You are on page 1of 12

ASM International

1059-9495/$19.00

JMEPEG (2013) 22:38903901


DOI: 10.1007/s11665-013-0663-3

Friction Stir-Welded Dissimilar Aluminum


Alloys: Microstructure, Mechanical Properties,
and Physical State
M. Ghosh, Md.M. Husain, K. Kumar, and S.V. Kailas
(Submitted March 22, 2013; in revised form July 11, 2013; published online August 8, 2013)
A356 and 6061 aluminum alloys were joined by friction stir welding at constant tool rotational rate with
different tool-traversing speeds. Thermomechanical data of welding showed that increment in tool speed
reduced the pseudo heat index and temperature at weld nugget (WN). On the other hand, volume of
material within extrusion zone, strain rate, and Zenner Hollomon parameter were reduced with decrease in
tool speed. Optical microstructure of WN exhibited nearly uniform dispersion of Si-rich particles, ne grain
size of 6061 Al alloy, and disappearance of second phase within 6061 Al alloy. With enhancement in welding
speed, matrix grain size became ner, yet size of Si-rich particles did not reduce incessantly. Size of Si-rich
particles was governed by interaction time between tool and substrate. Mechanical property of WN was
evaluated. It has been found that the maximum joint efciency of 116% with respect to that of 6061 alloy
was obtained at an intermediate tool-traversing speed, where matrix grain size was signicantly ne and
those of Si-rich particles were substantially small.

Keywords

aluminum alloys, friction stir welding, light microscopy,


mechanical characterization

1. Introduction
Nowadays, friction stir welding (FSW) of versatile materials
nds widespread application in automobile and aerospace
industries for fabricating primary and secondary components.
The technique was developed and patented in the UK in 1990
by The Welding Institute for welding of plates in solid state
(Ref 1). The procedure is a complex solid-state thermomechanical process, in which a rotating tool with a shoulder and
pin moves through rigidly clamped plates, placed in butting/
lap/llet joint conguration over metallic support (Ref 2).
Shoulder maintains intimate contact with top surface of
workpiece. Heat is generated by friction at shoulder and pin
surfaces. Material gets softened under severe plastic deformation, and ow occurs along welding direction with the
translation of tool. Material is thus transported from the front
end of tool to trailing edge, where it is forged to form a joint.
Along the welding line, the side, where the direction of tool
rotation is the same as that of traversing of tool, is called
advancing side, with other side being termed as retreating side.
Ultimate microstructure considering grain size, second-phase
M. Ghosh and MD.M. Husain, Materials Science & Technology
Division, CSIRNational Metallurgical Laboratory, Jamshedpur
831007, India; K. Kumar, Department of Materials Science &
Engineering, University of Northern Texas, Denton, TX 76203-5017;
and S.V. Kailas, Department of Mechanical Engineering, Indian
Institute of Science, Bengaluru 560012, India. Contact e-mails:
ghosh_mnk@yahoo.com and mainakg@nmlindia.org.

3890Volume 22(12) December 2013

fraction, dissolution and reappearance of precipitate in different


zones like heat-affected zone (HAZ), thermomechanically
affected zone (TMAZ), and weld nugget (WN) along with
joint efciency depend on total heat input, cooling rate, plastic
strain, material ow, and state of stress. The plastic strain and
strain rate are very high and substantially greater than
conventional metal-working processes like extrusion, rolling,
forging, etc. (Ref 3). Welding variables, for example, rotational
speed, traversing speed, tool tilt angle, plunging depth,
plunging speed, motor torque, normal load, and tool design,
have predominant effect on welding defects, residual stress, and
joint quality (Ref 1). Three types of material ow have been
identied during FSW. First is the rotation of plasticized
material around tool, which is governed by the revolution of
tool generating friction between tool and workpiece. Second is
the downward movement of material by pin nearby itself and
subsequent upward motion of an equivalent amount of material
away from pin. Third one is the relative motion of material
between tool and workpiece (Ref 2).
Till date, the process has made reasonable breakthrough in
respect of copper alloys, magnesium alloys, titanium alloys,
steels, nickel alloys, molybdenum alloys, Al-alloy matrix
composites, and thermoplastics (Ref 1, 3). However, major
thrust has been given for FSW of Al and its derivatives to
produce lap/butt joints consisting of similar materials. Such
type of studies highlighted microstructural-mechanical property
co-relation, mechanism of weld-zone formation, heat input and
temperature rises, materials ow prediction using marker
technique, fatigue response, fracture toughness, corrosion
resistance, and residual stress determination for transition joints
(Ref 2, 4-9).
It has been established that variation in weldability of Al
alloys occurs owing to the presence of various alloying
elements and heat-treatment condition. In general, precipitation
behavior in 2xxx and 7xxx series alloys have been matched

Journal of Materials Engineering and Performance

with precipitation hardenable 6xxx series alloys. Coarsening of


precipitates and solutionization of needle precipitates were
observed in HAZ of welded alloy 6063 alloy. Characterization
at WN of friction stir-welded 6061 Al alloy indicated the
presence of second phase, but their identity was not established
(Ref 10). In another attempt, same AA6061 Al alloy plates,
both in O and T6 temper conditions, were joined by FSW using
four different sets of weld parameters (Ref 10). Microstructural
and mechanical characterization of the joints were made by
detailed optical microscopy investigations, extensive hardness
measurements and tensile tests. The effect of temper condition
on joint performance was explored in addition to the effect of
weld parameters. For alloy AA 6082, it is found, that some
coarsening of precipitates occurred along with partial dissolution of them at high FSW temperature (Ref 2).
Microstructural investigation in nonheat-treatable friction
stir-welded Al alloys in cold worked,tempered condition has
described the loss of hardness across the weld. FSW of Al-5086
plates in H32 condition at tool rotational speed of 1600 rpm
and tool-traversing speeds of 175, 200, and 225 mm/min
exhibited failure through stir zone during tensile test (Ref 11).
Microhardness measurements and bend test results were in
accordance with tensile properties. This was attributed to the
loss of cold-work hardening within stirred zone due to heat
generation during welding. No cracking phenomenon was
observed at the time of bend testing of welds, although a little
porosity was present. FSW of 5082 Al alloy was also carried
out, which produced uniform hardness across the weld (Ref 2).
For heat-treatable Al alloys, FSW promoted the formation of
relatively softer region at WN with respect to other parts of
transition joint due to dissolution of second phases; however,
strength could be regained with subsequent growth of strengthening precipitates during thermal cycle (Ref 12, 13). For
nonheat-treatable Al alloys, FSW produced hardened region at
WN because of the development of ne grain structure and
homogeneous distribution of nondissolved second phase. HAZ
became weak mainly because of decrease in dislocation density
and increase in grain size (Ref 14, 15). Considering the diverse
responses as mentioned above, a few attempts have been made
to study weldability of two different Al alloys. One of the
examples is friction stir spot welding (FSSW) between AA
2024-T3 and 5754-H22 alloy sheets at tool rotation speed of
1500 rpm (Ref 16). Tool plunge depths were of 2.45,
2.55, and 2.65 mm from specimen surface with dwell times
of 2, 5, and 10 s. The maximum strength was achieved for the
joint produced by placing AA 5754-H22 sheet on the top with
the tool plunge depth of 2.65 mm and dwell time of 10 s with
the characteristic of pull-out nugget fracture. The minimum
strength was obtained for the joint produced by placing AA
2024-T3 sheet on the top with tool plunge depth of 2.45 mm
and dwell time of 2 s with typical cross-nugget failure. It has
been inferred that an increase in tool penetration depth up to a
certain limit ensured the increment in joint strength.
Challenge arises when welding consists of dissimilar alloys
because of their different physicochemical properties, precipitation behavior, trend in defects accumulation, and phase
transformation characteristic during welding. FSW of 5083
alloy to 6061 alloy under various tool rotation and traversing
speeds is one of the examples (Ref 17). For this couple,
microhardness near interface showed heterogeneous distribution, and the bond strength was 63% with respect to that of
6061 alloy with drop in elongation. In another endeavor,
Kumbhar and Bhanumurthy (Ref 18) have pointed out the

Journal of Materials Engineering and Performance

absence of rigorous mixing during FSW of 5052-6061 alloy.


They reported the highest bond strength of 71% of that of
6061 alloy with elongation of 3%. Effect of FSW parameters
on microstructure was evaluated for transition joint consisting
of 2024-T3 and 7075-T6 Al alloys (Ref 19). For this couple
stirring zone (SZ), microstructure was heterogeneous because
of shorter welding time and decorated with onion-ring-like
pattern illustrating differences in grain size and composition.
Microhardness at WN was close to the hardness of 2024 Al,
and weaker region appeared at the periphery of shoulder. Most
of the weld joints failed from HAZ of 2024 Al, and in few
cases, from SZ; however, overall strength and elongation of the
assemblies were lower than those of base alloy. Commercially
pure aluminum (CP Al) was also friction stir welded at different
welding speeds with 7039 aluminum alloy (Ref 20). For them,
onion-ring pattern was found at WN. Microhardness showed an
upward trend starting from CP Al side with a little zigzag near
SZ. FSW and FSW followed by heat treatment indicated UTS
values to the tune of 61 and 68% of that of CP Al,
respectively, for the same couple. Fracture occurred either
through CP Al or CP Al-HAZ interface. Palanivel and Mathews
(Ref 21) have attempted to join 6351-T6 to AA5083-H111
aluminum using different tool geometry at constant rpm but
various tool-traversing speeds. The highest bond strength was
275 MPa when welding was done with square cross-section
pin at 63 mm/min welding speed. FSW of 6061 Al to 2024 Al
reported formation of three regions within nugget; a dispersion
zone, stirring-induced plastically deformed zone containing
alternate lamellar structure, and equiaxed zone (Ref 22). Miles
et al. (Ref 23) have made dissimilar Al alloy joints consisting of
5182/5754, 5182/6022, and 5754/6022 combinations. Mechanical properties of welds were evaluated under biaxial strain, by
transverse tension test, and in stretching through OSU plane
strain testing. Tailor-welded blanks were produced by FSW of
AA 5182-H111 to AA 6016-T4 aluminum alloy (Ref 24).
These welds showed rupture during formability test because of
weld root defects. AA6082-T6 and AA6061-T6 aluminum
alloys were also friction stir welded in the recent past (Ref 25).
Microhardness values decreased both at WN and HAZ for the
same couple. Bond strength (32%) of assembly was lower
with respect to parent alloys, and fracture occurred through
TMAZ and HAZ. It has been concluded that the loss of T6
condition was responsible for deterioration in joint efciency. A
single alloy in two different heat-treated conditions was also
friction stir welded; for example, FSW of AA7075 Al alloy in
annealed and aged conditions with parameters 750/100, 1000/
150, 1250/200, and 1500/300 rpm/mmmin (Ref 26). Plates can
be satisfactorily friction stir welded with a large window of
parameters in O-temper condition, whereas the weld parameters
were stringent to obtain adequate joint strength in T6 condition.
Joints produced in O-temper condition displayed a hardness
increase in weld region with a tensile failure at base metal,
whereas those produced in T6-temper condition exhibited a
hardness drop in weld region with tensile fracture at weld zone.
FSW of 7xxx to 2xxx series Al alloy was also reported, where
complexity occurred because of phase transformation (Ref 27).
In few attempts FSW was compared with other joining
techniques. One of the examples is power beam and FSW of
5005_H14, 2024_T351, 6061_T6 and 7020_T6 Al alloys plates
with thickness 3 mm (Ref 28). Microhardness, microtensile,
macrotensile and fracture toughness were evaluated for all
joints. Fracture toughness at fusion zone was found higher for
all joints with respect to that of base alloy owing to reduction in

Volume 22(12) December 20133891

Table 1 Chemical compositions of substrates


Elements in wt.%
Alloy

Si

Mg

Ti

Fe

Cu

Zn

Mn

Al

Al-Si alloy (AS)


Al-Mg alloy (AM)

4.3
1.4

0.420
1.1

0.17
0.016

0.20
0.26

0.01
0.60

0.06
0.07

0.01
0.007

Bal
Bal

yield strength. It has been reported that difference in strength


levels for two types of joints has meager effect on toughness.
From above glimpses, it is evident that very negligible
efforts were made to produce welds consisting of cast and
wrought alloys. Tehyo et al. (Ref 29) has produced FSW joints
between SSM 356 and AA6061-T651 alloys. SZ exhibited ne
grain structure produced by mixing of two alloys. It has been
found that 6061 Al alloy moved from retreating side to upper
region of advancing side and SSM 356 alloy travelled vice
versa. TMAZ showed distorted elongated grain structure owing
to mechanical working by tool. Microhardness across weld was
low; however, at WN it was comparatively higher than TMAZ
and HAZ. Tensile strength was in the range of 180-197 MPa
with elongation being 7-9.4%, depending on welding parameters. Lee et al. (Ref 30) has carried out FSW of A356 and 6061
Al alloys at 1600 rpm with different tool speed. WN
microstructure was dominated by retreating side alloy with
onion-ring pattern at the edge of weld bead. At WN, dynamic
recrystallization resulted in ne dispersion of Si-rich particles,
dislocation cell structure, and ne equiaxed grain size. Hardness of SZ was dropped with respect to base alloy owing to
dissolution of second phase. Tensile strength in transverse
direction was close to A356 alloy, and the maximum longitudinal tensile strength was 192 MPa.
Thus, joining of the Cast to wrought aluminum alloy still
remains a partially explored area and needs more attention to
optimize welding parameters to fabricate joints with better
efciency. In this respect, WN plays a major role for
determining the quality of assembly. Therefore, in the current
investigation, major thrust has been given to quantitative
aspects of heat evolution, temperature rise, extent of strain
generation, and material transport during FSW of dissimilar Al
alloys, which are responsible for nal microstructural and
mechanical properties at WN.

2. Experimental
In the current endeavor, Al-Si (AS) and Al-Mg (AM) alloys
with dimensions 100 (l) 9 30 (w) 9 3 (t) mm3 were joined by
FSW. Chemical composition and tensile properties of base
materials are furnished in Table 1 and 2, respectively.
Welding was done in indigenously designed FSW equipment. Tool was made of high-speed steel with concave shoulder
diameter of 15 mm, pin diameter of 5 mm, and cylindrical
pin length of 2.6 mm. Tool tilt angle (3) and rotational
speed (1000 rpm) were kept constant at the time of joining.
During welding, data logger was used to record normal load,
traverse load, and spindle torque. AS and AM alloys were xed
at retreating and advancing sides, respectively. Before welding,
substrates were machined to obtain at surface along transverse
direction and then cleaned in acetone. Joining was done under
variable tool-traversing speeds as given in Table 3. It has been

3892Volume 22(12) December 2013

Table 2 Mechanical properties of substrates at room


temperature
Tensile properties

Al-Si alloy

Al-Mg alloy

UTS, MPa
Elongation, %
Hardness, VHN

233.4 11
3 0.5
71.4 1

351.8 16
12 1
94.7 1

Table 3 Friction Stir welding parameters and sample


nomenclature
Combination
Al-Si (AS) vs. Al-Mg (AM) alloy

Tool-traversing
speed, mm/min

Tool rotating
speed, rpm

70
80
130
190
240

1000
1000
1000
1000
1000

found in most reports that rotational speed of 1600 rpm was


preferable during FSW of dissimilar Al alloys (Ref 29, 30). In
the current investigation, the same has been reduced to decrease
heat generation and ash formation.
Microstructural investigation was done on transverse section
of welds. Both base alloys and welded specimens were
prepared by conventional metallographic technique, etched
with Kellers reagent and examined in optical microscope
(LEICA DM 2500M). Microhardness was evaluated across
weld line on transverse section nearly at mid thickness region
under a 50-g load with a 10-s dwelling time (LEICA VHMT
Auto). Subsize tensile specimens were prepared from transverse
section of weld as per ASTM E 8/8M-11 keeping weld line at
the center of gauge length. Gauge length and thickness of
tensile specimen were 10 and 1.5 mm, respectively. Test
was performed at a crosshead speed of 0.1 mm/min in tensile
testing machine (Hounseld) and repeated for four samples for
each set. Cross sections of failed samples were examined in
optical microscope to identify failure location.

3. Results
Microstructure of parent alloys is shown in Fig. 1. AS alloy
consisted of bright primary a-Al matrix and eutectic network of
Al-Si (Fig. 1a). Area fraction of network was small in
comparison with conventional eutectic Al-Si alloy owing to
less Si content than eutectic composition. Distribution of
Si-rich particles within grain body was scanty. AM alloy
exhibited quasi-polygonal/little elongated grains containing
dark spots within matrix indicating the presence of second
phase (Fig. 1b). As conjectured in literature, second phase was

Journal of Materials Engineering and Performance

Fig. 1

Optical microstructures of base materials: (a) AS and (b) AM alloy

Fig. 2

Optical microstructure of joint welded at 70 mm/min tool speed showing different regions across weld

presumably Mg2Si (Ref 30). Grain size of AM alloy was


30-35 lm. Grain boundary of AM alloy was decorated with
broken network of Si-rich particles.
WN microstructures after FSW are revealed in Fig. 2, 3, 4,
5, and 6. SZ was free from weld defect and decorated with
characteristics of both alloys (Ref 8). A curved line was
observed within WN. Region A in macroimage of weld was
enlarged in Fig. 2-6(c), presenting predominant characteristic
of retreating side. This area consisted of nearly homogeneous
distribution of fragmented Si-rich particles, arose from breaking
down of parent dendritic network of Al-Si. Breaking up and
redistribution of Si-rich particles at WN and TMAZ were also
indicated by Nandan et al. (Ref 2) in friction stir-welded cast

Journal of Materials Engineering and Performance

Al-Si alloys along with healing of casting defects. Region B of


macroimage exhibited development of ne-grained structure
because of dynamic recrystallization (Fig. 2-6a). The grain
sizes of this region were 23 lm (Fig. 2a) and 14 lm
(Fig. 6a) at the lowest and the highest tool-traversing speed,
respectively. The same region contained ne dispersion of
Si-rich particles, which came from parent alloys owing to
material churning by tool. At the slowest welding speed of
70 mm/min, within SZ, onion-ring pattern was found (Fig. 2d).
It was a lamellar-like structure, consisting of alternate bands of
AM and AS alloys. The stacked morphology has not been
observed for other welded specimens. Onion-ring formation
and its dependence on FSW parameters have been discussed in

Volume 22(12) December 20133893

Fig. 3

Optical micrographs of weld processed at 80 mm/min tool speed exhibiting different regions

Fig. 4

Optical micrographs of FSW joint made at 130 mm/min tool speed presenting various regions across weld

3894Volume 22(12) December 2013

Journal of Materials Engineering and Performance

Fig. 5

Optical photographs of weld fabricated at 190 mm/min tool speed revealing various regions across weld

details by Threadgill et al. (Ref 3) and Krishnan (Ref 5). The


onion-ring patterns for 7075 and/2014 Al alloys were the
outcome of etching response due to variations in grain size,
texture, and dislocation density between the rings (Ref 3).
According to their postulation during welding, extrusion of one
layer of semi-cylinder occurred with one rotation of tool. There
was a small time lag between the production of heat by tool
rotation and extrusion of hot metal by forward tool advancement (Ref 3). This cyclic process produced continuous set of
semicircular rings. Therefore, ring formation became the
function of tool geometry, tool rotation, and tool-traversing
speeds. It has been also indicated that increment in tool speed
may lead to disappearance of onion ring as it happened beyond
70 mm/min tool-traversing speed in the current study. At high
speed of tool beyond a certain value, though weld formation
was present, one of the processes became recessive owing to
too short a time gap between them resulting in the absence of
onion ring. Practical signicance of the phenomenon remained
nearly unexplored as mechanical properties of the nugget
became satisfactory and fracture path in mechanical tests was
seldom associated with this feature (Ref 3).
Microhardness distribution perpendicular to weld line has
been illustrated in Fig. 7. Welded assembly can be divided into
three regions as per microhardness proles, such as the region
outside of tool shoulder in retreating side, the zone near the
weld line, and the area outside of tool shoulder in advancing
side. The rst region at retreating side exhibited minimum
microhardness value. A maximum in microhardness prole was
obtained in shoulder-processed region. For all joints, this

Journal of Materials Engineering and Performance

middle zone revealed (dotted regions in Fig. 7) a number of ups


and downs owing to development of composite structure
through material mixing. The oscillation in hardness at nugget
has been also inferred by Threadgill et al. (Ref 3) for friction
stir-welded heat-treatable Al alloys because of contribution of
two constituent alloys.
Figure 8 shows the tensile properties of welds along with
broken tensile specimen. Ultimate tensile strength was the lowest
at the slowest tool-traversing speed (170 MPa), reached the
maximum at 80 mm/min tool-traversing speed (409 MPa), and
reduced upon further increment in tool-traversing speed (300330 MPa). Previously it has been reported that FSW joint
consisting of A356 and 6061 Al alloys exhibited tensile strength
to the tune of 185-208 MPa (Ref 30). Welded joints of AA5052AA6061 alloys reported maximum bond strength of 225 MPa
with 7% ductility, which was 73% with respect to that of AA6061
alloy (Ref 18). Ultimate tensile strength of friction stir-welded
SSM 356 to AA6061-T651 alloy joint was 180-191.3 MPa
(Ref 29). FSW of high strength 6061-5083 aluminum alloys
revealed reduction in bond strength in comparison with parent
alloy and was 63%with respect to that of 6061 alloy (Ref 17). In
all these illustrations, dimensions of tensile specimens with
respect to weld assembly, fracture location, and crack propagation have not been indicated clearly. Compared with these
reports, the current investigation showed substantial improvement in bond strength of WN for all friction stir welds except for
the joint, which was made at 70 mm/min tool-traversing speed.
Ductility of friction stir-welded joints was low in the current case
and close to the breaking strain of AS alloy.

Volume 22(12) December 20133895

Fig. 6

Optical images of FSW joint prepared at 240 mm/min tool speed illustrating different regions across weld

Failure location during tensile testing of FSW joints was


within A356 Al alloy with small shift according to weld
parameter. At the lowest bond strength (70 mm/min tool
speed), it was away from weld center (2.8 mm, Fig. 9a); at
the highest bond strength, fracture occurred close to weld center
line (Fig. 9b); and for the rest, fracture moved through region a
little away from weld center line (1.0-1.5 mm, Fig. 9c).
Fracture path did not propagate exactly through bond line. This
observation might be compared with inference of Threadgill
(Ref 31). According to that author, the interface, i.e., visible
after FSW of dissimilar material could be termed as joint-line
remnant, and the presence of the same did not affect tensile
properties of joint. Existence of sharp boundaries was also
reported for dissimilar welds like 2219-T87/7075-T6 and 5083H321/6082-T6 alloys (Ref 3).

4. Discussion
Microstructures at WN of friction stir-welded joints were
dependent on quantied values of thermal and physical states.
Thermal state could be described by pseudo heat index (PHI)
and peak temperature at WN. Physical state could be explained
by extrusion volume, strain rate, and Zenner Hollomon
parameter. These parameters can be elucidated as follows:
PHI is a signature of heat input during welding and could be
expressed as (Ref 32)

3896Volume 22(12) December 2013

PHI

RPM2
 EPL  PDcos a;
IPM

Eq 1

where RPM is the tool rotation/min, IPM the tool-traversing


speed (mm/min), EPL is the effective pin length (mm), PD is
the plunge depth (mm), and a is the tool tilt angle ().
For given tool geometry and plunge depth, peak temperature
at WN was inuenced by tool rotation and travel speed.
Ignoring the minor difference in temperatures at advancing and
retreating sides during welding, general expression of maximum temperature is (Ref 33)

a
T
x2
K
;
Tm
Vf  104

Eq 2

where T is the temperature at WN (C); Tm is the melting


point of alloy (C); x is the rotational speed (rev/s); Vf is the
forward travel speed (mm/s); and a and K are the constants
with values 0.05 and 0.70, respectively.
During welding, material in front of pin moved toward the
back of the tool and got consolidated. In unit time, the total
amount of material that traveled around was governed by
extrusion zone width. In case the material passing through the
retreating side is greater than the material travelling around the
advancing side, then volume of material passing through
extrusion zone in unit time could be given by the expression
(Ref 34):

Journal of Materials Engineering and Performance

Fig. 7

Microhardness proles along OO line of Fig. 2-4 for joints produced with (a) 70 mm/min, (b) 80 mm/min, and (c) 130 mm/min

Fig. 8 Tensile properties of friction stir-welded dissimilar Al alloy (a) variation in UTS-elongation with tool speed, and (b) broken tensile specimen (dimensions are not to scale)



V h  wr V t Vf 2Vf kd2 ;

Eq 3

where V is the volume (mm3), h is the pin length (mm), wr is


the width of extrusion zone (mm) measured from transverse

Journal of Materials Engineering and Performance

section after etching, Vt is the pin tangential velocity (mm/s),


d2 is the projected curved pin area (mm2), and k is the number of threads per unit length of pin. Now, Vt was related
with tool rpm as

Volume 22(12) December 20133897

Fig. 9

Microstructures near failure locations for different tool-traversing speeds (a) 70 mm/min, (b) 80 mm/min, and (c) 240 mm/min

Vt 2R: x;

Eq 4

where Rp is the pin radius (mm).


Second term of Eq 1 was zero in the current welding trials
because a cylindrical tool without thread was used.
Final grain size at WN obtained through continuous
dynamic recrystallization was the function of strain rate. High
strain at WN propelled ne grain structure. Strain rate has been
found to be the function of Zener Hollomon (ZH) parameter
and related as per the expression below (Ref 34):
 
Q
Eq 5
Z e_ exp
RT 0
where Z is the Zenner-Hollomon parameter, e is the strain
rate (s1), Q is the activation energy for process and considered to be 190 kJ/mol, and T is the WN temperature in K.
Now,
e
e_ ;
t

Eq 6

where e is the total strain at WN, and t is the time required


for deformation and can be calculated from the relation:
2R
t Vfp .
Total strain at WN during FSW was determined by
Reynolds (Ref 35) assuming material only passed around tool
in streamline path. In this respect, difference in ow behavior
of two Al alloys was ignored assuming minor variation at
elevated temperature. The derived expression for the same is
given by

3898Volume 22(12) December 2013

 

APR
l
:
ln

APR
l


e ln

Eq 7

where APR is the tool advance per revolution, and


2R x
l 22Rp cos1 2Rp p . In the above relationship, x is the
distance perpendicular to welding direction from retreating side
of tool. Strain would be maximum when l would reach its
maxima, and in that condition, x = 2Rp.
From the above equations PHI, T, V, e, and Z were evaluated
and presented in Fig. 10, 11-12. The inuence of individual
parameter on microstructural and mechanical properties of
joints is discussed in the following section.
It has been revealed in Fig. 10 that PHI and temperature at
WN were increased with drop in the tool-traversing speed
(Fig. 10). This trend supports the inference of Nandan et al.
(Ref 36). Moreover, the effect of temperature on second phase
has been discussed in the literature, where it has been inferred
that the temperature of 402 C helped in dissolving Mg2Si
and AlFeSi precipitates of 6061 Al alloy (Ref 30). Disappearance of hardening precipitate was also observed during FSW of
6XXX Al alloy by Nandan et al. (Ref 2). In the current study,
temperature at WN was in the range of 475-500 C (Fig. 10);
therefore, all second phases except Si-rich particles were
dissolved in solid solution.
Figure 11 exhibited a monotonic relation between volume of
material transport and tool-traversing speed. This nding
endorsed Arbegast Model (Ref 33) on extrusion zone
ow calculation, which proposed that higher tool speed
propelled more materials sweeping through extrusion zone.

Journal of Materials Engineering and Performance

Fig. 10 Variation in PHI and temperature at WN during FSW of


dissimilar aluminum alloy

Fig. 11 Volume of material passing through extrusion zone for different tool-traversing speeds during FSW of aluminum alloy

Fig. 12
alloy

Strain rate and ZH parameter during FSW of aluminum

This phenomenon did not inuence WN microstructure


directly; however, it was indirectly responsible for enhancing
strain rate at nugget (Ref 37).

Journal of Materials Engineering and Performance

The WN microstructure consisted of ne grain structure and


distribution of tiny Si-rich particles through severe plastic
deformation, and could be directly related with strain rate and
ZH parameter as shown in Fig. 12. Higher tool travelling speed
promoted faster cooling. Increased cooling rate, on the other
hand, enhanced the strain rate at WN (Ref 6, 38). Higher strain
rate propelled ner matrix grain size with increased ZH
parameter.
Microhardness prole can be co-related with welded joint
microstructure. Inherent low hardness of A356 Al alloy was
reected at retreating side. Increment in hardness near weld line
was due to composite microstructure where both the alloys
contributed. Further increment at advancing side was the
signature of higher strength of 6061 Al with respect to A356 Al
alloy. Though 6061 Al alloy lost its strength by precipitates
dissolution as mentioned before, yet it was compensated for
by matrix grain renement and appearance of ne dispersion of
Si-rich particles.
Tensile properties of WN exhibited an excellent microstructural dependence. At the lowest tool-traversing speed
(70 mm/min), heat input and peak temperature were the
highest, strain rate was the lowest, and ZH parameter was the
smallest. These resulted in the highest matrix grain size
(22 0.8 lm). Therefore, bond strength was the least for that
particular joint. Comparing with the average microhardness as
shown in Fig. 7(a), the same value also displayed minimum
among all welds (78 VHN). With increment in tool-traversing
speed (80 mm/min), WN temperature and heat input decreased,
strain rate increased, and ZH parameter enhanced. These
aspects reduced the grain size (18 1 lm) and produced tiny
Si-rich particle distribution (3-6 lm). Strength of welded
joint reached the highest level. Average microhardness of this
region also became the largest considering all transition joints
(94 VHN). Still further increment in traversing speed of tool
(130 mm/min) encouraged lowering of WN heat input and
temperature, and increments of strain rate and ZH parameter.
Matrix grain size became still smaller (15 0.6 lm). However,
size of Si-rich particles within WN was increased (4-12 lm)
owing to reduction in interaction time between material and
tool. Matrix strengthening by Si-rich particles was reduced.
This phenomenon was also revealed in the marginal drop in
average microhardness value (89 VHN) in that region
(Fig. 7c) with respect to the weld processed at 80 mm/min.
Moreover, this created local notch weakening through stress
concentration. Bond strength also decreased. Decrement in
bond strength continued with increment in traversing speed of
the tool beyond 130 mm/min, and at the highest speed of
240 mm/min, the size of Si-rich particles became 15 lm. In
that situation, positive contribution owing to grain size
decrement was overshadowed by matrix weakening through
stress concentration by large-sized Si-rich particles. The low
breaking strains of all welds (2-4%), might be because of
embrittlement due to heterogeneous microstructure at WN
restricting the deformation of grains.

5. Conclusion
In the current investigation, friction stir butt welding was
performed to join A356 and 6061 Al alloys under variable tooltraversing speeds in the range of 70-240 mm/min. Temperature rise at WN was in the range of 475-500 C during

Volume 22(12) December 20133899

welding, resulting in dissolution of precipitates of 6061 Alloy.


With increment in tool-traversing speed, strain rate and ZH
parameter increased gradually leading to ne grain structures of
6061 alloy within SZ. However, increase in tool speed was
responsible for the reduction in heat input and temperature rise
at WN. Eutectic network of Al-Si in A356 alloy and grain
boundary Si-rich phases of 6061 alloy were fragmented
because of severe deformation and became homogeneously
distributed within WN. Si-rich particle distribution was dependant on interaction time of tool with substrate; higher tool speed
led to lower transit time to produce relatively larger Si-rich
particles. Tensile strength and microhardness prole within WN
were governed by microstructure. To obtain maximum joint
efciency, ne dispersion of Si-rich particles with optimum
matrix grain size was preferable. In this respect, 80 mm/min
tool-traversing speed was found optimal to achieve joint
efciency of 116% with respect to that of 6061 Al alloy.

Acknowledgments
The authors are indebted to Director-NML for his kind support
during the study, as well as providing permission to publish the
research study. The cooperation received from Dr. A. K. Ray
during investigation is also gratefully acknowledged. The authors
are deeply indebted for the nancial support received from the
Department of Science & Technology, Govt. of India, New Delhi
through sanction letter no. SR/S3ME/028/2007 dated 08/11/2007
to carry out the investigation.

References
1. G. Cam, Friction Stir Welded Structural Materials: Beyond Al-Alloys,
Int. Mater. Rev., 2011, 56(1), p 148
2. R. Nandan, T. DebRoy, and H.K.D.H. Bhadeshia, Recent Advances in
Friction Stir WeldingProcess, Weldment Structure and Properties,
Prog. Mater. Sci., 2008, 53(6), p 9801023
3. P.L. Threadgill, A.J. Leonard, H.R. Shercliff, and P.J. Withers, Friction
Stir Welding of Aluminium Alloys, Int. Mater. Rev., 2009, 54(2), p 4993
4. Y.S. Sato, T.W. Nelson, C.J. Sterling, R.J. Steel, and C.O. Pettersson,
Microstructure and Mechanical Properties of Friction Stir Welded SAF
2507 Super Duplex Stainless Steel, Mater. Sci. Eng. A, 2005, 397,
p 376384
5. K.N. Krishnan, On the Formation of Onion Rings in Friction Stir
Welds, Mater. Sci. Eng. A, 2002, 327, p 246251
6. M. Ghosh, K. Kumar, and R.S. Mishra, Analysis of Microstructural
Evolution During Friction Stir Welding of Ultra High Strength Steel,
Scripta Mater., 2010, 63, p 851854
7. T.U. Seidel and A.P. Reynolds, Visualization of Material Flow in
AA2195 Friction Stir Welds Using Marker Insert Technique, Metall.
Mater. Trans. A, 2001, 32A, p 28792884
8. O. Frigaard, O. Grong, and O.T. Midling, A Process Model for Friction
Stir Welding of Age Hardening Aluminum Alloys, Metall. Mater.
Trans. A, 2001, 32A, p 11891200
9. M. Ghosh, K. Kumar, S.V. Kailas, and A.K. Ray, Optimization of
Friction Stir Welding Parameters for Dissimilar Aluminum Alloys,
Mater. Des., 2010, 31, p 30333037
_
lu, S. Erim, B. Goren Kral, and G. Cam, Investigation into
10. G. Ipekog
the Effect of Temper Condition on Friction Stir Weldability of AA6061
Al-Alloy Plates, Kovove Mater., 2013, 51(3) (in press)
11. G. Cam, S. Gucluer, A. C
akan, and H.T. Serindag, Mechanical
Properties of Friction Stir. Butt-Welded Al-5086 H32 Plate, Mat.-Wiss.
U. Werkstofftech., 2009, 40(8), p 638642
12. S. Benavides, Y. Li, L.E. Murr, D. Brown, and J.C. McClure, LowTemperature Friction Stir Welding of 2024 Aluminum, Scripta Mater.,
1999, 41, p 809815

3900Volume 22(12) December 2013

13. Y.S. Sato and H. Kokawa, Distribution of Tensile Property and


Microstructure in Friction Stir Weld of 6063 Aluminum, Metall. Mater.
Trans. A, 2001, 32, p 30233031
14. Y.S. Sato, S.H. Park, and H. Kokawa, Microstructural Factors
Governing Hardness in Friction-Stir Welds of Solid-Solution-Hardened
Al Alloys, Metall. Mater. Trans. A, 2001, 32, p 30333042
15. O.V. Flores, C. Kennedy, L.E. Murr, D. Brown, S. Pappu, B.M.
Nowak, and J.C. McClure, Microstructural Issues in a Friction-StirWelded Aluminum Alloy, Scripta Mater., 1998, 38, p 703708
16. Y. Bozkurt, S. Salman, and G. C
am, The Effect of Welding Parameters
on Lap-Shear Tensile Properties of Dissimilar Friction Stir Spot
Welded AA5754-H22/2024-T3 Joints, Sci. Technol. Weld. Joining,
2013, 18(4), p 337345
17. I. Shigematsu, Y.J. Known, K. Suzuki, T. Imai, and N. Saito, Joining of
5083 and 6061 Aluminum Alloys by Friction Stir Welding, J. Mater.
Sci. Lett., 2003, 22, p 353356
18. N.T. Kumbhar and K. Bhanumurthy, Friction Stir Welding of
Al 5052 and Al6061 Alloys, www.hindawi.com/journals/jm/aip/
303756.pdf
19. S.A. Khodir and S. Toshiya, Microstructure and Mechanical Properties
of Friction Stir Welded Similar and Dissimilar Joints of Al and Mg
Alloys, Trans. JWRI, 2007, 36, p 2740
20. R. Kumar and R. Singh, Effect of Friction Stir Welding on Mechanical
Properties of Dissimilar Aluminum Alloys, National Conference on
Innovative Paradigms in Engineering & Technology (NCIPET-2012);
proceedings published by Int. J. Comput. Appl. Jan 28th, 2012 (Nagpur,
India), p 1216: www.ijcaonline.org/proceedings/ncipet/number9/52561068
21. R. Palanivel and P.K. Mathews, The Tensile Behavior of Friction-StirWelded Dissimilar Aluminum Alloys, Mater. Technol., 2011, 45,
p 623626
22. J.H. Ouyang and R. Kovacevic, Material Flow and Microstructure in
the Friction Stir Butt Welds of the Same and Dissimilar Aluminum
Alloys, J. Mater. Eng. Perform., 2002, 11, p 5163
23. M.P. Miles, D.W. Melton, and T.W. Nelson, Formability of FrictionStir-Welded Dissimilar-Aluminum Alloy Sheet, Metall. Mater. Trans.
A, 2005, 36A, p 33353342
24. C. Leitao, B. Emilio, B.M. Chaparro, and D.M. Rodrigues, Formability
of Similar and Dissimilar Friction Stir Welded AA5182-H111 and
AA 6016-T4 Tailored Blanks, Mater. Des. (in press), doi:10.1016/
j.matdes.2008.12.005
25. P.M.G.P. Moreira, T. Santos, S.M.O. Tavares, V. Richter-Trummer, P.
Vilaca, and M.S.T. DeCastro, Mechanical and Metallurgical Characterization of Friction Stir Welding Joints of AA6061-T6 with AA6082T6, Mater. Des., 2009, 30, p 180187
_
lu, B. Goren Kral, S. Erim, and G. Cam, Investigation of the
26. G. Ipekog
Effect of Temper Condition on Friction Stir Weldability of AA7075 AlAlloy Plates, Mater. Technol., 2012, 46(6), p 627632
27. M.B. Prime, T. Gnaupel-Herold, J.A. Baumann, R.J. Lederich, D.M.
Bowden, and R.J. Sebring, Residual Stress Measurements in a Thick
Dissimilar Aluminum Alloy Friction Stir Weld, Acta Mater., 2006, 54,
p 40134021
28. A. Von Strombeck, G. Cam, J.F. Dos Santos, V. Ventzke, and M.
Kocak, A Comparison Between Microstructure, Properties, and
Toughness Behavior of Power Beam and Friction Stir Welds in
Al-Alloys, Proceedings of the TMS 2001 Annual Meeting Aluminum,
Automotive and Joining at New Orleans, S.K. Das, J.G. Kaufman, and
T.J. Lienert, February 1214, 2001 (Louisiana, USA), TMS, Warrendale, PA, p 249264
29. M. Tehyo, P. Muangjunburee, and S. Chuchom, Friction Stir Welding
of Dissimilar Joint Between Semi-Solid Metal 356 and AA6061-T651
by Computerized Numerical Control, Songklanakarin J. Sci. Technol.,
2011, 33, p 441448
30. W.B. Lee, Y.M. Yeon, and S.B. Jung, The Mechanical Properties
Related to the Dominant Microstructure in the Weld Zone of Dissimilar
Formed Al Alloy Joints by Friction Stir Welding, Weld. J. Mater. Sci.,
2003, 38, p 41834191
31. P. Threadgill, Terminology in Friction Stir Welding, Sci. Technol. Weld.
Joining, 2007, 12, p 357360
32. S.K. Chimbli, D.J. Medlin, and W.J. Arbegast, Minimizing Lack of
Consolidation Defects in Friction Stir Welds, 4th Symposium on
Friction Stir Welding and Processing, 25th February1st March 2007
(Orlando, FL), TMS Annual Meeting, p 135142

Journal of Materials Engineering and Performance

33. W.J. Arbegast, Modeling Friction Stir Joining as a Metal Working


Process, Hot Deformation Aluminum Alloys III, Z. Jin, A. Beaudoin,
T.A. Bieler, and B. Radhakrishnan, Ed., 26th March 2003, TMS, San
Diego, CA, p 313327
34. K.V. Jata and S.L. Semiatin, Continuous Dynamic Recrystallization
During Friction Stir Welding of High Strength Aluminum Alloys,
Scripta Mater., 2000, 43, p 743749
35. A.P. Reynolds, Flow Visualization and Simulation in FSW, Scripta
Mater., 2008, 58, p 338342

Journal of Materials Engineering and Performance

36. R. Nandan, G.G. Roy, and T. Debroy, Numerical Simulation of ThreeDimensional Heat Transfer and Plastic Flow during Friction Stir
Welding, Metall. Mater. Trans. A, 2006, 37A, p 12471259
37. N. Balasubramanian, B. Gattu, and R.S. Mishra, Process Forces During
Friction Stir Welding of Aluminum Alloys, Sci. Technol. Weld. Joining,
2009, 14, p 141145
38. C.C. Koch, K.M. Youssef, and R.O. Scattergood, Mechanical Properties of Nanocrystalline Materials Produced by In Situ Consolidation
Ball Milling, Mater. Sci. Forum, 2008, 579, p 1528

Volume 22(12) December 20133901

You might also like