You are on page 1of 19

11.

Forming of Friction Stir Welded Blanks


Vinayak R. Malik[1] R. Ganesh Narayanan[2]Satish V. Kailas[1]

11.1. INTRODUCTION
Friction stir welding is an emerging solid state joining process. Seeing its potential, it is being implemented extensively for
joining of aluminum alloys used in automotive and aerospace industries. In this context, a situation arises when the banks made
by this process need to be formed. This chapter discusses on the formability aspects of friction stir welded sheets, but prior to
discussing formability, we need to understand the friction star welding process in detail.

Friction stir welding (FSW) is a new process invented and patented by Welding Institute of Cambridge, UK in 1991. It is used for
applications where the original metal characteristics must remain unchanged as far as possible. This process is primarily used
on aluminum (originally intended for welding of aerospace alloys, especially aluminum extrusions), and most often on large
pieces which cannot be easily heat treated post weld to recover temper characteristics.

FSW is an extension of conventional friction welding (FRW). In conventional friction welding, heating of interfaces is achieved
through friction by rubbing two contacting surfaces, but in the FSW process, a third body is rubbed against the two surfaces to
be joined in the form of a rotating non-consumable tool that is plunged into the joint. The contact pressures cause frictional
heating, raising the temperature [1]. Friction stir welding (FSW) is a solid state welding process (without fusion/no melting)
where a cylindrical-shouldered tool, with a profiled threaded/unthreaded probe (nib or pin) is rotated at a constant speed and
fed at a constant traverse rate into the joint line between two pieces of sheet or plate material, which are butted together. The
parts have to be clamped rigidly onto a backing bar in a manner that prevents the abutting joint faces from being forced apart.
The length of the nib is slightly less than the weld depth required and the tool shoulder should be in intimate contact with the
work surface. The nib is then moved against the work, or vice versa. The rotation action and the specific geometry of the FSW
tool generate friction and mechanical working of the material. This in turn generates the heat and the mixing or stirring
necessary to transport material from one side of the joint line to the other and to form a joint. This heat, along with the heat
generated by the mechanical mixing process and the adiabatic heat within the material, cause the stirred materials to soften
without reaching the melting point (hence cited a solid-state process), allowing the traversing of the tool along the weld line in a
plasticized tubular shaft of metal. As the pin is moved in the direction of welding, the leading face of the pin, assisted by a
special pin profile, forces plasticized material to the back of the pin while applying a substantial forging force to consolidate the
weld metal. The welding of the material is facilitated by severe plastic deformation in the solid state, involving dynamic
recrystallization of the base material. The two technical terms which need to be understood in FSW are (a) advancing side, and
(b) retreating side. These require knowledge of the tool rotation and travel directions. Figure 11.1 shows the schematic of FSW.
The tool rotates in the clockwise direction and moves in the direction shown (away from the reader). When the direction of the
velocity vector of the tool and traverse direction are same, the side is called the advancing side of the weld, and when the
direction of the velocity vector is opposite to the traverse direction, it is called the retreating side [2].

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 11.1 Schematic diagram of friction stir welding (FSW) [2]; with permission from Elsevier,
Copyright [2008]

11.2. FSW PROCESS PARAMETERS


The three major process parameters identified during FSW of materials are as follows:

i. Tool rotation and traverse speeds

The tool speeds that influence the quality of weld are (a) the speed at which the tool is rotated, and (b) the speed at which
the tool is traversed (translated) along the joint. These two parameters have considerable importance and must be chosen
with care to ensure a successful and efficient welding cycle. The relationship is complex between the welding speeds and
the heat input during welding. In general, it can be said that increasing the rotation speed or decreasing the traverse speed
will result in a hotter weld. To produce a successful weld it is necessary that the material surrounding the tool is hot enough
to enable the extensive plastic flow required for welding and to reduce the forces acting on the tool. If the material is too
cool then voids or other flaws occur in the stir zone and in extreme cases this may lead to damage and breakage of tool. On
the other hand, high heat input may be detrimental to the final properties of the weld. This can even induce defects due to
the liquation of low-melting-point phases (just like liquation cracking in fusion welds). This leads onto the concept of a
'processing window'. This process window gives the range of process parameters that will produce a good quality weld.
Within this window the resulting weld will have a sufficiently high heat input to ensure adequate material plasticity but not so
high that will deteriorate the weld properties.

To prevent tool fracture and to minimize excessive wear and tear on the tool, the welding cycle needs to be modified so that
the forces acting on the tool are as low as possible. For the best combination of welding parameters it is likely that a
compromise must be reached, since the conditions that favor lower forces (example: high heat input, low travel speeds)
may be undesirable from the point of view of productivity and weld properties.

ii. Tool design

The increased speed of welding is necessary for improving the production rate and keeping this process competitive
compared to others. Proper design of tool plays a critical role in increasing the welding speed and improving the quality of
weld. To achieve the above mentioned objective the tool material should be sufficiently strong, tough and wear resistant at
the welding temperature. It should also have a good oxidation resistance and a low thermal conductivity to minimize heat
loss and thermal damage to other parts of welding machine which are in vicinity. Few tools such as hot-worked tool steel
(for example AISI H13) have been tried and tested for welding aluminum alloys within thickness range from 0.5 to 50 mm
[3]. They proved to be perfectly all right but for more demanding applications like highly abrasive metal matrix composites,
higher melting point materials such as steel or titanium, more advanced tool materials are required. Improvements in tool

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
design have shown to cause appreciable improvements in productivity and quality. Specifically designed tools are now
available to increase the depth of penetration and plate thickness that can be successfully welded. The 'whorl' design is an
example of a tool that uses a tapered pin with re-entrant features or a variable pitch thread in order to assist downward flow
of material. Other designs include the Triflute and Trivex series [4]. The Triflute design has a complex system of three
tapering, threaded re-entrant flutes that help in increasing material movement around the tool.

The Trivex tools use a simple non-cylindrical pin and it is observed that they reduce the forces acting on the tool during
welding. A concave shoulder profile is attributed to majority of tools for the reasons such as (a) it acts as an escape volume
for the material displaced by the pin, (b) it prevents material from extruding out of the sides of the shoulder, and (c) it
maintains downward pressure and hence good forging of the material behind the tool. To produce additional movement of
material in the upper layers of the weld a triflute tool is used which uses an alternative system with a series of concentric
grooves machined into the surface.

iii. Tool tilt and plunge depth

Plunge depth and tool tilt are critical parameters for ensuring good quality of weld (Fig. 11.2). The plunge depth is the depth
of the downward extreme point of the shoulder below the surface of the plate being welded. Plunging the shoulder below the
surface of the plate at the joint increases the pressure at the bottom of the tool and helps in ensuring adequate forging of
the material at the rear portion of the tool. Tilting the tool by an angle of 2–4°, such that the rear portion of the tool is slightly
lower than the front portion, helps in assisting the forging process required in this case of welding [4]. The plunge depth
needs to be appropriately set to ensure the necessary downward pressure of the tool on the material to be welded and for
full penetrations of tool probe and bottom into the material to be welded. On the other hand, if the plunge depth provided is
more, the high downward force generated may lead to deflection of welding machine and may result in rubbing of the pin on
the surface of the backing plate or a significant under-match of the weld thickness compared to the base material. If the
plunge depth provided is less it may even result in flaws in the weld.

Figure 11.2 Schematic diagram showing plunge depth and tool tilt. The movement of the tool is
from right to left

11.3. METALLURGY OF FRICTION STIR WELDS


Based on the microstructural characterization of grains and precipitates, the FS weld can be divided into three distinct zones

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
viz., (a) Stirred (Nugget) zone, (b) Thermo-mechanically affected zone (TMAZ), and (c) Heat-affected zone (HAZ) as shown in
Fig. 11.3.

Figure 11.3 Schematic diagram of cross-section of a typical FSW weld showing four distinct zones:
(A) base metal, (B) heat-affected, (C) thermomechanically affected, and (D) stirred (nugget) zone

i. Stirred (nugget) zone

Frictional heating and heavy plastic deformation during FSW result in generation of a recrystallized fine-grained
microstructure within stirred zone. This region is referred to as nugget zone (or weld nugget) or dynamically recrystallized
zone (DXZ). Under some FSW conditions, concentric rings are observed in the nugget zone. These concentric rings are
referred to as the 'onion rings' in FSW [5]. There is low dislocation density in the interior of the recrystallized grains.
However, from some investigations it is found that the small recrystallized grains of the nugget zone contain high density of
sub-boundaries, sub-grains, and dislocations. The interface between the recrystallized nugget zone and the parent metal is
relatively blurred on the retreating side of the tool, but quite sharp on the advancing side of the tool [6].

ii. Thermo-mechanically affected zone (TMAZ)

This is a unique zone observed in FSW process. It can be termed as a transition zone, thermo-mechanically affected zone
(TMAZ) between the parent material and the nugget zone. The TMAZ experiences both temperature and deformation during
FSW. The TMAZ is characterized by a highly deformed structure. The grains in the parent metal are deformed in an upward
flowing pattern around the nugget zone. Although the TMAZ undergoes plastic deformation, recrystallization does not occur
in this zone due to insufficient deformation strain. However, dissolution of some precipitates occurs in this zone for certain
alloys which show precipitates in their microstructure. The extent of dissolution depends on the thermal cycle experienced
by TMAZ. The grains in the TMAZ usually contain a high density of sub-boundaries.

iii. Heat-affected zone

Beyond the TMAZ there is a heat-affected zone (HAZ). This zone experiences a thermal cycle, but does not undergo any
plastic deformation. The HAZ retains the same grain structure as the parent material. In case of heat-treatable aluminum
alloys this zone experiences a temperature rise above 250°C [7]. But the thermal exposure above 250 °C exerts a significant
effect on the precipitate structure which in turn affects the properties. FSW process has relatively little effect on the size of
the sub-grains in the HAZ, but for certain alloys it results in coarsening of the strengthening precipitates and the precipitate-
free zone (PFZ) increases by an appreciable amount. These changes in HAZ need to be considered during welding as they
may bring undesirable changes in mechanical properties.

11.4. MECHANICAL PROPERTIES OF FRICTION STIR WELDS


As there is no melting involved in FSW, the quality of weld obtained is superior to other fusion welding processes which show a
cast structure. FSW produces a unique microstructure within and around the stirred zone, i.e., nugget zone, TMAZ, and HAZ.
The evolution of mechanical properties such as strength, ductility, fatigue, fracture toughness, and hardness during FSW are
discussed in this section. Some important points are highlighted on residual stresses.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
i. Strength and ductility

Tensile properties tested on FSW specimens along (longitudinal) the weld show better strength compared to across
(transverse) the weld. This is because longitudinal tensile specimens contain only fully recrystallized grains from the nugget
zone, whereas transverse tensile specimens contain microstructures from all four zones, i.e., parent material, HAZ, TMAZ,
and nugget zone. The strength and ductility in transverse orientation are substantially less than those in longitudinal
orientation. In general, shear failures occurs in HAZ. This is because the tensile specimens in the transverse orientation
cover four different microstructures. The observed ductility is an average strain over the gage length including various
zones. The different zones have different resistances to deformation due to differences in grain size, precipitate size and
distribution. The HAZ has the lowest strength due to significantly coarsened precipitates and the development of the
precipitate free zones. Thus, during tension, strain occurs mainly in the HAZ resulting in necking and fracture. Therefore,
fracture always occurs in the HAZ, resulting in a low strength and ductility along transverse orientation of the weld.

The transverse tensile strength of FSW welds and joining efficiency of FSW welds for various aluminum alloys are
summarized in Mishra and Ma [8]. It is seen that the joining efficiency of FSW welds ranges from 65 to 96% for heat-
treatable aluminum alloys and is 95–119% for non-heat-treatable aluminum alloy. The joining efficiency for FSW is
significantly higher than that for conventional fusion welding, particularly for heat-treatable aluminum alloys [8].

Normally the strength of FSW is around 80% of the base metal which is much better than other fusion welds. The retreating
side exhibits lower strength to its contrary advancing side, so failures are seen at the retreating side.

ii. Fatigue behavior

There is a growing interest in the structural use of aluminum alloys, for applications such as automotive and railway
vehicles, bridges, offshore structure topsides and high-speed ships. In all cases, welding is the primary joining method and
fatigue is a major design criterion. However, as is well known, welded joints can exhibit poor fatigue properties. Thus, clear
design guidelines are needed to ensure that fatigue failures are avoided in welded aluminum alloy structures. Apart from
basic design of new structures, there is also increasing interest in methods for assessing the remaining fatigue lives of
existing structures. A substantive work needs to be carried out in above mentioned area not only for aluminum but also for
other materials which can be joined using FSW. It is important to understand the fatigue characteristics of FSW welds due
to its variety of applications like aerospace structures, transport vehicles, bridge constructions, etc., where fatigue
properties are critical. Substantive work is being carried out in evaluating the fatigue behavior of FSW welds, which
importantly include stress—number of cycles to failure (S–N) behavior and fatigue crack propagation (FCP) behavior.

Based on the research carried out especially of S–N behavior and fatigue crack propagation (FCP) behavior of FSW welds,
following points are observed. First, the fatigue strength of the FSW welds at 107 cycles is lower than that of the base metal,
which indicates they are weaker against fatigue crack initiation.

The transverse FSW specimens have lower fatigue strength than the longitudinal FSW specimens. However, it is observed
that the fatigue strength of the FSW welds is higher than that of MIG and laser welds [8]. Such studies need to be carried out
for commercially important materials. A proper mechanism should be established to understand the fatigue failure in FSW
in order to t improve the fatigue life of FSW components. The study should be such that it can be applied to any material
joined using FSW. The reason behind better properties of FSW welds is its finer and uniform microstructure as compared to
fusion (laser and MIG) welds. Second, surface quality of the FSW welds exerts a significant effect on the fatigue strength of
the welds. The fatigue strength of the FSW weld decreases with increasing tool traverse speed/rotation rate (v/w) ratio due
to the formation and increase of non-welded groove at the root side of the weld [8]. Skimming of layers (up to 0.5 mm) from
both root and top sides removes all surface irregularities and results in fatigue strength, of both transverse and longitudinal
FSW specimens, comparable to that of the base metal. Third, the fatigue resistance of friction stir welded specimens in air
is inferior to that of the base metal.

In case of FCP (fatigue crack propagation), the study reveals following important observations. First, the quality of the FSW
welds exerts limited effects on the da/dN–DK curve, where da/dN–DK gives idea about fatigue crack propagation (the data
is presented in terms of crack growth rate per cycle of loading (Da/ DN or da/dN) versus the fluctuation of the stress-

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
intensity factor at the tip of the crack (DKI). DKI is representative of the mechanical driving force, and it incorporates the
effect of changing crack length and the magnitude of the cyclic loading, the most common form of presenting fatigue crack
growth data is a log–log plot of da/dN versus DKI). Second, at lower loads and lower R-ratio of 0.1, the FCP properties of the
FSW welds are superior to that of the base metal, whereas at higher loads or higher R-ratios of 0.7–0.8, base materials and
FSW welds exhibits similar da/dN–DK behavior. This is because of the presence of compressive residual stresses at the
crack tip region in the FSW welds, which decreases the effective stress intensity at the crack front. In this case, fatigue crack
propagation rates at lower loads and lower R-ratio are apparently reduced due to reduced effective stress intensity.
However, at higher loads or higher R-ratios, the effect of the compressive residual stress becomes less important and
similar base material and FSW da/dN–DK curves are observed. Third, specimen geometry exhibits a considerable effect on
the FCP behavior of the FSW welds. The nugget zone and HAZ of FSW welds exhibit lower fatigue crack growth rates and
higher fatigue crack growth threshold, DKth, at both R = 0.1 and 0.5, in air and in 3.5% NaCl solution, compared to the base
metal. Furthermore, the FCP properties of the nugget zone are higher than those of the HAZ. The fatigue crack growth rates
in 3.5% NaCl solution for the base metal, HAZ, and nugget zone, in the intermediate and high DK regions, are about two
times higher than those observed in air [8]. The crack growth behavior in the FSW joints is generally dominated by the weld
residual stress, while microstructure and hardness changes in the FSW welds have minor influence. Furthermore, fatigue
crack growth rates in FSW welds depends strongly on their location and orientation with respect to the weld centerline.
However, if the FSW welds are mechanically stress relieved by application of 2% plastic strain, crack growth rates are
almost identical to those of the base metal, irrespective of location and orientation.

iii. Fracture toughness

No welding process is perfect and defects are bound to occur in the welds. The weld is accepted or rejected based on the
intensity of defects. There are standards for acceptability of the welds pertaining to different inspection codes. The non-
acceptable flaws must be repaired before the weld is put into service. Most existing codes cater toward weldments made by
conventional welding techniques. FSW is generally found to produce defect-free welds. However, no established code exists
so far for FSW welds. Considering the potential applications of FSW, there is a critical need for proper evaluation of the
fracture behavior of the friction stir welds. The most commonly used parameters are the crack tip intensity factors (K) for
linear elastic loading, and the J integral or the crack tip opening displacement (CTOD) for elastic–plastic loading.

Generally it is observed that friction stir welds exhibit better fracture toughness compared to fusion welds. The reason for
higher fracture toughness associated with the FSW welds is attributed to the fracture and rounding of large primary
particles by the stirring process, and the softening of the matrix. No detailed microstructure–property correlation has been
established so far for the fracture toughness of FSW welds. Since there is a significant change in microstructures during
FSW, it is necessary to understand the influence of microstructural characteristics on the fracture toughness of friction stir
welds. For commercial precipitation-strengthened high-strength aluminum alloys, three types of particles are identified, i.e.,
large constituent particles (5–30 microns), dispersoids (0.2–0.5 microns), and precipitates in the nanometric size range. In
the absence of constituent particles and dispersoids, the deformation behavior becomes strongly influenced by shearing of
precipitates, thus, leading to strain localization. Like the conditions of strain localization in the matrix, it is found that low
fracture toughness results from the presence of a narrow and soft PFZ and a large grain size.

The most critical stage in controlling fracture toughness is the control of the initiation of voids. Particles provide interfaces
that are easy initiation sites for voids. Critical stress for particle cracking, is related to their size, and surface energy. The
two factors viz., (i) the increase in the bonding strength between matrix and particles, and (ii) the decrease in particle size
tend to increase critical stress for particle cracking, enhances the fracture toughness. Improvement in fracture toughness is
often achieved by reducing iron and silicon content in aluminum alloys, thus, reducing the volume fraction and size of
constituent particles. Apart from above-mentioned factors, the nature of grain boundaries influences fracture behavior of a
material. A large fraction of low energy grain boundaries might toughen the material by changing the failure mode from
intergranular to transgranular fracture. However, these concepts are still evolving and no quantitative relation is available to
predict fracture toughness based on grain boundary character distribution.

Based on above-mentioned microstructural analyses, the fracture toughness of FSW aluminum alloys can be rationalized.
FSW results in generation of a nugget zone characterized by: (a) very fine grain size (b) fine precipitates and constituent

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
particles, (c) lower yield stress, and (d) high ratio of high-angle boundaries. Fine grain structure and small particles tend to
enhance the fracture toughness of nugget zone, whereas low yield stress and high ratio of high-angle boundaries tend to
reduce the fracture toughness. The overall impact of these factors is that the fracture toughness of nugget zone is higher
than or comparable to that of base material, depending on the alloy chemistry and FSW parameters. The lower fracture
toughness of HAZ/TMAZ region compared to the nugget zone is due to widened PFZ (precipitate free zone) and coarsened
particles [8].

iv. Hardness

Aluminum alloys can be classified into heat-treatable (precipitation-hardenable) alloys, and non-heat-treatable (solid-
solution-hardened) alloys. The change in hardness in the friction stir welds is different for precipitation-hardened and solid-
solution-hardened aluminum alloys. FSW creates a softened region around the weld center in a number of precipitation-
hardened aluminum alloys. Such a softening is caused by coarsening and dissolution of strengthening precipitates during
the thermal cycle of the FSW. If hardness is measured across the weld at intermittent distances and a graph (distance from
weld centre in +ve and -ve direction on x-axis and hardness on y-axis) is generated using these hardness values, the profile
thus generated is called hardness profile. Hardness profile is strongly affected by precipitate distribution rather than grain
size in the weld. A typical hardness distribution across the weld of FSW 6063Al-T5 is described by Mishra and Ma [8].
Clearly, significant softening is produced throughout the weld zone, compared to the base material in T5 condition. Further,
it is shown that the lowest hardness does not lie in the center part of the weld zone, but is approximately 10 mm away from
the weld centerline. The hardness curve can be labeled by BM (the same hardness region as the base material), LOW (the
region of lower hardness than base material), MIN (the minimum-hardness region), and SOF (the softened region).

Two kinds of precipitates are observed in the BM, LOW, and MIN regions; needle-shaped precipitates of about 40 nm in
length, which are partially or completely coherent with the matrix, and rod-shaped precipitates approximately 200 nm in
length, which have low coherency with the matrix. The mechanical properties of 6063Al depend mainly on the density of
needle-shaped precipitates and only slightly on the density of rod-shaped precipitates [8]. The microstructure (type, size, and
distribution of precipitates) in the BM region is basically the same as that in the base material, which explains the same
hardness in the BM region and the base material. In the LOW region, the density of needle-shaped precipitates was
substantially reduced, whereas the density of rod-shaped precipitates increases. This results in a reduction in hardness of
the LOW region. For the MIN region, only low density of rod-shaped precipitates remains. Thus, not only the hardening effect
of needle-shaped disappears completely, but also solid-solution-hardening effect of solutes reduces due to the existence of
rod-shaped precipitates, which leads to the minimum hardness in the MIN region. In the SOF region, no precipitates are
detected due to complete dissolution of the precipitates. For the solid-solution-hardened aluminum alloys, generally, FSW
does not result in softening of the weld. For such alloys higher hardness is observed in the SOF region than in the base
material because of the smaller grain size, higher density of sub-boundaries dislocations, strain hardening. The combined
effect of the above mentioned factors leads to improved hardness of weld region in friction stir welded solid-solution-
hardened aluminum alloys.

Kumar et al. [2] studied the role of axial load on the tensile behavior and hardness distribution of FSW blanks made of 7020-
T6 material. A typical hardness distribution is shown in Fig. 11.4 for axial load of 8.1 kN and is observed that that the
minimum hardness is around 120VHN, which occurs 20 mm away from the weld. The hardness minima are observed in
both sides of the weld in the HAZ. It is also rationalized that in the HAZ precipitates are not dissolved in the matrix, and
consequently there is no reappearance of precipitates due to post weld thermal cycle. This causes the lowest hardness and
consequently fracture occurs during tensile testing in the HAZ.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 11.4 Hardness profile across the weld region for axial load of 8.1 kN [2]; with permission
from Elsevier, Copyright [2008]

v. Residual stress

The occurrence of large residual stress influences the postweld mechanical properties, particularly the fatigue properties
and its presence also shows undesirable corrosion effects if the weld is used in corrosive environment. Therefore, it is
necessary to understand the residual stress distribution in the FSW welds. Generally the residual stresses are low in friction
stir welds as it is a low temperature solid-state process. But, the rigid clamping used in FSW exerts a much higher restraint
on the welded plates. These restraints impede the contraction of the weld nugget and heat-affected zone during cooling in
both longitudinal and transverse directions, thereby resulting in generation of longitudinal and transverse stresses. Cut
compliance technique, X-ray diffraction, neutron diffraction and high-energy synchrotron radiation are some of the
techniques used to measure residual stresses in FSW. The residual stresses in all the FSW welds are quite low in
comparison to those generated during fusion welding. Higher residual stresses are observed at the transition between the
fully recrystallized and partially recrystallized regions of the weld. Generally, longitudinal (parallel to welding direction)
residual stresses are tensile and transverse (normal to welding direction) residual stresses are compressive. The low
residual stress in the FSW welds are attributed to the lower heat input during FSW and recrystallization accommodation of
stresses. Few other observations that are made related to residual stresses in FSW welds are, (i) the longitudinal residual
stresses are always higher than the transverse ones, independent of pin diameter, tool rotation rate and traverse speed, (ii)
both longitudinal and transverse residual stresses exhibit an "M"-like distribution across the weld. Maximum tensile residual
stresses are located at the HAZ region of the weld, (iii) residual stress distribution across the welds is similar at the top and
root sides of the welds, (iv) large-diameter tool widens the M-shaped residual stress distribution and with decreasing
welding speed and tool rotation rate, the magnitude of the tensile residual stresses decrease. It is observed that maximum
residual stresses for various friction stir welds of aluminum alloys are below 100 MPa [9]. The residual stress magnitudes
are appreciably lower than those observed in fusion welding, and also appreciably lower than yield stress of aluminum
alloys. This results in substantive reduction in the distortion of FSW components and an improvement in mechanical
properties.

However, for the stainless steels, the pattern of residual stress is different. This shows that the behavior of residual stress
changes with material. In case of FSW of stainless steels the residual stress patterns observed are typical of most welding
processes such as fusion welding, namely, high value of longitudinal tensile residual stress and very low transverse residual
stress. The maximum values of longitudinal residual stress are close to the base metal yield stress, and therefore similar in
magnitude to those produced by fusion welding processes in austenitic stainless steels. The longitudinal residual stress varies
only slightly with depth, whereas the transverse stress varies significantly through the thickness. The sign of the transverse
residual stress near the weld centerline is in general positive at the top and negative at the root. This is attributed to rapid
cooling experienced by the weld root due to the intimate contact between the weld root side and the backing plate. Therefore,

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
the distribution and magnitude of residual stress in friction stir welds are different for aluminum alloy and steel. This
occurrence of difference can be related to the temperature dependence of the yield strength of material welded and the
influence of final deformation by the trailing edge of the tool shoulder.

11.5. IMPORTANT PHENOMENON IN FSW


There are two important phenomenons in FSW, the knowledge of which is necessary in understanding and modeling of the
process at fundamental level. They are (i) metal flow and mechanism of joining, and (ii) heat generation and heat transfer.

i. Metal flow and mechanism of joining

The material flow during friction stir welding is quite complex and depends on the tool geometry, process parameters, and
material to be welded. For optimal tool design and to obtain high structural efficient welds, the knowledge of mechanism of
material flow is necessary. A number of investigations have been carried out to understand the material flow behavior
during FSW. A number of approaches, such as tracer technique by marker, welding of dissimilar alloys/metals, have been
used to visualize material flow pattern in FSW [10]. Understanding material flow is important for production of sound
dissimilar metal welding that control the intermixing of two alloys being welded and consequent formation of new
constituents which influences the weld properties. The material flow visualization study is normally utilized to analyze the
mechanism of weld formation and its usefulness in improving fatigue properties and for dissimilar metal welds. Few
innovative experiments had revealed that only a portion of material transported from the leading edge undergoes chaotic
flow and the remaining is deposited systematically in the trailing edge of the weld [11]. Apart from experimental
approaches, a number of studies have been carried out to model the material flow during FSW using different computational
codes, mathematical modeling tools, simple geometrical model, and metal working model (modeled using the concept of
metal forming). These attempts were made to understand the basic physics of the material flow occurring during FSW.

But the metal working model suggested seems to give results closer to experimental value. In this investigation an attempt
was made to understand the mechanism of friction stir weld formation and the role of friction stir welding tool in it. This
was done by understanding the material flow pattern in the weld produced in a special experiment, where the interaction of
the friction stir welding tool with the base material was continuously increased. The results showed that there are two
different modes of material flow regimes involved in the friction stir weld formation; namely "pin-driven flow" and "shoulder-
driven flow". These material flow regimes merge together to form a defect-free weld. The etching contrast in these regimes
gives rise to onion ring pattern in friction stir welds. In addition to that based on the material flow characteristics a
mechanism of weld formation was also proposed [12]. The resultant microstructure and metal flow features of a friction stir
weld closely resemble hot worked microstructure of typical aluminum extrusion and forging. Therefore, in this case the FSW
process is modeled as a metalworking process. The process is divided into five conventional metal working zones: (a)
preheat, (b) initial deformation, (c) extrusion, (d) forging, and (e) post heat/cool down [13]. In the preheat zone ahead of the
pin, there is a temperature rise due to the frictional heating of the rotating tool and adiabatic heating because of the
deformation of material. The thermal properties of material and the traverse speed of the tool govern the extent and heating
rate of this zone. As the tool proceeds, an initial deformation zone is formed when material is heated to above a critical
temperature and the magnitude of stress exceeds the critical flow stress of the material, resulting in material flow. In this
zone the material is forced both upwards into the shoulder zone and downwards into the extrusion zone. A small amount of
material is captured in the swirl zone beneath the pin tip where a vortex flow pattern exists. Material flows around the pin
from the front to the rear, in the extrusion zone with a certain width. The width of the extrusion zone is defined by critical
isotherm on each side of the tool where the magnitudes of stress and temperature are not sufficient to allow metal flow.
Forging zone follows the extrusion zone where the material from the front of the tool is forced into the cavity left by the
traveling pin under hydrostatic pressure conditions. Constraining the material in this cavity and application of downward
forging force is done with the help of shoulder of the tool. Material from shoulder zone is dragged across the joint from the
retreating side toward the advancing side. Post heat/cool zone is on the rear of forging zone where the material cools under
either passive or forced cooling conditions.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
The features of material flow in FSW and the joining process can be briefed into three steps as most of the material flow
occurs through the retreating side and the transport of the plasticized material behind the tool forms the welded joint. The
overall transport of plasticized materials during FSW is affected by three types of flows: (1) Near the tool, a slug of
plasticized material rotates around the tool. The rotation of the tool and the resulting friction between the tool and the
workpiece drives this movement of material, (2) Rotational motion of the threaded pin tends to push material downward
close to the pin which drives an upward motion of an equivalent amount of material somewhat farther away, and (3) There
is a relative motion between the tool and the workpiece. The simultaneous interaction of these three effects is responsible
for the overall motion of the plasticized material and the formation of the joint.

ii. Heat generation and heat transfer

The technique of friction stir welding (FSW) puts effective use of frictional heat for the purpose of joining metallic materials.
Experiments were conducted to determine the coefficient of friction during FSW to study the interaction between the FSW
tool and the base metal at various contact pressures and rotation speeds and it was found that there exists a critical
temperature at which point a steep increase in the coefficient of friction is observed. Below the critical temperature the
coefficient of friction is less, and above the critical temperature it reaches the maximum value. The steep increase in the
coefficient of friction is found to be due to the seizure phenomenon and the contact condition during FSW between the tool
and the workpiece (base metal) is found to be sticking [14]. To prevent or minimize the ill effects of welding on the
mechanical properties of the weld and to increase the productivity, it is desired to minimize the heat input and increase the
travel speed. At the same time it is necessary to ensure that the temperature around the tool is sufficiently high to permit
adequate material flow and prevent flaws or tool fracture. When the traverse speed is increased, for a given heat input, there
is less time for heat to conduct ahead of the tool and the thermal gradients are larger. At some point the speed will be so
high that the material ahead of the tool will be too cold and the flow stress is too high to permit adequate material
movement, resulting in flaws or tool fracture. If the 'hot zone' is too large then there is scope to increase the traverse speed
and hence productivity. The variable process window is responsible for the change in total heat input and cooling rate
during welding. Structural characterization of the bonded assemblies exhibits recovery-recrystallization in the stirring zone
and breaking of coarser particles (phases). Dispersion of fine particles, refinement of grain size, low residual stress level,
and high defect density within weld nugget contribute towards the improvement in bond strength.

Based on the heat flow and thermal profile the welding cycle can be divided into following four stages (i) Dwell: For smooth
traverse of the tool, the material needs to be heated. This preheat is provided by tool rotation at appropriate speed but without
linear movement in order to achieve a sufficient temperature ahead of the tool. This period also includes the plunge of the tool
into the workpiece, (ii) Transient heating: As the tool starts moving along the joint there will be a transient period where the heat
production and temperature around the tool will alter in a complex manner until a required steady-state condition is reached, (iii)
Pseudo steady-state: It is the state where thermal field at macroscopic level remains constant around tool though there are
fluctuations in heat generation, (iv) Post steady-state: It is the state where the heat reflection from the welded region leads to
additional heat around the tool [14].

There are two main sources for heat generation in FSW which are (a) friction at the surface of the tool, and (b) deformation of
the material around the tool. The heat generation predominantly occurs under the shoulder, due to its larger surface area, and is
approximately equal to the power required to overcome the contact forces between the tool and the workpiece. The contact
condition under the shoulder can be described by sliding friction, using a friction coefficient μ and interfacial pressure P, or
sticking friction, based on the interfacial shear strength and torque, at an appropriate temperature and strain rate. Mathematical
equations for the total heat generated by the tool shoulder during welding, i.e., Qtotal using sliding and sticking friction approach
are as given below.

(11.1)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(11.2)

where ω is the angular velocity of the tool, Rshoulder and Rpin are radius of the tool shoulder and tool pin, respectively. However,
there is a difficulty while applying these equations as the values for the friction coefficient or the interfacial shear stress which
are required for total heat calculations are difficult to measure in actual process. At present these parameters are used as
'fitting parameters' where the reverse calculations are performed based on measured thermal data to obtain a reasonable
simulated thermal field. This reverse approach may be useful for creating process models to predict variables like residual
stresses, but it is less useful for providing the insights into the process.

11.6. DIFFERENT JOINT CONFIGURATION FSW


Friction stir welding provides the advantages of being used in a variety of joint configurations. The joint configurations that can
be welded are butt welds, dissimilar thickness butt welds, double sided butt, tee welds, lap penetration and lap fillet [Fig. 11.5].
The configuration which can be welded conveniently are butt and lap welds as the advantage here is that no joint preparation is
required for these configurations unlike conventional welding processes. The major drawback is joining the traditional tee fillet
joint which is not possible with this process and which is commonly used in many fusion welding applications. However the
advantage of the FSW process can be taken by changing the product design [15].

Figure 11.5 Different joints possible using FSW (a) Dissimilar thickness butt (b) Lap penetration (c)
Double sided butt (d) Lap fillet (e) Corner (f) Tee (g) Full and partial penetration butt

11.7. FSW VARIANTS


The friction stir welding process has two significant variants that can be used for other applications. The first is Friction Stir
Spot Welding (FSSW), where the tool is plunged and retracted, without traversing the tool. This particular variant has parallels
with the Resistance Spot Welding (RSW) process and has significant advantages over RSW, and avoids many of the difficulties
of RSW of aluminum [16]. In many applications, FSSW can be readily substituted for RSW with little or no design modifications.
Another variant of FSW is Friction Stir Processing (FSP), where the friction stir tool is simply traversed through the material.
There is no joint in these applications. FSP can be used to significantly enhance the material properties of castings and wrought
material. In castings, porosity can be eliminated. Furthermore, a significant improvement in material properties can be realized.
In other applications, the material can be processed to improve the ductility of the material. This can be taken advantage of in
forming or bending applications of material that typically exhibits lower ductility, such as aluminum.

11.8. MERITS AND DEMERITS OF FSW


Friction stir welding offers a variety of advantages over traditional welding processes. Problems such as porosity, solute

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
redistribution, cracking during solidification and liquation cracking are not an issue during FSW. FSW produces a low
concentration of defects and is very tolerant to variations in parameters and materials. As the joint is characterized with fine
microstructure it results in number of advantages. These advantages include improved joint properties, in particular static
strength and fatigue properties, low deformation of the workpieces, smaller heat affected zone. The advantages from process
point of view are significant reduction in production costs and the freeing of skilled labor for other tasks. The process is also
very robust, not being sensitive to environmental conditions, unlike many traditional welding processes. This is especially true
for welding of lower melting point materials such as aluminum, magnesium, bronze, and copper [17]. It also yields significantly
less distortion than the fusion welding processes, allowing for dramatic cost reductions in many applications. The advantage of
the FSW process is, no consumables, no filler or gas shield is required, it is absolutely a safe process with no toxic fumes or
spatter of molten material during welding. The other advantages include ease of automating the process as simple milling
machines can be used to carry out the process which results in lower setup costs. It can be operated in all positions (horizontal,
vertical, etc.), as there is no weld pool. It exhibits a good weld appearance and minimal size difference between parent material
and weld, i.e., thickness under/over-matching, thus reducing the need for expensive machining after welding. In contrast, fusion
welding techniques are characterized with a cast microstructure that will lead to severe degradation in the mechanical and
physical properties of the joint.

On the other hand, FSW is associated with a few unique defects. Insufficient weld temperatures, due to low rotational speeds or
high traverse speeds may result in inability of weld material to accommodate the extensive deformation during welding. If the
material is too cool then voids or other flaws occur in the stir zone and in extreme cases this may lead to damage and breakage
of tool. This also results in long, tunnel-like defects running along the weld which may occur on the surface or subsurface. Low
temperatures also limit the forging action of the tool and thus reduce the continuity of the bond between the materials on each
side of the weld. The light contact between the materials is termed as 'kissing-bond' in FSW [17]. This defect is quite critical as
it is very difficult to detect using nondestructive methods like X-ray or ultrasonic testing. Other problem is longevity of pin, if the
pin is too short or if the tool goes slightly out of the plate then the interface at the bottom of the weld will not be properly forged
by the tool, resulting in a defect called insufficient-penetration of tool in material. This behaves like a notch in the material which
can be a potential source of fatigue crack and reduces the fatigue resistance of weld. Other drawback is occurrence of exit at
the weld hole during withdrawal of FSW tool.

11.9. APPLICATIONS OF FSW


The original application for friction stir welding (FSW) was the welding of large lengths of material in the aerospace,
shipbuilding, and railway industries. Examples include large fuel tanks, for space launch vehicles, cargo decks for high-speed
ferries, and roofs for railway carriages. In the last several years, the automotive industry has been aggressively studying the
application of FSW in its environment. The drive to build more fuel efficient vehicles has led to the increased use of aluminum
in effort to save on weight, which also improves recyclability when vehicles are scrapped [18]. The process has been
commercialized in many applications, including rail car, automotive, aerospace, heavy truck, medical applications, etc. Today,
the process is being transitioned into fabrication of complex assemblies, yielding significant quality and cost improvements. As
the process is maturing, designers are taking advantage of the process, by designing the product specifically for the FSW
process.

11.10. FORMABILITY OF FRICTION STIR WELDED BLANKS


The formability of friction stir welded blanks is important in the context of tailor welded blanks. As described elaborately in the
earlier chapter, tailor welded blanks are made usually by fusion welding processes like laser welding. Recently FSW, a solid
state welding process, is used for the same. In particular, it can be used to join high-strength aerospace aluminum alloys and
other materials that are difficult to weld by conventional fusion welding processes. Therefore, now-a-days FSW is frequently
used for making tailor welded blanks required for aerospace and automotive industries. But the weld region in the friction stir
welded blanks is relatively large as compared to weld regions produced by other welding processes like laser welding, gas

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
tungsten arc welding, etc. Also it is seen from the existing data (like hardness, strength, etc.) that, though the weld zone
mechanical properties are different as compared to the base materials, they are not very different from the base materials.
Therefore, the combined effect of the change in weld zone properties and dimensions on the formability of FSW blanks as a
whole becomes important. It is also worth to note the changes in formability of FSW blanks with respect to TWBs made by
other welding processes too. In this context, there are few research groups who have studied the formability of FSW blanks
using standard formability tests and modeling methods. The theoretical results are compared and validated with experimental
observations. Some of the important results pertaining to the formability of FSW blanks are discussed in this section.

For example, Miles et al. [19] compared the formability of friction stir welded aluminum sheets with gas tungsten arc welded
aluminum sheets under stretch forming conditions. The three aluminum alloys 6022-T4 (2.03 mm), 5754-O (1.98 mm), and
5182-O (2.03 mm) were used for formability studies. The same-thickness sheets of each alloy were welded together using both
friction stir welding and gas tungsten arc welding. The FSW samples were welded at feed rates of 13 to 60 cm/min and at
speeds of 1220 to 1500 rpm. The tool material is H13 steel and the tool has shoulder diameter of 10.2 mm and tool tip
diameter of 1.78 mm. The gas tungsten arc welds were produced at a feed rate of 114 cm/min. The operating parameters were
50 pct AC at 200 Hz and 195 amps. The joints were compared by doing tensile test, limit dome height (LDH) test and OSU test
and forming limit curves (FLC) were evaluated. The forming limit curves for FSW blanks and base materials are compared for
5182-O and 5754-O [19].

It is observed that the forming limit of base material and FSW blanks is almost same in plane strain conditions for both the 5xxx
FSW blanks, while some reduction is forming limit is witnessed in the bi-axial stretching region. In the case of 6022 FSW blank,
the forming limit is drastically reduced and is less in the case of bi-axial stretching region as compared to plane strain
conditions. The forming limits of the FSW blanks decrease rapidly as the strain paths move toward biaxial tension. This is
consistent with the formability results where the lowest dome heights for FSW 6022-T4 were obtained with a fully clamped
specimen that produced biaxial or near biaxial strain states during failure. For the 5182-O and the 5754-O alloys, there were no
significant differences in any of the average tensile properties of the FSW and GTAW samples. For 6022-T4, the yield strength
and ultimate tensile strength drops significantly in the welded specimens, although the FSW specimens retain more strength
than the GTAW specimens. In addition, the FSW samples had a greater average elongation than the GTAW samples, by about
50%. In this alloy, the fracture location was always in the HAZ for the FSW samples, while it was always in the weld for the
GTAW samples.

For LDH and OSU test the two 5xxx series alloys perform about the same regardless of the welding process used, while the
6022-T4 has a lower LDH for gas tungsten arc welding than for friction stir welding. The FSW specimens failed in the HAZ and
not in the weld for the LDH test, but several of the GTAW samples failed in the weld. For the 6022-T4, the differences between
the friction stir welding and gas tungsten arc welding results are not as great when the OSU test was performed. This is
because the OSU test causes stretching along the weld and not across it, resulting in a plane-strain fracture perpendicular to the
weld.

Rodrigues et al. [20] assessed the formability of TWB joined by two different conditions of welding named 'hot weld' (HW) and
'cold weld' (CW) by deep drawing axi-symmetric cups. The 'hot weld (HW)' is generated by a tool having conical shoulder with
welding speed, rotation speed, and tool tilt angle of 180 mm/min, 1800 rpm, and 2.5° respectively. The 'cold weld (CW)' is
generated by a tool having scrolled shoulder with welding speed, rotation speed and tool tilt angle of 320 mm/min, 1120 rpm,
and 0°, respectively. The deep drawing tests were performed using circular TWB specimens, with 200 mm diameter. No rupture
was observed in the welds for both types of TWBs, which confirms the good plastic deformation behavior of both the even-
matched (HW) and under-matched (CW) welds. Though no rupture was registered, strong wrinkling occurred for the HW blanks.
The punch force–displacement curves for the TWBs were monitored and it is observed that the punch force evolution for the
CW blank is very similar to that of the base material. In the case of even-matched HW blanks, the punch force behavior is very
close to that of the base material, until approximately 20 mm punch displacement. After this, strong wrinkling started at the cup
flange region, which was accompanied by a strong decrease in the punch force relative to the base material. The main cause
for wrinkling is the deformation taking place in the zone between the die and the punch, where the workpiece is unsupported
and tangential compressive stresses arise promoting buckling and folds. The absence of wrinkling in the case of the under-

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
matched CW blanks, which were drawn under the same blank-holder force used for the HW blanks, can be explained by the
plastic deformation of the softer weld material that avoids buckling.

Miles et al. [21] evaluated the forming behavior of the dissimilar welds of the different combination base materials like 5182-
H16, 5754-O, 6022-T4 and 6111-T4 using LDH test and OSU test and finally generated forming limit diagrams experimentally.
The dome testing was effective for discriminating between an acceptable weld and a bad weld. This test causes biaxial strain
in the weld area, resulting in stretching across the weld and subsequent splitting in the weld nugget, if the welding parameters
are not good. The OSU plane-strain testing showed that, when stretching is done only along the weld (without any transverse
stretching), the ductility of the welded dissimilar-alloy pairs are same as that of the base materials. The FLD of 5182/5754
appears to have a relatively flat FLC, exhibiting similar formability both in plane-strain and in biaxial stretching conditions. This
is caused by the strength difference in the two alloys, where more strain occurs in the weaker alloy, as is seen in the shifting of
the weld line during testing. This results in relatively more minor strain at a given major limit strain. The FLDs for 5182/6022
and 5754/6022 have curves that are maximum at strain-path corresponding to plane strain and that slope toward minimum
values as strain ratios approach biaxial tension. This behavior is caused by the localization of strain in the HAZ of the 6022
alloy, especially as biaxial-strain conditions cause stretching across the weld, and is similar to that of monolithic, friction-stir-
welded 6022.

Kim et al. [22] measured FLD experimentally for base materials utilizing the hemispherical dome stretching test. However, FLC
of weld zones were calculated utilizing Hill's bifurcation theory and the M–K theory, based on isotropic hardening rigid-plasticity
with the isotropic yield function. The forming limit criterion based on Hill's bifurcation theory was applied to the drawing side of
FLD (or negative minor strain) and the M–K model was followed on the stretching side of FLD (or positive minor strain). For the
M–K theory, the defect size was obtained such that the M–K theory would provide the same limit strain as Hill's theory in the
plane-strain strain path (with the minor strain zero). In their work, the same materials with the same thickness (similar gauges,
SG) and different thickness (dissimilar gauges, DG) were welded together. The thinner part of all (DG) materials was in the
retreating side of the FSW. The four different automotive sheets namely 6111-T4, 5083-H18, 5083-O and DP590 steel sheets,
each having one or two different thicknesses were used for the analyses [22].

For formability studies, three tests with varied complexity were performed: the simple tension test with three loading directions
namely—type I: along the weld, type II: across the weld and type III: at an angle of 45° to the weld for simple tension stretching,
the hemisphere dome stretching test (or the FLD test) for biaxial stretching and the cylindrical cup deep drawing test over a
cylindrical punch for draw forming. Three different simulation methods were selectively utilized for performance comparison as
—model-A: the perfect welding condition (the weld zone property is ignored), model-B: the measured average mechanical
properties of the weld zone with uniform average weld zone thickness, and model-C: the average weld zone property with
varying weld zone thickness. In simple tensile test, load-engineering strain profiles were obtained up to failure and engineering
strains at the maximum load and failure were measured [22].

For numerical simulations, the reduced four node shell elements S4R with five integration points through thickness were used
for specimens with parallel (Type I) and vertical (Type II) weld lines. For the sample with 45° weld line (Type III), triangular
elements S3R with five integration points through thickness were used within the gauge length and S4R was utilized for the
rest. In order to investigate the formability of FSW sheets in biaxial stretching deformation, hemispherical dome stretching
(HDS) tests were performed with six welded samples: 6111-T4 (SG, DG), 5083-H18 (SG, DG), AA5083-O (SG) and DP-steel (SG).
The hemispherical dome stretching (HDS) test was carried out on a 50 ton double action hydraulic type press. The punch speed
was 1.5 mm/s and blank holding force was applied just enough to completely clamp the blank, which was about 200 kN. The
lubricant WD-40 was applied on the punch only. The punch load profiles during forming were monitored and the limit dome
heights (LDH) at the failure were measured. The failure onset locations and patterns were also observed and compared with
simulation results. The FLDs of different FSW sheets (both SG, DG) predicted using Hollomon and Voce's strain hardening laws
are compared with each other for analysis. The failure pattern and location is also compared for all the FSW blanks in plane-
strain and bi-axial stretching strain paths. The load–stroke curves during HDS tests were also obtained from experiments and
simulations for all the FSW blanks and are compared [22].

The cylindrical cup drawing (CCD) test was carried out on the 50-ton double action hydraulic type press. The punch speed was

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
set at 1.5 mm/s and tool dimensions are: punch diameter = 50 mm; punch corner profile radius = 6 mm; die opening diameter =
53.68 mm; die corner profile radius = 8 mm. The failure for 5083-O (SG) sample during experimentation occurred normal to the
weld region at the bottom of the cup wall. Similarly, the simulated failure was initiated at the weld zone of the cup wall bottom
and propagated to the base along the bottom rim rapidly. In the case of DP590 (SG) sample, the experimental and simulated
failure lines were generated perpendicular to the weld line at the weld zone of the cup wall bottom [22].

Miles et al. [23] compared the forming behavior of friction stir welded and laser welded dual phase 590 steel sheets of both
same and different thickness using transverse tensile test, LDH test and OSU test. The LDH tests were performed to find out
some acceptable welding condition and using these conditions FSW were carried out which were compared with the laser
welded joints. The dissimilar gauge tension tests (1.6/1.0 mm) showed that transverse tensile ductility in the friction stir
welded specimens could reach the same level of performance as the laser welded specimens. When stretching along the weld
was carried out using the plane strain OSU formability test, the friction stir welded blanks had better formability than the laser
welded blanks by 20%. It is observed from the experiments that the FSW sample failed at an average punch stroke of 36.1 mm
compared to 30.1 mm for the laser welded sample [23]. The improved ductility of the FSW blank is due the lower peak
hardness in the weld region as compared to laser welded blank.

Lee et al. [24] studied the forming behavior of tailor welded blanks both experimentally and numerically. When the formability
performance of the weld samples are compared to that of the base sheets, the improved ductility of the weld promoted the
performance for AA5083-H18, AA5083-O, and AZ31 TWB sheets. However, the weld strength was inferior for all welded sheets,
except AA5083-O and hence the TWB samples were vulnerable to the strain localization. Therefore, aligning the weld zone
properly in an optimized angle to avoid strain localization and to take advantage of the improved (or almost sustained) weld
zone ductility would be important in the process design.

Gan et al. [25] analyzed the formability of 5083-H18 and 6111-T4 FSW sheets by tensile tests and one-dimensional analysis of
tensile test in both rolling direction (RD) and transverse direction (TD). The temperature measurements and stress–strain
behavior were monitored during the FSW process. The strain hardening behavior of 6111 base material and the weld zone are
same in both the transverse and longitudinal rolling directions. The hardening behaviors were fit into equations and the
resulting stress–strain behavior is also compared. The equation based stress–strain behavior agrees well with the
experimental curves [25]. A one-dimensional analysis of tensile behavior is also performed to study the formability and the
results are compared with the experiments. Since 1-D analysis is done, the rate effects and 3-D effects are not included and
hence the tensile behavior is valid till necking only. It is seen that the simulated stress–strain behavior and experimental results
correlates well with each other for both the 6111 and 5083 FSW blanks [25].

The deep drawing analysis of FSW blanks made of similar and dissimilar Al sheets was performed by Leitão et al. [26]. The
TWBs made of AA 5182-H111 and AA 6016-T4 were formed to round cups. The tool with a threaded probe of 3 mm in diameter
and 0.9 mm long and the scrolled shoulder of 14 mm diameter is used for welding. The tool was moved at 320 mm/min
travelling speed and 1120 rpm rotational speed. Both the similar (A5182–A5182 and A6016–A6016) and dissimilar (A5182–
A6016) TWBs were made. It is concluded that the formability of the TWBs, made of similar material, is influenced by the
difference in mechanical properties between the weld and the base materials, and also, by the initial size of the blanks. By using
a 200 mm diameter TWB, non-defective cups were formed, while defective cups were formed if 180 mm diameter TWB was
used. In the case of dissimilar TWBs, the presence of small defects at the weld root of the dissimilar welds induced failure of
some of the blanks during the formability tests. This is seen in TWBs with diameter of 200 and 180 mm tested with 8 kN blank-
holder force. However, it was possible to draw the 180 mm diameter blanks under higher blank-holder forces (16, 20, and 32
kN) which proves that the rupture resulted exclusively from the presence of the defects. The dissimilar TWBs have good
formability behavior as compared to similar TWBs and base material [26].

An important issue during the formability prediction of friction-stir welded blanks is the accurate implementation of weld zone
details like the mechanical and geometrical properties of nugget zone and HAZ. It is relatively easy to evaluate the global
mechanical properties of weld zone (including nugget zone and HAZ as a single entity) as compared to the evaluation of NZ
and HAZ properties separately. In this context, Zadpoor et al. [27] analyzed the effects of the implementation of the weld details
on the accuracy of the failure prediction, strain distribution, and springback behavior of FSW TWBs for the limiting dome height

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(LDH) test and the S-rail problem. The effects of implementing the weld detail on the simulation time are also considered. The
M–K model is used to predict the forming limit of FSW blanks. It is observed from the simulation results that neglecting the
mechanical properties of the weld zone can result in 38% over-prediction of the dome height in the LDH test and 27%
overestimation of the terminal punch travel in the S-rail simulation. In the case of LDH test, there is not much difference in the
punch travel obtained from the models without HAZ and with NZ and HAZ separately. Similar pattern is observed in the case of
'stamping' stage of spring back simulation. However, the simulation time of the 'springback' stage increased more than 120%
with implementation of the weld details in the FEM models. It is concluded by them that the weld details implementation is not
significant if springback is not simulated. The implementation of weld zone details is still important while predicting the strain
distribution and onset of failure [27].

Similar analysis was performed by Perumalla et al. [28–30] wherein the predictions from single zone model (without separate
NZ and HAZ) and double zone model (NZ and HAZ are modeled as separate entities) are compared to identify the domain of
weld conditions in which single zone assumption is valid. This is done by comparing the load–punch stroke behavior of FSW
blanks with single zone and double zone models. A criterion based on the difference between the load–punch stroke behavior
of single zone and double zone model is defined beyond which the double zone model is valid. It is found that multiple domains
wherein double zone model is valid are possible due to parametric changes in weld conditions and such changes affect the
local and global deformation influencing the model representation. When the weld conditions are equivalent to reference model
weld conditions, it represents single zone model. If the conditions are very different from the reference model, it breaks down
the single zone model and double zone is found to be valid. Figures 11.6 and 11.7 show the domain of weld conditions for
single zone and double zone model validity for FSW blanks made of similar thickness sheets (1.5:1.5 mm). The failure locations
also affect the weld zone representation based on the critical error percentage difference between reference model (single
zone) and double zone model parameters [28–30].

Figure 11.6 Domains of weld conditions for representing the double zone modeling validity with
maximum load at failure as criterion (longitudinal weld orientation) [29]; with permission from [29]
copyright [2011] Sage

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 11.7 Domains of weld conditions for representing the double zone model validity with
displacement at failure as criterion (longitudinal weld orientation) [29]; with permission from [29]
copyright [2011] Sage

11.11. SUMMARY
Despite considerable interests in the FSW technology in past, the basic understanding of the process is lacking. The
fundamental knowledge of the FSW process and the knowledge of evolution of the structure and properties needs to be
combined to build intelligent process control models with a goal to achieve, defect free, structurally sound and reliable welds.
Some important aspects include: (a) Tailoring weld structure and properties based on fundamental knowledge still remains an
important milestone in FSW requiring reliable, efficient and real time process modeling so that process can be optimized and
controlled. (b) Exact nature of material flow needs to be understood. (c) Tool geometry design. (d) Wear of welding tool. (e)
Micro structural stability. (f) Welding of dissimilar alloys and metals. (g) Joining of harder and/or high melting temperature
alloys and metals such as steel, nickel, titanium, etc. using FSW. (h) Corrosion studies after welding. (i) Work on variants of
FSW like FSP (Friction Stir Processing) and FSSW (Friction Stir Spot Welding). (j) Process Monitoring. (k) Identifying newer
applications where FSW can be applied successfully.

11.12. REFERENCES
1. Thomas, W.M, Nicholas, E.D., Needham, J.C., Murch, M.G., Temple-Smith, P., and Dawes, C.J. (1991). Friction stir butt
welding, International Patent Application No. PCT/GB92/02203.

2. Kumar, K. and Kailas, S.V. (2008). On the role of axial load and the effect of interface position on the tensile strength of a
friction stir welded aluminium alloy, Materials and Design, 29; 791.

3. DebRoy, T. and Bhadeshia, H.K.D.H. (2010). Friction stir welding of dissimilar alloys—a perspective, Science and Technology
of Welding and Joining, 15 (4); 260.

4. Ghosh, M., Kumar, K., Kailas, S.V., and Ray, A.K. (2010). Optimization of friction stir welding parameters for dissimilar
aluminum alloys, Materials and Design, 31; 3033.

5. Nandan, R., DebRoy, T. and Bhadeshia, H.K.D.H. (2008). Recent advances in friction-stir welding-process, weldment
structure and properties, Progress in Materials Science, 53; 980.

6. Murr, L.E., Li, Y., Trillo, E.A., Flores, R.D., and McClure, J.C. (1998). Microstructures in Friction-Stir Welded Metal, J. Mater.
Process. Manuf. Sci., 7; 146.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
7. Somasekharan, A.C. and Murr, L.E. (2005). Microstructural Details of Friction Stir Weld Interfaces, Friction Stir Welding and
Processing III, in K.V. Jata, (Ed.), The Minerals, Metals and Materials Society, Warrendale, PA, p. 261.

8. Mishra, RS. and Ma, ZY. (2005). Friction stir welding and processing, Material Science and Engineering, R 50; 1.

9. Murr, L.E. (2010). A review of FSW research on dissimilar metal and alloy systems,Journal of Materials Engineering and
Performance, Published online: 2 February, DOI:10.1007/s11665-010-9598-0.

10. Flores, R.D., Murr, L.E., Shindo, D.J., and Trillo, E.A. (2001). Friction-Stir Welding of Metals and Alloys: Fundamental Studies
of Solid-State and Intercalated Flow, J. Mater. Process. Technol., 117 (3); Sect. 1A.

11. Kumar, K. and Kailas, S.V. (2010). Positional dependence of material flow in friction stir welding: analysis of joint line
remnant and its relevance to dissimilar metal welding, Science and Technology of Welding and Joining, 15 (4); 305.

12. Kumar, K. and Kailas, S.V. (2008). The role of friction stir welding tool on material flow and weld formation, Materials
Science and Engineering A, 485; 367.

13. Reynolds, (2000). Visualisation of material flow in autogenous friction stir welds,Science and Technology of Welding and
Joining, 5 (2); 120.

14. Kumar, K., Kalyan, C., Kailas, S.V., and Srivatsan, T. S. (2009). An investigation of friction during friction stir welding of
metallic materials, Materials and Manufacturing Processes, 24; 438.

15. Lohwasser, D. and Chen, Z. (2010), Friction stir welding—From basics to applications, Chapter 5, Woodhead Publishing, p.
118.

16. Arbegast, W.J. (2006). Friction stir welding: After a decade of development,Welding Journal, March, 28.

17. DebRoy, T. and Bhadeshia, H.K.D.H. (2010). Friction stir welding of dissimilar alloys—a perspective, Science and Technology
of Welding and Joining, 15(4); 260.

18. Murr, L.E., Li, Y., Trillo, E.A., and McClure, J.C. (2000). Fundamental Issues and Industrial Applications of Friction-Stir
Welding, Mater. Tech. Adv. Performer. Mater, 15 (1); 37.

19. Miles, M. P., Decker, B. J. and Nelson, T. W. (2004). Formability and strength of friction-stir-welded aluminum sheets,
Metallurgical and Materials Transactions A, 35a; 3461.

20. Rodrigues, D.M., Loureiro, A., Leitao, C., Leal, R.M., Chaparro, B.M., and Vilaça, P. (2009). Influence of friction stir welding
parameters on the microstructural and mechanical properties of AA 6016-T4 thin welds, Materials and Design, 30; 1913.

21. Miles, M.P., Melton, D.W. and Nelson, T.W. (2005). Formability of friction-stir-welded dissimilar aluminum alloy sheets,
Metallurgical and Materials Transactions A, 36a; 3335.

22. Kim, D., Lee, W., Kim, J., Chung, K.H., Kim, C., and Okamoto, K. (2010). Wagoner, R.H., Chung, K., Macro-performance
evaluation of friction stir welded automotive tailor-welded blank sheets: Part II—Formability, International Journal of Solids
and Structures, 47; 1063.

23. Miles, M. P., Pew, J., Nelson, T. W., and Li, M (2006). Comparison of formability of friction stir welded and laser welded dual
phase 590 steel sheets, Science and Technology of Welding and Joining, 11 (4); 384 .

24. Lee, W., Chung, K.H., Kim, D., Kim, J., Kim, C., Okamoto, K., Wagoner, R.H., and Chung, K. (2009). Experimental and numerical
study on formability of friction stir welded TWB sheets based on hemispherical dome stretch tests, International Journal of
Plasticity, 25; 1626.

25. Gan, W., Okamoto, K., Hirano, S., Chung, K., Kim, C., and Wagoner, R. H. (2008). Properties of friction-stir welded aluminum
alloys 6111 and 5083, Journal of Engineering Materials and Technology, 130; 031007–1.

26. Leitão, C., Emílio, B., Chaparro, B.M., and Rodrigues, D.M. (2009). Formability of similar and dissimilar friction stir welded AA
5182-H111 and AA 6016-T4 tailored blanks, Materials and Design, 30; 3235.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
27. Zadpoor, A.A., Sinke, J. and Benedictus, R. (2009). Finite element modeling and failure prediction of friction stir welded
blanks, Materials and Design, 30; 1423.

28. Perumalla Janaki Ramulu, Nilesh. B., and Ganesh Narayanan, R. (2010). Weld zone representation during friction stir welded
blanks formability prediction with equal thicknesses, Proceedings of the Third International and Twenty Fourth All India
Manufacturing Technology, Design and Research conference, Andhra University, Visakhapatnam, India, December 13–15,
989.

29. Perumalla Janaki Ramulu and Ganesh Narayanan, R. (2011), Weld zone representation during the formability prediction of
friction stir welded blanks with similar thickness sheets, Journal of Strain Analysis for Engineering Design, 46; . 456.

30. Perumalla Janaki Ramulu and Ganesh Narayanan, R. (2011). Comparing the weld zone representation methods during the
formability prediction of friction stir welded blanks, The 14th International ESAFORM Conference on Material Forming,
Queen's University, Belfast, April 27–29, AIP Conference Proceedings,1353 213.

[1] Department of Mechanical Engineering, Indian Institute of Science, Bangalore 560 012, India
[2] Department of Mechanical Engineering, IIT Guwahati, Guwahati 781039, India

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.

You might also like