pol, A
2013;6:5923e41. https://doi.org/10.3390/ma6125923. 2017;132:892e5. https://doi.org/10.12693/APhysPolA.132.892.
[38] ASTM. Standard test methods for determining average grain [51] Mishra RS, Ma ZY. Friction stir welding and processing.
size. E112-10. West Conshohocken, PA: ASTM International; Mater Sci Eng R Rep 2005;50:1e78. https://doi.org/10.1016/
2010. https://www.astm.org/e0112-13r21.html. j.mser.2005.07.001.