Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Materials & Design 131 (2017) 127–134

Contents lists available at ScienceDirect

Materials & Design


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

Effect of friction spot welding (FSpW) on the surface corrosion behavior of MARK
overlapping AA6181-T4/Ti-6Al-4V joints
G.S. Vacchia, A.H. Plainea,b, R. Silvaa, V.L. Sordia, U.F.H. Suhuddinb, N.G. Alcântaraa, S.E. Kuria,
C.A.D. Roverea,⁎
a
Munir Rachid Corrosion Laboratory, Department of Materials Engineering, Federal University of São Carlos, Rodovia Washington Luis Km 235, 13565-905 São Carlos,
SP, Brazil
b
Helmholtz-Zentrum Geesthacht, Institute of Materials Research, Materials Mechanics, Solid-State Joining Processes, Max-Planck-Str.1, Geesthacht, Germany

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

Keywords: The microstructural, mechanical and corrosion behavior of the surface of an AA6181-T4 aluminum alloy overlap
Friction spot welding welded onto Ti-6Al-4V titanium alloy by friction spot welding (FSpW) was analyzed based on optical microscopy
FSpW (OM), scanning electron microscopy (SEM), microhardness and potentiodynamic polarization techniques. The
Dissimilar joint, aluminum alloys results indicated that the FSpW process modified the microstructure of the AA6181-T4 aluminum alloy, as
Hardness
indicated mainly by the size, type and quantity of precipitates in various welding regions. The microhardness
Pitting corrosion
tests showed similar hardness values in the stir zone (SZ) and base metal (BM), which was ascribed to the
breakdown and homogenization of precipitates in the SZ. On the other hand, the heat affected zone (HAZ)
showed the lowest hardness, which was attributed to the coalescence of Mg2Si precipitates caused by the thermal
cycle of the welding process. The potentiodynamic polarization tests indicated that the SZ showed the highest
pitting potential due to the refined microstructure in this zone. SEM images recorded after potentiodynamic
polarization testing indicated that preferential sites for pitting nucleation were regions adjacent to the Al (Fe, Si,
Mn, Mg) precipitates, and the mildest corrosive attack was found in the SZ.

G RA P H I C A L AB S T R A C T

1. Introduction environmental impact, there is a growing trend toward utilizing light-


weight components and structures in the transportation industry.
Today, in order to save money and energy as well as reduce Taking this scenario into account, lightened hybrid structures of


Corresponding author.
E-mail addresses: carlosdrovere@gmail.com, rovere@ufscar.br (C.A.D. Rovere).

http://dx.doi.org/10.1016/j.matdes.2017.06.005
Received 27 March 2017; Received in revised form 29 May 2017; Accepted 1 June 2017
Available online 02 June 2017
0264-1275/ © 2017 Elsevier Ltd. All rights reserved.
G.S. Vacchi et al. Materials & Design 131 (2017) 127–134

aluminum (Al) and titanium (Ti) alloys have attracted increasing at- approximately 49 mm × 26 mm. Table 1 describes the chemical com-
tention because they combine the individual properties of the two position of the alloys (information provided by the respective manu-
original alloys, i.e., the low weight and cost of Al with the excellent facturers).
properties of Ti [1,2,3,4,5]. However, due to a mismatch in their phy- The FSpW process was performed using an overlap aluminum-on-
sical properties, a limited mutual solubility and formation of brittle titanium configuration at the Helmholtz Zentrum Geesthacht [14]. The
intermetallic phases in Al-Ti system, one of the difficulties encountered welding tool used in this study comprised a clamping ring, sleeve and
in the production of such hybrid structures is joining these materials. In pin, with diameters of 18, 9 and 6.4 mm, respectively, and the sleeve
addition, current techniques employed in spot joining increase the and pin were threaded. The welded joint was produced using optimized
weight of the structure, such as riveting, or are expensive, such as re- process parameters: a tool rotational speed of 2500 rpm, a sleeve
sistance welding [6]. Thus, there is a natural tendency to seek a more plunge depth of 1.4 mm, a clamping force of 12 kN, and a dwell time of
efficient method to join these hybrid structures of Al-Ti. 3 s [7].
In this context, Friction Spot Welding (FSpW) arises as an alter- It should be noted that the sleeve plunge depth of 1.4 mm was
native to overcome the difficulties above mentioned. FSpW is an in- chosen to avoid reaching the Ti sheet, considering that the temperature
novative solid-state spot welding process developed by GKSS Research of about 550 °C produced in the welding process is not sufficient to
Centre in Germany (currently HZG), offering significant advantages plasticize the titanium, thus preventing excessive tool wear [15]. In a
over conventional welding techniques, e.g. environment-friendly (nei- previous study by our research group, Plaine et al. [16] found that this
ther gases nor fumes are produced), fast welding times, energy efficient, temperature does not suffice to promote microstructural changes in
no additional material is required, minimal or no waste products, no titanium; hence, in the present work, we will focus solely on the
pre-cleaning necessary as well as no post processing due to good surface changes that occurred on the surface of the aluminum sheet.
quality, ease of automation and the possibility of joining dissimilar Fig. 1(a–c) shows the AA6181 surface, the welding region and a
materials, such as Al-Mg, Al-steel and Al-Ti [7,8]. cross-sectional view of the welded joint.
It should be mentioned that the current literature contains studies
whose authors evaluated the mechanical properties and interface fea- 2.2. Hardness measurements
tures of joints welded by FSpW, and satisfactory results were obtained
[9,10,8,11]. However, there is little information about the effect of the To investigate the effects of thermomechanical cycling on the
FSpW on the corrosion behavior of dissimilar welded joints, so addi- hardness and to identify the transition of the different welding zones,
tional studies are essential to further develop and expand the industrial the Vickers hardness of the aluminum alloy surface was tested (in an
application of this relatively new technology. In addition, it is worth area of 25.0 mm by 10.0 mm) by applying a load of 300 g, a dwell time
keeping in mind that, in heat-treatable aluminum alloys, the micro- of 10 s, and spacing of 0.5 mm between two adjacent indentations.
structural evolution of the different welding regions due to the ther- Vickers hardness measurements were also performed in the cross-sec-
momechanical cycle of the FSpW process can modify the corrosion tion (in an area of 30.0 by 1.5 mm) using the same parameters as in the
behavior of Al-Ti welds, which may reduce the lifetime of the structure surface tests.
[12,13]. Paglia and Buchheit [12] have highlighted this issue for alu-
minum alloys joined by means of conventional friction stir welding 2.3. Microstructure characterization
(FSW). Very recently, Gharavi and coworkers [13] revealed that the FS
lap welding process makes the weld regions of an AA6061-T6 alu- For the microstructural characterization, the AA6181-T4 welding
minum alloy more susceptible to intergranular and pitting attacks than surface was sequentially sanded with SiC abrasive paper up to
the base metal. According to these authors, there is an increase in the #1500 grit and then polished with a 1.0 μm alumina suspension. To
intermetallic constituent particles during FS welding process (as a result reveal the grain structure and the different welding zones, the sample
of temperature excursions), which brings about a more intense galvanic was anodized in Barker's solution (5 mL of HBF4 in 200 mL H2O) at
corrosion coupling, and hence decreases the corrosion resistance of FS 20 V for 120 s and the resulting microstructure was examined by OM
weld regions. using polarized light. The average grain size for each region was de-
In view of the above and the paucity of studies, this work examined termined using the intercept method.
the effect of the FSpW process on the localized corrosion resistance of To identify the phases and their compositions in each welded re-
the aluminum surface of a welded joint of dissimilar AA6181/Ti-6Al-4V gion, the anodized sample was repolished to a mirror condition and
alloys based on potentiodynamic polarization tests of the different analyzed by SEM and energy dispersive spectroscopy (EDS).
welding regions, i.e., base metal (BM), heat affected zone/thermo-
mechanically affected zone (HAZ/TMAZ) and stir zone (SZ).
2.4. Localized corrosion tests
Microstructural changes were analyzed by optical microscopy (OM) and
scanning electron microscopy (SEM), and mechanical properties were
To evaluate the localized corrosion resistance of different welding
evaluated based on Vickers microhardness mapping.
regions, potentiodynamic polarization tests were performed in a natu-
rally aerated solution of 0.01 M sodium chloride (NaCl) and 0.1 M so-
2. Experimental procedure dium sulfate (Na2SO4) at 25 °C. Na2SO4 was added to improve the
identification of pitting potential (Epit) during the polarization tests of
2.1. Material aluminum alloys [17]. According to Lee et al. [17], sulfate (SO4− 2) ions
retard the breakdown of the aluminum oxide passive film by chlorides
The materials under study were AA6181-T4 and Ti-6Al-4V com- (Cl−), which promotes an increase of the passive region in the polar-
mercial alloy sheets, both 1.5 mm thick and with dimensions of ization curve and takes the visualization of Epit more evident. A glass

Table 1
Chemical composition (wt%) of AA6181-T4 and Ti-6Al-4V sheets.

Alloy Ti Al Si Fe Cu Mn Mg V C O

AA6181-T4 0.023 Bal. 0.85 0.25 0.060 0.090 0.74 – – –


Ti6Al4V Bal. 6.25 – 0.14 – – – 3.91 0.023 0.13

128
G.S. Vacchi et al. Materials & Design 131 (2017) 127–134

Fig. 1. (a) Surface of AA6181-T4; (b) macroscopic image of the friction spot welded aluminum surface; (c) cross-section of the welded joint.

conventional corrosion cell kit with platinum auxiliary electrode and the AA6181-T4/Ti-6Al-4V FSp welded joint. In this figure, note the four
silver/silver chloride (Ag/AgCl in saturated KCl) reference electrode different welding regions: BM (a), HAZ (b), TMAZ (c) and SZ (d), which
was used for the potentiodynamic polarization tests. comprise the sleeve stir region and the pin stir region.
The working electrodes, consisting of samples from different Fig. 3 shows in greater detail the regions (a), (b), (c) and (d) that
welding regions, were positioned at the glass corrosion cell kit, leaving were highlighted in Fig. 2. Note that the BM (Fig. 3(a)) exhibited an
a 0.215 cm2 circular area of the aluminum alloy surface exposed to the equiaxial structure, with randomly oriented grains of similar size (with
test solution. Samples were further carefully wet sanded to a #600-grit an average grain size of 49 ± 3 μm). In Fig. 3(b), which represents the
silicon carbide (SiC) paper finish, followed by distilled water washing HAZ region, a slightly larger average grain size (53 ± 2 μm) is ob-
and air drying before polarization measurements. After immersion in served. This is due to the effect of the thermal cycle that promoted grain
the solution, the samples were left in open circuit for 30 min to ensure growth in this region [11]. The TMAZ region in Fig. 3(c) exhibits a
that a steady potential would be reached, and the resulting potential distorted structure, with the grain oriented in the same direction as the
value was regarded as the open circuit potential (Eoc). rotation of the sleeve, i.e., from right to left, which is the result of the
The potentiodynamic polarization tests started at 300 mV below Eoc deformation process occurring at high temperatures in this region [18].
and ended immediately after the pitting potential (Epit) was reached, so The SZ (Fig. 3(d)) clearly shows a homogeneous microstructure re-
as to prevent a deeper corrosive attack. A scan rate of 1 mV/s was sulting from the dynamic recrystallization process that occurred due to
applied. The mean values of corrosion potential (Ecorr) and (Epit) were the higher intensity of high temperature deformation. Moreover, note
obtained from seven independent measurements, along with the stan- that the grains in the SZ are smaller (average grain size of 21 ± 6 μm)
dard deviation. After the corrosion tests, each welding region was than those in the BM [19,20,21].
analyzed by SEM in order to identify the morphology of the corrosion Fig. 4 depicts SEM images of the BM ((a) and (d)), HAZ ((b) and (e))
attack. and SZ ((c) and (f)). The images of BM (Fig. 4(a) and (d)) show two
types of precipitates: large light-colored and small dark precipitates
uniformly dispersed throughout the matrix. The EDS chemical analysis
3. Results and discussion (Table 2) revealed that the light-colored precipitates are rich in Fe, Si,
Mn and Mg, while the dark ones are rich in Si and Mg. The presence of
Fig. 2 shows the macrostructure of the aluminum alloy surface of

Fig. 2. Macroscopic image of the friction spot welded surface of AA6181-T4.

129
G.S. Vacchi et al. Materials & Design 131 (2017) 127–134

Fig. 3. OM micrograph of different welding regions: (a) BM; (b) HAZ; (c) TMAZ; (d) SZ.

the light-colored precipitates may be attributed to the material's che- during solidification and difficult to dissolve during subsequent heat
mical composition and the heat treatment it underwent. This alloy treatments due to their thermodynamic stability [22,23,24]. On the
contains 0.25 wt% of Fe, whose maximum solubility in aluminum alloy other hand, the dark precipitates are the Mg2Si phases that are formed
is 0.05 wt% at 655 °C, making it possible for these precipitates to form during the processing steps of the alloy [22,25].

Fig. 4. SEM micrographs of the various welding regions: (a) and (d) BM; (b) and (e) HAZ; (c) and (f) SZ.

130
G.S. Vacchi et al. Materials & Design 131 (2017) 127–134

Table 2 surface. This map indicates that the BM (yellow color in the range of
EDS analysis of micrograph 3 (d–f). 12.0 and 20.0 mm) presented a hardness of approximately 90 HV0.3.
The HAZ begins from 12.0 mm with a reduction in hardness, and pre-
Weld Region Point Composition (wt%)
sents the lowest value (75 HV0.3) around 7.5 mm. The end of this region
Al Mg Si Fe Mn O is around 6.0 mm, with a slight increase in hardness. These results are
in agreement with Figs. 3 and 4, and also with Shen et al.’s study [11] of
BM a 80.22 1.03 5.56 12.72 0.47 –
FSpW welded 6061-T4 alloy. Those authors attributed the lower
BM b 74.64 10.64 12.17 2.65 0 –
HAZ c 84.27 1.2 6.71 7.35 0.47 – hardness observed in the HAZ to the coarse grains and Mg2Si grain
HAZ d 75.38 8.78 10.92 0.72 0.26 3.94 dissolution or coarsening caused by the thermal cycle in the welding
SZ e 70.22 0.72 7.72 17.55 3.79 – process.
The TMAZ (green color in the range of 4.0 to 6.0 mm) shows a slight
increase in hardness, in the order of approximately 10 HV0.3, when
The HAZ, Fig. 4(b) and (e), contains the same precipitates as those compared to the HAZ. This increase is directly attributable to the in-
found in the BM, although the Mg2Si precipitates in this region are tense plastic deformation caused by the sleeve rotation, which induces
slightly larger. The results reported by Donatus et al. [25] and Gallais the appearance of a structure of elongated grains and, consequently, an
et al. [26], whose work focused on friction stir welded 6xxx series increase in hardness.
aluminum alloys, revealed a similar increase in the size of the pre- The hardness in the SZ (orange color in the range of −4.0 to
cipitates in the HAZ. This fact is due to the thermal cycle imposed by 4.0 mm) is slightly higher (95 HV0.3) than in the BM. This is due to the
the welding process in this region, which reaches temperatures of more intense deformation at high temperatures in this region than in
around 450 °C, causing the coalescence and loss of coherence of Mg2Si, the TMAZ, which promotes the appearance of a more refined structure
and consequently, its hardening [26]. with smaller grains and precipitates, as can be seen in the SEM and OM
The SZ, Fig. 4(c) and (f), contains only Al (Fe, Si, Mn, Mg) pre- micrographs.
cipitates, which are usually smaller and more uniformly distributed Fig. 5(b) shows the hardness in the cross section of the FSpW welded
than in the HAZ and BM regions. This is a result of the deformation joint. Note that the hardness profile is similar to that found on the
process at high temperature generated by the welding process, which aluminum surface.
causes these precipitates to break down into smaller sizes. Yang et al. Fig. 6 shows the potentiodynamic polarization curves of the dif-
[27], who evaluated aluminum alloys of the 2xxx series subjected to ferent welding regions (BM, HAZ/TMAZ and SZ), while Fig. 7 presents
friction stir welding (FSW), identified a reduction in the size of pre- the mean values of Ecorr and Epit. Note that the Ecorr does not vary
cipitates in the SZ, attributing it to the comminution of larger particles significantly as a function of the different welding regions, which may
of the BM, which corroborates our results. On the other hand, the ab- be attributed to the anodic and cathodic reaction rates remaining es-
sence of Mg2Si precipitates in the SZ may be ascribed to two factors: the sentially unchanged, despite the microstructural modifications ob-
temperature reached in this region, which is around 540 °C, is close to served by OM and SEM. El-Menshawy et al. [28], who evaluated the
or higher than the melting temperature of Mg2Si, or the breakdown of effect of aging time at low temperatures on 6061 aluminum alloy, re-
these precipitates into very small sizes, preventing their detection by ported that the volumetric fraction of precipitates changes the anodic
SEM [26,23]. and cathodic reaction rates, modifying the resulting Ecorr values. In our
Fig. 5(a) shows a microhardness map of the FSpW welded aluminum study, the Ecorr values were similar in the different welding regions,

Fig. 5. Microhardness mapping of (a) AA6181-T4 welded surface and (b) cross-section.

131
G.S. Vacchi et al. Materials & Design 131 (2017) 127–134

Fig. 6. Potentiodynamic polarization curves obtained in 0.01 M NaCl and 0.1 M Na2SO4 solution of various regions of FSpW welded AA6181.

suggesting that the volumetric fraction of the precipitates did not vary dissolution of the matrix adjacent to the precipitate. This result is in
significantly. agreement with the mechanism proposed by Park et al. [30], who
However, the SZ showed higher (nobler) Epit values than the BM and suggested that in the early stages of the process the dissolution of the
HAZ/TMAZ, whose Epit values were very similar. These differences in matrix is independent of precipitate size, while in later stages the dis-
Epit may be attributed to the type, size and quantity of precipitates solution of the aluminum matrix is more intense around larger pre-
found in each welding region. In the BM, the most abundant and largest cipitates. Our results also coincide with those reported by Jiang et al.
precipitates are Al (Fe, Si, Mn, Mg), as indicated in Fig. 4 (a and d). [31], who evaluated an Al-Si alloy processed by severe plastic de-
According to Aballe et al. [29], these precipitates exhibit a cathodic formation (SPD). They observed that the reduction in precipitate size by
behavior and form a galvanic cell with the matrix, resulting in pre- mechanical processing favored the formation of less intense galvanic
ferential dissolution of the aluminum. In addition, the passive film cells in these precipitates, reducing susceptibility to localized corrosion.
formed around these precipitates is less effective, thus reducing the The HAZ also shows Al (Fe, Si, Mn, Mg) precipitates with sizes si-
corrosion resistance in these regions. Furthermore, the size of the pre- milar to those in BM, as observed in Fig. 4(b and e). This explains why
cipitate significantly influences localized corrosion resistance. As the the Epit value found in the HAZ is similar to that of the BM. However,
precipitate grows in size, the galvanic cell becomes more intense and the HAZ microstructure shows a slight increase in the size of Mg2Si
the passive film formed becomes less effective, causing greater precipitates, although findings suggest that these precipitates appear

Fig. 7. Corrosion potential (Ecorr) and pitting potential (Epit) of different regions of FSpW welded AA6181 in 0.01 M NaCl and 0.1 M Na2SO4 solution.

132
G.S. Vacchi et al. Materials & Design 131 (2017) 127–134

Fig. 8. Pitting corrosion morphology of the surface of AA6181 samples after potentiodynamic polarization in 0.01 M NaCl and 0.1 M Na2SO4 solution in various regions: BM (a), HAZ/
TMAZ (b) and SZ (c–d).

Table 3 corrosion resistance in the SZ.


EDS analysis of micrograph 8 (a–c). Fig. 8 shows the surface of the BM (a), HAZ/TMAZ (b) and SZ (c-d)
after potentiodynamic polarization testing. Note that the corrosive at-
Weld region Point Composition (wt%)
tack clearly occurs in regions adjacent to the Al (Fe, Si, Mn, Mg) pre-
Al Mg Si Fe Mn O cipitates, which is confirmed by the EDS composition analysis (Table 3).
A more Intense localized corrosion attack is visible in the BM (a) and in
BM a 84.91 1.22 6.69 4.62 0.32 2.24
the HAZ/TMAZ (b), and pitting is deeper, which is consistent with the
HAZ/TMAZ b 68.22 1.28 14.58 7.23 0.33 7.75
SZ c 72.89 0.78 12.02 10.85 0.7 2.77 lower Epit value recorded in these regions. However, the SZ also shows
nucleated pits in regions adjacent to the precipitates Al (Fe, Si, Mn, Mg),
although they are much smaller than those observed in the BM and in
not to significantly affect the Epit. According to Donatus et al. [25], the the HAZ/TMAZ, indicating a more noble pitting potential in this region.
influence of small precipitates on localized corrosion, such as the Mg2Si The reduction in the size of the precipitates decreases the intensity of
precipitates found in this alloy, is negligible when compared to the the galvanic cell formed between the precipitate and the matrix, re-
contribution of larger and more abundant precipitates such as Al (Fe, Si, sulting in a less corrosive attack in the stir zone. This finding is in good
Mn, Mg). This explains the similar Epit values found in the BM and HAZ/ agreement with the model proposed by Jiang et al. [31], who proposed
TMAZ. that galvanic cell formed between Al matrix and intermetallics with
On the other hand, as mentioned earlier, the deformation process in smaller sizes acts as “small cathode-large anode”, therefore, the sus-
the SZ at high temperature promoted the breakdown and homo- ceptibility to pitting corrosion is reduced.
genization of Al (Fe, Si, Mn, Mg) precipitates, as can be seen in Fig. 4(c
and f). This microstructural change reduced the intensity of the gal- 4. Conclusions
vanic cells and the interaction between them, thereby increasing loca-
lized corrosion resistance. However, this finding differs in part from The following conclusions can be drawn from the experimental re-
that observed by Gharavi et al. [13], who worked on AA6061 alloy sults:
overlap welded by FSW and reported that the BM showed higher Epit
than the HAZ/TMAZ and SZ. According to the authors, the size of in- √ The OM micrographs indicate that the FSpW process promotes grain
termetallic particles in the FS weldments are bigger than that in the BM refinement in the SZ and a deformed structure with grains aligned in
region, which leads to an increase in the number of sites for the for- the direction of sleeve rotation in the TMAZ. However, micro-
mation of galvanic cells, reducing the localized corrosion resistance in structural variations in the HAZ were not visible in these micro-
the HAZ/TMAZ and SZ. On the other hand, our results are in agreement graphs.
with those reported by Qin et al. [32], Yoshimawa et al. [33] and Na-
vaser et al. [34], who stated that a decrease in the size of precipitates √ The FSpW process led to the breakdown and homogenization of Al
and their better homogenization favors an increase in localized (Fe, Si, Mn, Mg) precipitates in the SZ, resulting in a similar

133
G.S. Vacchi et al. Materials & Design 131 (2017) 127–134

hardness to that observed in the BM. However, the thermal cycle 766–778, http://dx.doi.org/10.1016/j.matdes.2013.08.021.
[12] C.S. Paglia, R.G. Buchheit, A look in the corrosion of aluminum alloy friction stir
generated by this process reduced the hardness in the HAZ due to welds, Scr. Mater. 58 (2008) 383–387, http://dx.doi.org/10.1016/j.scriptamat.
the coalescence of Mg2Si precipitates. 2007.10.043.
[13] F. Gharavi, K.A. Matori, R. Yunus, N.K. Othman, F. Fadeifard, Corrosion evaluation
of friction stir welded lap joints of AA6061-T6 aluminum alloy, Trans. Nonferrous
√ The potentiodynamic polarization tests showed that Ecorr in the Metals Soc. China 26 (2016) 684–696, http://dx.doi.org/10.1016/S1003-6326(16)
different welding regions (BM, HAZ/TMAZ and SZ) did not vary 64159-6.
significantly in response to the welding process. However, the mi- [14] S.T. Amancio-Filho, A.P.C. Camillo, L. Bergmann, J.F. dos Santos, S.E. Kury,
N.G.A. Machado, Preliminary investigation of the microstructure and mechanical
crostructural changes triggered by FSpW in the SZ increased the behaviour of 2024 aluminium alloy friction spot welds, Mater. Trans. 52 (2011)
pitting corrosion resistance in this zone when compared to the BM 985–991, http://dx.doi.org/10.2320/matertrans.L-MZ201126.
and HAZ. [15] A.H. Plaine, A.R. Gonzalez, U.F.H. Suhuddin, J.F. dos Santos, N.G. Alcântara,
Process parameter optimization in friction spot welding of AA5754 and Ti6Al4V
dissimilar joints using response surface methodology, Int. J. Adv. Manuf. Technol.
√ The SEM images recorded after potentiodynamic polarization tests 83 (2015) 36–41, http://dx.doi.org/10.1007/s00170-015-8055-5.
showed that pits nucleate preferentially in regions adjacent to Al [16] A.H. Plaine, U.F.H. Suhuddin, N.G. Alcântara, J.F. dos Santos, Fatigue behavior of
(Fe, Si, Mn, Mg) precipitates in all the regions under study. friction spot welds in lap shear specimens of AA5754 and Ti6Al4V alloys, Int. J.
Fatigue 91 (2016) 149–157, http://dx.doi.org/10.1016/j.ijfatigue.2016.06.005.
However, this localized corrosive attack is much more intense in the [17] W.-J. Lee, S.-I. Pyun, Effects of sulphate ion additives on the pitting corrosion of
BM and HAZ due to the larger size of the precipitates found in these pure aluminium in 0.01 M NaCl solution, Electrochim. Acta. 45 (2000) 1901–1910,
regions. http://dx.doi.org/10.1016/S0013-4686(99)00418-1.
[18] R.S.S. Mishra, Z.Y.Y. Ma, Friction stir welding and processing, Mater. Sci. Eng. R.
Rep. 50 (2005) 1–78, http://dx.doi.org/10.1016/j.mser.2005.07.001.
Acknowledgements [19] C. Gao, R. Gao, Y. Ma, Microstructure and mechanical properties of friction spot
welding aluminium-lithium 2A97 alloy, Mater. Des. 83 (2015) 719–727, http://dx.
doi.org/10.1016/j.matdes.2015.06.013.
The authors gratefully acknowledge the Brazilian research funding [20] S. Sheikhi, J.F. dos Santos, Effect of process parameter on mechanical properties of
agencies CAPES (Federal Agency for the Support and Improvement of friction stir welded tailored blanks from aluminium alloy 6181-T4, Sci. Technol.
Higher Education) and CNPq (National Council for Scientific and Weld. Join. 12 (2007) 370–375, http://dx.doi.org/10.1179/174329307X173698.
[21] F. Gharavi, K.A. Matori, R. Yunus, N.K.N.K. Othman, Corrosion behavior of friction
Technological Development – grant no. 460659/2014-6), as well as the
stir welded lap joints of AA6061-T6 aluminum alloy, Mater. Res. 17 (2014)
Postgraduate Program in Materials Science and Engineering of the 672–681, http://dx.doi.org/10.1590/S1516-14392014005000053.
Federal University of São Carlos (PPGCEM/UFSCar), for their financial [22] I. Polmear, The light metals, Light Alloy. Met. Light Met, 1995.
[23] C. Gallais, A. Simar, D. Fabregue, A. Denquin, G. Lapasset, B. de Meester,
support of this work. We also thank the Helmholtz Association of
Y. Brechet, T. Pardoen, B. Meester, Y. Brechet, T. Pardoen, Multiscale analysis of the
German Research Centers for its technical support. strength and ductility of AA 6056 aluminum friction stir welds, Metall. Mater.
Trans. A 38 (2007) 964–981, http://dx.doi.org/10.1007/s11661-007-9121-x.
References [24] R. Ambat, A. Afseth, A.J. Davenport, G.M. Scamans, A. Afseth, Effect of iron-con-
taining intermetallic particles on the corrosion behaviour of aluminium, Corros. Sci.
48 (2006) 3455–3471 http://dx.doi.org/10.1016/j.corsci.2006.01.005.
[1] P.K. Mallick, Materials, Design and Manufacturing for Lightweight Vehicles, (2010), [25] U. Donatus, G.E. Thompson, X. Zhou, J. Wang, A. Cassell, K. Beamish, Corrosion
http://dx.doi.org/10.1533/9781845697822.2.309. susceptibility of dissimilar friction stir welds of AA5083 and AA6082 alloys, Mater.
[2] J.-P. Immarigeon, R.T. Holt, A.K. Koul, L. Zhao, W. Wallace, J.C. Beddoes, Charact. 107 (2015) 85–97, http://dx.doi.org/10.1016/j.matchar.2015.07.002.
Lightweight materials for aircraft applications, Mater. Charact. 35 (1995) 41–67, [26] C. Gallais, A. Denquin, Y. Bréchet, G. Lapasset, Precipitation microstructures in an
http://dx.doi.org/10.1016/1044–5803(95)00066-6. AA6056 aluminium alloy after friction stir welding: characterisation and modelling,
[3] M. Busse, A. Herrmann, K. Kayvantash, D. Lehmhus, Structural Materials and Mater. Sci. Eng. A 496 (2008) 77–89, http://dx.doi.org/10.1016/j.msea.2008.06.
Processes in Transportation, Ed, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, 033.
2013. [27] B. Yang, J. Yan, M.A. Sutton, A.P. Reynolds, Banded microstructure in AA2024-
[4] K. Matsuyama, Trend of automobile vehicles and the joining technologies, Riv. Ital. T351 and AA2524-T351 aluminum friction stir welds, Mater. Sci. Eng. A 364 (2004)
Della Saldatura. 59 (2007) 683–693, http://dx.doi.org/10.1007/BF03266560. 55–65, http://dx.doi.org/10.1016/S0921-5093(03)00532-X.
[5] Y. Wei, J. Li, J. Xiong, F. Huang, F. Zhang, S.H. Raza, Joining aluminum to titanium [28] K. El-Menshawy, A.-W.W.A. El-Sayed, M.E. El-Bedawy, H.A. Ahmed, S.M. El-Raghy,
alloy by friction stir lap welding with cutting pin, Mater. Charact. 71 (2012) 1–5, Effect of aging time at low aging temperatures on the corrosion of aluminum alloy
http://dx.doi.org/10.1016/j.matchar.2012.05.013. 6061, Corros. Sci. 54 (2012) 167–173, http://dx.doi.org/10.1016/j.corsci.2011.09.
[6] X.W. Yang, T. Fu, W.Y. Li, Friction stir spot welding: a review on joint macro- and 011.
microstructure, property, and process modelling, Adv. Mater. Sci. Eng. 2014 (2014) [29] A. Aballe, M. Bethencourt, F.J. Botana, M.J. Cano, M. Marcos, Localized alkaline
11, http://dx.doi.org/10.1155/2014/697170. corrosion of alloy AA5083 in neutral 3.5% NaCl solution, Corros. Sci. 43 (2001)
[7] A.H.H. Plaine, A.R.R. Gonzalez, U.F.H.F.H. Suhuddin, J.F.F. dos Santos, 1657–1674, http://dx.doi.org/10.1016/S0010-938X(00)00166-9.
N.G.G. Alcântara, The optimization of friction spot welding process parameters in [30] J.O. Park, Influence of Fe-rich intermetallic inclusions on pit initiation on aluminum
AA6181-T4 and Ti6Al4V dissimilar joints, Mater. Des. 83 (2015) 36–41, http://dx. alloys in aerated NaCl, J. Electrochem. Soc. 146 (1999) 517, http://dx.doi.org/10.
doi.org/10.1016/j.matdes.2015.05.082. 1149/1.1391637.
[8] U. Suhuddin, L. Campanelli, M. Bissolatti, H. Wang, R. Verastegui, J.F. dos Santos, A [31] J.-H.H. Jiang, A.-B.B. Ma, F.-M.M. Lu, N. Saito, A. Watazu, D. Song, P. Zhang,
review on microstructural and mechanical properties of friction spot welds in Al- Y. Nishida, Improving corrosion resistance of Al-11mass%Si alloy through a large
based similar and dissimilar joints, Proc. 1st Int. Jt. Symp. Join. Weld, 2013, pp. number of ECAP passes, Mater. Corros. 62 (2011) 848–852, http://dx.doi.org/10.
15–21, , http://dx.doi.org/10.1533/978-1-78242-164-1.15. 1002/maco.200905521.
[9] A.H.H. Plaine, U.F.H.F.H. Suhuddin, C.R.M.R.M. Afonso, N.G.G. Alcântara, [32] H. Long Qin, H. Zhang, D. Tong Sun, Q. Yu Zhuang, Corrosion behavior of the
J.F.F. dos Santos, Interface formation and properties of friction spot welded joints of friction-stir-welded joints of 2A14-T6 aluminum alloy, Int. J. Miner. Metall. Mater.
AA5754 and Ti6Al4V alloys, Mater. Des. 93 (2016) 224–231, http://dx.doi.org/10. 22 (2015) 627–638, http://dx.doi.org/10.1007/s12613-015-1116-9.
1016/j.matdes.2015.12.170. [33] D.S. Yoshikawa, M. Terada, S.L. Assis, I. Costa, A.F. Padilha, et al., Mater. Corros.
[10] T. Rosendo, B. Parra, M.A.D. Tier, A.A.M. da Silva, J.F. dos Santos, (2015) n/a, http://dx.doi.org/10.1002/maco.201508442.
T.R. Strohaecker, N.G. Alcântara, Mechanical and microstructural investigation of [34] M. Navaser, M. Atapour, Effect of friction stir processing on pitting corrosion and
friction spot welded AA6181-T4 aluminium alloy, Mater. Des. 32 (2011) intergranular attack of 7075 aluminum alloy, J. Mater. Sci. Technol. (2016), http://
1094–1100, http://dx.doi.org/10.1016/j.matdes.2010.11.017. dx.doi.org/10.1016/j.jmst.2016.07.008.
[11] Z. Shen, X. Yang, S. Yang, Z. Zhang, Y. Yin, Microstructure and mechanical prop-
erties of friction spot welded 6061-T4 aluminum alloy, Mater. Des. 54 (2014)

134

You might also like