Está en la página 1de 12

Int J Adv Manuf Technol (2010) 50:941952

DOI 10.1007/s00170-010-2560-3

ORIGINAL ARTICLE

Effect of process parameters on friction stir welding


of aluminum alloy 2219-T87
Kanwer S. Arora & Sunil Pandey & Michael Schaper &
Rajneesh Kumar

Received: 2 December 2009 / Accepted: 31 January 2010 / Published online: 20 February 2010
# Springer-Verlag London Limited 2010

Abstract In this work, successful friction stir welding of


aluminum alloy 2219 using an adapted milling machine is
reported. The downward or forging force was found to be
dependent upon shoulder diameter and rotational speed
whereas longitudinal or welding force on welding speed
and pin diameter. Tensile strength of welds was significantly affected by welding speed and shoulder diameter
whereas welding speed strongly affected percentage elongation. Metallographic studies revealed fine equiaxed
grains in weld nugget and microstructural changes in
thermo-mechanically affected zone were found to be the
result of combined and interactive influences of frictional
heat and deformation. A maximum joining efficiency of
75% was obtained for welds with reasonably good
percentage elongation. TEM studies indicated coarsening
and/or dissolving of precipitates in nugget. For the gas
metal arc weld, SEM investigations revealed segregation of
copper at grain boundaries in partially melted zone.
K. S. Arora : S. Pandey
Mechanical Engineering Department,
Indian Institute of Technology,
Delhi 110016, India
K. S. Arora : M. Schaper
Institute fr Werkstoffwissenschaft,
Technische Universitt Dresden,
Dresden 01069, Germany
R. Kumar
Engineering Division, National Metallurgical Laboratory,
Jamshedpur 831007, India
K. S. Arora (*)
Institute of Safety Research,
Forschungszentrum Dresden-Rossendorf,
Dresden, Germany
e-mail: k.arora@fzd.de

Keywords Friction stir welding . AA2219 . Process forces .


Mechanical characteristics

1 Introduction
Friction stir welding (FSW) is a relatively new joining
technique; initially conceived and patented by The Welding
Institute (TWI), UK [1]. Being a new process, specialized
machines termed as Friction Stir Welders are not so
common in applications and are rather expensive. This
problem has been circumvented by adaptation of conventional milling machines for the purpose of FSW [2]. Successful
implementation of FSW using adapted milling machines have
been reported for high [3], medium [4], and low strength [5]
aluminum alloys. This paper discusses the application of FSW
to AA2219 used by the Indian Space Research Organization
for fabrication of cryogenic fuel tanks. Age hardenable 2xxx
aluminum alloys exhibit mechanical properties as required for
structural application in aerospace industry, but the loss of
strength in components fabricated by fusion welding imposes
limitations on their use. This has led to the exploration of
FSW, a solid state process, which can provide engineers a
better alternative for welding of these alloys.

2 Experimental and analysis procedure


Workpieces in this research were two identical aluminum
alloy 2219 plates; measuring 250505 mm. FSW was
performed using an adapted vertical milling machine. In order
to measure the process forces, a fixture with load cells, was
designed, developed, and then interfaced to a computer
through a DAQ card using LabVIEW. Final experimentation
was done according to Taguchi L9 orthogonal array. Table 1

942

Int J Adv Manuf Technol (2010) 50:941952

Table 1 The L9 orthogonal array with parameters and their levels


Run

1
2
3
4
5
6
7
8
9

Tool
rotations
(RPM)

Welding
speed
(mm/min)

Shoulder
diameter
(mm)

Pin
diameter
(mm)

250
250
250
325
325
325
400
400
400

60
120
180
60
120
180
60
120
180

18
20
22
20
22
18
22
18
20

7
8
9
9
7
8
8
9
7

shows the orthogonal L9 array and the three levels of input


parameters. The selected independent process parameters
were tool rotational speed (N), welding speed (S), shoulder
diameter (D), and pin diameter (d). Three replications per
row of the matrix, i.e., 27 (39) plates were welded during
the experimentation using a step pin profile (Fig. 1). The
downward and longitudinal forces, i.e., Fz and Fx, respectively, were measured using the experimental setup. A
graphical plot of the recorded Fz and Fx forces for one of
the runs is shown in Fig. 2. The two surges in Fz readings
during the pre-welding phase of the process correspond to
the start of penetration of the tool pin and shoulder into the
workpieces. Similar graphical plots of forces versus time
have been recorded in previous works using dynamometers
[68]. Force readings recorded during the welding phase of
the process runs were averaged to obtain a single value,
which was used for further analysis namely pooled analysis
of variance (ANOVA) and for developing empirical models
[9].
2.1 Empirical model of process
Using the average values of the output parameters empirical
models were developed to evaluate the relation between

Fig. 1 FSW tool-pin profile

input and output process parameters, as defined by the


following equation
Y AX1 p X2 q X3 r X4 s

where, Y is the output process parameter and X1, X2, X3,


and X4 are the input parameters and p, q, r, and s are the
exponents of the input parameters. Application of log
transformation linearizes the above non-linear equation
log Y log A p log X1 q log X2 r log X3 s log X4

2
the above equation can now be written as
y b 0 b 1 x1 b 2 x2 b 3 x3 b 4 x4

where, y and x1, x2, x3, and x4 are the logarithms of the
output and input parameters, respectively; 1, 2, 3, and
4 are the corresponding coefficients. Minitab statistical
software was used to calculate the parameters of the above
mentioned first order equation for the different responses.

3 Results and discussion


3.1 Effect of welding parameters on Fx
The main effects of welding process parameters on Fx are
shown in Fig. 3. Pooled ANOVA analysis for mean effects
is presented in Table 2 and the developed empirical model
for Fx is given below
Fx 478:2N 0:468 S0:646 D0:078 d0:523

For the above model R2 =94.9% and R2adj 94:0%. Fx


represents the resistance to tools forward movement during
FSW and it increases with increasing welding speed and
shoulder and pin diameter of the tool. It is inversely
proportional to the tool rotational speed. As seen from
ANOVA analysis, welding speed is the most significant
process parameter affecting Fx. The model also represents

Int J Adv Manuf Technol (2010) 50:941952

943

Fig. 2 Fx and Fz recorded


during an experimental run

the same, with the highest coefficient being associated with


welding speed, S. With increasing welding speed, the
amount of material to be plasticized per unit time increases;
this in turn exerts higher longitudinal force on the tool.
Also, Fx increases with increasing pin diameter, this can be
attributed to the increasing area of contact between the pin
and parent material. The drag force due to tool shoulder
constitutes more than 50% of the measured Fx [7].
Therefore, Fx should increase with increasing shoulder
diameter. However, Zhang et al. [10] have reported a
substantial increase in temperature beneath the tool shoulder with increasing shoulder diameter. This decreases the
frictional coefficient and hence Fx. Therefore, two contradicting factors govern the variation of Fx with shoulder
diameter. In the present study, a slight increase in Fx is

observed with increase in shoulder diameter. Accordingly, it


can be concluded that the drag force due to tool shoulder is
the significant of the above mentioned two factors. On the
contrary, increasing tool rotational speed decreases Fx; and
similar observation has been reported in a study including
both experiments and FEM analysis [11, 12]. This trend can
be attributed to increasing heat input per unit length of
material and subsequent decrease in flow stress caused by
more number of tool rotations per unit advance of tool,
which generates higher amount of frictional heat [12, 13].
3.2 Effect of welding parameters on Fz
The downward/axial force is a key factor for successful
FSW [14]. This force exerted by the rotating tool forges the
material, stirred from leading edge at the rear end, and plays
a significant role in weld formation [14, 15]. Insufficient or
excessive Fz can result in failure of FSW. But, as long as
Fz is within the range needed to produce defect free
welds, its influence on weld properties is modest [13]. For
the present study, the main effects of process parameters
on Fz are shown in Fig. 4. The pooled ANOVA analysis
for mean effects is shown in Table 3 and the empirical
model is
Fz 9:68N 0:312 S0:146 D2:78 d0:121

Fig. 3 Effect of process parameters on Fx

For the above model R2 =96.1% and R2adj 95:4%. As


seen from Eq. 5, Fz is proportional to the shoulder diameter
and the welding speed but is inversely proportional to the
rotational speed and the pin diameter. Additionally, from
Table 3 it can be inferred that the shoulder diameter is the
most significant parameter followed by welding and tool

944
Table 2 Pooled ANOVA for FX
(means)

Significant at 95% confidence


level

Int J Adv Manuf Technol (2010) 50:941952


Source

SS

Total rotational speed


Welding speed
Shoulder diameter
Pin diameter
Error
Total

1885224
15083860
106777
849764
808385
18734010

rotational speed. The axial thrust exerted by the tool


shoulder can be calculated from the following equation
Fz pD2 s Y =4

This equation indicates a strong dependence of axial


thrust on shoulder diameter. Quantitatively, an increase of
50% in the tool shoulder area (from 18 to 22 mm) resulted in
a 75% increase in Fz. Also, Fz increases with welding speed
because the tool has to stir more material per unit time. This
increase in Fz with welding speed is less than that observed
for Fx, due to the absence of a direct dependence of Fz on
welding speed. Contrarily, Fz decreases with an increase in
the tool rotational speed due to increased material softening
beneath the tool shoulder caused by greater frictional heat.
A similar reduction in both Fx and Fz with increasing
rotational speed was also observed by other researchers
[16, 17]. Since the tool pin is fully immersed in to the
workpieces, it bears most of the Fx but the effect of pin
diameter on Fz is insignificant. Accordingly, it has been
pooled during ANOVA analysis (Table 3).
3.3 Effect of process parameters on tensile strength
After welding, three sub-size tensile specimens for each of
the welds were machined-transverse to the welding direc-

Fig. 4 Effect of process parameters on Fz

DOF
2
2
2
2
20
26

F ratio

SS

P%

942612
7541930
53388
424882
40419
-

20.60a
164.82a
Pooled
9.46a
-

1793708
14992343
758247
1167667
18734010

9.57
80.03
4.06
6.23
100.00

tion, in conformance with American Society for Testing and


Materials (ASTM) standard E8M. Tensile tests were
conducted at a constant cross-head speed of 1 mm/min
and the tensile stress-strain curves were recorded using the
testXpert software. The empirical model for ultimate tensile
strength (UTS) is as follows
UTS 1; 064:2N 0:102 S0:117 D0:382 d0:0248

For the above model R2 =87.1% and R2adj 84:8%. The


tensile strength of the joints decreases with increasing tool
rotational speed (Fig. 5). The base material AA2219 is a
heat treatable alloy which gains strength from second phase
precipitates. During FSW, the temperature rises to about
400-500C in the nugget zone [18]. This rise causes
dissolution and/or overaging of these precipitates in the
welds. Since, the heat input during welding increases with
increasing tool rotational speed therefore the amount of
dissolution of precipitates and degree of overaging also
increase proportionally. Also, increasing tool rotations leads
to turbulence in the weld joint and can result in insufficient
consolidation of the plasticized material and thereby
degradation of strength of joints formed with higher tool
rotations [19]. Besides, simulation of FSW process has also
indicated degradation of joint quality with increasing tool
rotational speed due to a serious increase in weld flash [10].
In contrast, heat input decreases with increasing welding
speed as the time of interaction between tool and workpieces is reduced and hence, frictional heat generated per
unit length of weld is also reduced. Therefore, an increase
in strength is observed. Also, welding speed is a more
significant factor than rotational speed as can be seen from
ANOVA analysis (Table 4).
The frictional heat generated depends on the area of
contact between the tool and the workpieces. Shoulder
geometry has been shown to have a greater influence of
resultant weld properties as compared to pin geometry [20].
An increasing shoulder diameter increases the heat input
and reduces strength values (Fig. 5). The shoulder diameter
is the second most significant factor affecting tensile
strength followed by tool rotational speed. This can be
attributed to the fact that with increasing shoulder diameter,
the area/width of the softened nugget region increases in
addition to the additional heat input. These two factors then

Int J Adv Manuf Technol (2010) 50:941952


Table 3 Pooled ANOVA for Fz
(means)

Significant at 95% confidence


level

945

Source

SS

Tool rotational speed


Welding speed
Shoulder diameter
Pin diameter
Error
Total

7375262
12726067
135162846
314261
8913340
164177518

combine to give a predominant role to shoulder diameter


over tool rotational speed. The insignificant effect of tool
pin diameter on tensile strength can be understood by the
small heat input contribution of the FSW pin. This heat
input has been reported to be as little as 0.1% [21] or 2%
[22]. Therefore, it can be said that insignificant heating
effect is caused by increasing tool pin diameter. This is also
confirmed in ANOVA analysis (Table 4) where being the
least significant factor, pin diameter is pooled.
3.4 Effect of process parameters on percentage elongation
The transverse tensile specimens gage length is an
inhomogeneous zone where strain localization occurs in
the minimum hardness zone. All but few of the tested
tensile specimens fractured at the thermo-mechanically
affected zone (TMAZ)/nugget zone boundary, where
minimum hardness was observed. Additionally, it can be
argued that the parameters which increase the nugget area
and the corresponding heat input, lead to better elongation
values as shown by the result of ANOVA analysis (Table 5
and Fig. 6). This is because, with increasing nugget size,
the degree of homogeneity (uniform grain structure) of the
gage length increases. This desist the confinement of strain

Fig. 5 Effect of process parameters on tensile strength

DOF
2
2
2
2
20
26

F ratio

SS

P%

3687631
6363033
67581423
157130
445667
-

8.27a
14.28a
151.64a
Pooled
-

6483928
11834733
134271512
11587343
164177518

3.95
7.21
81.8
7.06
100

localization to a small region. Hence, the gage length


deforms uniformly during the tensile tests resulting in
evenly distributed elongation along the gage length and
higher elongation values.
3.5 Metallography
Samples for metallographic examination were cut transverse to the welding direction from mid-length of the
friction stir joints. These were cold mounted, wet grounded,
and polished to a mirror finish. Macroscopic examination
was carried out by etching using a caustic etch, i.e., 10 g
NaOH in 90 ml H2O with desmutting in a 50% HNO3
solution and a final rinse in water. Macrograph of an FS
weld is shown in Fig. 7. Microscopic examination was
done using Dix and Kellers reagent. The weld microstructure was captured using a Leica light microscope coupled
with a Qwin image analysis. Figure 8 shows the microstructure observed at five different locations in an FS
weld.
3.5.1 Nugget
The weld nugget is formed where the tool pin penetrates
the joint. It consists of recrystallized microstructure of
very fine grains, formed as a result of the severe plastic
deformation and high temperature. Onion rings are
observed here (Fig. 8), in a plane 90 to the rotation plane
of the tool and transverse to the weld direction. It has been
postulated [23] that each rotation of the tool during FSW
process, results in extrusion of semi-cylinder layers of
material which when viewed on a sliced cross-section,
displays the familiar onion ring structure. The peculiar
macrostructure observed in Fig. 8 can be attributed to the
step profile pin (Fig. 1). The upper half of the plate
thickness worked by cylindrical portion of the pin shows no
onion rings whereas these are clearly visible for the
triangular profile worked lower half. This observation
highlights the benefit of using the peculiar pin profile as
the authors obtained defective welds with cylindrical tool
pin at higher welding speeds. The inadequate amount of
heat in lower half of joint causes insufficient plasticization
and stirring of the material. With the step pin profile, higher

946
Table 4 Pooled ANOVA for
tensile strength (means)

Significant at 95% confidence


level

Int J Adv Manuf Technol (2010) 50:941952


Source

SS

Total rotational speed


Welding speed
Shoulder diameter
Pin diameter
Error
Total

842.32
7028.81
2365.55
358.52
1309.19
11545.87

degree of material flow is achieved in the lower thickness


halves of workpieces and defects are eliminated. Additionally, the pulsating action associated with triangular portion
of tool profile [24] also helps explain the elimination of
defects in the lower half of weld joints.
3.5.2 Thermo-mechanically affected zone
TMAZ is the zone between the nugget and heat-affected
zone and it may or may not show signs of recrystallization,
depending on the material being welded. In optical micrographs of the present welds, the TMAZ is characterized by
a highly deformed structure which may result from
insufficient deformation strain and temperature and recrystallization resistance of the base alloy [25]. The advancing
side of the weld (position 5 in Fig. 8) is characterized by an
abrupt transition from pancake-shaped grains to fineequiaxed grains. Here, the material is subjected to elevated
temperature deformation leading to bending of grains (in
the plane of the metallographic section) towards TMAZ/
nugget boundary. On the other hand, there is an obscure
transition on the retreating side of the weld (position 4),
where subgrains nucleate at the grain boundary of the larger
base material grains (Fig. 8). This region can be thought of
as the nucleation phase of recrystallization, which is
experienced to full extent in weld nugget. The different
microstructures observed on either side of the weld nugget
can be caused by the torsion (rotating motion of the tool)
and the linear (translational motion of the tool) velocity
fields having opposite direction on the advancing side but
same on the retreating side. Hence, the simultaneous effect
of the generated heat and deformation induced by the tool
result in microstructural changes in TMAZ.
Table 5 Pooled ANOVA for
percent elongation (means)

Significant at 95% confidence


level

Source

SS

Total rotational speed


Welding speed
Shoulder diameter
Pin diameter
Error
Total

6.04
30.77
2.32
3.11
3.62
45.85

DOF
2
2
2
2
20
26

F ratio

SS

P%

421.16
3514.40
1182.77
179.26
65.46
-

6.43a
53.69a
18.07a
Pooled
-

711.41
6897.89
2234.63
1701.95
11545.87

6.16
59.74
19.35
14.74
100

3.5.3 Heat-affected zone


Heat-treatable aluminum alloys develop a heat-affected
zone (HAZ) around the nugget. Theoretically, it extends
up to the shoulder diameter of the FSW tool. This zone is
similar to the one observed during fusion welding where
the material undergoes a thermal degradation due to the
heat of welding. In case of AA2219, a precipitation
hardened alloy, the degradation of HAZ occurs due to
overaging of the material and results in poor mechanical
properties. With increasing distance from the joint line, a
continuous increase in hardness up to the level of parent
material is observed (Fig. 9).
3.6 Microhardness
Vickers hardness profile of the weld joint was carried with a
100 gf load along the mid-thickness of the joint perpendicular to the welding direction. The indents were spaced at
distance of 0.3 mm in accordance with ASTM E384. Figure 9
shows the hardness variations and the corresponding
microstructural zones of a FS weld.
3.7 Grain size
FSW process can be considered as a hot-working process,
wherein the workpieces are subjected to unusually high
strains [26]. Strain in excess of 40 and strain rate of about
10 s1 during FSW process have been reported and the
residual microstructure in nugget has been found to be the
function of Zenner-Holloman parameter [27]
Z " expQ=RT

DOF
2
2
2
2
18
26

F ratio

SS

P%

3.02
15.38
1.16
1.55
0.20
-

15.02a
76.53a
5.76a
7.73a
-

5.64
30.36
1.91
2.71
5.23
45.85

12.29
66.23
4.18
5.90
11.40
100

Int J Adv Manuf Technol (2010) 50:941952

947

grain size in the weld nugget was 2.5 m, which is a


reduction by a factor of 10 as compared to the base alloy.
3.8 Transmission electron microscopy

Fig. 6 Effect of process parameters on elongation

Similarly, other postulated models such as Eqs. 9 [28]


and 10 [29] also highlight the dependence of grain size on
strain rate and temperature
D2  D2o A expQ=RT:t n
DCDRX C1 "j "k Dho expQ=RT

10

In above equations, " is the deformation strain, " is


deformation strain rate, D and Do are the residual and
initial nugget grain size, T is absolute temperature, Q is
activation energy for grain growth, R is gas constant, t is
time, and n is an exponent usually taken as unity. Grain size
was estimated according to ASTM E112, using the line
intercept procedure. The measured grain size of the base
material was 25 m in the traverse direction. For the weld
nugget, a ubiquitous reduction of grain size was observed
(Fig. 8). This observation points to the fact that the tool
shape was an inconsequential factor as far as recrystallization in the nugget zone is concerned. In this work, separate
quantification of strains and temperatures was not done. But
with help from Eqs. 8, 9, and 10 an overal reduction of
grain size in the weld nugget, it can be said that the
combined effect of strain, strain rate, and temperature on
microstructure evolution was similar for the whole of the
weld nugget.
Additionally, generation of major heat at the tool
shoulder/workpieces interface [18, 30] and the presence of
a backing plate near the root of the joint (acting as a heat
sink) created a decreasing temperature gradient from top to
the root face of the joint. Consequently, grain growth
decreased progressively from positions 1-3 and a slight
reduction of grain size is observed from the top surface to
the root face of the joint (Fig. 8). The average measured

To determine the reasons for the decrease in hardness,


transmission electron microscopy (TEM) was employed to
examine the shape and size of the precipitates present in the
various zones of the weld joint. For this purpose, thin foils
were prepared with the help of a conventional electro
polisher using an electrolyte of 30% (by volume) of nitric
acid in methanol at 20C. The foils were examined under
a Zeiss TEM operated at 200 kV. The hardening observed
in 2219 alloy at room temperature is attributed to localized
concentrations of Cu atoms forming GPZs (Guinier Preston
zones). These consist of two-dimensional Cu-rich regions
of disk-like shape, oriented parallel to {100} planes. Zone
diameter is estimated to be 30-50 [31] The precipitation
sequence can be stated as
Super saturated solid solution SSSS
! GP1 ! GP2q^ ! q_ ! q
The strengthening mechanism in age-hardening alloys is
explained in terms of obstacles to the motion of dislocations. With an increase in particle spacing, hardening
achieved in the material decreases. Inter-particle spacing
can increase due to dissolution of finer particles. Additionally, the solid solution strengthening effect of copper is
reduced by the coarsening of second phase particles as it
removes copper from the matrix. The optimum situation is
when precipitates can resist cutting by dislocations and are
too close to permit by-passing by dislocations [31]. Fig. 10
shows the presence of GP zones in the base material.
Analysis of TEM micrographs from nugget zone (Fig. 11)
shows that fine precipitates were absent and only few
coarse precipitates are distributed throughout the matrix.
This indicates that the weld nugget experienced high
temperatures resulting in coarsening/dissolution of precipitates leading to lower hardness/strength.
3.9 Confirmation experiment
The purpose of the confirmation experiment is to validate
the conclusions drawn during the analysis phase [32]. The

Fig. 7 Macrograph of a FS weld

948

Fig. 8 Macrograph and microstructures in a FS weld

Fig. 9 Hardness profile across a


FS weld

Int J Adv Manuf Technol (2010) 50:941952

Int J Adv Manuf Technol (2010) 50:941952

949

The predicted optimum range (for three confirmation


runs) can be calculated as
345  17:78 < Fx < 345 17:78
327 < Fx < 363
The average tensile strength for the three confirmation
runs were 337, 336, and 349 MPa, which was within the
confidence interval of the predicated optimal tensile
strength. The corresponding percentage elongation values
were 7.1%, 7.9%, and 9.1 %.
3.10 Comparison with gas metal arc welded joint

Fig. 10 TEM image of base material

confirmation also serves the purpose of testing specific


combination of factors and levels. The preferred levels of
significant factors depend on the quality characteristics
type, i.e., lower is better; nominal is best or higher is better.
In the present work, tensile strength of the welds were
considered as an important response characteristics and
hence, it was decided to run confirmation experiments that
may yield an optimum value of tensile strength. The
significant parameters and their optimum levels selected
are tool rpm at level 1(A1), welding speed at level 3 (B3),
and shoulder diameter at level 1 (C1). The insignificant
parameter, i.e., pin diameter was kept at level 2 (C2) so as
to obtain defect-free weld. The estimated mean for tensile
strength can be computed as
Fx A1 B3 C1  2T 318 327 322  2  311 345

9
The 95% confidence interval for the confirmation
experiments was calculated by the following equation [33]

CICE

s


1
1
Fa 1; fe Ve

neff R

GMAW was carried out using 1.2 mm wire of AA2319 as


filler metal. The welding parameters were 6.8 mm/s travel
speed, 28 V arc voltage, 235 A average current, Ar shielding,
wire feed rate 16.1 cm/s and contact tube-to-workpiece
distance of 12 mm. Metallographic specimens were prepared
as discussed above. A joint efficiency of 60 % (UTS
272 MPa and yield strength 156 MPa) was achieved. The
tensile specimen fractured parallel to the fusion boundary.
One probable reason for this and the reduced tensile strength
can be the occurrence of partially melted zone (PMZ)
cracking. Partially melted zone cracking in aluminum alloys
has also been reported by several researchers [3436]. For an
Al-Cu system, the eutectic temperature is 548C and the
eutectic concentration is at 33.2 % Cu. The Al2Cu is
tetragonal with 12 atoms in the unit cell and vickers hardness
of the phase at room temperature is about 500 HV [37].
Figure 12 is an optical micrograph showing a MIG weld
with PMZ on the left. According to the Al-Cu phase
diagram, the liquation zone is in the narrow region
immediately outside the fusion zone, where the maximum
temperature experienced during welding ranges from
liquidus temperature of about 642C on the fusion zone

10

where, Ve is error due to variance, is the confidence level,


F(1,fe) is the F ratio at the confidence level of (1) at 1
and error degree of freedom (fe).
neff

N
1 total DOF associated with the estimation of mean

11
where, N is the total number of experiments, 27 (9 trials3
replications) and R the number of confirmation runs, i.e.,
three. So, according to Table 4, error variance Ve =65.46 and
error d.o.f=20. Hence, F0.05(1,20)=4.35 and CICE =17.78.

Fig. 11 TEM image of weld nugget

950

Int J Adv Manuf Technol (2010) 50:941952

Fig. 12 Optical micrograph of GMA weld

Fig. 14 SEM image of a eutectic particle in PMZ

side (right) to eutectic temperature of 543C on the base


metal side (left) [31]. A scanning electron (SE) micrograph
of the base metal is shown in Fig. 13 with large particles. An
EDX analysis indicates that the Al/Cu weight ratio (53/46)
of these particles is close to that of about 53/47 for
(Al2Cu). An SE micrograph of a eutectic particle (compositelike structure) in PMZ is shown in Fig. 14. This is confirmed
by EDX with an Al/Cu ratio of 33/67. Figure 15 shows an
EDX line scan across a grain boundary in the PMZ
indicating that liquation takes place at GBs.

significantly by shoulder diameter and slightly by both


tool rotational and welding speeds. Whereas, Fx is
affected strongly by welding speed and slightly by tool
rotational speed and pin diameter.
2. Deterioration of tensile properties is experienced in
case of welded specimens as compared to the base
material values. Tensile strength of the welds is
significantly affected by welding speed and shoulder
diameter and slightly by welding speed. Welding speed
is the most significant parameter effecting percentage
elongation.

4 Conclusions
1. Process forces (Fz and Fx) are critical for the selection
of a suitable milling machine. Axial thrust is affected

Fig. 13 SEM image of base material

Fig. 15 EDX line scan across a grain boundary in PMZ

Int J Adv Manuf Technol (2010) 50:941952

3. Vickers hardness value is lowest in the nugget where


TEM studies showed the dissolution and coarsening of
second phase particles.
4. Microstructure in nugget consisted of recrystallized grain
structure with an average decrease in grain size by a
factor of 10. Microstructural changes in TMAZ resulted
from the combined effect of heat and deformation.
5. GMA-weld microstructure showed liquation in the
PMZ of the weld. This led to the embrittlement of
grain boundaries and subsequent decrease in strength of
the GMA weld joint.

951

14.

15.

16.

17.
Acknowledgement The authors gratefully acknowledge Indian
Space Research Organization (ISRO) for providing material for
research work. One of the authors (KSA) gratefully acknowledges
Deutscher Akademischer Austausch Dienst (DAAD) for financial
assistance and opportunity to carry out a part of the research work in
TU Dresden, Germany as a DAAD Fellow.

18.

19.

20.

References
1. Thomas WM, Nicholas ED, Needham JC, Murch, MG, TempleSmith P, Dawes CJ (1991) G. B. Patent, Application No. 9125978.8
2. Williams SW (2001) Welding of airframes using friction stir. Air
Space Eur 3(34):6466
3. Lockwood WD, Tomaz B, Reynolds AP (2002) Mechanical
response for friction stir welded AA2024: experiment and modeling. Mater Sci Eng, A 323:348353, PII: S0921-5093(01)01385-5
4. Minton T, Mynors DJ (2006) Utilization of engineering workshop
equipment for friction stir welding. J Mater Process Technol 177
(13):336339. doi:10.1016/j.jmatprotec.2006.03.227
5. Deqing W, Shuhua L, Zhaoxia C (2004) Study of friction stir
welding of aluminum. J Mater Sci 39:16891693
6. Johnson R (2001) Forces in friction stir welding of aluminum
alloys. Third International Symposium on Friction Stir Welding,
Japan
7. Sorensen CD, Stahl AL (2007) Experimental measurements of
load distribution on friction stir weld pin tools. Metall Mater
Transaction B 38:451459. doi:10.1007/s11663-007-9041-6
8. Cavliere P, Campanile G, Panella F Squillace A (2006) Effect of
welding parameters on mechanical and microstructural properties
of AA6056 joints produced by friction stir welding. J Mater Process
Technol 180:263270. doi:10.1016/j.jmatprotec.2006.06.015
9. Chattopadhyay KD, Verma S, Satsangi PS, Sharma PC (2009)
Development of empirical models for different process parameters
during rotary electric discharge machining of copper-steel (EN-8)
system. J Mater Process Technol 209(3):14541465. doi:10.1016/
j.jmatprotec.2008.03.068
10. Zhang Z, Zhang HW (2009) Numerical studies on controlling of
process parameters in friction stir welding. J Mater Process
Technol 209:241270. doi:10.1016/j.jmatprotec.2008.01.044
11. Chen C, Kovacevic R (2004) Thermomechanical modeling and
force analysis by friction stir welding by finite element method.
Proceedings of the Institution of Mechanical Engineers, Part C. J
Mech Eng Sci 218:509519
12. Zhang Z, Zhang HW (2007) Numerical studies on effect of axial
pressure in friction stir welding. Sci Technol Weld Join 12
(3):226248
13. Yan J, Sutton MA, Reynolds AP (2005) Process-structureproperty relationships for nugget and heat affected zone regions

21.

22.

23.

24.

25.

26.

27.

28.

29.

30.

31.
32.

for nugget and heat affected zone regions of AA2524-T351


friction stir welds. Sci Technol Weld Join 10(6):725736
Lombard H, Hattingh DG, Steuwer A, James MN (2009) Effect of
process parameters on the residual stresses in AA5083-H321
friction stir welds. Mater Sci Eng, A 501:119124. doi:10.1016/j.
jmsea.2008.09.078
Kumar K, Kailas SV (2008) On the role of axial load and effect of
interface position on the tensile strength of friction stir welded
aluminum alloy. Mater Des 29:791797. doi:10.1016/j.jmatdes.
2007.01.012
Ding RJ, Oelgetz PA (1999) Mechanical property analysis in the
retracted pin tool (RPT) region of friction stir welded (FSW)
Aluminum-lithium 2195. Proceedings of First International Symposium on Friction Stir Welding, USA
Muthukumaran S, Mukherjee SK (2008) Multi-layered metal flow
and formation of onion rings in friction stir welds. Int J Adv
Manuf Technol 38:6873. doi:10.1007/s00170-007-1071-3
Nandan R, DebRoy T, Bhadeshia HKDH (2008) Recent advances in
friction stir weldingprocess weldment structure and properties.
Prog Mater Sci 53:9801023. doi:10.1016/j.jpmatsci.2008.05.001
Li Y, Murr LE, McClure JC (1999) Solid state flow visualization
in the friction stir welding of 2024 Al to 6061 Al. Scr Mater
40:10411046
De Filippis SA, LAC CP (2007) Influence of shoulder geometry
on microstructure and mechanical properties of friction stir welded
6082 alloy. Mater Des 28:11241129. doi:10.1016/j.jmatdes.2006.
01.031
Peel MJ, Steuwer A, Wither PJ, Dickerson T, Shi Q, Shercliff H
(2006) Dissimilar friction stir welds on AA5083-AA6082. Part 1
process parameters effects thermal history and weld properties.
Metall Mater Trans A 37:21832193
Russell MJ, Shercliff HR (1999) Analytical modeling of microstructure development in friction stir welding. Proceedings of the
First International Symposium in Friction Stir Welding, USA
Krishnan KN (2002) On the formation of onion rings in friction
stir welds. Mater Sci Eng A 327:246251, PII: S0921-5093(01)
01474-5
Elangovan K, Balasubramanian V (2007) Influences of pin profile
and rotational speed of the tool on the formation of friction stir
processing zone in AA2219 aluminum alloy. Mater Sci Eng A
459:718. doi:10.1016/j.jmsea.2006.12.124
Paglia CS, Jata KV, Buchheit RG (2006) A cast 7050 friction stir
weld with scandium: microstructure, corrosion and environmental
assisted cracking. Mater Sci Eng A 424:196204. doi:10.1016/j.
jmsea.2006.03.065
Su JQ, Nelson TW, Mishra R, Mahoney M (2003) Microstructural
investigation of friction stir welded 7050-T651 aluminum. Acta
Mater 51:713729. doi:10.1016/S1359-6454(02)00449-4
Jata KV, Semiatin SL (2000) Continuous dynamic recrystallization during friction stir welding of high strength aluminum alloys.
Scr Mater 43:743749, PII: S1359-6462(00)00480-2
Benavides S, Li Y, Murr LE, Brown D, McClure JC (1999) Low
temperature friction-stir welding of 2024 aluminum. Scr Mater 41
(8):809815
Fratini L, Buffa G (2007) Continuous dynamic recrystallization
phenomena modelling in friction stir welding of aluminium alloys:
a neural network-based approach. Proc IMechE Part B: J Engg
Manuf 221:857864. doi:10.1243/09544054JEM674
Zhang Z, Liu YL, Chen JT (2009) Effect of shoulder size on the
temperature rise and the material deformation in friction stir
welding. Int J Adv Manuf Technol 45:889895. doi:10.1007/
s00170-009-2034-7
Russell AM, Lee KL (2005) Structureproperty relations in
nonferrous metals. Wiley, New York
Kumar S, Kumar P, Shan HS (2008) Optimization of tensile
properties of evaporative pattern casting process through Tagu-

952
chis method. J Mater Process Technol 204(1-3):5969.
doi:10.1016/j.jmatprotec.2007.10.075
33. Ross PJ (1995) Taguchi techniques for quality engineering: loss
functions, orthogonal experiments, parameter and tolerance
design. Mc-Graw Hill Professional
34. Katoh M, Kerr HW (1987) Investigation of heat-affected zone
cracking of GTA welds of AlMgSi alloys using the Varestraint
test. Weld J 66:360-s368-s

Int J Adv Manuf Technol (2010) 50:941952


35. Miyazaki M, Nishio K, Katoh M, Mukae S, Kerr HW (1990)
Quantitative investigation of heat-affected zone cracking in
aluminum Alloy 6061. Weld J 69:362-s371-s
36. Huang C, Kou S (2004) Liquation cracking in full-penetration Al
Cu welds. Weld J 83:50-s58-s
37. Mondolfo LM (1976) Aluminum-copper system. In. Aluminum
alloys: structure and properties, pp. 253-278, Butterworths,
London, England

También podría gustarte