Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Corrosion Science xxx (xxxx) xxxx

Contents lists available at ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

Macro-galvanic effect and its influence on corrosion behaviors of friction stir


welding joint of 7050-T76 Al alloy
Yong Chena, Yanqiu Wanga,b,*, Li zhouc, Guozhe Menga, Bin Liua, Junyi Wanga, Yawei Shaoa,
Jiantang Jiangd
a
Key Laboratory of Superlight Materials and Surface Technology (Harbin Engineering University), Ministry of Education, Harbin, 150001, China
b
Key Laboratory of Preparation and Application of Environmental Friendly Materials (Jilin Normal University), Ministry of Education, Changchun, 130103, China
c
Shandong Provincial Laboratory of Special Welding Technology, Harbin Institute of Technology at Weihai, Weihai, 264209, China
d
School of Materials Science and Engineering, Harbin Institute of Technology, Harbin, 150001, China

A R T I C LE I N FO A B S T R A C T

Keywords: The corrosion behavior of various zones of the 7050-T76 aluminum alloy friction stir welding (FSW) joint and
Al alloy their interaction were studied. The wire beam electrode (WBE) analysis reveals that heat-affected zone (HAZ)
Friction stir welding adjacent to nugget zone (NZ) receives the most intense negative current and thus presents the intensified sus-
Corrosion potential ceptibility to corrosion. Both pitting and intergranular corrosion are aggravated greatly in HAZ while suppressed
Galvanic corrosion
in base material (BM) when the whole FSW joint was put into NaCl solution, indicating the dominating influence
from macro-galvanic effect. The coupling relationship between macro-galvanic and micro-galvanic effects is
revealed based on the electrochemical analysis.

1. Introduction microstructure in FSW joints.


Many researches have been carried out to investigate the corrosion
The development of friction stir welding (FSW) has greatly pro- behavior of the FSW joints [7–16] and some methods for improving the
moted the fabrication of complicated components from aluminum corrosion resistance of local regions were developed [17–20]. The high
alloy, as the hot cracking sensitivity that derived in conventional susceptibility of FSW joint is generally believed to originate from the
welding was substantially eluded [1]. Components fabricated from enhanced galvanic effect, either in micro scale that related to the mu-
7xxx and 2xxx Al alloys via FSW process have thus been widely used in tated microstructure in local regions, or in macro scale that related to
aircrafts, spacecrafts and ships in the past a few decades and excellent the microstructure heterogeneity that developed during the FSW pro-
mechanical properties were thereafter achieved [2]. Although FSW cess. Micro-galvanic corrosion between the matrix and second phases,
endues the aluminum alloy in welding zone better retention of baseline grain boundaries (GBs) or precipitation free zones (PFZs) is common
mechanical properties and fewer defects, comparing with the conven- mechanism for pitting and intergranular corrosion (IGC) in aluminum
tional welding process, it still produces intensively modified micro- alloys [13]. For instance, Al7Cu2Fe, one of commonly identified parti-
structure quite different from that of the base materials and thus bring cles in aluminum alloys, can cause the dissolution of the surrounding
out heterogeneity in FSW joints and the neighboring region. The pre- matrix and thus induce continuously accelerating pitting because of its
sence of heterogeneous microstructure influences the performances of noble electrochemical potential and the initiative de-alloying process
the components including mechanical properties and corrosion re- [21]. Similarly, the presence of Al2Cu precipitate is believed the main
sistance as well [3–5]. Many studies noted that FSW structures exhibit cause for the accelerated anodic-cathodic reactivity in 2xxx alloys [22].
intensified susceptibility to localized corrosion (LC) to compare with The localized corrosion occurred at the GB/PFZ regions was attributed
the base materials in local regions [6–16]. The aroused LC susceptibility to the non-uniform distribution of MgZn2 phase within the TMAZ in
in joint then restricts the applications of FSW components that may FSW joint of AA7108-T79 alloy [23]. The mutation in microstructure
exposed to corrosive environment such as marine service and so on that occurs during the FSW can enhance the micro-galvanic effect and
[6,13]. It is thus of technical significance to investigate the corrosion thus leads to deteriorated corrosion resistance. For instance, previous
behavior and to address the influences from the modified studies note that the enhanced corrosion susceptibility of local regions


Corresponding author at: Key Laboratory of Superlight Materials and Surface Technology (Harbin Engineering University), Ministry of Education, Harbin,
150001, China.
E-mail address: qiuorwang@hrbeu.edu.cn (Y. Wang).

https://doi.org/10.1016/j.corsci.2019.108360
Received 27 May 2019; Received in revised form 25 September 2019; Accepted 19 November 2019
0010-938X/ © 2019 Elsevier Ltd. All rights reserved.

Please cite this article as: Yong Chen, et al., Corrosion Science, https://doi.org/10.1016/j.corsci.2019.108360
Y. Chen, et al. Corrosion Science xxx (xxxx) xxxx

of FSW joints in most cases is correlated with wide PFZs and coarsened Table 1
precipitates such as η (MgZn2) and S(Al2CuMg)for 7075 Al alloy and Chemical composition of 7050 aluminum alloy (wt.%).
2024 Al alloy, respectively [24]. It was also noted that the corrosion Zn Cu Mg Fe Si Mn Zr Al
behavior of the FSW joint was largely determined by the distribution of
η phases that developed during the FSW process [23]. A most recent 5.7-6.7 2.0-2.6 1.9-2.6 ≤0.15 ≤0.12 0.1 0.08-0.15 Balance
study on 2014 Al alloy FSW joints demonstrated that, the grain re-
finement of Al matrix in the NZ was positive to the passivation and thus
improved the corrosion resistance [22]; meanwhile the grain coar- sample surface. The schematic drawing of the welding process is shown
sening of the Al matrix in the HAZ is negative to the passivation and in Fig. 1.
should be responsible for the reduction of the corrosion resistance [22]. The joint was divided into the base material (BM), the heat-affected
Similar influence from grain refinement was also observed in AA6005 zone (HAZ), the thermo-mechanically affected zone (TMAZ) and the
Al alloy FSW joints [25]. Additionally, reducing the size of second nugget zone (NZ) according to the characteristic microstructure ob-
phases was found beneficial to the passivation and thus can lead to served on the cross section. The advancing HAZ side is marked 'HAZ-A'
improved corrosion resistance [25–27]. and the retreating HAZ side as 'HAZ-R'. The welded plate was naturally
In these previous studies, much attentions have been focused on the aged for more than 3 months before putting into the subsequent testing.
localized corrosion in FSW and the micro-galvanic effect was specifi- Small samples that were supposed to represent the local characteristics
cally addressed considering the dispersoids, precipitates and micro- of each zone were cut respectively from the zones of BM, HAZ and NZ
chemistry in local regions. The interactions between welding zone and according to microstructure regionalization.
the base materials, namely the macro-galvanic effects, however was
somewhat ignored. The circumstance deviates apparently from the re- 2.2. Microstructure characterization
quirement from the actual service of components from the FSW process,
since the heterogeneous microstructure in FSW joints cause not only Metallographic observation was carried out on samples cut from the
micro-galvanic corrosion but also macro-galvanic effect that strides joint transversely to welding direction. Samples were waterish grinded
across different regions. Wadeson [23] observed that the open circuit using successively low-to-high grades of SiC abrasive paper and then
potentials (OCP) and the corrosion resistance changed greatly with mechanically polished to mirror surface before etching in Keller’s re-
various location across the thermo-mechanically affected zone (TMAZ) agent for 30 s.
of FSW joints of AA7108-T79 alloy. Bousquet revealed that the zone Scanning electron microscope (SEM, FEI Quanta 200 F) equipped
close to TMAZ was the most corrosion susceptible region in FSW joint with energy dispersive X-ray (EDS, EDAX Genesis 2000) was used, in
fabricated from 2024-T3 alloy [3]. The increased susceptibility that the mode of secondary electron (SE) imaging or backscattered electron
observed in this region was believed related to the macro-galvanic (BSE) imaging, to identify the distribution and composition of the
coupling between the base material and the welded zone. Another study coarse dispersoids located in different zones, and to observe the mor-
also notes that the HAZ is the most corrosion susceptible zone for phology of corroded samples.
AA2014 alloy [22]. Padgett depicted the difference in OCP between Foils 80 μm in thickness were extracted from zone of NZ, BM and
sub-regions leads to significant macro-galvanic effect when the FSW HAZ (adjacent to TMAZ) respectively for TEM observing. The thin disc
joint was exposed to aggressive environment [28]. This macro-galvanic samples 3 mm in diameter acquired from the foils were prepared by
phenomenon was emphasized by Donford using the scanning reference twin jet electropolisher in a solution containing 30 vol.% nitric acid and
electrode technique on a FSW joint of aluminum alloy in 3.5 % NaCl 70 vol.% methanol at −25 °C. During TEM observation, bright field
solution [29]. The investigation on corrosion behavior of FSWed (BF) images were taken to characterize the morphology, EDS-line
AA2024 and AA2195 alloys however indicated that the welding zone scanning or EDS-mapping was carried out to examine the micro-
and the base metal exhibits similar corrosion potentials [30]. chemistry in selected regions.
Although the corrosion behavior was investigated considering the
galvanic effect in micro- and macro-scale as well, the study on corrosion 2.3. Corrosion test
behavior of FSW joints remains inadequate. Specifically, the relevance
as well as the cause of macro/micro galvanic effect remain to be clar- Samples for open circuit potential (OCP) test and potentiodynamic
ified. The present work was inspired to investigate the corrosion be- polarization (PDP) analysis were taken from the joint surface of each
havior of FSW joints fabricated from 7050-T76 Al alloy, aiming to zone and sealed by epoxy resin with area of 5 mm × 5 mm exposed,
clarify the interaction between galvanic effects in micro- and macro- then progressively ground to 2000 grit using SiC abrasive paper. The
scale, and to address the cause that resulted in the microstructure PDP analysis was conducted in 3.5 % NaCl solution in the range −300
heterogeneity. The analysis based on wire electrode beam test and the mV to 200 mV relative to the OCP with a scanning rate of 0.333 mV/s,
systematic comparison of corrosion configuration have revealed the using a three-electrode system: a platinum auxiliary electrode, a satu-
dominating influence of macro-galvanic effect on micro-galvanic effect. rated calomel reference electrode (SCE) and a working electrode.
The microstructure evolution combined with the redistribution of so- Samples were immersed in 3.5 % NaCl solution until the OCP was
lutes that occur during FSW process were observed to induce galvanic stable, whereat the PDP test was started. The OCP test for samples
effects both in micro- and macro-scales. extracted from each zone (BM, HAZ and NZ) lasted for 1 h and the OCP-
time curves were recorded.
2. Experimental The immersion corrosion tests in 3.5 % NaCl solution were con-
ducted on the whole joint and the separated samples extracted from
2.1. Samples preparation HAZ according to ASTM G44. Intergranular corrosion (IGC) tests were
conducted on the samples extracted from each zone (BM, HAZ and NZ)
The raw material used in the current work was 7050-T76 Al alloy and the whole FSW joint via immersing in 57 g/L NaCl +10 ml/L H2O2
plate, the nominal chemical composition of which is listed in Table 1. solution. The surface morphology or cross-sectional morphology of the
The welding samples (250 mm × 75 mm × 6.8 mm in size) was cut samples was observed using SEM after immersion corrosion and IGC
from the plate and cleaned thoroughly before putting into the FSW tests.
process. During the FSW process, the rotation rate of the tool was set at A wire beam electrode (WBE) was used for measuring the galvanic
800 rpm and the traverse speed at 200 mm/min. The tilt angle of the current and corrosion potential of the upper-surface of FSW joint. The
rotating tool was set at 2.5° with respect to the normal line of the WBE was composed of 30 micro electrode units each of them has an

2
Y. Chen, et al. Corrosion Science xxx (xxxx) xxxx

Fig. 1. Schematic diagram of friction stir welding. L for longitudinal (rolling), T for transverse and S for short-transverse direction of the raw plate, respectively.

Fig. 2. Diagrams of wire beam electrode used for measuring galvanic current and potential profile of the FSW joint surface (a), and for measuring galvanic current
and potential profile of the cross section of FSW joint (b).

exposed area of 2 mm × 5 mm, as shown in Fig. 2a. The WBE used for stirring action is believed to cause the grain coarsening [10,15,19].
measuring galvanic current and potential of the cross section of FSW Metallograph in Fig. 3d reveals that elongated grains 5∼20 μm in the
joint was composed of 30 × 3 micro electrode units, and each of them width distribute along the rotating direction of welding tool in the
was sealed by epoxy resin side by side with an exposed area 2 mm × 2 TMAZ, suggesting the severe deformation resulted from the stirring
mm (Fig. 2b). The galvanic current and potential were recorded with an action of the pin. The NZ, that located in the center zone of FSW joints,
interval of 9 s. The measurements were conducted in a thermostat bath consists of fine equiaxial grains 5−15 μm in the diameter, as seen in
containing 3.5 % NaCl solution at 25 ± 1 °C. Fig. 3e. These grains are reborn from the dynamic recrystallization
(DRX) that occurred in NZ [31]. Additionally, some fine particles are
observed to distribute linearly along the streamline of the matrix at the
3. Results
bottom of NZ, as indicated by white arrows in Fig. 3f. Besides, the
grains in this region, 4−7 μm in the diameter, are obviously finer than
3.1. Microstructure of the FSW joint
those in the center of NZ. The bottom of NZ has lower temperature rise
and shorter cooling time than the center of NZ and thus suffers lower
The microstructure of the various zones of the FSW joint, namely
centrifugal force, which then leads to incomplete developed re-
the BM, HAZ, TMAZ and NZ, was observed via metallurgical micro-
crystallization [32].
scope and the representative morphology was shown in Fig. 3. Fig. 3a
The microstructure in samples extracted from various zones in the
displays the 3D optical micrograph of the whole weld joint, revealing
joint were further examined via SEM observation using backscattered
the heterogeneous feature of the metallurgical structure in the FSW
electron (BSE) model. As demonstrated in Fig. 4, the segregation of
joint. Fig. 3b shows that elongated grains 10∼15 μm in the width
solutes outlines the grain boundaries (GBs) in all three zones. The shape
distribute along the rolling direction of the plate, indicating the fibrous
and size of grains displayed in SEM-BSE images are consistent with that
characteristic of grains in the BM. Grains in HAZ are also fibrous but
in metallograph. The SEM-BSE observation also reveals the morphology
larger (10∼25 μm in the width) than those in the BM (Fig. 3c), sug-
and the distribution of dispersoids and precipitates. Precipitates a few
gesting that the grains within this zone coarsen slightly during the FSW
ten nanometers in diameter can be observed on GBs in the BM (Fig. 4a).
process. The intense temperature rise in HAZ induced by the friction-

3
Y. Chen, et al. Corrosion Science xxx (xxxx) xxxx

Fig. 3. Optical observation of FSW joint of 7050-T76 Al alloy after etching in Keller’s reagent: (a) 3D image of the whole welded joint, (b) the BM, (c) the HAZ, (d) the
TMAZ, (e) the middle of NZ, (f) the bottom of NZ.

Comparative observation suggests that GB precipitates (GBPs) in HAZ in Fig. 5a, dispersoids distribute along the rolling direction in the zone
are apparently larger than those in BM, as shown in Fig. 4b. Precipitates of BM. The corresponding EDS maps demonstrate that these particles
or solute segregation can also be recognized on GBs of recrystallized are composed mainly of Al, Cu and Fe, and thus verified as Al7Cu2Fe
grains in NZ (Fig. 4c), confirming the solutes’ enrichment. Dispersoids particles referring to previous studies [33,34]. The particles in HAZ and
can be observed in each zone but the morphology varies from zone to NZ present the similar composition to those in BM (Fig. 5b and c), in-
zone. Precipitates are presumably believed to exist on the matrix; dicating that Al7Cu2Fe particles retains in all zones of FSW joint despite
however, they cannot be recognized via the SEM-BSE observation due the FSW process. The morphology of Al7Cu2Fe dispersoids in HAZ
to the limitation in magnification. (Fig. 5b) is similar to that of particles in BM, however Al7Cu2Fe dis-
SEM-EDS analysis was carried out to examine the composition of persoids in NZ present quite different morphology, as demonstrated in
dispersoids. Fig. 5 shows the micrographs and the EDS maps of the Fig. 5c. The pre-existing Al7Cu2Fe particles in NZ were fragmented and
particles in different zones on the upper surface of FSW joint. As shown redistributed due to severe plastic deformation during FSW and the

Fig. 4. BSE images of microstructure in the various zones of FSW joint: (a) BM, (b) HAZ, (c) NZ.

4
Y. Chen, et al. Corrosion Science xxx (xxxx) xxxx

Fig. 5. EDS maps of dispersoids in various zones of FSW joint: (a) BM, (b) HAZ, (c) NZ.

5
Y. Chen, et al. Corrosion Science xxx (xxxx) xxxx

Fig. 6. TEM bright-field images of precipitates within the matrix or on GBs in various zones of FSW joint, taken along < 011 > Al: (a) and (b) BM, (c) and (d) HAZ, (e)
and (f) NZ.

Fig. 7. The 3D microhardness profile of the FSW joint of 7050-T76 Al alloy.

crushed particles distribute uniformly in the recrystallized matrix. with the coarsened precipitates on GBs. The similar characteristics are
The precipitates, either within the matrix or on GBs were examined also found in some previous studies [31,35,36]. These originally existed
via TEM observation and the BF images taken from representative site precipitates, both within the matrix and on GBs, evolve dramatically
of each zone are shown in Fig. 6. As exhibited in Fig. 6a and b, the into coarsened ones in zone of HAZ (Fig. 6c and d), due to the strong
7050-T76 Al alloy, used as BM in the study, is found to contain high temperature rise induced by the frication and stirring within the zone.
density of disc-like η' or η phases (MgZn2) within the matrix together In addition, PFZs in HAZ are obviously wider to compare with those in

6
Y. Chen, et al. Corrosion Science xxx (xxxx) xxxx

Fig. 8. The distribution of solutes across representative GBPs in BM (a) and HAZ (b), obtained via EDS-line scanning in TEM analysis.

BM. The coarsening of precipitates contributes to the decreases of 3.2. Corrosion behavior of the FSW joint
hardness in HAZ, as demonstrated in Fig. 7, which is consistent with
that previously noted [37]. Fig. 6e shows that precipitate can hardly be 3.2.1. Corrosion behavior of the various zones of the joint
observed within grains in NZ. Also, the number of GBPs is much less The electrochemical corrosion behavior of different zones of FSW
than that in HAZ and BM, as shown in Fig. 6f, despite the evident solute joint, namely BM, HAZ and NZ, was evaluated separately via the testing
segregation (as seen in Fig. 4c). The pre-existing precipitates are be- on samples that extracted from various zones of joint. The OCP and PDP
lieved to dissolve now that exposed to the violently elevated tempera- curves of BM, HAZ, and NZ zone are shown in Fig. 10a and 10b, re-
ture in the NZ. On the other hand, solutes can transfer quickly with spectively. The OCP of sample extracted from BM, HAZ and NZ is
dislocations’ slipping and deposit eventually on the newly generated around -0.51 V, -0.55 V and -0.53 V respectively. The fluctuation of
GBs during the recrystallization [38]. Despite the re-dissolution and OCP is intense in case of BM sample but much slighter for HAZ and NZ
segregation of solutes, precipitation can hardly occur in the NZ, either sample. PDP curves test from samples extracted from different regions
within recrystallized grains or on GBs since the cooling is so quickly and presents similar overall profile (Fig. 10b). The cathodic current density
the temperature window for precipitation is very narrow. The climbing was approximately 6☓10−6A·cm-2 for each region despite the diversity
back of microhardness observed in NZ (Fig. 7) is mainly attributed to in the microstructure. This similarity in PDP indicates similar corrosion
the grain refinement strengthening and the high sensitivity to natural behavior for these three regions in FSW joint when tested separately.
aging [6,37]. The HAZ sample shows the lowest corrosion potential while the BM
The coarsening of precipitates in HAZ is accompanied by the re- possesses the highest one; this is compatible with the sequence that
distribution of solutes, especially the segregation of Cu into GBPs. EDS observed in OCP.
analysis in Fig. 8 indicates that the content of Cu in GBPs in HAZ (up to The IGC behavior of samples extracted from the zones of BM, HAZ
13.5 at.%) is much higher than that in BM (below 5.7 at.%). Similar and NZ was examined by immersing these samples individually into
behavior of Cu enrichment in precipitates was previously observed in Al IGC solution. The cross-section of samples immersed for 6 h was ob-
alloys when exposed to elevated temperature [9]. The enrichment of Cu served using SEM and the representative morphology is shown in
in GBPs inevitably leads to Cu’s depletion in the PFZ since the short- Fig. 11. Attacking along GBs together with the spalling of grains can be
range diffusion in local region can be evident [39]. Although the de- observed on the corroded surface of each sample, which suggests that
pletion of Cu can hardly be recognized by the EDS analysis, it can all zones in FSW joint are susceptible to IGC when put into IGC test
contribute to the decrease in corrosion potential. On the other hand, the individually. The susceptibility, however, is different for these three
enrichment of Cu into η/η’ precipitates within the matrix can also be regions. The IGC is around 60, 140 and 110 μm in depth for the samples
evident since the diffusion of Cu can be evidently accelerated when extracted from BM, HAZ and NZ, respectively, indicating that the HAZ
elevated temperature (200 ∼ 300 °C) emerges in HAZ [40]. This en- exhibits the highest susceptibility and the BM exhibits the lowest.
richment of Cu into η/η’ phases depletes Cu in the matrix. The IGC susceptibility is dominated by the micro galvanic effect that
As mentioned in Fig. 3f, some fine particles are found to distribute induced by the inhomogeneous distribution of solutes within the
linearly at the bottom of NZ, which then was further examined. Fig. 9a narrow regions along GBs. The intense temperature rise in HAZ during
shows the cross-sectional micrograph of NZ. A representative zone, FSW can induce GBPs’ coarsening and the solutes’ segregation to GBs,
labeled as A in Fig. 9a, was selected to further examine the micro- as the EDS analysis suggested (as seen in Fig. 8). The coarsening of GBPs
structure in the bottom of NZ. The TEM image in Fig. 9c presents the and the resulted depletion of Cu in adjacent regions leads to increased
characteristics of fine uniaxial grains as well as the fine particles dis- potential difference between the matrix and the GBs and thereafter
tributed along GBs. The EDS maps of various elements suggest that the enhanced the local micro galvanic effect. Similar phenomenon was
particle labeled as B is rich in Cu and Fe and thus determined to be previously observed in 7075 alloy in Mahoney’s research [39]. The
Al7Cu2Fe dispersoids. The Al7Cu2Fe dispersions observed here on GBs research notes that the precipitate coarsening and resulted Cu depletion
in NZ are smaller in size than those observed in HAZ and BM (Fig. 4). along GBs promoted by the thermal excursion during welding should
These fine particles are believed to be fragmentized from pre-exiting contribute to the enhanced sensitivity to IGC [43]. On the other hand,
Al7Cu2Fe dispersoids and then accumulate along the treamline of alloy the intense thermal-mechanical process in the center region of FSW,
during the FSW. The schematic diagram showed in Fig. 9b [41,42] namely the NZ, induces sufficient recrystallization and the accom-
reflects the material flow in FSW process, in which the accumulation of panying redistribution of solutes. Especially, the deposition of solutes
fragmentized dispersoids can be inferred. The particle marked C mainly on GBs is evident in NZ, as indicated in the Fig. 4. The increased dif-
contains elements of Mg, Zn, Cu indicating the presence of η phase. ference in solute content between the GBs and grain interior then in-
duces aggravated corrosion along GBs.

7
Y. Chen, et al. Corrosion Science xxx (xxxx) xxxx

Fig. 9. Metallographic image of the cross-section of NZ (a), schematic diagram of flow model of plastic material in the NZ (b), TEM image of zone A in the bottom of
NZ and the EDS maps of various elements (c).

Fig. 10. OCP-time curves (a) and potentiodynamic polarization curves (b) tested on the separated samples extracted from BM, HAZ and NZ of FSW joint, respectively.

8
Y. Chen, et al. Corrosion Science xxx (xxxx) xxxx

Fig. 11. The cross-sectional morphology of the various separated zones after IGC immersion test for 6 h in 57 g/L NaCl +10 ml/L H2O2 solution.

3.2.2. Corrosion behavior of the whole FSW joint effect spanning across zones of FSW joint. Fig. 14 shows the variation of
3.2.2.1. The three-dimensional potential profile of FSW joint. The 3D OCP OCP and galvanic current over time at various zones. The potential of
profile of the FSW joint was figured out based on the WEB testing. As the BM fluctuates in the range of -460~-510 mV (vs. SCE) while the
shown in Fig. 12, the HAZ presents the lowest OCP (−580 mV, vs. potential of the NZ fluctuates slightly and keeps at the level of around
SCE), which is apparently lower than that of BM (−500 mV, vs. SCE) −530 mV (vs. SCE). The HAZ presents the lowest potential and
and NZ (−530 mV, vs. SCE) suggesting the highest corrosion specifically very low potential of −580 mV is observed in the narrow
susceptibility. In addition, OCP as low as −580 mV is also observed regions adjacent to NZ, as seen in Fig. 14a. The potential in the center of
at the bottom of NZ, which is similar to that observed in Paglia’s NZ is near to that of BM, but slightly lower in the boundary regions in
research [9]. The OCP of HAZ tested in the whole joint is apparently both sides. The potential of each region fluctuates slightly as the test
lower than that tested in the separated HAZ sample, while the OCP of proceeds, but doesn’t vary apparently all through the test. The galvanic
BM tested in the whole joint is higher than that tested in the separated current increases first to a peak level and then decreases slightly and
BM sample. This oppositely directed shift in OCP in HAZ and BM remains at a level all through the test (Fig. 14b). A strong negative
suggests a macro-galvanic effect that spans across the different zones in current as high as 0.03 mA·cm−2 is observed at the narrow region along
FSW joint. HAZ/NZ boundaries on both sides, revealing the intense anodic effect.
The 3D morphology of the FSW joint after immersing for 3 h in NaCl The lower potential together with the strong negative current in these
solution is shown in Fig. 13. It can be seen that the white corrosion boundary regions suggest an evident and long-lasting macro galvanic
products cover some local zones of the joint, indicting the preferential effect that occurs in the FSW joint.
occurrence of corrosion in HAZ and the bottom of NZ. It is obvious that The surface morphology of the post-immersed FSW joint is shown in
the distribution of preferential corrosion is highly consistent with the Fig. 15. Apparently, serious corrosion is located at the regions 10 mm
distribution of low potential regions that shown in the 3D potential from the center, which is corresponding to the HAZ that possesses the
profile (Fig. 12). This consistency in distribution also suggests an effect lowest potential. Meanwhile, white spots 50 ∼ 200 μm in diameter can
of macro galvanic. be observed in NZ, which are actually corrosion pits. Obvious change
can hardly be observed in the region of BM, which suggests that this
3.2.2.2. Macro-galvanic corrosion effect. The WEB was employed to test region is protected from corrosion. The distribution of corrosion on the
the local OCP and galvanic current and to examine the macro-galvanic whole welded joint is highly consistent with the distribution of the

Fig. 12. The 3D OCP profile of the FSW joint after immersing in 3.5 % NaCl solution.

9
Y. Chen, et al. Corrosion Science xxx (xxxx) xxxx

Fig. 13. The 3D morphology of the FSW joint immersed in 3.5 % NaCl solution for 3 h. The corrosion products were not removed before observation.

lowest potential and negative galvanic current regions in Fig. 14. This morphology observed on the corroded surface and cross-section are
indicates a dominating influence from macro-galvanic effect on the shown in Fig. 17. Characteristics of IGC can be observed on the cor-
corrosion when the whole FSW joint is exposed to the corrosive en- roded surface in region of BM, but can hardly be observed on the ap-
vironment. positional cross-section (displayed in the left column of Fig. 17), in-
In-situ SEM observation was carried out to record the pitting evo- dicating the slight IGC in the BM. Obvious characteristics of IGC are
lution and reveal the influence of macro-galvanic effect in different seen on both the surface and the cross section in zone of HAZ (shown in
zones of FSW joint. Fig. 16 shows the surface morphology of the dif- the middle column of Fig. 17) indicating the severe IGC in the zone. The
ferent zones of the whole FSW joint and separated HAZ sample after IGC is found to extend horizontally along the grain boundaries; and
immersion. It can be seen that few pitting can be observed in the BM of that, the depth of IGC reaches to about 200 μm. The propagation of
whole-Joint before the immersion proceeds to 2 h. Very small corrosion corrosion along GBs together with spalling is also observed in the re-
pits with diameter below 1 μm can be observed after immersing for 0.5 gion of NZ, as shown in the right column of Fig. 17. The depth of IGC in
h in the BM, indicating the initiation of pitting at the early stage. These the NZ is about 110 μm and the sectional profile is of semi-ellipse shape,
pits develop very slowly as time prolongs, as shown in the first line of which is different from that in HAZ. To compare Fig. 17 with Fig. 11, it
Fig. 16. The corrosion in HAZ however is quite different. Severe pitting can be found that the susceptibility to IGC decreases apparently in BM,
can be observed in HAZ after the whole FSW joint was immersed for decreases slightly in the NZ but intensifies evidently in HAZ, when the
only 0.5 h. Besides the matrix adjacent to Al7Cu2Fe dispersoids, pits are FSW joint was put into immersion test as whole. The change of IGC
also observed in adjacent matrix, which can be related to the coarsened susceptibility in the two cases suggests the significant influence from
η’/η precipitates (seen in Fig. 6c and 6d). The pitting in HAZ develops the macro galvanic effect.
continuously as the immersion proceeds with a lot of corrosion products
generated. Additionally, a few cracks shape up on the matrix around the 4. Discussion
Al7Cu2Fe particles when the immersion time is prolonged to 6 h. The
corrosion behavior in HAZ that observed on the whole FSW joint is As the BM in this study, the 7050-T76 Al alloy contains high density
quite different from that occurs in the separated HAZ sample (shown in of η'/η (Mg(Zn, Cu, Al)2) [31], as demonstrated in Fig. 6a and b. These
the last line of Fig. 16), where pitting initiates only at sites of Al7Cu2Fe precipitates coarsen in HAZ (Fig. 6c and d) but dissolve in NZ (Fig. 6e
particles and develops much more slowly. The difference in corrosion and f) due to stirring-induced temperature rise in local regions. The
behavior of HAZ, in separated sample and in whole joint, indicates the enrichment of copper into precipitates, either in matrix within grains or
evident influence from the macro galvanic effect. Pitting can also be on GBs, accompanied by the intense coarsening, occurs due to the
recognized in region of NZ when the FSW joint was immersed for 0.5 h, temperature rise in HAZ. The migration of copper results in the de-
as shown in the third line of Fig. 16, which indicates that the suscept- pletion of Cu in the matrix, which then induces the decrease in OCP in
ibility to pitting is higher in the NZ than in the BM, but far below that in HAZ (Figs. 10 and 12) since the electrode potential of Cu is much higher
HAZ. than that of aluminum [44,45]. For the NZ, the dissolution of pre-
The IGC behavior of the whole FSW joint was examined via the cipitates releases Mg, Zn and Cu simultaneously into the matrix, which
immersion test in the 57 g/L NaCl +10 ml/L H2O2 solution. The results in a moderate potential between that of HAZ and BM due to the

Fig. 14. Variation curves of local OCP (a) and galvanic current (b) with the immersing time in 3.5 % NaCl solution.

10
Y. Chen, et al. Corrosion Science xxx (xxxx) xxxx

Fig. 15. The surface morphology of the whole FSW joint after immersion in 3.5 % NaCl solution for different duration.

combined effect from Mg/Zn and Cu on the corrosion potential. The the oxygen reduction and diffusion. For this particular case, all of their
lower OCP detected in the bottom of NZ (as seen in Fig. 12) can be corrosion rates are determined by oxygen diffusion and the ratio of
attributed to the solute segregation to GBs that related to the stirring- anodic area to cathodic area [46,47]:
induced dynamic recrystallization. Icorr1 = IL (1)
From the typical polarization curves of those separated zones
(Fig. 10b), it can be seen that all of the potentiodynamic polarization Ia1 A
γ= =1+ 2
curves present the same overall shape with a cathodic plateau related to Icorr1 A1 (2)

Fig. 16. SEM morphology of the separated HAZ sample and the different zones in the whole FSW joint, after immersing in 3.5 % NaCl solution for different duration.

11
Y. Chen, et al. Corrosion Science xxx (xxxx) xxxx

Fig. 17. Surface and cross-sectional morphology of BM, HAZ and NZ after the whole FSW joint was immersed in 57 g/L NaCl +10 ml/L H2O2 solution for 6 h for IGC
test.

A2 is also enhanced by macro-galvanic effect. Therefore, pitting corrosion


Ia1 = Ig (1 + )
A1 (3) also occurs at the region that is not adjacent to the Al7Cu2Fe phase, as
shown in Fig. 16. For IGC process, IGC in the cathodic region (BM) is
where γ is the contact corrosion effect, Ia1 is the anode dissolves current greatly suppressed while IGC in the anodic region (HAZ) is enhanced
density after contact, Icorr1 is the anode dissolves current density before greatly. As a brief summary, the micro-galvanic corrosion induced by
contact, A1 is the area of lower potential, A2 is the area of higher po- heterogeneous microchemistry is greatly enhanced by the macro-gal-
tential, IL is limited diffusion current density, Ig is the measured value of vanic effect across the whole FSW joint.
galvanic current.
According to Eqs. (1)–(3), if the cathodic reaction is fully controlled
5. Conclusions
by diffusion-reaction during corrosion, the effect of impurity phase on
the corrosion rate can be neglected. In other words, the cathodal im-
(1) The pre-existing microstructure mutates greatly when the 7050-T76
purity phase Al7Cu2Fe in different zones has little influence on accel-
Al alloy was subjected to FSW process. Precipitates in HAZ coarsen
erating corrosion of the metallic matrix due to their small area
evidently, and this induces enrichment of Cu in precipitates but
(equivalent to A2). This can be further demonstrated by the fact that
depletion in the matrix. Precipitates’ dissolving as well as DRX in-
there is no serious pitting between Al7Cu2Fe and matrix from SEM
duce solute segregation on the recrystallization GBs in NZ. Al7Cu2Fe
observation (Fig. 16). However, when the area of BM is considered as
dispersoids in NZ are fragmented and located at the recrystalliza-
A2 (which is far larger than that of Al7Cu2Fe phase) and the area of HAZ
tion GBs.
as A1, the macro-galvanic corrosion rate of HAZ (Ia1) will be greatly
(2) The HAZ has the lowest OCP and the BM presents the highest one,
accelerated according to Eq. (3).
according to OCP test on the separately extracted samples as well as
From Eq. (3), the corrosion rate Ia1 is proportional to the Ig when the
the test on the whole joint. The difference in potential thus induce
area of A1 and A2 is constant. That is, the galvanic current is propor-
intensive macro-galvanic effect between the different zones in the
tional to the corrosion rate. The larger negative galvanic current in the
joint; the boundary region of HAZ adjacent to NZ receives the most
HAZ close to NZ indicates a higher corrosion rate, leading to severe
intense negative current and thus presents the highest susceptibility
corrosion (Figs. 15–17). Therefore, both pitting corrosion and IGC in
to corrosion.
HAZ were promoted greatly by the macro-galvanic effect. Pitting cor-
(3) The pitting, basically induced by Al7Cu2Fe dispersoids, aggravates
rosion in 7xxx Al alloy by itself is generally induced by the micro-gal-
greatly when the whole joint is exposed to immersion test. The
vanic effect between the second particles (Al7Cu2Fe) and the adjacent
coarsened MgZn2 precipitates cannot induce pitting in the case of
aluminum matrix, in which Al7Cu2Fe phase serves as cathode and the
separately extracted HAZ, however they can induce severe pitting
surrounding matrix as anode [17]. IGC in 7xxx Al alloy is generally
when the joint is put into immersion test as a whole.
induced by the micro-galvanic effect between precipitates on GBs and
(4) The macro-galvanic effect greatly aggravates IGC in HAZ but ap-
matrix/PFZ, where precipitates serve as anode and the surrounding
parently suppresses IGC in BM, and it has a slight influence on IGC
matrix/PFZ as cathode. However, both of these micro-galvanic effects
in NZ.
can be influenced by the macro-galvanic effect induced by the potential
(5) The pitting and IGC resulted from micro-galvanic corrosion will
difference between various zones of the FSW joint. The HAZ adjacent to
change greatly by macro-galvanic effect, both in intensity and in
NZ that presents the lowest potential serves as the anodic region, and
path. The macro-galvanic effect is believed more dominating com-
the BM that presents the highest potential serves as the cathodic region.
paring with the micro-galvanic effect for corrosion of FSW weld-
Therefore, an intense current in the HAZ/NZ boundary was induced
ment.
once the FSW joint was immersed into NaCl solution as a whole. This
macro-galvanic current can be merged into the current of micro-gal-
vanic corrosion, and thus leads to enhanced corrosion. For pitting Data availability statement
process, beside the susceptibility related to Al7Cu2Fe phase, the micro-
galvanic corrosion induced by the coarsened anodic MgZn2 precipitates The raw/processed data required to reproduce these findings cannot

12
Y. Chen, et al. Corrosion Science xxx (xxxx) xxxx

be shared at this time as the data also forms part of an ongoing study. nanoparticles for improved corrosion resistance of AA 6061, Trans. Inst. Met.
Finish. 89 (2011) 320–324.
[21] F. Andreatta, H. Terryn, J.H.W. de Wit, Corrosion behaviour of different tempers
Declaration of Competing Interest ofAA7075 aluminum alloy, Electrochim. Acta 49 (2004) 2851–2862.
[22] S. Sinhmar, D.K. Dwivedi, A study on corrosion behavior of friction stir welded and
The authors declare that there are no conflicts of interest to this tungsten inert gas welded AA2014 aluminum alloy, Corros. Sci. 133 (2018) 25–35.
[23] D.A. Wadeson, X. Zhou, G.E. Thompson, P. Skeldon, L.D. Oosterkamp, G. Scamans,
work. Corrosion behaviour of friction stir welded AA7108 T79 aluminum alloy, Corros.
Sci. 48 (2006) 887–897.
Acknowledgments [24] C.G. Rhodes, M.W. Mahoney, W.H. Bingel, R.A. Spurling, C.C. Bampton, Effects of
friction stir welding on microstructure of 7075 aluminum, Scripta Mater. 36 (1997)
69–75.
The authors wish to acknowledge the financial supports from the [25] M. Esmaily, N. Mortazavi, W. Osikowicz, H. Hindsefelt, J.E. Svensson,
National Natural Science Foundation of China (No. 51971073, M. Halvarsson, G.E. Thompson, L.G. Johansson, Corrosion behaviour of friction stir-
welded AA6005-T6 using a bobbin tool, Corros. Sci. 111 (2016) 98–109.
51771061 and U1737206), the Fundamental Research Funds for the
[26] M. Esmaily, N. Mortazavi, W. Osikowicz, H. Hindsefelt, J.E. Svensson,
Central Universities (No. HEUCFG201838), the Open Project Program M. Halvarsson, G.E. Thompson, L.G. Johansson, Influence of multi-Pass friction stir
of Key Laboratory of Preparation and Application of Environmental processing on the corrosion behavior of an Al-Mg-Si alloy, J. Electrochem. Soc. 163
Friendly Materials (Jilin Normal University), Ministry of Education, (2016) C124–C130.
[27] M. Esmaily, D.B. Blücherb, J.E. Svensson, M. Halvarsson, L.G. Johansson, Achieving
China (No. 2017012) superior corrosion resistance in friction stir processed AA6005-T6 aluminum alloy
joints, ECS Trans. 64 (2015) 29–43.
References [28] Z.Y. Ma, R.S. Mishra, M.W. Mahoney, Friction Stir Welding and Processing II, TMS,
Warrendale, PA, 2003, pp. 221–230.
[29] M. Donford, R. Ding, A Study of Friction Stir Welded 2195 A1-Li Alloy by the
[1] R.S. Mishra, Z.Y. Ma, Friction stir welding and processing, Mater. Sci. Eng. R 50 Scanning Reference Electrode Technique, NASA/TP-1998-207399, Marshall Space
(2005) 1–78. Flight Center, AL, 1998.
[2] G. Cam, S. Mistikoglu, Recent developments in friction stir welding of Al-alloys, J. [30] J. Corral, E.A. Trillo, Y. Li, L.E. Murr, Corrosion of friction-stir welded aluminum
Mater. Eng. Perform. 23 (2014) 1936–1953. alloys 2024 and 2195, J. Mater. Sci. Lett. 19 (2000) 2117–2122.
[3] E. Bousquet, A. Poulon-Quintin, M. Puiggali, O. Devos, M. Touzet, Relationship [31] J.Q. Su, T.W. Nelson, R. Mishra, M. Mahoney, Microstructural investigation of
between microstructure, microhardness and corrosion sensitivity of an AA 2024-T3 friction stir welded 7050-T651 aluminum, Acta Mater. 51 (2003) 713–729.
friction stir welded joint, Corros. Sci. 53 (2011) 3026–3034. [32] W.F. Xu, J.F. Liu, G.F. Luan, C.L. Dong, Temperature evolution, microstructure and
[4] S. Li, H.G. Dong, L. Shi, P. Li, F. Ye, Corrosion behavior and mechanical properties mechanical properties of friction stir welded thick 2219-O aluminum alloy joints,
of Al-Zn-Mg aluminum alloy weld, Corros. Sci. 123 (2017) 243–255. Mater. Des. 30 (2009) 1886–1893.
[5] P. Bala Srinivasan, K.S. Arora, W. Dietzel, S. Pandey, M.K. Schaper, Characterisation [33] J.C. Bertoncello, S.M. Manhabosco, L.F. Dick, Corrosion study of the friction stir lap
of microstructure, mechanical properties and corrosion behaviour of an AA2219 joint of AA7050-T76511 on AA2024-T3 using the scanning vibrating electrode
friction stir weldment, J. Alloys Compd. 492 (2010) 631–637. technique, Corros. Sci. 94 (2015) 359–367.
[6] P.L. Threadgill, A.J. Leonard, H.R. Shercliff, P.J. Withers, Friction stir welding of [34] J.A. Moreto, C.E.B. Marino, W.W. Bose Filho, L.A. Rocha, J.C.S. Fernandes, SVET,
aluminium alloys, Int. Mater. Rev. 54 (2009) 49–93. SKP and EIS study of the corrosion behaviour of high strength Al and Al–Li alloys
[7] C.S. Paglia, K.V. Jata, R.G. Buchheit, A cast 7050 friction stir weld with scandium: used in aircraft fabrication, Corros. Sci. 84 (2014) 30–41.
microstructure, corrosion and environmental assisted cracking, Mater. Sci. Eng. A [35] W.C. Yang, S.X. Ji, Q. Zhang, M.P. Wang, Investigation of mechanical and corrosion
424 (2006) 196–204. properties of an Al-Zn-Mg-Cu alloy under various ageing conditions and interface
[8] O. Hatamleh, P.M. Singh, H. Garmestani, Corrosion susceptibility of peened friction analysis of η’ precipitate, Mater. Des. 85 (2015) 752–761.
stir welded 7075 aluminum alloy joints, Corros. Sci. 51 (2009) 135–143. [36] J.T. Jiang, W.Q. Xiao, L. Yang, W.Z. Shao, S.J. Yuan, L. Zhen, Ageing behavior and
[9] C.S. Paglia, L.M. Ungaro, B.C. Pitts, The corrosion and environmentally assisted stress corrosion cracking resistance of a non-isothermally aged Al–Zn–Mg–Cu alloy,
cracking behavior of high strength aluminum alloys friction stir welds: 7075-T651 Mater. Sci. Eng. A 605 (2014) 167–175.
vs. 7050-T7451, Friction Stir Welding and Processing II, (2003), pp. 65–75. [37] M. Dumont, A. Steuwer, A. Deschamps, M. Peel, P.J. Withers, Microstructure
[10] C.S. Paglia, R.G. Buchheit, The time-temperature-corrosion susceptibility in a 7050- mapping in friction stir welds of 7449 aluminum alloy using SAXS, Acta Mater. 54
T7451 friction stir weld, Mater. Sci. Eng. A 492 (2008) 250–254. (2006) 4793–4801.
[11] J.E. Kertz, P.I. Gouma, R.G. Buchheit, Localized corrosion susceptibility of Al-Li-Cu- [38] S.S. Wang, J.T. Jiang, G.H. Fan, A.M. Panindre, G.S. Frankel, L. Zhen, Accelerated
Mg-Zn alloy AF/C458 due to interrupted quenching from solutionizing tempera- precipitation and growth of phases in an Al-Zn-Mg-Cu alloy processed by surface
ture, Mater. Trans. A 32 (2001) 2561–2573. abrasion, Electrochim. Acta 131 (2017) 233–245.
[12] V. Proton, J. Alexis, E. Andrieu, J. Delfosse, M.C. Lafont, C. Blanc, Characterisation [39] J.B. Lumsden, M.W. Mahoney, G. Pollock, C.G. Rhodes, Intergranular corrosion
and understanding of the corrosion behaviour of the nugget in a 2050 aluminum following friction stir welding of aluminum alloy 7075-T651, Corrosion 55 (1999)
alloy friction stir welding joint, Corros. Sci. 73 (2013) 130–142. 1127–1135.
[13] M. Guérin, E. Andrieu, G. Odemer, J. Alexis, C. Blanc, Effect of varying conditions of [40] J.F. Chen, X.F. Zhang, L.C. Zou, Y. Yu, Q. Li, Effect of precipitate state on the stress
exposure to an aggressive medium on the corrosion behavior of the 2050 Al-Cu-Li corrosion behavior of 7050 aluminum alloy, Mater. Charact. 114 (2016) 1–8.
alloy, Corros. Sci. 85 (2014) 455–470. [41] Y.Q. Mao, L.M. Ke, F.C. Liu, Y.H. Chen, L. Xing, Effect of tool pin-tip profiles on
[14] K. Surekha, B.S. Murty, K. Prasad Rao, Effect of processing parameters on the material flow and mechanical properties of friction stir welding thick AA7075-T6
corrosion behaviour of friction stir processed AA 2219 aluminum alloy, Solid State alloy joints, Int. J. Adv. Manuf. Technol. 88 (2017) 949–960.
Sci. 11 (2009) 907–917. [42] Y. Morisada, T. Imaizumi, H. Fujii, M. Matsushita, R. Ikeda, Three-dimensional
[15] R.K. Gupta, H. Das, T.K. Pal, Influence of processing parameters on induced energy, visualization of material flow during friction stir welding of steel and aluminum, J.
mechanical and corrosion properties of FSW butt joint of 7475 AA, J. Mater. Eng. Mater. Eng. Perform. 23 (2014) 4143–4147.
Perform. 21 (2012) 1645–1654. [43] C.S. Paglia, R.G. Buchheit, A look in the corrosion of aluminum alloy friction stir
[16] M. Jariyaboon, A.J. Davenport, R. Ambat, B.J. Connolly, S.W. Williams, D.A. Price, welds, Scripta Mater. 58 (2008) 393-387.
The effect of welding parameters on the corrosion behaviour of friction stir welded [44] F.X. Song, X.M. Zhang, S.D. Liu, Q. Tan, D.F. Li, The effect of quench rate and
AA2024–T351, Corros. Sci. 49 (2007) 877–909. overageing temper on the corrosion behaviour of AA7050, Corros. Sci. 78 (2014)
[17] P.A. Dick, G.H. Knörnschild, L.F.P. Dick, Anodising and corrosion resistance of AA 276–286.
7050 friction stir welds, Corros. Sci. 114 (2017) 28–36. [45] S.D. Liu, B. Chen, C.B. Li, Y. Dai, Y.L. Deng, X.M. Zhang, Mechanism of low ex-
[18] K. Prasad Rao, G.D. Janaki Ramb, B.E. Stucker, Improvement in corrosion resistance foliation corrosion resistance due to slow quenching in high strength aluminium
of friction stir welded aluminum alloys with micro arc oxidation coatings, Scripta alloy, Corros. Sci. 91 (2015) 203–212.
Mater. 58 (2008) 998–1001. [46] F. Mansfeld, Area relationships in galvanic corrosion, Corrosion 27 (1971)
[19] R.D. Fu, Z.Q. Sun, R.C. Sun, Y. Li, H.J. Liu, L. Liu, Improvement of weld temperature 436–442.
distribution and mechanical properties of 7050 aluminum alloy butt joints by [47] F. Mansfeld, J. Kenkel, Galvanic corrosion of Al alloys-III. The effect of area ratio,
submerged friction stir welding, Mater. Des. 32 (2011) 4825–4831. Corros. Sci. 15 (1975) 239–250.
[20] S.L. Zhang, M.M. Zhang, Y. Yao, F. Sun, Use of silane films modified with Y2O3

13

You might also like