You are on page 1of 14

Heliyon 10 (2024) e23835

Contents lists available at ScienceDirect

Heliyon
journal homepage: www.cell.com/heliyon

Nanomechanical, mechanical and microstructural


characterization of electron beam welded Al2219-T6 tempered
aerospace grade alloy: A comprehensive study
Ghulam Hussain a, *, Tauheed Shehbaz b, Mohammed Alkahtani c,
Usman Abdul Khaliq d, Hongyu Wei e
a
Mechanical Engineering Department, College of Engineering, University of Bahrain, Isa Town, 32038, Bahrain
b
Faculty of Materials and Chemical Engineering, Department of Materials Science, Ghulam Ishaq Khan Institute of Engineering Sciences and
Technology, 23640, Topi, Pakistan
c
Department of Industrial Engineering, College of Engineering, King Saud University, P.O. Box 800, Riyadh, 11421, Saudi Arabia
d
Department of Mechanical Engineering, University Malaya, Malaysia
e
College of Mechanical & Electrical Engineering, Nanjing University of Aeronautics & Astronautics, Nanjing, 210016, PR China

A R T I C L E I N F O A B S T R A C T

Keywords: As compared to traditional fusion welding processes, electron beam welding (EBW) is known to
EBW produce structurally robust microstructures and narrow heat-affected zone (HAZ) in metals. The
SEM process becomes more significant for the tempered alloys vulnerable to heat exposure. In the
Nanoindentation
present investigation, Al 2219-T6 alloy was joined using the EBW process. The microstructural,
Pile-ups
Mechanical properties
mechanical, and nanomechanical characteristics of the resulting joint were investigated. EBW
resulted in a narrow HAZ (22 μm) with a 430 mm fusion zone (FZ). A dendritic structure was
observed in the FZ zone, while second-phase particles were absent indicating their dissolution
during welding and interesting formation of Al2Cu mixture around the dendrites. The limited
content of Cu in the base metal (BM) resulted in the formation of a solid solution in the FZ, along
with the presence of fine equiaxed grains in the HAZ and equiaxed dendritic grains in the FZ zone.
The X-ray diffraction analysis confirmed the absence of peaks corresponding to incoherent phases
in the FZ. Compared to the BM, micro-hardness measurements revealed a 12.7 % increase in the
hardness in the HAZ, while a significant decrease of approximately 19 % was observed in the FZ.
The joint exhibited reduced tensile strength, ultimate strength by 42.2 %, and yield strength by
47.3 % when compared to the BM. The fracture analysis indicated a ductile failure mode with the
presence of microvoids. Nano-indentation tests at various loads demonstrated a decrease in the
nanohardness from the BM to the HAZ and FZ regions. Atomic force microscopy (AFM) analysis
revealed significant pile-ups in the FZ, indicating the occurrence of plastic deformation during the
welding process. The presented findings are valuable for the joint and structure design of Al
− 2219T6 alloy in particular and other Al alloys in general.

* Corresponding author.
E-mail address: ghussain@uob.edu.bh (G. Hussain).

https://doi.org/10.1016/j.heliyon.2023.e23835
Received 12 July 2023; Received in revised form 19 October 2023; Accepted 13 December 2023
Available online 16 December 2023
2405-8440/© 2023 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
G. Hussain et al. Heliyon 10 (2024) e23835

1. Introduction

Aluminium is widely employed in various industries due to its favorable chemical, physical, and mechanical characteristics,
rendering it the second most utilized metal [1,2]. Aluminium alloys find widespread use in the construction of chemical plants and
appliances used in the food industry. Moreover, certain aluminum-based compositions are utilized in aircraft construction, whereas
others are employed to produce automobile components and lightweight structures [3]. Yet some kinds are utilized as a middle layer to
promote mechanical bonding between incompatible materials [4]. The utilization of aluminum alloys in the automotive industry offers
a significant advantage in terms of reducing weight and pollution. By substituting 1 kg of aluminum for 2 kg of steel, a total of 10 kg of
CO2 equivalents can be reduced over the average lifespan of a vehicle [5,6].
The Al–Cu alloy known as Al 2219 has been extensively utilized in the automotive, aerospace, and defense industries due to its
exceptional mechanical properties. These properties include a high strength-to-weight ratio, resistance to fatigue, and thermal stability
[7]. Applications of this alloy include cryogenic fuel tanks, rocket casings, and aircraft structural components [8]. Welding similar
aluminum alloys is crucial in industries for achieving strong, lightweight, and durable joints, enhancing the performance and longevity
of various applications while maintaining the material’s inherent properties. However, the welding of Al 2219 remains challenging due
to its susceptibility to hot cracking, porosity, and other weld-related defects [9–12].
Numerous researchers have utilized a variety of welding methods to join Al 2219 alloy. Conventional fusion welding techniques,
such as gas metal arc welding (GMAW), gas tungsten arc welding (GTAW), and plasma arc welding (PAW), are widely used in joining
aluminum alloys due to their versatility and ease of implementation. However, one significant drawback associated with these
techniques is the formation of a large HAZ, owing to the high heat inputs, thermal conductivity, low melting point phases, and thermal
expansion mismatch of aluminum alloys [13,14]. Moreover, cracks during solidification, porosity in the weld zone, and liquation
cracking are among the specific types of defects commonly observed in the fusion welding process of aluminum alloys. Li. Hui et al.
[15] conducted experiments using variable polarity TIG (VPTIG) and activated flux TIG (A-TIG) techniques on 2219 alloy and observed
the occurrence of both micro and macro porosity in the VPTIG process, while the A-TIG process proved to be effective in eliminating
porosity. In addition, the strength of the weldment in this study was measured to be 276.4 MPa. Zhang et al. [16] observed the ex­
istence of cracks within the weld zone of the 2219 TIG welded joint. The formation of these cracks was attributed to the existence of
coarsened eutectic particles and an increased thickness of the grain boundary eutectic structure Chen et al. [17] conducted pulsed-gas
metal arc welding (P-GMAW) on Al–Si (4047) alloy. A notable presence of porosity in the FZ was observed as a result of air entrapment
and alloy shrinkage during the solidification process. Wang et al. [18] employed the variable polarity vertical plasma arc welding
(VPPAW) technique to weld Al 2219 and documented the occurrence of hot cracks, owing to the high heat input during the welding
process, which resulted in the melting of low melting point phases situated at the boundaries of the crystalline structure. Consequently,
these molten phases prolonged the liquid state of the crystal boundaries, leading to the formation of hot cracks.
Electron beam welding (EBW) has emerged as a promising technique for welding Al 2219, owing to its promising features such as
low heat input, high energy density and precise control of the welding process [19]. These features help mitigate many issues asso­
ciated with conventional fusion welding methods, such as hot cracking and loss of alloying elements [20]. Fujii et al. [21] described the
pore formation process when welding 2219 aluminium alloy with an electron beam, reporting that the pores produced by EBW are
fewer than those produced by TIG welding. Rao et al. [22] investigated the EBW of the 2219-T87 alloy and compared the results to TIG
welding. The EBWed joint was found to be 42 % stronger than the TIG welded joint. The increased strength was attributed to the
matrix’s homogenous distribution of Cu, as opposed to the TIG welded joint, where Cu segregated at the grain boundaries Hayat et al.
[23] conducted electron beam welding (EBW) on 7075-T6 alloy using current settings of 20mA and 25mA. They reported a
comparatively higher hardness value of 175HV in the FZ zone and a tensile strength of 256 MPa, when employing a current of 25mA,
which can be ascribed to the formation of brittle intermetallic compounds (IMCs) within the FZ region of the weldment.
To mitigate the distortion caused by post-weld T6 treatment, the welding is preferred in T6 condition. Welding Al 2219 alloy in the
T6 condition poses several challenges and complexities. One of the primary difficulties arises from strengthening precipitates, such as
the metastable phase η′(eta-prime), which can be susceptible to over-aging during welding. This over-aging can lead to a reduction in
the alloy’s mechanical properties and toughness in the heat-affected zone (HAZ) adjacent to the weld [24,25]. Additionally, the T6
condition typically involves the formation of a dispersion of fine, coherent precipitates throughout the aluminum matrix, which can
affect the weldability and heat transfer characteristics of the alloy. Controlling the welding process parameters, including heat input,
cooling rates, and inter-pass temperature, becomes crucial to minimize the potential for over-aging and preserving the desired T6
properties. These difficulties underscore the importance of careful process planning, advanced welding techniques, and a thorough
understanding of the alloy’s metallurgical behavior to achieve successful welding in 2219 aluminum alloy in the T6 condition.
The performance of welded joints is primarily influenced by the solidified microstructures that emerge from the complex heat
transfer processes occurring during the solidification of the molten pool. To achieve welded joints with outstanding mechanical
properties, it is crucial to gain a deeper comprehension of the microstructure evolution mechanisms in electron beam welding (EBW).
Nanoindentation, a technique used for nanoscale characterization, has emerged as a powerful tool for characterizing the local me­
chanical properties of weldments [26]. This method is particularly useful for studying the microstructural changes effects induced by
welding processes, such as grain refinement, precipitation, and phase transformations known to significantly influence the overall
mechanical behavior of the welded joints [27,28].
From the above presented literature, it can be identified that EBW can offer several merits as compared to the competing traditional
fusion welding processes, such as reduced defects, improved strength, and narrow HAZ. Moreover, the pre-welding condition of
material affects the resulting joint characteristics, and these should be studied at the nano-scale besides at the micro-scale for a
comprehensive understanding of the mechanical behavior of the joint. As literature lacks on EBW of 2219 alloy in T6 condition,

2
G. Hussain et al. Heliyon 10 (2024) e23835

research work is needed to quantify the joint performance for structural applications. The AA 2219-T6 alloy is successfully joined
through the EBW process, and then the joint is subjected to several tests to reveal the nano-mechanical behavior, microstructure
evolution, and mechanical properties. The mechanical properties that the current article will particularly detail include tensile
properties, microhardness, and elastic modulus of weld joint.

2. Materials and Methods

The 2219-T6 alloy comprises the following chemical composition (by weight) measured by EDS attached with a scanning electron
microscope: 4.2 % Cu, 0.3 % Fe, 0.2 % Si, 0.3 % Mn, 0.1 % V, 0.1 % Zn, 0.15 % Zr, and trace amounts of Mg and Ti, with the remaining
balance being aluminum. The alloy sheets, with dimensions of 200 × 100 × 2 mm, underwent ultrasonic cleaning using acetone.
Subsequently, the sheets were clamped against a flat surface and butt welded along their longest dimension. The welding process was
carried out with an electron beam (EB) welding equipment (Model THDW-3C, Jiangsu, China) as shown in Fig. 1. The optimized
process parameters, as determined through preliminary trials, for achieving a weld bead without defects were an accelerating voltage
of 60 kV, beam current of 19 mA, Beam diameter of 0.3 mm, and a sample travel speed of 1300 mm/min.

2.1. Microstructural characterization

The metallographic specimens were obtained by sectioning the welded plate using electric discharge machining (EDM). The
metallographic samples underwent grinding with SiC grit papers and polishing with 3–0.25 μm diamond paste. Keller’s reagent was
used to perform the etching procedure. The sample was examined using an Olympus BH2-UMA polarizing microscope. The micro­
structures and chemical composition of the weldment were examined and analyzed using a scanning electron microscope (SEM)
equipped with an energy-dispersive X-ray spectroscopy (EDS) attachment. X-ray diffractometry was employed to analyze the phases
within the joint. The measurements were conducted using a Proto X-ray diffractometer from Canada, operating at 25 mA and 40 kV,
with a Cu K-alpha target.

2.2. Mechanical characterization

Tensile test specimens (03) were fabricated using wire EDM following the guidelines outlined in ASTM standard E8M-0446. Tensile
tests were conducted on the samples using a 30 kN load Instron machine at a constant strain rate of 2 mm/min, and the average results
were reported. The microhardness profile of the weld zone was measured using a Tukon Model 300 Vickers hardness tester. The test
subject experienced a 200-gf load for a duration of 15 s.
The experiment involved the utilization of a nanoindenter (iMicro, Nanomechanics, USA) to conduct nanoindentation at ambient
temperature. During the nanoindentation test, using a three-sided Berkovich tip and a load control mode with a maximum load of 100,
150, and 200 mN, the hardness and elastic modulus at various locations of the joint were measured. A 2 × 2 and 3 × 3 grid was utilized
to determine each zone of the weldment. To prevent inter-indentation impacts, a separation of 40 μm was implemented between the
indents. The indented surface was imaged using Atomic Force Microscopy (AFM) with a Nanosurf C3000 instrument. The acquired
images were obtained with a resolution of 256 × 256 pixels. The scan area covered was 27 mm × 27 mm, and the scan rate was 0.5 Hz.

Fig. 1. The schematic of EBW process applied for welding of Al2219-T6 alloy.

3
G. Hussain et al. Heliyon 10 (2024) e23835

3. Results and discussion

3.1. Physical appearance

Fig. 2(a) depicts the visual test (VT test) conducted on the upper and bottom surface of the weld joint. The absence of contaminants
in the weldment suggests that the vacuum conditions have been maintained at a stable level and the welding parameters have been
optimized. The welded joint exhibits complete penetration, as evidenced by the measurements of 4.3 mm at the upper portion (arc
facing side), 3.60 mm at the middle, and 2.35 mm at the lower side of the weldment, as depicted in Fig. 2(b). The reason for the wider
width of the welding zone at the top and narrower at the bottom is primarily due to the effects of gravity during the welding process.
The molten metal accumulates more at the bottom of the joint, leading to a narrower width of the welding zone. On the other hand, the
top side of the joint receives less molten metal due to gravitational forces, resulting in a wider width of the welding zone at the top [29].

3.2. Microstructure of the weldment

Fig. 3(a) depicts the grain structure of the parent material. It is evident that the microstructure of BM is characterized by longi­
tudinally elongated and equiaxed grain as a result of the deformation caused during rolling. Furthermore, the grain and subgrain
structures exhibited a substantial presence of finely dispersed second-phase particles, specifically Al2Cu and S-phase (Al2CuMg), as
shown in Fig. 3(b). These precipitates are believed to be responsible for the observed age hardening, which is attributed to the for­
mation of Guinier-Preston zones (GPZ) within the grains. Similar observations are reported by other researchers [30–32]. The for­
mation of GP zones originates due to T6 treatment, which is responsible for the age-hardening process. These precipitates (Al2Cu and
S-phase) are associated with age hardening developed from the Guinier-Preston zone (GPZ) within the grain.
Moreover, Fig. 4(a) depicts the lower temperature region close to the unaffected base metal during over-aging. The “lower tem­
perature region” refers to the part of the microstructure that is relatively cooler or experiences lower temperatures compared to other
regions during the welding process. This temperature was favorable for the over aging of the region highlighted. It signifies a specific
temperature range or zone within the material where distinct microstructural changes, such as precipitation, occur as a result of the
aging treatment. Moreover, this region demonstrates that the precipitates grew to a size where they no longer maintained coherency
with the matrix. During over aging the precipitates lost their coherency from the matrix, and the Al2Cu precipitates transformed into
theta phase. These phases cannot be observed at lower magnification. However, grain coarsening is indicative of the over aging
phenomenon. Similar observations have been reported by various researchers [33,34].
Furthermore, it is important to note that the HAZ, depicted in Fig. 4(a), serves as a transitional area between the unaffected BM and
the FZ zone. Consequently, the microstructure of the HAZ differs from that of the typical Al 2219 observed near the BM region. Instead,
it resembles that of the solution-treated zone located adjacent to the FZ zone. The.
Highly uniform and well-defined HAZ is comprised of relatively small equiaxed grains with a width of approx. 22 μm. This region
has an average grain size of 5 μm. As shown in Fig. 4(a), the FZ zone of the weldment is coarse dendrites and equiaxed grains. The

Fig. 2. (a) The EBW joint of Al-2219T6 alloy: (a) complete joint, and (b) macroscopic cross-sectional view of the joint.

4
G. Hussain et al. Heliyon 10 (2024) e23835

Fig. 3. Optical micrographs of Al2219-T6 base metal at different magnifications.

Fig. 4. Optical micrographs showing: (a) narrow HAZ with equiaxed grain morphology and (b) equiaxed and dendrites in FZ of welded joint.

formation of equiaxed grains can be attributed to the rate of solidification and welding speed [35]. The distribution of Marangoni
convection in the top half of the molten pool puts the heterogeneous point towards the fusing line, where the metal nucleates and
increases as the temperature declines. Owing to the higher temperature differential from the fusion line to the weld centre, the crystal
grains develop towards the weld centre, generating columnar dendrites [36].
The SEM micrograph of the overall microstructure of the FZ zone is presented in Fig. 5(b), which exhibited the equiaxed dendritic
morphology compared to Fig. 5(a). Additionally, it is observed that the grain boundaries in the fusion zone (FZ) of the weldment
exhibit an increase in thickness. This suggests that the precipitates present in the FZ are fully dissolved at elevated temperatures.
However, upon cooling, some of these precipitates re-precipitate and accumulate along the grain boundaries. This phenomenon
counteracts the benefits of age hardening and adversely affects the mechanical properties of the material.
The EDS analysis at location 1 (interdendritic spacing), as mentioned in Fig. 6, showed the presence of 9.56 % Cu in Al. The Cu
contents more than 5.8 % (maximum solubility of Cu in Al) lead to the formation of eutectic mixture of Al and Al2Cu, according to the
binary phase diagram of Cu and Al [37–40]. Furthermore, the presence of 94.86 % Al and 5.14 % Cu at location 2 clearly indicated the
redistribution of Al and Cu in the form of a solid solution, owing to the solubility of Cu in the Al matrix. A significant decrease in the
proportion of the second phase particle can be observed in comparison to BM, which can be responsible for the reduction of joint

5
G. Hussain et al. Heliyon 10 (2024) e23835

Fig. 5. SEM micrographs showing: (a) BM, HAZ, and FZ, and (b) magnified image of FZ.

Fig. 6. EDS analysis of FZ of welded joint: (a) dendrite (location1), and (b) inter-dendrite (location 2).

strength.

3.3. X-ray diffraction

The X-ray diffraction (XRD) pattern observed on the BM and FZ zone of the weldment is shown in Fig. 7. In base alloy (BM), distinct
intensities corresponding to distinct α phase matrix are produced by constructive interference on various crystallographic planes. The
peaks appeared at 2θ = 38◦ , 46.5◦ , 66.5◦ , and 78.5◦ , corresponding to diffraction on (111), (200), (220), and (311) plane, respectively.
In addition, at 2θ = 47.5◦ , a non-equilibrium coherent θ’ precipitates correspond to (202) is identified, which clearly indicates the age
hardening treatment of BM.
In welded joint, the intensity of majority of the peaks corresponding to the α phase is deceased. However, the peak corresponding to
the θ’ is diminished in the FZ zone, which can be attributed to the dissolution of precipitates in the FZ zone, owing to high heat input
during welding [41].
The peak shift can be observed in the XRD spectra of FZ and BM, which can be attributed to the thermal cycling that the FZ zone
experiences during welding process. These thermal cycles caused lattice distortions, which, in turn, can result in peak shifts. These
distortions are primarily due to the non-uniform temperature distribution across the FZ during welding [42].

6
G. Hussain et al. Heliyon 10 (2024) e23835

Fig. 7. Pattern of X-ray diffraction obtained on the FZ and BM of the specimen.

3.4. Microhardness

Fig. 8 illustrates the profile of microhardness variations along the transverse cross-sections of the weldment. The BM metal (BM)
exhibits an approximate average hardness value of 73.7 HV0.1. However, in the heat-affected zone (HAZ), a significant increase in
hardness of approximately 12.8 % is observed. This increase can be attributed to multiple factors, including grain refinement, high
density dislocations, and the enhanced solid solution strengthening effect [43]. These phenomena contribute to the strengthening of
the material in HAZ.
On the other hand, a sudden decrease in the hardness values to 59.7 HV0.1 is observed in the fusion zone (FZ) of the weldment. This
decrease can be attributed to the dissolution of incoherent second phase particles that occur during the welding process [44]. These
particles, present in the BM, tend to dissolve when subjected to the high temperatures experienced during welding, resulting in a
reduction in hardness in the FZ.
It is widely documented that the hardness of a precipitation-hardened aluminum alloy joint is synthetically regulated by various
factors, including precipitate dispersion, dislocation density, grain size, and solid solution [45–47]. These factors are influenced by the
welding process in the present study, thereby affecting the overall hardness distribution within the weldment.

3.5. Tensile properties

The stress-strain curves depicted in Fig. 9 illustrate the mechanical behavior of both the BM and the weldment during a tensile test.
Notably, the yield stage of both the BM and the welded joint is clearly evident throughout the entire testing process. The yield strength
of the BM was 380 MPa, and the tensile strength was 490 MPa. However, a substantial reduction in both tensile strength (42.2 %) and
yield strength (47.3 %) is observed in the joint compared to the base alloy. This decline in strength can be attributed to several factors

Fig. 8. Microhardness profile across the welded joint.

7
G. Hussain et al. Heliyon 10 (2024) e23835

Fig. 9. Tensile stress-strain curves of BM and welded specimen.

associated with the welding process.


One contributing factor is the dissolution of second phase particles within the weld joint. These particles, present in the base alloy,
play a crucial role in strengthening the material. During welding, the elevated temperatures cause the dissolution of these particles,
thereby reducing the overall strength of the joint [48]. Coarse grained structure of the FZ zone, as a result of rapid cooling during the
welding process, further contributed to the decrease in strength of the joint. Moreover, rapid cooling during welding induces high
thermal gradients, leading to the development of residual stresses within the joint. These residual stresses can act as internal forces that

Fig. 10. Fractographic images of the Al2219-T6 welded joint: (a) fractured surface, (b) magnified image as indicated in (a), (c) dimples and tear
ridges, and (d) EDS analysis of the specified area.

8
G. Hussain et al. Heliyon 10 (2024) e23835

weaken the material, causing a reduction in strength [49].


Furthermore, comparing the elongation properties of the weld joint to the BM, a substantial reduction of 68.7 % is evident. This
diminished elongation can be attributed to the broad width of the weld zone, resulting in a non-uniform strain distribution during
deformation, affecting the overall elongation of the joint. Additionally, the uneven grain morphology within the weldment, along with

Fig. 11. p-h curves and corresponding hardness and elastic modulus of various zones at 100 mN,150 mN and 200 mN.

9
G. Hussain et al. Heliyon 10 (2024) e23835

the relatively coarser grain size compared to the BM, might have contributed to the reduced elongation properties. It is worth
mentioning here that the fracture occurred in HAZ/FZ interface, which can be attributed to the fine-grained structure and dislocation
pile-up [50].

3.6. Fractography

The fracture surfaces of the tested tensile specimen were examined with a scanning electron microscope to comprehend the fracture
mode and characterize the fracture features. The micrographs were taken from the uppermost region of the fractured surface.
Fig. 10(a) shows the fractured surface of 2219-T6 welded joint. The fractured surface of electron beam welded 2219-T6 exhibited
typical behavior associated with ductile fracture. Upon examination, the fracture surface displayed distinct features such as dimples,
second phase particles, and tear ridges, as shown in Fig. 10(b) and (c).
Dimples are rounded or elongated depressions on the fracture surface, indicating plastic deformation and energy absorption during
fracture. At higher magnification, dimples of various sizes were found, as highlighted in Fig. 10(c); fine equiaxed dimples and elon­
gated dimples. The presence of defects or stress concentrations areas might cause the strain localization, leading to the formation of
fine dimples. However, the formation of elongated dimples can result from the strain along specific directions, reflecting the aniso­
tropic deformation behavior of the alloy and elongated [51]. The presence of these dimples signifies the ability of the material to
undergo significant plastic deformation before failure [52]. Additionally, the fracture surface reveals the presence of second phase
particles, which are dispersed throughout the microstructure of the alloy. These particles act as obstacles to crack propagation,
contributing to the material’s resistance against fracture. Furthermore, tear ridges are observed along the fracture surface, repre­
senting the tearing and separation of the material. The combination of dimples, second phase particles, and tear ridges on the fracture
surface indicates a ductile fracture behavior in the electron beam welded 2219-T6 alloy.
Moreover, EDS analysis carried out at the location marked in Fig. 10(d) indicates the presence of Al matrix and Al2Cu as a eutectic
mixture, in which the Al matrix caused more plastic deformation, resulting in significant elongation of the fractured specimen.

3.7. Nanoindentation

Nanoindentation tests were performed on different zones, including BM, HAZ, and FZ, formed during the welding process. The
load-displacement relationship was analyzed using the approach proposed by Oliver and Pharr [53], which considers the elastic re­
covery of the surface after unloading. Fig. 11 (a, c, d) presents the load vs. displacement plots of the various zones (BM, HAZ, and FZ) at
100 mN, 150 mN, and 200 mN.
The deformation caused by indentation comprises both elastic-plastic loading and pure elastic unloading curves, as indicated by
arrows in Fig. 11(a). It can be observed that the BM at a load of 100 mN exhibits the lowest depth of penetration, which increases to
2282 nm and 2471 nm in the HAZ and FZ zones, respectively. This suggests that the BM at 100 mN undergoes the least plastic
deformation. Fig. 11(b) displays the corresponding hardness and elastic modulus at a load of 100 mN. It shows that the BM has the
highest hardness (1.65 GPa), which decreases to 3.47 % in the HAZ and 17.39 % in the FZ zone.
Moreover, the elastic modulus of the different zones ranges from 66.5 GPa to 71.5 GPa, which is in agreement with the typical
modulus of aluminum alloy (68–71 GPa) [54]. At 150 mN in Fig. 11(c), the penetration depth increases by 7.43 % and 24.18 %, while
the hardness decreases by 16.36 % and 37.57 % in the HAZ and FZ zones, respectively. Furthermore, at a load of 200 mN, all three
zones exhibit similar trends of variation in hardness and elastic modulus. The reduction in hardness from the BM to the HAZ and FZ can
be attributed to the dissolution of second phase particles and precipitates during the welding process. The variations in depth of
penetration, hardness, and elastic modulus indicate the differing levels of plastic deformation and mechanical response among the
zones. The dissolution of second phase particles and precipitates during welding contributes to the observed reduction in hardness.
It can be observed from Figs. 8 and 11 that the nanohardness of BM is lower, whereas the same region showed the highest
microhardness. This can be associated with the welding process, where the rapid heating and cooling in the HAZ can lead to the
formation of smaller and less organized grains compared to the Base Metal. These changes in grain structure can result in lower
nanohardness as nanoindentation is highly sensitive to grain boundaries and defects. The redistribution of alloying elements can lead
to the formation of softer phases or precipitates in the HAZ, further reducing the nanohardness. Microhardness measurements are less
sensitive to fine-scale microstructural changes. The indentation load applied during microhardness testing is significantly larger than
that used in nanoindentation, which can average out small variations. The presence of smaller and more disorganized grains in the HAZ
can lead to an increase in dislocation density, which can contribute to higher microhardness values. Nanohardness measurements excel
at detecting fine-scale changes in microstructure, resulting in lower values in the HAZ. In contrast, microhardness measurements,
which operate at a larger scale, are less sensitive to these subtle changes and therefore yield higher values, as supported by various
studies [55,56].

3.8. Atomic force microscopy

In the present study, atomic force microscope analysis was conducted to investigate the pile-up behavior in the BM, HAZ, and FZ of
the weldment. The atomic microscope results revealed distinct differences in the pile-up characteristics among these regions, as shown
Fig. 12(a), (b), and (c). The BM exhibited a relatively minimal pile-up, with a relatively uniform distribution of atoms along the grain
boundaries, indicating a limited pile-up effect. This observation can be attributed to the lower levels of plastic deformation experi­
enced by the BM of Al 2219-T6 alloy during the welding process [57]. Previous studies have also reported similar findings in terms of

10
G. Hussain et al. Heliyon 10 (2024) e23835

Fig. 12. AFM images showing 2D and side view of indentation impression on the Al2219-T6 weldment: (a) BM, (b) HAZ, and (c) FZ.

reduced pile-up in the BM due to its more stable atomic arrangement and lower dislocation density [58].
Moving on to the HAZ, a moderate pile-up effect was observed. The thermal cycles experienced by the HAZ during welding induced
localized changes in the atomic arrangement. This resulted in a slightly pronounced pile-up effect compared to the BM. The plastic
deformation induced by the heat input during welding leads to the rearrangement of atoms and the formation of localized dislocations.

Table 1
Comparison of tensile strength with previous studies.
Tensile strength (MPa) Elongation [Ref]

Ultrasonic TIG welding (Al 2219) 239 – [60]


P-GMAW (Al 6082) 150 2.9 [61]
TIG welding (Al 2219-T87) 229 4.7 [16]
TIG welded (Al 2024-O) 196 8.3 [62]
EBW (Al 2219-T6) 290 5.5 Present

11
G. Hussain et al. Heliyon 10 (2024) e23835

Several studies have reported similar observations, highlighting the role of thermal cycles in generating a moderate pile-up effect in the
HAZ [59]. In contrast, the FZ exhibited a more distinguished pile-up effect compared to both the BM and HAZ. The atomic microscope
images clearly showed significant pile-ups along the grain boundaries in the FZ, indicating a higher accumulation of atoms. This
enhanced pile-up effect can be attributed to the higher levels of plastic deformation experienced by the FZ during the welding process.
The more plastically deformed material in the FZ contributes to the increased pile-up phenomenon. Previous research on welded
specimens has reported similar findings, emphasizing the correlation between plastic deformation and the extent of pile-up in the FZ.
Table 1 compares the tensile strength of the EB welded Al2219-T6 with previous studies on various fusion welding techniques. Our
results revealed a significantly higher tensile strength of 290 MPa for the EBW joint of 2219-T6 alloy. This notable improvement in
tensile strength surpasses the values reported in previous investigations. Specifically, Ultrasonic TIG welding of Al 2219 yielded a
tensile strength of 239 MPa, P-GMAW on Al 6082 resulted in 150 MPa, TIG welding of Al 2219-T87 produced 229 MPa, and TIG welded
AA 2024-O exhibited 196 MPa. The findings demonstrate the superior performance of the EBW process in achieving enhanced tensile
strength in the 2219-T6 alloy. However, the reduction in elongation can be attributed to the residual stress produced during welding
owing to the rapid cooling. The controlled cooling during welding operation or post weld annealing can be helpful in minimizing the
residual stresses. These results highlight the potential of EBW as an effective method for producing high-strength fusion-welded joints
in aluminum alloys, particularly in the case of 2219-T6 with post weld heat treatment.

4. Conclusions

The joining of T6 tempered 2219 Al alloy poses numerous difficulties, thereby demanding advanced techniques. Therefore, in the
present study, EBW characterized by low current and narrow HAZ was applied to weld this alloy. In addition to micro scale char­
acterization, nano scale characterization was also applied to gain deeper insights into local mechanics of the weldment. The following
are the important findings:

1. The electron beam welding (EBW) of Al 2219-T6 alloy resulted in distinct microstructural features in different zones. The fusion
zone (FZ) exhibited a dendritic structure, while the heat-affected zone (HAZ) displayed a combination of equiaxed grains and fine
equiaxed grains.
2. The microhardness measurements revealed a 12.7 % increase in the hardness in the HAZ compared to the base metal (BM). In
contrast, the FZ demonstrated a significant decrease of approximately 19 % in hardness when compared to the BM, owing to the
dissolution of second phase particles during the welding process.
3. As a result of EBW, the joints experience a loss in tensile strength, namely 42.2 % in the ultimate strength and 47.3 % in the yield
strength as compared to the base alloy. This loss is caused by the dissolution of second phase particles.
4. Nano-indentation tests conducted at various loads demonstrated a decreasing trend in the nanohardness from the BM to the HAZ
and FZ regions. The HAZ exhibited a decrease of 3.47 % in nanohardness, while the FZ showed a larger decrease of 17.39 %
compared to the BM.
5. The presence of pile-ups in the FZ, as observed through atomic force microscopy (AFM) analysis, indicates a higher degree of plastic
deformation in this zone. This suggests that the FZ experienced more significant plastic deformation during the welding process
compared to the other regions.

Funding

This work was supported by Researchers Supporting Project Number (RSP2023R274), King Saud University, Riyadh, Saudi Arabia.

CRediT authorship contribution statement

Ghulam Hussain: Conceptualization, Resources, Supervision, Visualization, Writing – review & editing, Project administration.
Tauheed Shehbaz: Formal analysis, Investigation, Writing – original draft, Writing – review & editing, Methodology. Mohammed
Alkahtani: Funding acquisition, Project administration, Resources, Supervision, Writing – review & editing. Usman Abdul Khaliq:
Formal analysis, Methodology, Validation. Hongyu Wei: Conceptualization, Project administration, Supervision, Writing – review &
editing.

Declaration of competing interest

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

Acknowledgements

This work was supported by Researchers Supporting Project Number (RSP2023R274), King Saud University, Riyadh, Saudi Arabia.

12
G. Hussain et al. Heliyon 10 (2024) e23835

References

[1] D. Ashkenazi, How aluminum changed the world: a metallurgical revolution through technological and cultural perspectives, Technol. Forecast. Soc. Change
143 (March) (2019) 101–113, https://doi.org/10.1016/j.techfore.2019.03.011.
[2] F. Radkovský, et al., Determination of linear expansion of AlSi10Mg aluminium alloy depending on external conditions during solidification, Heliyon 8 (no. 11)
(2022), https://doi.org/10.1016/j.heliyon.2022.e11363.
[3] M. N. Ilman Tiwan, Kusmono, Microstructure and mechanical properties of friction stir spot welded AA5052-H112 aluminum alloy, Heliyon 7 (2) (2021),
e06009, https://doi.org/10.1016/j.heliyon.2021.e06009.
[4] Y. Fouad, Characterization of a high strength Al-alloy interlayer for mechanical bonding of Ti to AZ31 and associated tri-layered clad, Alex. Eng. J. 53 (2) (2014)
289–293, https://doi.org/10.1016/j.aej.2014.03.005.
[5] N. Ding, F. Gao, Z. Wang, J. Yang, Life cycle energy and greenhouse gas emissions of automobiles using aluminum in China, J. Ind. Ecol. 20 (4) (2016) 818–827,
https://doi.org/10.1111/jiec.12298.
[6] H.J. Kim, C. McMillan, G.A. Keoleian, S.J. Skerlos, Greenhouse gas emissions payback for lightweighted vehicles using aluminum and high-strength steel, J. Ind.
Ecol. 14 (6) (2010) 929–946, https://doi.org/10.1111/j.1530-9290.2010.00283.x.
[7] S. Pantelakis, Historical Development of Aeronautical Materials (2020), https://doi.org/10.1007/978-3-030-35346-9_1.
[8] X. Lei, Y. Deng, Z. Yin, G. Xu, Tungsten inert gas and friction stir welding characteristics of 4-mm-thick 2219-T87 plates at room temperature and -196 ◦ C,
J. Mater. Eng. Perform. 23 (6) (2014) 2149–2158, https://doi.org/10.1007/s11665-014-1013-9.
[9] P. Kah, R. Rajan, J. Martikainen, R. Suoranta, Investigation of weld defects in friction-stir welding and fusion welding of aluminium alloys, Int. J. Mech. Mater.
Eng. 10 (1) (2015), https://doi.org/10.1186/s40712-015-0053-8.
[10] Y. Kang, X. Zhan, T. Liu, Effect of welding parameters on porosity distribution of dual laser beam bilateral synchronous welding in 2219 aluminum alloy T-joint,
J. Adhes. Sci. Technol. 33 (23) (2019) 2595–2614, https://doi.org/10.1080/01694243.2019.1650991.
[11] J.Y. Bai, C.L. Yang, S.B. Lin, B.L. Dong, C.L. Fan, Mechanical properties of 2219-Al components produced by additive manufacturing with TIG, Int. J. Adv.
Manuf. Technol. 86 (1–4) (2016) 479–485, https://doi.org/10.1007/s00170-015-8168-x.
[12] D. Radi, M. Eldin A Abo-Elsoud, F. Khalifa, Segmenting welding flaws of non-horizontal shape, Alex. Eng. J. 60 (4) (2021) 4057–4065, https://doi.org/10.1016/
j.aej.2021.02.052.
[13] Z. Lu, J. Xu, L. Yu, H. Zhang, Y. Jiang, Studies on softening behavior and mechanism of heat-affected zone of spray formed 7055 aluminum alloy under TIG
welding, J. Mater. Res. Technol. 18 (2022) 1180–1190, https://doi.org/10.1016/j.jmrt.2022.03.074.
[14] H. Rojas, et al., The impact of heat input on the microstructures, fatigue behaviors, and stress lives of TIG-welded 6061-T6 alloy joints, Mater. Res. Express 7
(12) (2020), https://doi.org/10.1088/2053-1591/abd136.
[15] H. Li, J. Zou, J. Yao, H. Peng, The effect of TIG welding techniques on microstructure, properties and porosity of the welded joint of 2219 aluminum alloy,
J. Alloys Compd. 727 (2017) 531–539, https://doi.org/10.1016/j.jallcom.2017.08.157.
[16] D. Zhang, et al., Study on the inconsistency in mechanical properties of 2219 aluminium alloy TIG-welded joints, J. Alloys Compd. 777 (2019) 1044–1053,
https://doi.org/10.1016/j.jallcom.2018.10.182.
[17] C. Chen, G. Sun, W. Du, Y. Li, C. Fan, H. Zhang, Influence of heat input on the appearance, microstructure and microhardness of pulsed gas metal arc welded Al
alloy weldment, J. Mater. Res. Technol. 21 (2022) 121–130, https://doi.org/10.1016/j.jmrt.2022.09.028.
[18] H.X. Wang, C.L. Yang, Y.H. Wei, S.B. Lin, H.Y. Shen, Study of keyhole closure and analysis of microstructure and mechanical performance of weld joints for
variable polarity vertical up plasma arc welding process, Sci. Technol. Weld. Join. 11 (3) (2006) 315–325, https://doi.org/10.1179/174329306X101436.
[19] M.S. We, Electron beam welding e Techniques and trends e Review * 130 (2016) 72–92, https://doi.org/10.1016/j.vacuum.2016.05.004.
[20] S.A. Hosseini, A. Abdollah-zadeh, H. Naffakh-Moosavy, A. Mehri, Elimination of hot cracking in the electron beam welding of AA2024-T351 by controlling the
welding speed and heat input, J. Manuf. Process. 46 (September) (2019) 147–158, https://doi.org/10.1016/j.jmapro.2019.09.003.
[21] H. Fujii, H. Umakoshi, Y. Aoki, K. Nogi, Bubble formation in aluminium alloy during electron beam welding, J. Mater. Process. Technol. 155 (1–3) (2004)
1252–1255, https://doi.org/10.1016/j.jmatprotec.2004.04.141. –156.
[22] S.R. Koteswara Rao, G. Madhusudhan Reddy, K. Srinivasa Rao, M. Kamaraj, K. Prasad Rao, Reasons for superior mechanical and corrosion properties of 2219
aluminum alloy electron beam welds, Mater. Char. 55 (4–5) (2005) 345–354, https://doi.org/10.1016/j.matchar.2005.07.006.
[23] F. Hayat, Electron beam welding of 7075 aluminum alloy: microstructure and fracture properties, Eng. Sci. Technol. an Int. J. 34 (2022), 101093, https://doi.
org/10.1016/j.jestch.2022.101093.
[24] N.J. Luiggi, M. del V. Valera, Kinetic study of an AA7075 alloy under RRA heat treatment, J. Therm. Anal. Calorim. 130 (3) (2017) 1885–1902, https://doi.org/
10.1007/s10973-017-6683-8.
[25] W. Xu, J. Liu, G. Luan, C. Dong, Microstructure and mechanical properties of friction stir welded joints in 2219-T6 aluminum alloy, Mater. Des. 30 (9) (2009)
3460–3467, https://doi.org/10.1016/j.matdes.2009.03.018.
[26] Y.I. Golovin, Nanoindentation and Mechanical Properties of Materials at Submicro- and Nanoscale Levels: Recent Results and Achievements 63 (1) (2021),
https://doi.org/10.1134/S1063783421010108.
[27] P. Maier, A. Richter, R.G. Faulkner, R. Ries, Application of nanoindentation technique for structural characterisation of weld materials, Mater. Char. 48 (4)
(2002) 329–339, https://doi.org/10.1016/S1044-5803(02)00274-7.
[28] S. Sabooni, F. Karimzadeh, M.H. Enayati, A.H.W. Ngan, H. Jabbari, AC SC. Elsevier B.V. (2015), https://doi.org/10.1016/j.matchar.2015.08.009.
[29] J. Li, Q. Sun, K. Kang, Z. Zhen, Y. Liu, J. Feng, Process stability and parameters optimization of narrow-gap laser vertical welding with hot wire for thick stainless
steel in nuclear power plant, Opt. Laser Technol. 123 (2) (2020), 105921, https://doi.org/10.1016/j.optlastec.2019.105921.
[30] M. Sobih, Z. Elseddig, K. Almazy, M. Sallam, Experimental Evaluation and Characterization of Electron Beam Welding of 2219 AL-Alloy, vol. 2016, 2016.
[31] H. He, Y. Yi, S. Huang, Y. Zhang, Effects of deformation temperature on second-phase particles and mechanical properties of 2219 Al-Cu alloy, Mater. Sci. Eng. A
712 (December 2017) (2018) 414–423, https://doi.org/10.1016/j.msea.2017.11.124.
[32] C. Zhang, M. Gao, G. Li, C. Chen, X.Y. Zeng, Strength improving mechanism of laser arc hybrid welding of wrought AA 2219 aluminium alloy using AlMg5 wire,
Sci. Technol. Weld. Join. 18 (8) (2013) 703–710, https://doi.org/10.1179/1362171813Y.0000000153.
[33] A. Bachmann, T. Gigl, C.P. Hugenschmidt, M.F. Zaeh, Characterization of the microstructure in friction stir welds of EN AW-2219 using coincident Doppler-
broadening spectroscopy, Mater. Char. 149 (November 2018) (2019) 143–152, https://doi.org/10.1016/j.matchar.2019.01.016.
[34] Y.T. Lin, D.P. Wang, M.C. Wang, Y. Zhang, Y.Z. He, Effect of different pre-and post-weld heat treatments on microstructures and mechanical properties of
variable polarity TIG welded AA2219 joints, Sci. Technol. Weld. Join. 21 (3) (2016) 234–241, https://doi.org/10.1179/1362171815Y.0000000087.
[35] H.L. Wei, J.W. Elmer, T. Debroy, Origin of grain orientation during solidification of an aluminum alloy, Acta Mater. 115 (2016) 123–131, https://doi.org/
10.1016/j.actamat.2016.05.057.
[36] S. He, L. Liu, Y. Zhao, Y. Kang, F. Wang, X. Zhan, Comparative investigation between fiber laser and disk laser: microstructure feature of 2219 aluminum alloy
welded joint using different laser power and welding speed, Opt. Laser Technol. 141 (November 2020) (2021), 107121, https://doi.org/10.1016/j.
optlastec.2021.107121.
[37] Y. Zhou, X. Lin, N. Kang, W. Huang, Z. Wang, Mechanical properties and precipitation behavior of the heat-treated wire + arc additively manufactured 2219
aluminum alloy, Mater. Char. 171 (2021), 110735, https://doi.org/10.1016/j.matchar.2020.110735.
[38] Q. Li, A. Wu, Y. Li, G. Wang, D. Yan, J. Liu, Influence of temperature cycles on the microstructures and mechanical properties of the partially melted zone in the
fusion welded joints of 2219 aluminum alloy, Mater. Sci. Eng. A 623 (2015) 38–48, https://doi.org/10.1016/j.msea.2014.11.047.
[39] F. Lefebvre, S. Ganguly, I. Sinclair, Micromechanical aspects of fatigue in a MIG welded aluminium airframe alloy. Part 1. Microstructural characterization,
Mater. Sci. Eng. A 397 (1–2) (2005) 338–345, https://doi.org/10.1016/j.msea.2005.02.051.
[40] Z. Wang, X. Lin, L. Wang, Y. Cao, Y. Zhou, W. Huang, Microstructure evolution and mechanical properties of the wire + arc additive manufacturing Al-Cu alloy,
Addit. Manuf. 47 (September) (2021), 102298, https://doi.org/10.1016/j.addma.2021.102298.

13
G. Hussain et al. Heliyon 10 (2024) e23835

[41] H. Zhu, et al., Strengthening mechanism in laser-welded 2219 aluminium alloy under the cooperative effects of aging treatment and pulsed electromagnetic
loadings, Mater. Sci. Eng. A 714 (December 2017) (2018) 124–139, https://doi.org/10.1016/j.msea.2017.12.081.
[42] M. Meyers, et al., Mechanical behavior of materials 46 (12) (2009), https://doi.org/10.5860/choice.46-6830.
[43] Q. Chu, et al., On the association between microhardness, corrosion resistance and microstructure of probeless friction stir spot welded Al–Li joint, J. Mater. Res.
Technol. 14 (2021) 2394–2405, https://doi.org/10.1016/j.jmrt.2021.07.120.
[44] K.S. Arora, S. Pandey, M. Schaper, R. Kumar, Microstructure evolution during friction stir welding of aluminum alloy AA2219, J. Mater. Sci. Technol. 26 (8)
(2010) 747–753, https://doi.org/10.1016/S1005-0302(10)60118-1.
[45] C. Genevois, A. Deschamps, A. Denquin, B. Doisneau-Cottignies, Quantitative investigation of precipitation and mechanical behaviour for AA2024 friction stir
welds, Acta Mater. 53 (8) (2005) 2447–2458, https://doi.org/10.1016/j.actamat.2005.02.007.
[46] M.M. Attallah, C.L. Davis, M. Strangwood, Microstructure-microhardness relationships in friction stir welded AA5251, J. Mater. Sci. 42 (17) (2007) 7299–7306,
https://doi.org/10.1007/s10853-007-1585-y.
[47] C. Gallais, A. Denquin, Y. Bréchet, G. Lapasset, Precipitation microstructures in an AA6056 aluminium alloy after friction stir welding: characterisation and
modelling, Mater. Sci. Eng. A 496 (1–2) (2008) 77–89, https://doi.org/10.1016/j.msea.2008.06.033.
[48] G. Chen, J. Liu, Z. Dong, X. Shu, B. Zhang, Underlying reasons of poor mechanical performance of thick plate aluminum-copper alloy vacuum electron beam
welded joints, Vacuum 182 (April) (2020), 109667, https://doi.org/10.1016/j.vacuum.2020.109667.
[49] W. Xu, J. Liu, H. Zhu, Analysis of residual stresses in thick aluminum friction stir welded butt joints, Mater. Des. 32 (4) (2011) 2000–2005, https://doi.org/
10.1016/j.matdes.2010.11.062.
[50] J. Shen, et al., Microstructure and mechanical properties of gas metal arc welded CoCrFeMnNi joints using a 410 stainless steel filler metal, Mater. Sci. Eng. A
857 (September) (2022), 144025, https://doi.org/10.1016/j.msea.2022.144025.
[51] Z.X. Lu, H. Zhang, Tensile mechanical properties and failure mechanism of microcellular polycarbonate, Zhongguo Suliao/China Plast. 17 (1) (2003) 39.
[52] P.J. Noell, J.D. Carroll, B.L. Boyce, The mechanisms of ductile rupture, Acta Mater. 161 (2018) 83–98, https://doi.org/10.1016/j.actamat.2018.09.006.
[53] W.C. Oliver, G.M. Pharr, An improved technique for determining hardness and elastic madulus using load and displacement, J. Mater. Res. 7 (1) (1992)
1564–1583.
[54] N. Alatorre, R.R. Ambriz, B. Noureddine, A. Amrouche, A. Talha, D. Jaramillo, Tensile properties and fusion zone hardening for gmaw and miea welds of a 7075-
T651 aluminum alloy, Acta Metall. Sin. (English Lett. 27 (4) (2014) 694–704, https://doi.org/10.1007/s40195-014-0103-x.
[55] Y. Liu, F, Y. Chen, J. Mai, C. Wang, Q. Wang, Liu, Optimizing microstructure and mechanical properties of Ti-5Al-2Sn-2Zr-4Mo-4Cr alloy electron beam welded
joint through post-weld heat treatment, J. Mater. Res. Technol. 26 (2023) 7052–7071.
[56] S. Baskutis, V. Vasauskas, A. Žunda, Nano and microhardness testing of heterogeneous structures. Mechanics, Mechanics 22 (2) (2016) 85–99.
[57] M. Hassaan, M. Junaid, T. Shahbaz, M. Ilyas, F.N. Khan, J. Haider, Nanomechanical response of pulsed tungsten inert gas welded titanium alloy by
nanoindentation and atomic force microscopy, J. Mater. Eng. Perform. 30 (2) (2021) 1490–1503, https://doi.org/10.1007/s11665-020-05413-5.
[58] X. Zhang, N. Hansen, A. Godfrey, X. Huang, Dislocation-based plasticity and strengthening mechanisms in sub-20 nm lamellar structures in pearlitic steel wire,
Acta Mater. 114 (2016) 176–183, https://doi.org/10.1016/j.actamat.2016.04.040.
[59] N.V. Nguyen, T.H. Pham, S.E. Kim, Microstructure and strain rate sensitivity behavior of SM490 structural steel weld zone investigated using indentation,
Constr. Build. Mater. 206 (2019) 410–418, https://doi.org/10.1016/j.conbuildmat.2019.02.013.
[60] X. Cai, S. Lin, X. Wang, C. Yang, C. Fan, Characteristics of periodic ultrasonic assisted TIG welding for 2219 aluminum alloys, Materials 12 (24) (2019), https://
doi.org/10.3390/ma1224081.
[61] R. Kumar, U. Dilthey, D.K. Dwivedi, S.P. Sharma, P.K. Ghosh, Welding of thin sheet of Al alloy (6082) by using Vario wire DC P-GMAW, Int. J. Adv. Manuf.
Technol. 42 (1–2) (2009) 102–117, https://doi.org/10.1007/s00170-008-1568-4.
[62] A. Şik, M. Önder, Comparison between mechanical properties and joint performance of AA 2024-O aluminium alloy welded by friction stir welding and TIG
processes, Kov. Mater. 50 (2) (2012) 131–137, https://doi.org/10.4149/km_201_22_131.

14

You might also like