You are on page 1of 10

Materials and Design 90 (2016) 777–786

Contents lists available at ScienceDirect

Materials and Design

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

Hybrid welding of AA5754 annealed alloy: Role of post weld heat


treatment on microstructure and mechanical properties
Paola Leo a,⁎, Sonia D'Ostuni a, Giuseppe Casalino b
a
Innovation Eng. Dept., University of Salento, Via per Arnesano s,n, 73100 Lecce, Italy
b
DMMM, Politecnico di Bari, Viale Japigia, 182, Bari, Italy

a r t i c l e i n f o a b s t r a c t

Article history: In this paper hybrid welding of the annealed AA5754 alloy was studied and the role of post weld heat treatment
Received 23 July 2015 (PWHT) on microstructure and mechanical properties was investigated. The microstructure and hardness of the
Received in revised form 27 October 2015 base material and welds were analyzed. Due to the segregation, the lowest microhardness value was observed at
Accepted 28 October 2015
the welding centerline. High microhardness was registered in a narrow zone at the interface between the fusion
Available online 29 October 2015
zone and the heat affected zone where the welding cycle promotes both the strengthening by Mg solid solution
Keywords:
and the lowest grain size. The hardness of the heat affected zone was higher than that of the base material due to
Hybrid welding dissolution of soluble particles (Mg2Si and Mg2Al3). PWHT at 350 °C increased the microhardness in the fusion
AA5754 zone and the tensile strength with respect to the untreated joint, whereas the strain at fracture was similar show-
Welding defects ing that the fracture was promoted mainly by the presence of porosity, coarse grain size and insoluble segregated
Microstructure second phases.
Hardness © 2015 Elsevier Ltd. All rights reserved.
Post weld heat treatments

1. Introduction lowers from base material towards the joint fusion zone. Because usual-
ly Al–Mg alloys are joined in work hardening condition [4], the loss of
AA5754 aluminum alloy shows good formability, high fatigue hardness is offset by selecting an alloy that has good strength in the
strength and fair machinability. It is within the alloys of the 5xxx series annealed condition. Moreover, PWHT is not applied to Al–Mg joints
of higher strength. It is highly suited to flooring applications, shipbuild- welded in work-hardening condition in order to avoid recovery of the
ing or chemical and nuclear structures. Automotive structural parts and base material microstructure and its subsequent drop of strength.
inner body panels are commonly made of 5xxx alloy. Kumar et al. inves- For Al–Cu–Mg, Al–Mg–Si, Al–Mg–Zn–Cu alloys, which strongly
tigated the solidification microstructure of AA5754 alloy with a high harden by aging (or second phase particles), some heat treatments are
level of impurities [1]. The dominant strengthening mechanisms are performed after welding even if those alloys have been joined in the
work hardening and solute strengthening [2]. In fact for Al-Mg alloys work hardened condition [4,7,8,9,10,11]. In this case, the loss of hard-
the hardening coming from second phase particles (aging heat treat- ness due to the recovery of the work hardened microstructure promot-
ment) is very low in comparison with both solution and work harden- ed by the heat treatment is balanced or even overcame by the strong
ing [3,4]. In as cast state microstructure is dendritic with a strong rise of hardness induced by the second phase particles [7].
microsegregation development and several particles such as Mg2Si Generally with conventional autogenous welding both thermal dis-
and iron-bearing intermetallic (Fe, Mn) Al6 [1]. Homogenization treat- tortions and residual stresses limit the quality of welds. Innovative
ment leads to dissolution of soluble particles like Mg2Si and so increases welding methods are available for extending the range of applications
the hardness by solid solution strengthening [1,5,6]. and improving the overall performance of AA5754. Liu discussed the ef-
When 5000 series alloys in the work-hardened condition are joined fect of double pulsed gas metal arc welding (DP-GMAW) on metal drop-
by autogenous welding, the strength and hardness of the joint are sim- let transfer, weld pool profile, weld bead geometry and weld joint
ilar to those of the alloy in the annealed condition [3]. The loss of mechanical properties of Al alloy AA5754 [12]. The results indicate
strength/hardness is due to the solidification microstructure of the that the metal transfer, weld pool profile and weld bead geometry in
fusion zone and to the annealing of the zone adjacent to the fusion DP-GMAW significantly differs from P-GMAW. In particular, the grain
line defined heat affected zone (HAZ). The welding thermal cycle does size of the weld bead decreases with increasing of thermal pulse fre-
not affect the zone far from the fusion line (base material). The hardness quency and the eutectic Mg2Si precipitates are distributed homoge-
neously in the fusion zone. Senkara and Zhang studied the crack
⁎ Corresponding author. behavior of AA5754 HAZ during resistance spot welding [13]. Another
E-mail address: paola.leo@unisalento.it (P. Leo). joining method that has been studied for this alloy is the friction stir

http://dx.doi.org/10.1016/j.matdes.2015.10.150
0264-1275/© 2015 Elsevier Ltd. All rights reserved.
778 P. Leo et al. / Materials and Design 90 (2016) 777–786

spot welding. After T.S. Mahmoud and T.A. Khalifa [14], the tool rota- Table 2
tional speed and stir (dwell) time affect the weld structure and static Laser parameters.

strength of welds. Diameter fiber BPP Wavelength Focal distance Spot diameter
Hybrid laser arc welding (HLAW) recently has established itself as a [μm] [mm] [nm] [mm] [mm]
leading welding process. It combines conventional arc welding and laser 200 6.3 1070.6 250 0.4
beam welding. The coupled heat sources interact at the same time on a
single weld pool having a synergistic effect. Due to the actions of laser
beam and welding arc, hybrid welding offers many advantages over of PL and three of PMIG were used according to the experimental plan
laser welding and arc welding stand alone, such as high welding shown in Table 4.
speed, deep penetration, excellent gap bridging ability as well as good The microstructure of the cross weld section was analyzed with
process stability and efficiency [15,16,17,18]. The laser-to-arc power Nikon Epiphot 200 Optical Microscope (OM) and Zeiss EVO scanning
ratio (R) plays a significant role in the hybrid welding process. The electron microscope (SEM), equipped with a Bruker energy-dispersive
ratio determines which between the laser and the arc has the greater in- X-ray spectrometer (EDS). The samples were prepared by standard
fluence on the weld pool formation. The role of R-value on joint geom- metallographic procedure, which involved etching with Keller's reagent
etry and solidification rate has been investigated in different studies [19, or NaOH solution. The grain structure was revealed by electro-polishing
20,21]. Particularly, the joint solidification rate increases with R leading (20% perchloric acid and 80% ethanol at 0 °C, electro-polishing parame-
to a finer welded microstructure [19]. ter: 15 V and 60 s), anodic oxidation (Barker etching, anodizing param-
There are not many studies on hybrid laser arc welding of Al–Mg eters: 20 V and 80 s) and subsequent investigation under polarized light
alloys. Yan and others investigated the microstructure, mechanical in OM.
properties and fatigue strengths of AA5083-H111 [19]. They observed The dimension of the fusion zone (FZ) and HAZ were evaluated
an uncommon softening in the HAZ and a loss of microhardness in the using NIS software for imaging analysis. NIS-Elements is a NIKON soft-
fusion zone owing to the Mg burning during welding. Leo et al. studied ware supplied with Epiphot 200 OM. The software is tailored to facili-
the HLAW of annealed AA5754 [21]. Porosity and cracks were observed tate image capture, object measurement and counting.
in the fusion zone. Casalino et al. realized a comparison between laser In order to study the particle distribution in the HAZ, 3 zones 300 μm
and arc leading configuration [22]. In recent years promising disk and each thick were collected by tiling of macrographs (100×). The particle
fiber active medium laser system have been developed. They simulta- area was evaluated using NIS software for imaging analysis. Percentage
neously offer high optical output powers, high conversion efficiencies, particles area for each zone (particle total area divided by the area of the
high beam qualities and a short emission wavelength around 1 μm. singular zone) was determined.
Particularly the process parameters were optimized to obtain a joint Three different post weld heat treatments (at 350 °C) were per-
without defects with good mechanical properties [23]. formed on the s2 sample (20 × 5 × 3mm) using a salt bath.
This paper presents the effect of the post weld heat treatment of The holding times and temperatures are summarized in Table 5.
AA5754, which was welded in annealing condition by fiber laser arc Sample s2 was selected because it exhibited the best hardness profiles.
hybrid welding (FLAHW). In the present study, the hardness profile in Moreover, the BM was heat treated for 20 s at 550 °C and then air
the HAZ was not known and, moreover, the base material hardness cooled. Particularly the heating time segment was 20 s. The size of the
was not expected to drop due to PWHT [3]. So, the role of the PWHT sample was 5 × 3 × 10mm.
on mechanical properties was evaluated by microhardness, tensile Vickers microhardness (0.5/15 s) was measured using a LEICA
test, and fracture surface analysis. VMHT at a fixed distance (equal to 1 mm) from the top and the bottom
surface of the weld cross section. The number of indentations at a fixed
2. Experimental setup distance from top/bottom surface was equal to 17. Distance between in-
dentations was 500 μm. This test was carried out both for all the joints as
AA5754 (rolled sheet) alloy 3 mm thick was supplied in annealed con- welded and for the T350 heat treated joint.
dition. Table 1 reports the chemical composition of the alloy (weight %). Tensile strength test was conducted on the s2 T350_50 rectangular
The heat source was a fiber laser Ytterbium Laser System (IPG YLS- specimens (20 × 5 × 3mm) taken across the weld seam. A universal
4000) with a maximum available power output of 4 kW, coupled with electromechanical tensile test machine (INSTRON 5959) equipped
an MIG generator (GENESIS 503 PSR). The laser source was fiber deliv- with pneumatic grips was used. The mechanical properties and the frac-
ered. Particularly fiber diameter, beam parameter product (BPP), wave- ture surface of the s2 joint were compared with those of the BM and
length, focal distance and spot diameter are reported in Table 2. with the untreated joint.
The torch tilting angle was 40° and the laser beam was focused per-
pendicularly on the surface of the sheet. The experiment configuration
was laser leading, which means that the laser beam is the primary 3. Base material characterization
heat source while the arc is the assisting heat source providing filler
material. Being a combined process, the shape of the weld depends on The investigated Al–Mg alloy was supplied in annealed and recrys-
several factors, especially on the power of the heat sources. The distance tallized state. The OM micrograph (Fig. 1) shows the aluminum matrix
between the two sources was equal to 2.5 mm, the shielding gas (solid solution phase) together with a series of intermetallic precipi-
adopted was Argon, owing to the stable ignition of the arc in that tates. Based on previous work [1,24,25], it can be concluded that the
environment. acicular shape light gray particles are (Fe, Mn) Al6 (Fig. 1a,b), while
The filler material was ER 5356 and its chemical composition is in the rounded shape dark gray particles are the very fragile
Table 3. (Fe,Mn)3SiAl12 (Fig. 1a,b). The last ones can exhibit some cracks due
Four joints with different laser power (PL) and arc power (PMIG) and to fracture during the machining of rolling. The greater black particles
with a constant welding speed (3.5 m/min) were processed. Two levels are Mg2Si (Fig. 1a,b) while the smaller ones are Mg2Al3 [24,25]. After

Table 1 Table 3
Chemical composition of the AA5754 (wt.%). Chemical composition of the ER5356 (wt.%).

Si Fe Cu Mn Mg Cr Zn Ti Al Si Fe Cu Mn Mg Cr Zn Ti Al

0.40 0.40 0.10 0.50 2.6–3.6 0.3 0.20 b0.15 Bal. 0.25 0.40 0.05 0.1–0.2 4.5–5.5 0.1–0.2 0.10 0.15 bal.
P. Leo et al. / Materials and Design 90 (2016) 777–786 779

annealing the Mg2Si particles can nucleate also on the particles contain- Table 4
ing iron (Fig. 1c), in agreement with [26]. Experimental plan.

The element mapping (Fig. 2a) of the BM (Fig. 2b) confirmed the Sample PL MIG MIG PMIG Filler wire speed PTOT R
presence of Mg and Si in the dark particles and the presence of iron in [W] voltage current int. [W] [mm/min] [W] (PL/PMIG)
the white ones. Magnesium distribution was uniform in the solid [V] [A]

solution. s1 3000 28 140 3920 12.2 6920 0.77


The BM microstructure (Fig. 3) exhibited equiaxed recrystallized s2 3000 26.5 128 3392 11.4 6392 0.88
s3 3500 28 140 3920 12.2 7420 0.89
grains induced by the annealing heat treatment. The recrystallized
s4 3500 26.5 128 3392 11.4 6892 1.03
grains are clearly observable in Fig. 3 after metallographic anodizing of
the sample. The average diameter of the BM grains in the cross weld
section was 20 ± 3 μm. The microhardness value was 60 ± 1 HV, the occurrence of black particles (Mg2Si and/or Mg2Al3) was strongly re-
ultimate tensile strength was 244.3 N/mm2 and the elongation 16.7%. duced in the HAZ.
Within the HAZ and BM, the distribution (Fig. 9a) and the area of
particles (Fig. 9b) having equivalent diameter equal to or higher than
4. Results 2 μm was measured in the cross section of s2. The investigated area
was divided into three zones 300 μm thick (Fig. 9a). Zone 1 and the
4.1. Weld shape and microstructure zone 2 pertained to the HAZ; zone 3 was the BM. Zone 1 was the closest
to the melting zone. The analysis of the particles distribution in the HAZ
The weld geometry depends on the process parameters. All the joint (Fig. 9a) shows that the number of particles decreased moving from BM
cross sections presented the classic “wine-cup” shape with an upper to FZ. The total area of particles was calculated in each zone using NIS
zone denoted as “Crown Zone” (CZ), and a narrow lower Laser Zone software and divided by the area of the singular zone (Fig. 9b). Reduc-
(LZ) [27] (Fig. 4a). The two zones presented different aspect and mor- tion of particle area obtained in zones 1 and 2 was 50% higher with
phology due to the influence of the two heat source powers respect to that in zone 3, which indicates that the welding thermal
(Fig. 4b,c). The LZ shape was controlled by the laser beam and the CZ cycle leads to particle dissolution (Fig. 9b).
shape by the MIG torch [20,27]. The size of the FZ changed mainly
with PMIG (Fig. 4b). As PMIG increased at fixed PL the molten pool size in-
4.2. Mechanical properties and PWHT
creased because the base material melting was under the electric arc
control (Fig. 4b,c). By contrast, at constant PMIG, the FZ area decreased
Microhardness was measured along the weld cross section both site
when PL rose (Fig. 4b,c) due to the improved arc stability, which was
at 1 mm from the top (CZ) and from the bottom (LZ). Fig. 10 shows
enhanced by the higher power laser [28].
microhardness for the two sites. The profiles exhibit symmetrical trends
The FZ exhibited a dendritic type microstructure (Fig. 5a). After me-
(Fig. 10) with respect to the welding centerline. The highest microhard-
tallographic anodizing, columnar grains were identified. Starting from
ness values for each profile were observed at the FGZ. The hardness
the HAZ interface the columnar grains grew towards the inner of the
gradually decreased from the FGZ to welding centerline and from FGZ
FZ (Fig. 5b). Moreover, a transition columnar to coarse equiaxed grains
to the BM. The last result indicates that the welding thermal cycle led
close to the welding centerline was observed (Fig. 5b). The equiaxed
to hardness increase in HAZ with respect to BM. The main difference
grains occupied an area that was much smaller in the LZ than in the
between all the analyzed profiles was in the welding centerline
CZ (Fig. 6). The average size of equiaxed grains was significantly greater
hardness value.
than that of the base material (Fig. 5b). In addition, there was a small
For each microhardness profile the lowest value was obtained in the
band of fine grains, called “Fine Grain Zone” (FGZ), at the interface be-
LZ centerline and it was lower than the BM hardness.
tween the fusion zone and the HAZ (Fig. 5d). The band average size
Particularly s4 and s2 samples presented the highest microhardness
was 85 ± 5 μm. The HAZ extended between the FGZ and BM and its
at the welding centerline both in LZ and in CZ. The lowest value was
size was lower than 600 μm. After metallographic etching with NaOH
measured for the s1 sample. Considering both CZ and LZ hardness, the
solution, another narrow area close to the FGZ (in the HAZ) character-
welding parameters used for processing the s2 joint were better. There-
ized by average grain size larger than that of the base material (28 ±
fore, the s2 joint was subjected to the PWHT in order to improve its
3 μm Vs 20 ± 3 μm) was identified. The slight grain coarsening with re-
mechanical properties.
spect to BM was probably due to the welding thermal cycle (Fig. 5d).
The hardness profiles after T350 heat treatments on s2 joint show
The FZ microstructure was affected by porosity that was probably
that the hardness values in the FZ greatly increased both in CZ
due to Mg vaporization together with evident solute segregation at
(Fig. 11a) and in LZ (Fig. 11b). The minimum hardness at the welding
interdendritic site (black color in Fig. 5a) and second phase segregation
centerline disappeared after 50 min holding time (Fig. 11 c,d) and the
(Fig. 5c) mostly close to the welding centerline. Second phase segrega-
FZ hardness was higher than that of BM (Fig. 11d).
tion was mainly insoluble white phases containing iron and black solu-
The values of the yield stress (σy), ultimate tensile strength (UTS)
ble Mg2Si phases in Chinese-script shape (Fig. 5c) [1,29,30,31].
and strain at fracture (elongation ef) for BM and specimen s2 both in un-
Some solidification cracks, which run along the edges of the colum-
treated and treated T350_50 state are reported in Table 6. The stress strain
nar grains, were observed in the LZ (Fig. 6).
curves are shown in Fig. 12. The yield stress (σy) and UTS of the heat
Due to the high temperature (T), the HAZ underwent a severe
treated T350_50 sample were higher with respect to the untreated one,
welding thermal cycle [4], which altered the microstructure in terms
whereas the strain at fracture was similar. The highest σy and elonga-
of grain size and particles. Particularly, the highest T was reached near
tion were observed in BM.
the fusion line. The macrograph in Fig. 7 clearly shows how the HAZ dif-
fered from the BM. The microstructure of the heat treated T550_20 BM
was analyzed by scanning electron microscope. Fig. 8 compares the
Table 5
back scattered electron (BSE) micrographs of the untreated BM, the PWHT on s2 sample.
T550_20 BM and the HAZ. These images show that both the heat treated
Temperature Holding time Cooling
BM (Fig. 8b) and the HAZ (Fig. 8c) had cleaner matrices with respect
to the untreated BM (Fig. 8a). The heat treatment T550_20 BM argued T350_5 350 °C 5 min Air cooling
that the soluble smaller particles were able to dissolve at high T T350_20 350 °C 20 min Air cooling
T350_50 350 °C 50 min Air cooling
even if the alloy was heated for a short holding time. Particularly the
780 P. Leo et al. / Materials and Design 90 (2016) 777–786

Fig. 1. (a, c) OM and (b) SEM image of AA5754 BM. Light gray acicular shape particles (Fe,Mn)Al6, rounded shape dark gray particles (Fe,Mn)3SiAl12, black particles Mg2Si. (c) Noteworthy
is the presence of fine Mg2Si particles (arrow) that, following treatment of annealing, grew on particles containing iron [26].

Fig. 13 shows the microscopic appearance of the fracture surface for the intensity of which varied with the distance from the molten zone.
the treated and untreated joints. Both the specimens cracked in the FZ. The heat treatment imposed a steep thermal gradient (from T close to
Particularly fracture arose from the bottom side and propagated melting point to room temperature) in a very short time. Due to high
throughout the entire section mainly along the welding centerline. Frac- temperature in the welding cycle, small Mg particles (both Mg2Al3
ture surfaces were rough and characterized by both small and large and Mg2Si) were able to dissolve, which led to atomic diffusion in the
dimples. Fracture was principally ductile although the moderate elon- solid state. Moreover, due to the fast cooling rate, probably the solute
gation for both samples. No evident differences between the two frac- atoms remained trapped in Al matrix without precipitating particles
ture surfaces were observed. Indeed the tensile test results (Table 6) as shown by [5,6]. Indeed the OM observation of the HAZ (Fig. 7)
indicated that both the samples present similar elongation showing showed a matrix much cleaner than that of the BM. The subsequent
that probably the fracture was promoted mainly by the presence of SEM observation of the heat treated BM at 550 °C-20 s and air cooled
porosity, coarse grain size and insoluble segregated second phases. (Fig. 8b) showed that, due to heat treatment, Mg particles were reduced
so Mg atoms diffused in the aluminum matrix (in solid solution). More-
5. Discussion over, the percentage total area of particles in HAZ decreased gradually
moving from BM to FZ interface showing the dissolution of the Mg par-
As received the Al–Mg microstructure exhibited equiaxed recrystal- ticles (Fig. 9) and confirming the previous results (Fig. 7). Due to the
lized grains as a result of the annealing heat treatment (Fig. 3). The dissolution of Mg particles, a solid solution strengthening occurred in
average diameter of BM grain in the cross weld section was 20 ± the HAZ leading to an increasing of hardness with respect to the BM
3 μm. The microhardness value is 60 ± 1 HV. Magnesium was distribut- (Fig. 10).
ed inside the aluminum matrix (solid solution) (Fig. 2). Moreover Mg Joint microstructure, affected by both undercooling and flow turbu-
atoms form intermetallic particles both soluble and insoluble with lence [32,33], originated from the BM interface and developed towards
other alloying elements (Figs. 1,2). the welding centerline. The FGZ (Fig. 5d) area was due to the strong
The HAZ lay between the BM and the FZ. The HAZ began where the thermal undercooling. In that zone the melted metal undergoes a con-
peak temperature of the welding cycle was just below the solidus siderable amount of heat followed by a fast cooling and the new grains
temperature [4]. The HAZ was due to a real heat treatment of the BM, had an average size smaller than that of the BM. Close to the FGZ, in the

Fig. 2. (a) Alloy elements mapping of (b) BM. The analysis confirms the chemical composition of the particles and the presence of Mg in solid solution, typical for the alloys AA5xxx.
P. Leo et al. / Materials and Design 90 (2016) 777–786 781

grains originated at the fusion edge by the growth of some crystals


of the FGZ oriented in the direction opposite to the heat flux (perpendic-
ular to the interface). The columnar growth was favored by steep T gra-
dients. LZ was characterized by a steeper T gradient with respect to the
CZ. Therefore the LZ microstructure was prevalently columnar with re-
spect to the CZ and the LZ solidification rate was higher. On the contrary
in the CZ the shallow thermal gradient promoted the transition from co-
lumnar to equiaxed growth [4,32,34].
Solidification cracks ran along the edges of the columnar grains in
LZ (Fig. 6). They formed in the interdendritic or intergranular spaces
that, in the columnar microstructure have a direction orthogonal to
that of the application of thermal stress [33]. Otherwise, equiaxed
dendritic microstructures were characterized by either intergranular
or interdendritic space oriented in many directions, which created
long and tortuous energy-consuming paths [4].
The power ratio (R) indicated whether the laser or the arc power
was dominant during the welding process, respectively R N 1 and
R b 1. Particularly, at a given total power, the joint solidification rate in-
creased with R leading to a finer welded microstructure [19].
The hardness profiles in Fig. 10 differed from one another mainly in
Fig. 3. Polarized light micrograph of the BM showing equiaxed recrystallized grains. the centerline hardness values where the effect of different solidifica-
tion rates on the microstructure was more significant [32]. The hardness
HAZ, the grains had a greater average size with respect to that of the BM profiles in Fig. 10a indicate that in the CZ the hardness value along
(Fig. 5d). Grains coarsened due to the welding thermal cycle. welding centerline strongly increased (compare s1 and s4) when, at
The FZ exhibited a dendritic microstructure (Fig. 5a) and presented very close PTOT (Table 4), the R-value increased from the lowest value
both columnar and equiaxed grain zone (Fig. 5b). The area of columnar (R1) to the highest one (R4), due to finer microstructure promoted by

Fig. 4. (a) OM macrograph of the join shape (Keller etching). The classic shape “wine cup” is evident due to the CZ shape and the narrower LZ. (b) Joints fusion zones size at PL equal to
3000 W (red line) and 3500 W (black line) as a function of PMIG. (c) Joints FZ and HAZ elaborated by NIS software. (For interpretation of the references to color in this figure legend, the
reader is referred to the web version of this article.)
782 P. Leo et al. / Materials and Design 90 (2016) 777–786

Fig. 5. (a) OM micrograph of the FZ (Keller etching) showing dense dendritic microstructure and interdendritic solute segregation. (b) OM micrograph of anodized sample showing transition
from columnar to coarse equiaxed grains close to welding centerline. (c) SEM micrograph of the FZ centerline with evident second phase segregation. (d) OM micrograph (NaOH etching) with
indication of the thickness of both the FGZ and the zone close to FGZ that is characterized by increased grain size with respect to BM. The average grain size in each area is also indicated.

the faster solidification rate. When the R-value changes from R1 to R4, solidification rate in the s4 sample (R4). Centerline hardness values
hardness increased in the LZ less than in the CZ and it could be explained for s2 and s3 samples were in between those of s1 and s4 (Fig. 10a)
by the effect of some LZ solidification cracks induced by the higher where R2 and R3 values were in between those of R1 and R4. In fact

Fig. 6. OM micrograph of the LZ showing pores and cracks in s4. Fig. 7. OM macrograph of the HAZ.
P. Leo et al. / Materials and Design 90 (2016) 777–786 783

Fig. 8. BSE image sequence of the (a) untreated BM, (b) heat treated T550_20 BM, (c) HAZ. Both treated T550_20 BM and HAZ matrix are cleaner than the untreated BM where dark gray
particles are evident.

Fig. 9. Distribution (a) and percentage particle area (b) in the HAZ. The particle number decreases from BM to FZ. Particularly a big reduction of particle percentage and particle area was
obtained in zones 1 and 2 with respect to zone 3 indicating that the welding thermal cycle leads to particle dissolution.

both the s1/s3 and s2/s4 couples were processed at fixed PMIG (Table 6, was observed in the welds (Fig. 5a,c) and it is well known that Mg
Fig. 4b,c) and with PL that rose from s1 to s3 and from s2 to s4 leading to segregation reduces the hardness of Al–Mg alloys [27]. Particularly,
higher R-values in s3 with respect to s1 and s4 respect to s2. So, as PL in- the amount of segregation was maximum along the welding axis due
creased at fixed PMIG (from s1 to s3 and from s2 to s4) both the steeper to the solute growth in the last consolidated liquid [35,36,37]. Moreover,
gradient and the reduction of molten pool size [28], which were pro- segregation happened less close to the interface. Therefore, as the segre-
moted by the increasing of R values (Fig. 4b,c), lead to faster solidifica- gation level lowered, FZ hardness values increased from the weld
tion rate and a finer microstructure. Hence, the centerline hardness centerline to the HAZ interface (Fig. 10), which justifies the characteris-
values of s3 (s4) sample were higher than those of s1 (s2) ones (Fig. 10). tic shape of the profile. Moreover, the microstructure predominantly
In order to explain the characteristic shape of the FZ microhardness columnar in LZ induced greater segregation near the axis because the
profiles, the role of solute and second phase segregation must be consid- columnar grains facilitated the solute rejection towards the welding
ered. In fact due to non-equilibrium solidification, some segregation centerline. By contrast, the equiaxed dendrites in the CZ, rejected the

Fig. 10. Cross weld section microhardness profile in the CZ at 1 mm from the top surface (a) and LZ at 1 mm from the bottom surface (b). The profiles exhibit symmetrical trends with
respect to the welding centerline.
784 P. Leo et al. / Materials and Design 90 (2016) 777–786

Fig. 11. Cross weld section microhardness profiles in the (a) CZ and (b) LZ of as welded and heat treated s2 samples for different holding times at 350 °C showing that the fusion zone
hardness increases with holding time at 350 °C. The CZ and LZ microhardness profiles in untreated (c) and treated T350_50 state (d) are compared.

solute in different growth directions and the solute rejection towards The CZ and LZ hardness profiles for s2 resulted the best and therefore
the centerline was reduced. Consequently, the hardness of all the joints underwent the post weld heat treatment, which aimed to improve
at the LZ welding centerline was always lower with respect to the CZ. hardness by reducing segregation [32,4].
P. Leo et al. / Materials and Design 90 (2016) 777–786 785

Table 6 [39,40] have shown that σy and UTS for Al–Mg alloys increase with
Tensile test for the BM, the untreated and heat treated T350_50 joint. Mg content in solid solution. The highest σy observed in the BM
UTS [N/mm2] σy [N/mm2] ef [%] could be probably due to the smaller grain size with respect to that
Untreated joint 201.4 70 11.1
observed in the fusion zone of welded samples (Figs. 3,4b) according
T350_50 joint 239 89 8.1 to the Hall–Petch relation [40]. Despite the welds' being character-
BM 244.3 117 16.7 ized by different strength values, they exhibited very similar elonga-
tion and fracture without necking evidence. Particularly it is
interesting to observe that the joints exhibited a great loss of ductil-
ity respect to the BM. The loss of ductility with respect to BM was due
to porosity, coarse grain size and insoluble segregated second
phases. Porosity and coarse segregated phases led to stress concen-
tration so the crack easily started from there and propagated
throughout the sample thickness. Moreover, the coarse grain size
of the fused zones favored the mobility of the crack tip. Coarse
grain size, porosity and segregation observed in the fusion zone en-
hanced both crack initiation and mobility of the crack tip. Fracture
surfaces showed no significant difference since the rupture was
mainly due to stress concentration around the defects (Fig. 13).

6. Conclusions

This paper has dealt with the effect of the post weld heat treatment
of AA5754, which was welded in annealing condition by FLAHW. The
conclusions involve the following points:

• Cross weld section microhardness of the untreated joints exhibited


axi-symmetric profile characterized by a minimum at welding center-
Fig. 12. Tensile stress–strain curves line and a maximum at FGZ. The high microhardness for each profile
was detected in FGZ and mainly depended on dissolution of the
small Mg particles (both Mg2Al3 and Mg2Si) as confirmed by a gradual
The role of segregation on hardness response of the FZ was con- reduction of particle percentage from the BM to the FGZ.
firmed by analyzing the hardness profile of the s2 joint, heat treated at • Joint cross weld microstructure was dendritic and was characterized
350 °C (Fig. 11a,b) for 5, 20 and 50 min and then air cooled. Hardness by equiaxed grains close to the centerline and columnar grains at
values in FZ greatly increased both in the CZ (Fig. 11a) and in the LZ the HAZ interface. In the CZ the equiaxed grain area is higher with
(Fig. 11b), particularly after 50 min holding time. Moreover, comparing respect to that of the LZ due to the shallow thermal gradient. FZ hard-
CZ and LZ profiles of the untreated s2 joint (Fig. 11c) with the 50 min ness values increased with R.
treated ones (Fig. 11d), it can be observed that the minimum hardness • Solute segregation and second phase segregation were observed in
values at the welding centerline disappeared with the heat treatment. the melted zone. Particularly, the prevalent columnar structure in
The explanation is that elimination of segregation led to a rise in the the LZ promotes a higher segregation with respect to the CZ.
amount of Mg atoms inside the aluminum matrix and hardening by • The PWHT led to the reduction of the segregation and therefore to the
Magnesium solid solution strengthening [2,38]. Mg atoms available for strengthening by Mg solutionizing. Microhardness of the heat treated
solid solution strengthening were more numerous in the FZ than in FZ was quite constant and higher than that of the BM.
the BM due to the filler material addition. Therefore, the reduction of • Ultimate tensile strength of s2 weld in T350_50 state was 19% greater
segregation induced a relevant hardening. with respect to the untreated weld and similar to that of BM. There-
Yield strength and tensile strength of the T350_50 heat treated joint fore, a targeted PWHT can improve the mechanical strength of
were greater with respect to the as welded joint. Several other papers AA5754 laser hybrid weld.

Fig. 13. Fracture surface of the (a) untreated joint and (b) heat treated T350_50 joint.
786 P. Leo et al. / Materials and Design 90 (2016) 777–786

• Untreated and PWHT weld showed low and similar elongation prop- [16] A.M. Mazar, J. Ma, G. Yang, R. Kovacevic, Hybrid laser/arc welding of advanced high
strength steel in different butt joint configurations, Mater. Des. 64 (2014) 573–587.
erty. Low elongation was probably due to porosity, coarse grain size [17] C. He, C. Huang, Y. Liu, J. Li, Q. Wang, Effects of mechanical heterogeneity on the ten-
and insoluble segregated second phases which promoted the rupture sile and fatigue behaviours in a laser-arc hybrid welded aluminium alloy joint,
as confirmed by no necking evidence. Mater. Des. 65 (2015) 289–296.
[18] Z. Zhang, S. Dong, Y. Wang, B. Xu, J. Fang, P. He, Microstructure characteristic of thick
aluminum alloy plate joints welded by fiber laser, Mater. Des. 84 (2015) 173–177.
[19] F. Olsen, Hybrid Laser-arc Welding, Woodhead Publishing, Cambridge UK, 2009.
[20] M. El Rayes, C. Walz, G. Sepold, The influence of various hybrid welding parameters
References on bead geometry, Weld. J. 83 (2004) 147s–153s.
[21] P. Leo, G. Renna, G. Casalino, A.G. Olabi, Effect of power distribution on the weld
[1] S. Kumar, H.B. Nadendla, G.M. Scamans, D.G. Eskin, Z. Fan, Solidification behavior of quality during hybrid laser welding of an Al–Mg alloy, Opt. Laser Technol. 73
an AA5754 alloy ingot cast with high impurity content, Int. J. Mater. Res. 103 (2012) (2015) 118–126.
E1–E7. [22] G. Casalino, S.L. Campanelli, M.U. Dal, A.D. Ludovico, Arc leading versus laser leading
[2] Ø. Ryen, O. Nijs, E. Sjölander, B. Holmedal, H. Ekström, E. Nes, Strengthening mech- in the hybrid welding of aluminum alloy using a fiber laser, Procedia CIRP 12 (2013)
anism in solid solution aluminum alloys, Metall. Mater. Trans. A 37A (2006) 151–156.
1999–2006. [23] G. Casalino, M. Mortello, P. Leo, K.Y. Benyounis, A.G. Olabi, Study on arc and laser
[3] I.J. Polmear, Light Alloys: Metallurgy of the Light Metals, Butterworth Heinemann, powers in the hybrid welding of AA5754 Al-alloy, Mater. Des. 61 (2014) 191–198.
1995. [24] W.F. Smith, Structure and Properties of Engineering Alloys, McGraw-Hill, 1992.
[4] R.W. Messler, Principles of Welding, Wiley-VCH, 2004. [25] Aluminum and Aluminum Alloys, ASM International, 1993.
[5] H. Jiang, L. YE, X. Zhang, G. Gu, P. Zhang, Y. Wu, Intermetallic phase evolution of [26] O. Engler, Z. Liu, K. Kuhnke, Impact of homogenization on particles in the Al–Mg–Mn
5059 aluminum alloy during homogenization, Trans. Nonferrous Metals Soc. China alloy AA 5454 — experiment and simulation, J. Alloys Compd. 560 (2013) 111–122.
23 (2013) 3553–3560. [27] M. Gao, Y. Cao, X.Y. Zeng, T.X. Lin, Mechanical properties and microstructures of
[6] T. Radetic, M. Ppovic, E. Romhanji, Microstructure evolution of a modified AA5083 hybrid laser MIG welded dissimilar Mg–Al–Zn alloys, Sci. Technol. Weld. Join. 15
aluminum alloy during a multistage homogenization treatment, Mater. Charact. (7) (2010) 638–645 (09).
65 (2012) 16–27. [28] Y.B. Chen, Y.B. Zhao, Z.L. Lei, L.Q. Li, Effects of laser induced metal vapour on arc plas-
[7] J. Ding, D. Wang, Y. Wang, H. Du, Effect of post weld heat treatment on properties of ma during laser arc double sided welding of 5A06 aluminium alloy, Sci. Technol.
variable polarity TIG welded AA2219 aluminum alloy joints, Trans. Nonferrous Weld. Join. 17 (2012) 1.
Metals Soc. China 24 (2014) 1307–1316. [29] S. Kumar, B.N. Hari, G.M. Scamans, Z. Fan, Influence of intensive melt shearing on the
[8] R. Ahmad, M.A. Bakar, Effect of a post-weld heat treatment on the mechanical and microstructure and mechanical properties of an Al–Mg alloy with high added impu-
microstructure properties of AA6061 joints welded by the gas metal arc welding rity content, Metall. Mater. Trans. A 42A (2011) 3141–3149.
cold metal transfer method, Mater. Des. 32 (2011) 5120–5126. [30] S. Kumar, B.N. Hari, G.M. Scamans, Z. Fan, Microstructural evaluation of melt condi-
[9] V. Balasubramanian, V. Ravisankar, R. Madhusudhan, Influence of pulsed current tioned twin roll cast Al–Mg alloy, Mater. Sci. Technol. 27 (No. 12) (2011) 1833–1839.
welding and post weld aging treatment of fatigue crack growth behavior of [31] L.F. Mondolfo, Aluminium Alloys: Structure and Properties, Butterworth, 1976.
AA7075 aluminum alloy joints, Int. J. Fatigue 30 (2008) 405–416. [32] D.A. Porter, K.E. Easterling, Phase Transformations in Metals and Alloys, Chapman &
[10] Z.Y. Zhu, C.Y. Deng, Y. Wang, Z.W. Yang, J.K. Ding, D.P. Wang, Effect of post weld heat Hall, 1992.
treatment on the microstructure and corrosion behavior of AA2219 aluminum alloy [33] M. Baricco, R. Montanari, Solidificazione, AIM, 2012.
joints welded by variable polarity tungsten inert gas welding, Mater. Des. 65 (2015) [34] J.D. Verhoeven, Fundamentals of Physical Metallurgy, Wiley, 1975.
1075–1082. [35] Y.L. Liu, S.B. Kang, Solidification and segregation of Al–Mg alloys and influence of
[11] R.R. Ambritz, G. Barrera, R. Garcia, V.H. Lopez, The microstructure and mechanical alloy composition and cooling rate, Mater. Sci. Technol. 13 (1997) 331–336.
strength of Al-6061 T6 GMA welds obtained with the modified indirect electric [36] S. Das, N.S. Lim, H.W. Kim, C.G. Park, Effects of heat treatment on microstructure and
arc joint, Mater. Des. 31 (2010) 2978–2986. mechanical properties of twin roll casted Al–5.5 Mg–0.02 Ti alloy, Mater. Des. 31
[12] A. Liu, X. Tang, F. Lu, Study on welding process and prosperities of AA5754 Al-alloy (2010) 3111–3115.
welded by double pulsed gas metal arc welding, Mater. Des. 50 (2013) 149–155. [37] E.L. Huskins, B. Cao, K.T. Ramesh, Strengthening mechanisms in an Al–Mg alloy,
[13] J. Senkara, H. Zhang, Cracking in spot welding aluminum alloys, Weld. Res. Suppl. Mater. Sci. Eng. A 527 (2010) 1292–1298.
(2000) 194–201. [38] S. Kou, Welding Metallurgy, Wiley-VCH, 2003.
[14] T.S. Mahmoud, T.A. Khalifa, Microstructural and mechanical characteristics of alumi- [39] E.-D.H. Nasr, S.F. Moustafa, A.N. Abdel-Azim, A. Ismail, The deformation of alumini-
num alloy AA5754 friction stir spot welds, Mater. Eng. Perform. 23 (3) (2014) um–magnesium alloys, Mater. Sci. Technol. 10 (1) (1992) 16–20.
898–905. [40] G.E. Dieter, Mechanical Metallurgy, McGraw-Hill, 2001.
[15] J. Zhou, H.L. Tsai, Welding process, InTech, Change 1 (2012) 3–32.

You might also like