Sunteți pe pagina 1din 45

Friction stir welding of aluminium alloys

P. L. Threadgill1, A. J. Leonard2, H. R. Shercliff3 and P. J. Withers*4


The comprehensive body of knowledge that has built up with respect to the friction stir welding
(FSW) of aluminium alloys since the technique was invented in 1991 is reviewed. The basic
principles of FSW are described, including thermal history and metal flow, before discussing how
process parameters affect the weld microstructure and the likelihood of entraining defects. After
introducing the characteristic macroscopic features, the microstructural development and related
distribution of hardness are reviewed in some detail for the two classes of wrought aluminium alloy
(non-heat-treatable and heat-treatable). Finally, the range of mechanical properties that can be
achieved is discussed, including consideration of residual stress, fracture, fatigue and corrosion.
It is demonstrated that FSW of aluminium is becoming an increasingly mature technology with
numerous commercial applications. In spite of this, much remains to be learned about the
process and opportunities for further research and development are identified.
Keywords: Friction stir welding, Aluminium alloys, Microstructure evolution, Plastic flow, Residual stress, Mechanical properties, Thermomechanically
affected zone

Introduction
Historical background and principles
Radically new joining processes do not come along very
often: friction stir welding (FSW) was one such event,
being invented by the TWI in 1991.1,2 Since then
research and development in FSW and associated
technologies has mushroomed, with many companies,
research institutes and universities investing heavily in
the process and international conference series dedicated
to its study. By the end of 2007, TWI had issued 200
licences for use of the process, and 1900 patent
applications had been filed relating to FSW.3 The
number of research papers has also grown exponentially.
In essence, FSW is very simple, although a brief
consideration of the process reveals many subtleties. The
principal features are shown in Fig. 1. A rotating tool is
pressed against the surface of two abutting or overlapping plates. The side of the weld for which the
rotating tool moves in the same direction as the
traversing direction, is commonly known as the advancing side; the other side, where tool rotation opposes the
traversing direction, is known as the retreating side{.
An important feature of the tool is a probe (pin) which
protrudes from the base of the tool (the shoulder), and is
of a length only marginally less than the thickness of the

TWI, Granta Park, Great Abington CB21 6AL, UK


BP International, Compass Point, 7987 Kingston Rd, Staines, Middx
TW18 1DY, UK
3
Department of Engineering, University of Cambridge, Trumpington
Street, Cambridge CB2 1PZ, UK
4
School of Materials, University of Manchester, Grosvenor Street,
Manchester M1 7HS, UK
2

*Corresponding author, email philip.withers@man.ac.uk


{
Other terms have been used in the literature, namely shear side and
flow side, but these are ambiguous and have been discouraged.4

2009 Institute of Materials, Minerals and Mining and ASM International


Published by Maney for the Institute and ASM International
DOI 10.1179/174328009X411136

plate. Frictional heat is generated, principally due to the


high normal pressure and shearing action of the
shoulder. Friction stir welding can be thought of as a
process of constrained extrusion under the action of the
tool. The frictional heating causes a softened zone of
material to form around the probe. This softened
material cannot escape as it is constrained by the tool
shoulder. As the tool is traversed along the joint line,
material is swept around the tool probe between the
retreating side of the tool (where the local motion due to
rotation opposes the forward motion) and the surrounding undeformed material. The extruded material is
deposited to form a solid phase joint behind the tool.
The process is by definition asymmetrical, as most of the
deformed material is extruded past the retreating side of
the tool. The process generates very high strains and
strain rates, both of which are substantially higher than
found in other solid state metalworking processes
(extrusion, rolling, forging, etc.).
Friction stir welding is therefore both a deformation
and a thermal process, even though there is no bulk
fusion. The maximum temperature reached is a matter
of some debate. Thermocouple measurements during
FSW of aluminium alloys suggest that, in general, the
temperature stays below 500uC.57 These values must be
treated with some care, as the position of the thermocouple in the rapidly moving nugget can be difficult to
ascertain. Microstructural evidence seems to corroborate the thermocouple based conclusion that unless
extreme processing parameters are chosen, the maximum temperature usually lies between 425 and
500uC.6,8 It has been suggested that the temperature of
the material in contact with the pin may reach the
solidus temperature,9 although experimental validation
is difficult due to the intense deformation at the
interface. There is evidence of incipient melting for
some aluminium alloys (e.g. 7010) for fast weld

International Materials Reviews

2009

VOL

54

NO

49

Threadgill et al.

Friction stir welding of aluminium alloys

1 Schematic diagram of FSW process

speeds.1012 It can also be argued that the peak


temperature is inherently self-limiting. The workpiece
flow stress will fall rapidly as the solidus is approached,
so that heating of the nugget at the tool/workpiece
interface limits the available heat generation by reducing
the torque.13,14
To date, the predominant focus of FSW has been for
welding aluminium alloys, although the process has been
well developed for both copper alloys1521 and magnesium alloys.2229 Work is under way to develop the
process for materials such as titanium alloys,3034
steels,3543 nickel alloys4446 and even molybdenum.47
The welding process in these materials takes place at
considerably higher temperatures, and although the
feasibility of the process has been demonstrated, further
work is needed to improve the performance and
longevity of tool materials. In addition considerable
work has focused on using FSW to join dissimilar
aluminium alloys.4861 Furthermore the steady push to
lightweight vehicles has largely been responsible for
research in joining aluminium alloys to other metals,

including aluminium to magnesium,6265 aluminium to


metal matrix composites,66 aluminium to steel6769 and
aluminium to copper.7073
Coverage of the present review is confined to the FSW
of aluminium alloys. A summary of the AWS designations for wrought Al alloy groups and AWS basic
temper designations applicable to heat-treatable Al
alloys is contained within Table 1. Since FSW is a solid
state process, it can be used to join all common
aluminium alloys, including the 2xxx, 7xxx and 8xxx
series which are normally challenging or impractical to
weld by fusion processes. A key distinction is between
non-heat-treatable and heat-treatable alloy series. In
work hardened alloys (e.g. 5xxx), the heat from the
friction welding process will allow thermal recovery and
recrystallisation of dislocation substructures, although
this is partly countered in the intensely deformed region
where new dislocation structures are generated. In age
hardened alloys, the weld will normally be heated well
above the dissolution temperature of the initial precipitates, enabling dissolution, reprecipitation and
overaging to occur. Friction stir welded aluminium
alloys can therefore contain microstructures covering
the entire spectrum of normal tempers.
Since its inception, many papers and articles have
been published on FSW of aluminium alloys, many of
them dealing with microstructure and properties.
Recently there have been excellent general reviews of
FSW covering a wide range of materials by Mishra and
Ma,74 which also includes friction stir processing, and by
Nandan et al.,75 which concentrates on the heat
generation, heat transfer and tool/material flow interactions of FSW. A recent ASM speciality handbook also
covers FSW and friction stir processing.76 Nevertheless,
no critical and comprehensive review focusing specifically on aluminium FSW is available in the public
domain. It is therefore considered timely to correct this
omission. The present review draws on a wide selection
of published data to summarise current understanding
of the complex relationship between welding parameters, microstructure and properties for FSW of many
aluminium alloys. Process modelling of FSW has
evolved in parallel with empirical process development,

Table 1 AWS designations for wrought Al alloy groups and basic temper designations applicable to heat-treatable Al
alloys

50

Wrought alloy groups

Basic temper designations

1xxx Unalloyed 99% Al


2xxx Copper principal alloying element: gives substantial
increases in strength, permits precipitation hardening,
reduces corrosion resistance, ductility and weldability
3xxx Manganese: increases strength through solid solution
strengthening and improves work hardening
4xxx Silicon: increases strength and ductility, in combination
with magnesium produces precipitation hardening
5xxx Magnesium: increases strength through solid solution
strengthening and improves work hardening ability
6xxx Magnesiumsilicon
7xxx Zincmagnesium: substantially increases strength,
enables precipitation hardening, can cause stress corrosion
8xxx Other elements Li, for example, substantially increases
strength and Youngs modulus, provides precipitation
hardening, decreases density

F As fabricated
O Annealed: there may be a suffix to indicate the specific
heat treatment.
H Strain hardened (cold worked): it is always followed by
two or more digits to signify the amount of cold work
and any heat treatments that have been carried out
W Solution heat treated: applied to alloys that precipitation
harden at room temperature (natural aging) after a solution
heat treatment. The designation is followed by a time
indicating the natural aging period, e.g. W 1 h
T Thermally aged:
T1: cooled and naturally aged
T2: cooled, cold worked and naturally aged
T3: solution heat treated, cold worked and naturally aged
T4: solution heat treated and naturally aged
T5: cooled and artificially aged
T6: solution heat treated and artificially aged
T7: solution heat treated and overaged or stabilised
T8: solution heat treated, cold worked and artificially aged
T9: solution heat treated, artificially aged and cold worked

International Materials Reviews

2009

VOL

54

NO

Threadgill et al.

2 a commercial scale FSW machine designed to weld underground train bodies.


clamps (one is obscured) between which welding head passes from right to left;
of three FSWs shown; b typical welding head as it is removed from the plate; c
Triute MX tool it is possible to use this tool with zero tilt angle; d a 50mm
tool designed specically for welding thicker sections

and provides physical insight into all of these relationships. Since FSW modelling has been reviewed elsewhere,77,78 this aspect is not explicitly covered in the
present review, except where modelling helps to interpret
and complement the experimental observations, or to
clarify issues debated in the literature.

Advantages/disadvantages of FSW for


aluminium joining
The advantages of FSW for welding aluminium can be
summarised as follows:
(i) as a solid state process it can be applied to all the
major aluminium alloys and avoids problems of
hot cracking, porosity, element loss, etc. common to aluminium fusion welding processes
(ii) as a mechanised process (Fig. 2a), FSW does not
rely on specialised welding skills; indeed manual
intervention is seldom required
(iii) no shielding gas or filler wire is required for
aluminium alloys
(iv) the process is remarkably tolerant to poor
quality edge preparation: gaps of up to 20% of
plate thickness can be tolerated, although this
leads inevitably to a reduction in local section
thickness since no filler is added
(v) the absence of fusion removes much of the
thermal contraction associated with solidification and cooling, leading to significant reductions in distortion; however, it is not a zero
distortion technique79
(vi) it is very flexible, being applied to joining in one,
two and three dimensions, being applicable to
butt, lap and spot weld geometries; welding can
be conducted in any position
(vii) excellent mechanical properties, competing
strongly with welds made by other processes
(see the section on Comparison with other
joining processes)
(viii) workplace friendly: there are no ulraviolet or
electromagnetic radiation hazards as the absence

Friction stir welding of aluminium alloys

Workpiece is held by two hydraulic


head has just completed furthermost
25 mm shoulder (8 mm pin) diameter
shoulder diameter (50 mm pin) Triat

of an arc removes these hazards from the


process; the process is no noisier than a milling
machine of similar power, and generates virtually
zero spatter, fume and other pollutants
(ix) the energy required at the weld for FSW lies
between laser welding (which requires less
energy) and metal inert gas (MIG) welding
(which typically needs more)*
(x) high welding speeds and joint completion rates:
in single pass welds in thinner materials (down to
0?5 mm thickness), FSW competes on reasonable
terms with fusion processes in terms of welding
speed; in thicker materials, FSW can be accomplished in a single pass (e.g. the 50 mm tool in
Fig. 2d), whereas other processes need multiple
passes. This leads to higher joint completion
rates for FSW, even though the welding speeds
may be lower. Thick plates can also be joined by
FSW on either side43,80
(xi) various mechanical and thermal tensioning
strategies can be applied during welding to
engineer the state of residual stress in the weld
(see the section on Residual stress control).
There are of course disadvantages to FSW; indeed, some
of the advantages listed above can be viewed in a less
positive light in certain circumstances. For example, the
absence of a filler wire means that the process cannot
easily be used for making fillet welds. Similarly, the fully
mechanised nature of the process prevents its use for
applications where access or complex weld shape is best
suited to a manual process. The presence of a hole at the
end of the weld from which the probe was withdrawn is
often quoted as a disadvantage. In practice, this has

*Note that a distinction is needed between energy required to make the


weld and the total energy required to operate the process. The latter
depends very much on the specific equipment used, and comparisons are
difficult. However, it would be expected that the total energy requirement
for FSW would be greater than MIG, but less than laser for single pass
welds of the same thickness.

International Materials Reviews

2009

VOL

54

NO

51

Threadgill et al.

Friction stir welding of aluminium alloys

seldom been a significant problem, as there are many


possible solutions, which have been considered elsewhere.81,82 The workpiece also needs to be restrained in
well designed support tooling (Fig. 2a), both to react to
the forces applied, and to prevent the probe from
pushing the workpiece materials apart. Although the
process may reduce the strength of aluminium alloys,
this can be compensated for if necessary by appropriate
design of the joint, for example by locally increasing the
thickness, but in most cases no changes are made.
Process economics are generally considered favourable,
but specific published data are lacking. However, it is
known that the process drastically reduces weld preparation costs, skilled welder requirements and repair
rates. Efficient power consumption is dependent on
matching the size of machine being used to the size of
weld being made, although this is not always a practical
option.

Applications
Commercial applications have been reported across
many industries, and some selected examples are shown
below which illustrate the widening appeal of the
process. This list is representative rather than exhaustive, and it should be emphasised that new applications
are appearing all the time. It should be noted that FSW
does not restrict the operating temperature range of
aluminium alloys, with applications ranging from
cryogenic temperatures (e.g. liquid oxygen and liquid
hydrogen rocket fuel tanks) to mildly elevated temperatures (e.g. heat exchangers in heating systems). Most
FSWs used in production are butt welds, although lap
welds and friction stir spot welds are also being applied
with increasing frequency.
Marine

It is believed that the first commercial application of


FSW was the joining of 6xxx series alloy extrusions for
use in fish freezing plants for fishing vessels.83 There
have been numerous applications of the process for
joining 6xxx extrusions for incorporation in bulkheads
and decks in various high speed aluminium vessels, and
in large steel cruise ships which now often have
lightweight aluminium superstructures.84 In such applications, the FSW panels are very flat due to the low
distortion, and are cut up and welded into larger
structures, usually by MIG welding. Friction stir
welding has been used extensively in the aluminium
superstructures of cruise ships such as the Seven Seas
Navigator which contain many kilometres of friction
stir welds, mostly in 6xxx grade extrusions. The worlds
largest aluminium vessel, the Japanese fast ferry
Ogasawara, launched in 2004, makes extensive use of
FSW in its superstructure.85
Aerospace

The first major application was the use of the process to


replace fusion welding in fuel tanks for unmanned Delta
II and later Delta IV rockets. 8688 The manufacturer
(Boeing) has reported virtually zero defect incidence,
and significant cost savings over the previous variable
polarity plasma arc (VPPA) process. The process has
also been adopted for the large fuel tank for the Space
Shuttle.8991
Almost all the major airframe manufacturers are
investigating the use of FSW (alongside other welding

52

International Materials Reviews

2009

VOL

54

NO

processes such as laser welding) to replace many of the


rivets in current structures. The first aircraft to make
extensive use of FSW in its airframe, the Eclipse 500
business jet, has recently completed certification and is
now in production.92 In this aircraft, over 7300 fasteners
(approximately 60% of the total) are replaced by 263
friction stir welds.
Rail

High speed aluminium railcars such as the Japanese


Shinkansen are normally built from complex double
skin extrusions in 6xxx alloys.93 Since the welds which
join these are long (up to 25 m) and straight, FSW is an
ideal process, and the very low distortion is cited as a
major benefit94,95 (see also Fig. 2a).
Automotive

There are few long straight welds in road vehicles, and


so adoption of FSW has primarily been for components
such as suspension parts, wheels, seat components, crash
boxes, etc. where several leading companies are already
using the process in production. The needs of the
automotive sector have driven the development of
robotic FSW, to cope with the complex shapes and
high volume/low cost culture of this market. Significant
interest is now being shown in friction stir spot welding,
where the linear translation of the tool is either very
small or zero. Friction stir spot welding is rapidly
gaining acceptance as an efficient method of joining
aluminium sheet, and is already in production, for
example on the Mazda Rx-8 sports car, where it is used
on the aluminium bonnet and rear doors.96 Friction stir
welding is also being developed for lightweight
armoured vehicles, where the ability of the process to
weld material of around 2540 mm thickness in one pass
is being exploited.97,98
Process variants

In the past five years, two variants have emerged as


significant technologies in their own right, namely
friction stir processing74,99102 and friction stir spot
welding.103105 Mishra et al.,101 Mahoney et al.106 and
Charit et al.107 have demonstrated the capability of the
former to produce fine scale microstructures in superplastic 7xxx alloys, greater than 5 mm thick. Similar
results have also been demonstrated in an AlLiCu
alloy108 and in commercial purity aluminium alloy
1050.109 The technique has also been demonstrated as
effective for homogenising powder metallurgy processed
alloys.102,110 The ability to process aluminium alloys in
this manner raises a range of possible applications for
friction stir processing. These include tailoring microstructures for subsequent deep drawing and superplastic
forming operations, and the ability to refine locally the
microstructure of castings (for example, around stress
concentrations, where a superior wrought microstructure would be preferable). Friction stir processing is not
considered further in the present review.

Friction stir welding process


Flow mechanisms and tool design
The metal flow and heat generation in the softened
material around the tool are fundamental to the friction
stir process. Material deformation generates and redistributes heat, producing the temperature field in the
weld. But since the material flow stress is temperature

Threadgill et al.

Friction stir welding of aluminium alloys

3 a typical generic ow path of plate material round clockwise rotating pin in FSW, taken from two-dimensional CFD
model with cylindrical tool moving from left to right (after Seidel and Reynolds),111 b,c effect of interfacial boundary
conditions (b stick; c slip) on predicted ow from two-dimensional CFD model with proled tool (after Colegrove and
Shercliff):112 change in thickness of streamlines indicates nal location of points initially forming straight line transverse to weld line (analagous to Cu foil in d) and d metallographic marker experiment using transverse copper foil,
illustrating ow induced by pin (after Reynolds)113

and strain rate sensitive, the distribution of heat is itself


governed by the deformation and temperature fields. In
fact their control lies at the core of almost all aspects of
FSW, for example, the optimisation of process speeds
and machine loading, the avoidance of macroscopic
defects, the evolution of the microstructure, and the
resulting weld properties.
As noted above, almost all the material in the weld is
extruded between the rotating pin on the retreating side
and the surrounding material which is too cold and too
lightly stressed to deform (see Fig. 1). In its simplest
form, this essential flow mechanism can be illustrated by
two-dimensional simulations depicting streamlines
round a rotating tool placed in a steady flow of material.
Figure 3a shows streamlines past a cylindrical tool,
predicted by computational fluid dynamics (CFD).111 A
longitudinal weld seam is formed behind the advancing
edge of the tool where the two flows come together.
Further modelling studies have investigated how this
two-dimensional flow is perturbed by:
(i) the addition of tool features such as flats and
flutes
(ii) changes in the contact conditions between tool
and workpiece, from sticking friction to slipping
at a lower interfacial shear stress.
Predicted streamlines round a fluted tool are shown in
Fig. 3b and c.112 Complete sticking generates a dead
metal zone round the tool, whereas the flow interacts
closely with the tool features when slipping takes place.
Another characteristic of the process a line initially
perpendicular to the welding direction is swept into a
backwards bulge in the wake of the tool can also be
seen in Fig. 3b and c. Marker experiments113 have confirmed this behaviour (Fig. 3d). One way of quantifying

the mixing effect of the tool is the ratio of the swept


volume to the pin volume.114 For 25 mm thick plates
this has found to be 1?1 : 1 for a cylindrical pin, 1?8 : 1 for
the Whorl and 2?6 : 1 for the MX-Triflute pin (see
Fig. 2c), each having similar root diameters and
lengths114 with the Triflute giving the more parallel
sided weld zone. Further refinements include the Trivex
tool115 which was designed to reduce the down and
traverse forces required and the Triflat tool for thicker
section materials (Fig. 2d).
Experience has shown that it is advantageous to
develop a vertical component to the flow of material,
and most tools therefore contain threads, helical flutes,
or similar features to force material adjacent to the pin
to flow away from the shoulder. Further variants have
emerged as the process matures, such as tools in which
the pin and shoulder rotate independently (including
non-rotating shoulders),33 retractable pin tooling,94 as
well as a bobbin tool with a shoulder on both ends of a
pin of length equal to the plate thickness.116 Tool design
has been reviewed in detail by Fuller117 and Dubourg
and Dacheux.118
The capture of three-dimensional flow greatly complicates the modelling challenge.78,119 This is not simply
because of the geometric complexity, but also because of
the inherent sensitivity in the flow response to the
interfacial conditions and the temperature and strain
rate sensitivity of the material flow stress. A further
feature of the flow that is not captured by current
models is the formation of a stable void immediately
behind the tool (see the section on Formation of voids).
This has been identified by stop-action experiments, in
which the traverse is abruptly arrested, with tool
extraction being synchronised with rotation to preserve

International Materials Reviews

2009

VOL

54

NO

53

Threadgill et al.

Friction stir welding of aluminium alloys

4 Longitudinal vertical section in travel direction after


synchronised pin retraction, showing trailing void on
left hand side (after Colligan)120

the material in contact with tool features. An early


example of this technique is shown in Fig. 4.120 The
threads on the leading edge were full and left intact,
while on the trailing edge the upper threads are
incomplete. The conditions that lead to formation of
this void, and whether it has any relevance in producing
a sound void free joint, remain open research questions.

Heat generation and process operating regimes


Friction stir welding differs from competing processes
such as arc and laser welding, since these use an external
heat source of specified power, whereas in FSW the
joining process itself generates the heat. The heat input
is therefore a complex function of the process variables
(traverse and rotation speeds, and down force), the alloy
being welded, and the tool design. The effect of tool
rotation and traverse speed on the heat input per second
and per mm are shown in Fig. 5a and b respectively.53
Analytical estimates of heat input have assumed sliding
Coulomb friction at the tool/workpiece interface with a
constant coefficient of friction, or sticking friction using
an estimate of the limiting shear yield stress, or have
inferred contact conditions and/or heat input from
measurements of machine torque.77,78,121127 Thermocouple measurements coupled to heat flow analysis also
provide a means to infer net power inputs. However,
Peel et al.53 found no simple correlation between the
temperature and the input power or heat. Although the
heat input is commonly considered in fusion welding, it
is a poor indicator of the temperature of the material
surrounding an FSW tool, at least for the joining of thin
plates of aluminium. It is likely that when the traverse
speed is reduced, much of the additional heat is
conducted into the backing plate, as evidenced by the
observed correlation between the heat input and the
backing plate temperature53 as well as through the tool.
The down force in FSW provides intimate thermal
contact between workpiece and backing plate, but this
contact evolves with position during the welding
process, requiring complex calibration.78,127129 The
growing recognition of the importance and complexity
of heat input has lead to the routine instrumentation of
welding equipment, with spindle rotation speed and
torque measurements now providing detailed power
data.

54

International Materials Reviews

2009

VOL

54

NO

5 Rate of heat input a per millimetre of weld line and b


per second for like to like and AA5083/AA6082 welds53

Heat is produced primarily by viscous dissipation in


the workpiece material close to the tool, driven by high
shear stresses at the tool/workpiece interface. The
temperature and normal contact stresses vary widely
over the tool, so it is unlikely that a single contact
condition will be valid. Material at the interface may
stick or it may slip, or there may be a combination of the
two. As discussed above, local melting may occur as
peak temperatures reach the solidus temperature. There
may be then be oscillating stick slip behaviour, as local
melting rapidly reduces the shear stress, leading to a
steep drop in local heat input and temperature, and selfstabilising behaviour. Process modelling using CFD has
been used to explore the sensitivities of the heat
generation, tool forces and size of deformation zone as
a function of tool design and process conditions.78
Recently the heat input has been calculated directly from
the hot deformation constitutive response of the alloy,
using a fully coupled deformation and heat flow
model.14,130 This approach is currently limited by the
quality of the constitutive data, particularly near the
solidus, but has the potential to reveal the deformation
regime that corresponds to the production of sound
welds with least machine load for any particular alloy.
For example, Fig. 6 shows the predicted variation of
power generation, peak temperature and traverse force
with rotation speed. Note that the power and peak
temperature saturate as the interface temperature and
heat generation are limited by rapid material softening,
while the minimum force required is predicted to occur
at an intermediate rotation speed. This result correlates
with very early Russian work on rotary friction welding
by Vill,131 which shows that the time to complete a
rotary friction weld goes through a minimum as the
rotation speed is increased. The effect has been
confirmed experimentally by TWI.132
For a given alloy and plate thickness, with a particular
tool, the operators remaining process variables are
down force, tool tilt angle, tool plunge, rotation speed

Threadgill et al.

Friction stir welding of aluminium alloys

8 Microstructural zone classication in a friction stir weld


in Al 2024 alloy (A: parent material, unaffected by process; B: HAZ, thermally affected but with no visible
plastic deformation; C: TMAZ, affected by heat and
plastic deformation)

6 Schematic illustration of variations, predicted by CFD,


of heat generation, peak temperatures and traversing
force with rotation speed for FSW of 2024 aluminium
alloy (after Shercliff and Colegrove)78

and traverse speed. Down force is only a preset variable


when in force control, and plunge depth is only a preset
in position control. Empirical trials invariably explore a
matrix of these variables, thereby defining the process
window that produces sound welds without tool breakage or macroscopic defects. Figure 7 shows the relationship between the welding parameters and the FSW
process window for an aluminium die casting alloy.133
With increasing tool down force the process window is
enlarged mainly towards lower rotation speeds and
higher welding (traverse) speeds. This pattern is broadly
typical of aluminium alloys generally. The process
operating window is commonly described as being
limited by hot welds and cold welds the former
associated with high rotation and low traverse speeds,
leading to excessive flash production, the latter with low
rotation and high traverse speeds, leading to tool
breakages. The nature of the defects associated with
unsound welds is discussed further below.

7 Range of optimum FSW conditions for various tool


plunge down forces for 4 mm thick ADC12 AlSi casting alloy welded using 15 mm shoulder, 5 mm diameter, 3?9 mm long threaded pin:133 ash style aws
were associated with excessive heat input, whereas
tunnel style voids (Fig. 11) were associated with insufcient heat input and abnormal stirring

Macroscopic weld features


Definition of macroscopic weld zones
The first attempt at classifying FSW microstructures was
made by Threadgill.134 This work was focused solely on
aluminium alloys, and was limited to features distinguishable by light microscopy. However, work on other
metallic materials has demonstrated that the behaviour
of aluminium alloys is not typical of most metals and
alloys, and this initial classification was inadequate.
Consequently, a revised set of terms was suggested135
and then subsequently revised4 and adopted in the
American Welding Society Standard D17?3M.136 These
microstructural terms are illustrated in Fig. 8, and are
defined below along with alternative terms commonly
found in the literature:
(i) unaffected material or parent metal: material
remote from the weld, which has not deformed
and which, although it may have experienced a
thermal cycle from the weld, is not affected by
heat in terms of detectable changes in microstructure or properties
(ii) heat affected zone (HAZ): the region close enough
to the weld for the weld thermal cycle to have
modified the microstructure and/or properties,
but no apparent plastic deformation is detected by
light microscopy although it is recognised that
some plastic deformation will have occurred, as is
typically the case in any weld HAZ (some
researchers have preferred the term thermally
affected zone; however, by analogy with other
welding processes, there is little justification for a
distinct terminology for friction stir welds where
the term HAZ is well understood)
(iii) thermomechanically affected zone (TMAZ): in
this region, the material has been plastically
deformed by the FSW tool, and heat from the
processing has also affected the material. In the
case of aluminium, it is possible to generate
considerable plastic strain without recrystallisation in this region, and there is generally a distinct
boundary, at least at a macroscopic level, between
the recrystallised and deformed zones of the
TMAZ. (The recrystallised region is often called
the nugget, which is a descriptive term, though
not very scientific. Terms such as dynamically
recrystallised region have been suggested136 and
used extensively in the literature, but the term
nugget remains widely used and understood.)
When presenting micrographs it is also conventional to
show the advancing side of welds on the right and this

International Materials Reviews

2009

VOL

54

NO

55

Threadgill et al.

Friction stir welding of aluminium alloys

a AA6082 on retreating side (left side lighter); b 6082 on advancing side (right side lighter)
9 Macrographs showing stir zone/TMAZ of 6082/5083 dissimilar welds with a 6082 on retreating side (left side lighter); b
6082 on advancing side (right side lighter) for welds made at traverse speed of 200 mm min21 and various rotation
speeds (dotted lines on 560 rev min21 welds show approximate size and position of 6 mm dia. pin)53

convention is used here except where specific mention is


made in the figure caption.

Mixing across dissimilar welds


Friction stir welding has been used with notable success
to join dissimilar aluminium alloys in a number of
configurations (for references, see the introductory
section). It would be reasonable to assume that a
process with so much shear strain would result in very
effective mixing of the alloys, but experience has shown
this is seldom the case. The handedness of the weld is
an important factor, i.e. which alloy is placed on the
advancing side in a dissimilar combination.53 Both the
heat generation and appearance of the weld crosssection change significantly when the weld handedness is
reversed, for the same rotation and traverse speed.
Figure 9 shows macrographs of the stir zone in the
dissimilar 6082/5083 welds produced using a range of
rotation speeds, with AA6082 on the retreating and
advancing side respectively. Changing the traverse and
rotation speeds can have a significant effect on the flow
of material within the stir zone. Generally speaking, the
extent of mixing and interface disruption increases as the
rotation speed is increased, or the traverse rate is
decreased.53 Unsurprisingly the rotation speed was
found to have a significantly greater impact than the
traverse speed. For a given combination of weld
parameters, the welds produced with AA6082 on the
advancing side (Fig. 9b) exhibited a significantly lower
level of mixing in the stir zone than those with the
materials reversed. This is in contrast to the results of
Larsson et al.,49 but corroborates those obtained by
Tanaka and Kunagi.50 Further examples are shown in
Fig. 10, which shows examples of AlMg and AlSi
castings welded to 5xxx and 6xxx series plates respectively. The alternating bands of material originating
from the casting and the plate material are clearly seen.
It has been suggested that the distance in the welding
direction between dissimilar bands corresponds with the
pitch distance (the travel speed divided by the tool
rotations per second).137

56

International Materials Reviews

2009

VOL

54

NO

Flaws in friction stir welds


Flaws arise in most materials joining processes. For
example, when arc welding aluminium alloys, weld
metal porosity138 and, depending on the particular alloy,
weld metal solidification cracking and HAZ liquation
cracking139 are among the most common flaw types. The
occurrence of such problems has contributed to the
widely held view that some aluminium alloys, in

a AlMgFe casting (CC601T6) to 5484T34 alloy; b


LM6 AlSiMg casting to 60682T651 alloy
10 Macrosections of friction stir welds

Threadgill et al.

Friction stir welding of aluminium alloys

a volumetric flaw in 2014;140 b tunnel (wormhole) defect at base of Trivex tool when welding 7449 at 120 rev min21/
60 mm min21;130 c surface defect located under shoulder in 2014A (Ref. 140)
11 Characteristic void aws in friction stir welds:

particular some of the high strength 2xxx and 7xxx series


alloys, are difficult, or indeed impossible, to fusion weld
successfully. Being a solid state joining process, FSW
obviates the problems of porosity and hot cracking. In
this respect it is worthwhile to make a distinction
between flaws and defects, although the two terms are
often used interchangeably within the literature. The
usual distinction is that a flaw or imperfection is a
feature that one would prefer not to be in the weld, but it
may or may not compromise the integrity of the weld. If,
after evaluation, the flaw is deemed unacceptable, then it
becomes a defect. If it does not compromise the
integrity, then it is a tolerable flaw. Flaws or discontinuities should be characterised as defects only when
specific acceptance criteria, related to the engineering
application, are exceeded, and the presence of the flaw
compromises the integrity of the structure. Table 2
summarises the characteristic flaw types in butt and lap
welds in friction stir welds and their principal causes. In
fact the most common flaw types are caused by use of

under optimised parameters or a lack of process control.


Since understanding of the causes of these flaws/defects
is good, it is usually possible to rectify these problems by
changes to parameters, tool designs or operating
practice.
Formation of voids

Figure 11a shows a typical void on the advancing side


of a weld. Similar features are sometimes observed near
the base of the pin (Fig. 11b). The formation of a
continuous tunnel void in this location was a common
observation in early FSW, but was eliminated by the use
of tool tilt or the redesign of pin features. Figure 11c
shows a near surface defect left by the trailing edge of
the shoulder.140 Since voids are not in general surface
breaking, they cannot usually be detected by visual or
surface inspection but are relatively easy to pick up by
NDE (see below). A number of factors have been
identified as contributing to void formation, including
inadequate welding pressure,140 high travel speed and

Table 2 Summary of characteristic aw types encountered in friction stir welds


Flaw type

Location

Cause

Void (Fig. 11a)

Advancing side at edge of nugget

Void (Fig. 11c)


Joint line remnant (butt weld)
(Fig. 12a)

Beneath top surface of weld


Weld nugget, extending from the root of the weld
at the point where the original plates butted together

Root flaw (Fig. 14)

Weld nugget, extending from the root of the weld


at the point where the original plates butted together

Joint line remnant (lap weld)

Plate interface

Hooking (lap joint) (Fig. 18)

Advancing side of weld, in unbonded TMAZ region,


normally extending upwards
Retreating side of weld, in unbonded TMAZ region.

Low forging pressure


Welding speed too high
Plates not clamped close enough
together. Joint gap too wide
Welding speed too high
Inadequate removal of oxide from
plate edges
Inadequate disruption and dispersal
of oxide by tool
Tool pin too short
Incorrect tool plunge depth.
Poor joint to tool alignment.
Inadequate removal of oxide from
plate edges
Inadequate disruption and dispersal
of oxide by tool
Ineffective tool design

Plate thinning (lap weld)


(Fig. 18)

Ineffective tool design.

International Materials Reviews

2009

VOL

54

NO

57

Threadgill et al.

Friction stir welding of aluminium alloys

slow tool rotation speed141,142 as well as inadequate


control of the joint gap.140,143 If the welding pressure is
inadequate, the weld will receive insufficient forging
action from the tool shoulder to achieve full consolidation.140,144 When welding at high travel speeds, and also
slow tool rotation speeds, the material receives less work
per unit of weld length, i.e. fewer tool rotations per millimetre. Under such conditions, the plasticised material
may not reach a sufficiently high temperature.142,145 One
view is that aluminium alloys can withstand only a
certain shear strain rate, which is dependent on temperature. When subjected to lower shear strain rates,
they will flow and recover, but at higher strain rates they
cannot flow/recover fast enough to keep flowing, and
they will break up, forming voids, etc. Colegrove and
Shercliff130 suggested that the void defect shown in
Fig. 11b formed near the base of a Trivex pin was due to
inadequate forging and consolidation due to the low
temperatures there. Kumar et al.146 found that a
rounded tool was less likely to produce void defects.
The presence of voids on the advancing side of the
weld (Fig. 11a) has been investigated by mathematical
modelling of metal flow during the welding process.9,119,130 These models predict that transitional
volumes of material will form, between regions of
rotational motion of material and regions in which
material is extruded past the rotating tool pin. Bendzsak
et al.9 described the motion in the transition region as
chaotic. The models predict flow singularities (stagnation) in this region of the advancing side of the weld,
which have been considered to be the source of tunnel
defects (Fig. 11b).130 These defects form under inappropriate welding parameters, such as high welding
speed or low pressures. Further increasing the welding
speed or decreasing the welding pressure would cause
extension of the tunnel defect up to the weld surface,
ultimately forming a groove type defect.147 It is noticeable that the tunnel defect is often observed near the
bottom of the weld at the advancing side. From these
observations, FSW could be considered as a process
where a cavity formed behind the welding tool is later
filled by plasticised material flowing from the front of
the pin to its rear (see also Fig. 4). If the cavity is not
filled, then a cavity will remain in the weld. This
explanation is supported by the study of Zhao et al.148
who found that under identical welding conditions
tapered and cylindrical pins with threads did not show
wormhole defects while those without did, because the
thread helps to transport material around the tool back
to the advancing side, leaving a void there. Crawford
et al.149 related the lack of adequate flow and wormhole
formation to low weld pitch, i.e. insufficient rotational
speed relative to the weld travel speed.
Prediction of void formation is a particularly difficult
modelling problem, due in part to the limitations of the
numerical methods used for flow modelling. Computational fluid dynamics solvers treat the deforming
metal as a hot, viscous fluid, neglecting elasticity. As
noted above, CFD gives a good representation of flow
patterns and the internal generation of heat, but cannot
readily describe a free metal surface the deforming
metal fills all the space between the tool and the backing
plate. Void formation necessarily implies a free surface,
and is also strongly influenced by the hydrostatic pressure and the mechanism of cavitation. Computational

58

International Materials Reviews

2009

VOL

54

NO

fluid dynamics models indicate that a region of


hydrostatic tension forms on the tool wake on the
advancing side, but the neglect of elastic stresses means
that the predicted pressures are unreliable quantitatively, and there is in any case no criterion available for
the formation of a stable void in hot deforming metals.
Finite element methods are generally better suited to
elasticplastic analysis and therefore to predicting
residual stress (see below), but not to material flow
because the strains are very high, due to the computational demands of continuous, fine scale remeshing.150
Some success has been achieved in modelling FSW using
the arbitrary LagrangianEulerian formulation.151,152
However, computational run times mean that only very
short weld traverses can be simulated, and the models
have great difficulty in capturing the complex weld
closure behind the tool, due to the sensitivity to the local
stress state of separation between workpiece and the
tool.
Joint line features

Joint line remnants are features that extend from the


weld root, at the location of the original parting line
between the butting plates to be welded, through the
weld (Fig. 12a). Joint line remnants distributed through
the bulk of the weld have been reported by several
workers in many alloys.140,153157 In some instances,
such features can be little more than a string of
inadequately dispersed oxide particles, which originated
from the surfaces of the butting plates. A recent study by
Sato et al.158 using transmission electron microscopy has
confirmed that these strings are indeed Al2O3 particles.
These are a consequence of the presence of the natural
aluminium oxide on the surfaces of the butting
plates.154,155 The original joint line is clearly discernable
in Fig. 12a, while at higher magnification, the clusters of
oxide along the prior interface are visible (Fig. 12b).
Such features are affected by the welding speed; increasing the welding speed reduces the disruption of the oxide
per unit advance of the tool. Tool shoulder size and tool
geometry140,154,155 may also have an effect. In many
cases such bands of oxide can be innocuous provided
good mixing occurs.158,159 Reynolds113 suggested that
the curving line arises from remnants of the oxide layer,
because the final position of the initial butt surface after
material flow of FSW exhibits a curving line on the
cross-section. This has been verified using marker
material experiments in which a thin copper layer is
trapped in the joint line, and its final position identified
using X-ray tomography (Fig. 3d).160 Figure 13 shows a
tomograph of a stop action weld, illustrating the
breakup and redistribution of the joint line. In crosssection (lower image), it can be seen that the copper is
predominantly distributed in a characteristic curve
bordering the weld nugget on the retreating side.
From the above discussion, it is important that the
oxide interface between the butting plates is adequately
disrupted in order to form a bond, and in the vast
majority of cases this is exactly what happens (Fig. 12c).
Indeed the original joint line is normally increased in
length some 35 times, and the oxide is broken up and
scattered, so that most of the original joint line is a metal
to metal bond, and therefore strong. In some cases, the
interface is only partially disrupted and remains
identifiable in a cross-section. It is unusual for such
joint line remnants to be a problem.

Threadgill et al.

Friction stir welding of aluminium alloys

12 a oxide defect in 5083 alloy which can in severe cases lead to kissing bond at the base,161 b magnied view showing increased oxide inclusion level in 2014 along prior joint line140 and c schematic of joint line remnant161

Naturally, the correct depth of penetration of the tool


pin is essential to ensure that mixing occurs over the full
plate thickness. Where a shortened pin is used, where the

tool plunge depth is incorrectly set, or where there is


poor alignment of the tool relative to the joint line, a
root flaw can be produced, indicative of a complete lack
of bonding (Fig. 14). In such cases the bond quality
improves towards the nugget. As a result there can be a
transition region where a bond exists, but is weak. This
is generally known as a kissing bond, a rather
unscientific name which has become widely adopted. It
is not possible to identify a kissing bond from a
microstructural examination. Zhou et al.161 have found
that kissing bonds can show 2040 times shorter fatigue
lives for Al 5083 welds and 3080 times shorter lives for
Al 2024 welds. The fatigue fracture surfaces observed by
SEM were consistent with a kissing bond in that the
fractured specimens failed from the root tip of the oxide
array. An example of failure from such a kissing bond
under tensile loading is depicted in Fig. 15. It should be
emphasised that kissing bonds are rare and can be
eliminated by appropriate tool/joining parameters.
Other flaw types

Flash style features can occur on the top surface:


excessive flash is not necessarily a bad thing; in some

13 X-ray tomographic and corresponding metallographic


interpretation
of
FSW
ow
mechanism
(after
Dickerson et al.):160 three-dimensional X-ray tomography image (top) showing break-up of Cu foil placed
on joint line; bottom: virtual cross-section revealed by
tomography

14 Characteristic root aw in FSW 2014A caused by


using too short a pin in 5083 alloy

International Materials Reviews

2009

VOL

54

NO

59

Threadgill et al.

Friction stir welding of aluminium alloys

15 Cross-section of 5454 FSW specimen with crack after


tensile test along kissing bond156

products high flash levels are deliberately introduced to


ensure adequate penetration, especially in welds where
the fit-up is poor. Flash is normally caused primarily by
large plunge depths.
An unusual chevron surface feature has been
observed on the root surface of welds in 5xxx series
alloys (Fig. 16). Pryzdatek162 noted that the features
were coincident with small root line imperfections in
welds in alloy 5083 and concluded that they were the
result of the expulsion of material from the joint during
welding lending evidence to the possibility that some
liquation may occur during welding. Leonard163
observed similar features in both 5083 and 5251 alloys.
To date the features have not been reported to affect
either the mechanical or corrosion behaviour of the
joints and it is possible that the features will remain a
curiosity rather than a point of concern. Pryzdatek162
has noted, in any case, that light dressing will remove
them.
Local melting during FSW has been a topic of
continual debate. However, conclusive evidence of local
melting in aluminium alloys has been rarely reported.
Figure 17 shows a region of the TMAZ beneath the tool
shoulder of a weld in alloy 7050T7451, produced at
TWI,164 providing clear evidence that liquation can
indeed occur. While this demonstrates that liquation is
possible, the paucity of reported cases of liquation
cracking in the literature suggests that the occurrence is
less widespread than in aluminium arc welds. The

16 Chevron markings on underside of friction stir weld in


5083 alloy (scale in mm)

60

International Materials Reviews

2009

VOL

54

NO

17 Evidence of partial melting on retreating side of nugget in FSW of 6 mm thick AA7050T7451 alloy164

apparent absence of extensive liquation during welding


may be explained as a consequence of the extensive
mechanical deformation experienced during welding.
Any liquated area is likely to be difficult to identify in a
metallographic section of the completed weld. It is also
worthy of note that liquation has been found in the
TMAZ region of friction stir welds in magnesium alloy
ZK60 by Johnson.165 The formation of grain boundary
films was observed, but these could be prevented by
using a lower heat input welding procedure.
Sato et al.64 have described liquation when welding
aluminium alloy 1050 to a magnesium alloy. In this case,
the maximum temperature reached can exceed the
melting point of the Al12Mg17 intermetallic, with clear
evidence of a cast structure being visible. However, this
is a special case resulting from the welding of two very
different materials which form a low melting point
eutectic.
Lap weld defects

Friction stir lap welds may contain some of the flaw


types encountered in butt welds, in particular voids and
oxide joint line remnants, the latter often appearing in a
horizontal orientation through the joint. Owing to the
joint geometry, root flaws are not encountered.
However, lap welds do exhibit features at the edges of
the bonded region which affect joint properties. Any
friction stir lap weld, like a riveted lap joint, contains
notches at the faying surfaces of the sheets. Cederqvist
and Reynolds166 have shown that such features in
friction stir welds can have a marked effect on joint
performance. The notch feature can deviate towards
either the top or bottom surface of the weld, depending
on welding conditions, reducing the effective sheet
thickness. This feature is sometimes described as
hooking, involving a significant rotation of the
unwelded interface, sometimes by 90uC. It is found on
the advancing side of the weld. It is also not unusual to
find upper plate thinning on the retreating side of the
joint, and this can significantly compromise the load
bearing capacity of the joint. Typical examples are
shown in Fig. 18. Thomas et al.167 have shown that
welding tool design can have a strong influence on the
occurrence of such features and consequently joint
properties, in particular fatigue. CFD modelling has
also been used to show how the tool profile modifies the
extent of hooking in lap welds, enabling a degree of
prior computer based optimisation of tool profiles.168

Threadgill et al.

Friction stir welding of aluminium alloys

18 Lap weld defects showing hooking on advancing side and plate thinning on retreating side in lap welds between
7075 (upper) and 2024 (lower) alloys

Non-destructive monitoring of defects

Currently, only limited standards exist for evaluating the


quality of friction stir welds136,162 and further work is
required to assess the significance of flaws and their
means of detection. It should be noted however that
all welds, whatever process is used, are likely to
contain some imperfections, but most do not seriously
compromise performance. Indeed the overwhelming
majority of friction stir welds do no undergo nondestructive inspection, partly because of cost and partly
because the incidence of significant flaws is small, and
the risk of a catastrophe minute, as most structures have
considerable redundancy. As a result, hundreds of
kilometres of friction stir welds have been safely
employed commercially in most cases, with no more
than visual inspection. In areas where a flaw is
absolutely unacceptable (e.g. rocket fuel tanks) very
detailed NDE is used, although defects are extremely
rare. In other cases, a statistical process control method
is used. If the QA system is tight, and the parameters fall
within certain narrow ranges, then the risk of a defect in
a fully mechanised process is extremely small.
Traditional NDE techniques do not work well with
FSW because the techniques were developed for fusion

welds, where different flaws/defects are found, e.g.


porosity, solidification cracks, hydrogen cracks, reheat
cracks, lack of fusion, slag entrapment etc., none of
which is found in FSWs. The NDE industry is responding by developing/adapting techniques to cope with
this challenge. Using techniques such as phased array
ultrasonics and meandering winding magnetometers,
voids can now be readily detected. A good example of
the capability of the ultrasonic method as applied to
FSWs is given in Fig. 19.
Current NDE techniques are not totally reliable for
detecting root flaws. The only definitive method is a
destructive bend test with the root in tension.136 Efforts
are being applied to both advanced ultrasonic inspection
techniques (for example phased array techniques), and
sophisticated eddy current techniques (meandering
winding magnetometers, pulsed eddy currents).169176
Kissing bonds (see the section on Joint line features),
however, are particularly insidious, as they are difficult
to detect using non-destructive techniques which rely on
an interruption in the microstructure. While the oxide
stringers are not directly sensed by ultrasonic attenuation, the associated reduction in grain size can be
detected.176 By comparing the mean noise level inside

19 a aws in cross-section from 6061 FSW tailor welded blank specimen (cylindrical threaded pin, welding speed
1200 mm min21, spindle speed 1500 rev min21, shoulder penetration 0 mm) and b corresponding synthetic aperture
focusing technique ultrasonic image169

International Materials Reviews

2009

VOL

54

NO

61

Threadgill et al.

Friction stir welding of aluminium alloys

a nugget; b nugget/TMAZ boundary; c HAZ region177 (scale bar corresponds to 50 mm in each case)
20 Microstructure of 2199 alloy FSW displayed using inverse pole gure map obtained by EBSD, showing renement of
microstructure in nugget

the root to that of the weld nugget, the operator can


estimate the pin depth and therefore the likelihood of
having a kissing bond.
In summary, the FSW process is now mature. Various
tools have been designed to meet differing geometrical
demands. In most cases relatively wide processing
windows have been established for the production of
large quantities of defect free welds using automated
commercial welding systems.

Microstructural features and hardness


profiles
Introduction
In any welding process, the properties and performance
of the weld are dictated by the microstructure, which in
turn is determined by the thermal cycle of the welding
process, which can normally be varied by changing the
welding parameters. Therefore welding parameters must
be selected that give the best possible microstructure and
that allow welds to be made free from defects and other
undesirable features. With most materials, it is well
understood that welding has some adverse effects on
microstructure and properties, and thus the optimised
weld parameters are often a compromise between
making sound welds at economical production rates
and producing acceptable, rather than ideal, microstructures and properties.
Friction stir welds in aluminium alloys contain a wide
variety of microstructures, which is hardly surprising
when the extreme range of strains, strain rates and
thermal cycles to which different regions of the weld are
exposed is considered. The microstructural variations
were first characterised by Threadgill134 (see Fig. 8). In
the HAZ, remote from the centre of the weld, there is no
obvious change to the grain structure (Fig. 20c), and the
HAZ is detected only by a change in hardness and
generally by a change in etching response. In precipitation hardened alloys it is widely accepted that some
coarsening of precipitates is occurring, and possible
dissolution at higher temperatures. In work hardened
alloys, dislocation networks may recover, and this may
cause some low angle cell boundaries to form.
Furthermore as the weld centre is approached, clear
evidence of plastic deformation can be seen in the grain
structure. In the outer part of the TMAZ, the original
grains remain identifiable in the deformed structure,
with the formation of subgrain structures and significant

62

International Materials Reviews

2009

VOL

54

NO

associated rotation of the parent grains as evidenced by


the pole figure in Fig. 20b. Closer still to the weld line,
the strains, temperatures and time at elevated temperature all increase, allowing the formation of the
recrystallised nugget with a fine equiaxed structure
(Fig. 20a). The microstructural characteristics will first
be discussed for the nugget region, in which deformation
dominates. Evolution of microstructure in the heat
affected zone is thermally controlled, and this will be
discussed separately for non-heat-treatable and heattreatable alloys. The microstructure of welds made
between dissimilar alloys is also discussed briefly.

Deformation microstructure in weld nugget


Onion ring structure

A common observation from the nugget region in FSW


is the appearance of a series of circular or elliptical
features in etched metallographic sections (see, for
example, Fig. 8), often termed onion rings (as the
sections reveal a slice through a set of nested layers of
roughly hemispherical shape, like an onion). The
significance of this structure in the weld nugget remains
an occasional topic of interest in the literature. Mahoney
et al.6 and Leonard178 have shown for alloys 7075 and
2014A that the ring patterns are an etching response to
variations in grain size between the rings. Other
characteristics of the rings include texture effects179,180
and variations in dislocation density.181 The nugget may
also contain fractured constituent particles178,182 and the
structure has been attributed to a variation in their
distribution.178,183 This is turn may be a consequence of
the banded distribution of the constituent particles
present in the base metal, a characteristic that is strongly
alloy dependent.184 These factors primarily relate to the
strength of contrast in microstructure observed in the
weld nugget, but do not offer a complete explanation of
the mechanism of formation, which has not yet been
formulated. There seems to be strong argument that
there is a purely kinematic basis for the formation of
each ring, associated with one rotation of the tool (or the
rotation between positions of tool symmetry, i.e. three
per revolution for a Triflute tool). Cyclic fluctuations in
the amount of material extruded past the tool and being
deposited are to be expected with profiled tools.185 It has
therefore been postulated that ring formation may be a
function of the tool geometry, tool rotation and forward
travel speeds.134 Computational fluid dynamics modelling112 and marker experiments120,186 have also made a
modest contribution to the discussion to date. The

Threadgill et al.

practical significance of the phenomenon remains rather


limited as the mechanical properties of the nugget are
generally good, and the fracture paths in mechanical
tests are seldom associated with the onion rings.
Recovery versus recrystallisation

A feature of the microstructure of friction stir welds in


aluminium alloys is the development of a fine grain
structure in the centre of the nugget region, as shown in
Fig. 20a. Typically, equiaxed grain sizes of the order of a
few micrometres have been measured in the nugget
region. The precise nature of this fine grained region has
been the subject of much research and debate in the
literature. TEM studies by several workers, examining a
range of different alloys, have shown that the fine grain
structure in this region in the light microscope comprises
fine grains possessing predominantly high angle grain
boundaries and a low dislocation density.6,187,188 On the
basis of these observations, it has been concluded that
the nugget consists of dynamically recrystallised grains,
and not subgrains.6,8,187,189 A similar observation has
also been made in a dissimilar weld between copper and
alloy 6061,71 in which fine recrystallised grains rich in
both aluminium and copper were observed. One study
of alloy 6061T651, has contrasted the grain structure of
the nugget region with that of the remainder of the
TMAZ.189 Tilting studies showed that the nugget
comprised dynamically recrystallised grains, whereas
the remainder of the TMAZ comprised deformed
subgrains, separated by low angle grain boundaries.
Rhodes et al.190 proposed that the final equiaxed nugget
grains in 7050 are formed by grain growth from much
finer grains nucleated by the dynamic recrystallisation
process, thus accounting for the low dislocation density.
It is also likely that, before recrystallisation, extensive
recovery occurred, as there will be significant plastic flow
in the material about to be welded. However, other
studies191 of 7050T7451 have shown a high dislocation
density in grains in the nugget, and a study by Sato
et al.192 of 6063T5 showed that whilst most grains
exhibited a low dislocation density, some grains
exhibited a much higher density. These variations in
dislocation density may be associated with the welding
process conditions, in particular the forward tool
movement per revolution, but this hypothesis has not
been tested. High values of forward tool motion per
revolution produce harder microstructures, but generally similar grain size. The influence of grain size and
dislocation density on strength is difficult to isolate in
the heat-treatable alloys, due to the simultaneous effects
of changes in precipitation hardening, requiring detailed
TEM studies. Furthermore the presence of precipitates
has a direct influence on the processes of recovery and
recrystallisation. Hassan et al.193 attempted to simulate
the formation of the nugget in AA7010 alloy using high
strain rate torsion tests. This study confirmed the effects
of a high precipitate density in increasing the resistance
to recrystallisation, which only occurred at strains of
.20. The process of recrystallisation was also aided by
heterogeneous plastic flow.
In a study carried out in the region of the tool pin exit
hole in a sample of alloy 7475, it has been argued that
the structure of the weld nugget is one of dynamically
recovered subgrains.194 This was argued on the basis of
rapid and massive grain growth observed during
annealing experiments at temperature above 500uC.

Friction stir welding of aluminium alloys

This postulation that grain growth is due to the presence


of a dynamically recovered subgrain structure in the
TMAZ has been supported by a study of friction stir
welds in alloys 6082T6 and 7108T79, using scanning
electron microscopy and electron backscattered diffraction.195 Using this technique, which can quantify grain
orientations on a polished surface, low angle boundaries
were reported, indicative of a subgrain structure. Fonda
et al.196 have also presented evidence from stop action
tests which indicates that recovery based mechanisms
could be of importance in grain refinement during FSW,
although they do not dismiss the possibility of recrystallisation. However, Mishra et al.197 have reported
abnormal grain growth during heat treatment of friction
stir processed alloys 7050 and 2519 in the temperature
range about 450470uC. The authors attributed the
phenomenon not to the presence of a subgrain structure,
but to a number of possible factors. These included dissolution and growth of precipitates, regions of localised
strain differences, regions of non-uniform grain size distribution and the existence of boundaries with different
mobility. However, it was recognised that the exact
origin of the phenomenon was not clear. Abnormal
grain growth is discussed in more detail below.
It has been suggested that the differences in microstructural observations may be resolved by considering a
mechanism of continuous dynamic recrystallisation in
the TMAZ.182,191,198 The deformation process associated with welding introduces a large quantity of
dislocations, while at the same time grain growth occurs
as the temperature rises. Subgrains, which are very small
and exhibit low angle boundaries, begin to form by a
process of dynamic recovery. Continuous dynamic
recrystallisation then occurs as dislocations are continuously introduced to the subgrains by further
deformation. The subgrains grow and rotate as they
accommodate more dislocations into their boundaries,
forming equiaxed recrystallised grains with high angle
grain boundaries. Plastic deformation continues with the
repeated introduction of dislocations and the process
continues until the end of the thermomechanical cycle,
at which point partial recovery takes place. The precise
mechanism remains unresolved, as recent experiments in
which the welding tool was retracted rapidly and the
material quenched have shown that very fine recrystallised grains, of the order of 25100 nm, from Ref. 190.
These are smaller than the 25 mm grains observed in the
weld nugget under normal welding conditions, suggesting that these arise from a nucleation and growth
mechanism. On the basis of such experiments, Prangnell
and Heason199 suggest that there is no evidence of
continuous recrystallisation by grain rotation, but rather
bands of fine nugget scale grains are first formed, from
closely spaced parallel high angle grain boundaries that
develop from finer scale deformation bands. A mixed
microstructure develops comprising a matrix of nugget
scale grains containing high aspect ratio fibrous grains
having more stable orientations. Finally, even these
fibrous grain fragments become unstable when they thin
to subgrain dimensions and break up to form a full
nugget-like microstructure comprised of low aspect ratio
ultrafine grains. This mechanism would explain the
bands of similarly oriented fine grains in Fig. 20b.
Another point which has been debated is whether all
of the recrystallisation in the nugget occurs during the

International Materials Reviews

2009

VOL

54

NO

63

Threadgill et al.

Friction stir welding of aluminium alloys

21 Microstructure in 6 mm thick 2195 alloy a before and b after post-weld heat treatment202

deformation process (i.e. dynamic recrystallisation) or


whether the process continues after deformation has
ceased. The answer is probably more of academic
interest than practical significance, but it is reasonable
to assume that static recrystallisation (and subsequent
grain growth) is more likely to occur in thick section
welds, where time at elevated temperature will be longer.
Despite some evidence for subgrain formation, the
consensus is that recovery is more important in the
highly deformed areas of the TMAZ outside the nugget,
where the original grain structure is retained. The fine
equiaxed microstructures in the nugget are the result of
recrystallisation processes, presumed to be predominantly dynamic.
Effect of post-weld heat treatment on grain structure

As noted above, a feature observed in many alloys is


that of massive grain growth in the nugget area during
post-weld heat treatment. This not only serves as a guide
to the interpretation of the nugget microstructure that
leads to this behaviour, but may also have practical consequences for welds subjected to post-weld heat treatment. The phenomenon has been reported in 1xxx,200
2xxx,201203 6xxx,121,204,205 and 7xxx201,206,207 alloys.
However, the mechanism may differ between alloys, as
the 1xxx alloys are not precipitation hardened.
Sato et al.200 have observed that massive grain growth
in 1100H24 alloy occurs only when the post-weld heat
treatment temperature exceeds the maximum temperature experienced during welding, and may be associated
with the formation of small grains with high angle
boundaries as a result of primary recrystallisation.
Typical examples are shown in Fig. 21, from work by
Litwinski202 on 6?4 mm 2195T8A3 alloy. Solution
treatment at 510uC (950uF) after welding resulted in
massive grain growth in the lower half of the nugget, and
also some less spectacular growth just below the upper
surface. The problem was solved by modifying the
welding procedure to produce a higher welding temperature which reduced the amount of cold work.
Attallah and Salem Hassan203 showed that the risk of
rapid grain growth in 2095 was reduced by high rotation
speeds and lower travel speeds, both of which will
increase the heat input. These authors observed rapid
grain growth both at the upper surface, and in the lower
part of the weld.
Work by Hassan et al.207 on 7010T7651 alloy
showed that post-weld heat treatment could lead to
massive grain growth of generally similar appearance to
that shown by Litwinski,202 i.e. concentrated in the
lower portion of the weld. Hassan et al. have suggested
that very fine grain sizes in the as welded nugget are

64

International Materials Reviews

2009

VOL

54

NO

more at risk, as there is insufficient dispersoid available


to stabilise the grain boundaries. Grain growth may also
be encouraged by the dissolution of these dispersoids
during heating. The use of a high heat input welding
cycle will give coarser as-welded grains, which should be
more stable, although conditions can exist where rapid
grain growth can still occur, and it is suggested that this
is associated with the formation of planar fronts on the
growing grain. It is also suggested that a mean grain size
of at least 10 mm would be required to provide grain
stability during solution heat treatment. Both Hassan
et al.207 and Attallah and Salem Hassan203 have related
the risk of rapid grain growth to theories of cellular
microstructures proposed by Humphreys.208210
Although the precise mechanism may not be fully
understood, the above studies indicate strongly that the
risk of rapid grain growth during post-weld heat
treatment can be reduced by using high heat input
welding procedures. These will help to remove cold work
and increase the as-welded grain size, and hence the
efficiency of a finite number of grain boundary pinning
particles. Presumably minimising post-weld heat treatment times and temperatures will also be helpful.
Grain size control

Certain alloying elements can be added to aluminium


alloys to restrict grain growth during high temperature
operations, notably scandium and zirconium. There are
claims that adding a small quantity of scandium to
aluminium alloys will improve fusion weldability (resistance to hot cracking, etc.), with the presence of the
thermally stable Al3Sc precipitate limiting grain growth.
Likewise, Al3Zr precipitates have also been identified as
beneficial. There are few published data on FSW of
scandium bearing alloys but there is some variance
among observations of the effect of Sc on nugget grain
size. Gittos and Bridges211 (studying two AlZnCuSc
alloys, similar to 7010 and 7050), Huneau et al.212
(studying an AlMgSc alloy), and Paglia et al.213
(studying cast 7050 containing Sc) found only marginal
effects on nugget grain size in the as-welded condition.
The Sc addition was ,0?12% (Refs. 211 and 213) or
0?26% (Ref. 212) and Zr was present to a similar level in
the work of Gittos and Bridges211 and Huneau et al.212
However, unpublished work has shown a beneficial
effect of scandium on grain size in the weld. Sato et al.214
found that in a binary AlZr alloy, the presence of Zr
had no significant effect on the nugget microstructures,
implying that recrystallisation occurs above the Al3Zr
solution temperature. However, the presence of the Zr in
solid solution limits dislocation movement and hence
recovery mechanisms in the non-recrystallised TMAZ,

Threadgill et al.

thus limiting grain growth. Charit and Mishra215 have


shown very fine grain structures in a friction stir
processed AlZnMgSc casting, quoting a mean grain
diameter of 0?68 mm. This value is significantly less than
found in friction stir welds in other alloys, and supports
the suggestion of Rhodes et al.190 that all welds initially
produce such fine grain sizes. This demonstrates that in
the absence of the stabilising effect of the Al3Sc or Al3Zr
precipitate there will be some grain growth. This view is
supported by later work from Hsu et al.216 who
introduced fine dispersions of Al3Ti into a pure
aluminium matrix, achieving mean grain sizes of
between 0?30 and 1?53 mm after friction stir processing.
In summary, it seems that the inclusion of grain growth
inhibiting elements can be beneficial, but further work is
needed to fully understand the mechanisms by which
they operate, and to achieve the full potential of these
alloy additions.
Texture

A number of studies have examined the development of


texture in friction stir welds.153,196,217219 Studies by
Field et al.218 indicated that local textures were largely
alloy independent, which is supported by the work of
others. There are two regimes of texture, one in the area
where the process is dominated by the shoulder, and
another in the area further down where the shoulder
plays no part, as evidenced by Ahmed et al.80 for thick
section 6082 welds. The texture is very simple in the pin
dominated region, as might be expected. Work to date
has reported textures in the weld nugget that are
consistent with a shear deformation process. Jin
et al.153 studying welds in 5xxx series alloys, have
predicted the deformation characteristics of the weld
nugget, concluding that they appeared to be more
isotropic than that of the parent plate. This finding has
implications for the deformation behaviour of friction
stir welds, and for potential applications in, for example,
tailor welded blanks. However, the significance of
texture development in relation to mechanical properties
and sheet forming operations has yet to be rigorously
tested.

Microstructure and hardness evolution in nonheat-treatable alloys


Limited microstructural studies have been performed on
non-heat-treatable alloys, covering alloys 1100,204,220
5083,180,221 5754, 5251,222,223 and 5182.153 Welds were
made in both the annealed (O) and various cold worked
conditions. Macroscopically, welds in these alloys
appear similar to welds in heat-treatable alloys, exhibiting a TMAZ and recrystallised nugget.153,220 Hardness
traverses in work hardened non-heat-treatable alloys
(e.g. 5xxx alloys in the H1xx, H2xx or H3xx conditions)
normally resemble that shown in Fig. 22. As the weld is
approached, the heat from the process causes annealing
and recovery to take place, leading to a drop in hardness. The minimum hardness is typically in the nugget,
where the fine grained, fully recrystallised structure
(discussed above) is formed. However, welds made in
annealed material (e.g. 5xxx in the O condition) do not
exhibit an HAZ. Hardness traces show little or no
variation in hardness between the parent metal and weld
(Fig. 22). Sometimes the weld nugget may be slightly
harder than the O condition, due to modest (hot) work
hardening and grain refinement. Since the nugget

Friction stir welding of aluminium alloys

22 Hardness traverses across friction stir welds in


5083O and 5083H321 alloys showing effect of heat
treatment

formation eliminates the prior deformation microstructure in cold worked material, the hardness of the nugget
region is independent of the original condition, as seen
for alloy 5083 in Fig. 22.
Since few finely dispersed second phase particles exist
to pin grain boundaries, the effects of recovery and
recrystallisation are, unsurprisingly, different from those
in precipitation hardened alloys. Thus the extremely
elongated and deformed grains found in the nonrecrystallised TMAZ in precipitation hardened alloys
are not generally seen in precipitate free alloys. The
transition from non-recrystallised to recrystallised is far
less distinct, implying that recrystallisation is much
easier in the absence of precipitates. Experiments by
Genevois et al.222 have shown that at 350uC, heavily
strained 5251 will recrystallise in 15 s, whereas identically treated 2024 required 1800 s to recrystallise.
Microstructural modelling in non-heat-treatable alloy
FSW has been limited to predicting the loss of hardness
across the weld in initially cold worked tempers.52 The
minimum hardness corresponds to complete recrystallisation, while the base metal hardness corresponds to no
recrystallisation. The problem is therefore to predict the
positions between which the volume fraction recrystallised varies from 0 to 100%. The resulting hardness is
then estimated using a linear rule of mixtures of the
limiting hardness values. The extent of recrystallisation
is primarily determined by the peak temperature reached
in the weld thermal cycle, together with the duration of
the time at temperature. A common approach to
modelling microstructural change in a thermal cycle is
to replace the cycle with an isothermal hold of duration
teq at the peak temperature Tp of the cycle. The duration
of the hold is defined to be that which provides the same
kinetic strength, I, as the thermal cycle, defined as

I~ dt=t T ~teq =t T


The function t*(T) captures the temperature dependence
of the transformation time for isothermal holds (e.g.
obtained via saltbath experiments). In practice, it is
common to use hardness data from annealing experiments directly, fitting the data semi-empirically as a
function of isothermal temperature and hold time. At
each position in the weld, Tp is evaluated from a thermal
model, the kinetic strength and teq found by integrating
the equation above, and the values substituted into the
hardness function.

International Materials Reviews

2009

VOL

54

NO

65

Threadgill et al.

Friction stir welding of aluminium alloys

24 Hardness proles showing natural aging in 6 mm


7075T6 alloy178

23 Schematic plots showing generic hardness responses


across friction stir welds

Microstructure and hardness evolution in heattreatable alloys


Heat-treatable aluminium alloys derive much of their
strength from the presence of fine precipitates, formed
during prior heat treatment (age hardening). The thermal
cycles experienced during welding can lead to precipitate
coarsening or dissolution, and further precipitation during
or after cooling, depending on the peak temperature and
duration of the cycle. Given the technological importance of
these alloys, FSW has been studied in a wide range of 2xxx,
6xxx and 7xxx series alloys, in naturally aged and artificially
aged tempers. The microstructural development of heattreatable alloys, and the corresponding evolution of hardness (Fig. 23), during FSW have been studied extensively.6,8,48,51,121,142,168,178,179,182,187189,191,192,204,221,222,224232
Hardness profiles in heat-treatable alloys

Transverse hardness profiles are a common starting


point for interpreting some of the changes that occur
during welding. Repeat measurement after a period of
natural aging is also useful, indicating that supersaturated solute remains immediately after welding and
the final profile may be enhanced. Indeed, since postweld natural aging starts immediately and may continue
for many months, reported results are subject to some
uncertainty as the interval after welding is frequently not
controlled or documented.
Interpretation of hardness data merits a degree of
caution for other practical reasons. For example, lower
loads may be chosen to enable greater spatial resolution
of rapidly varying profiles, but this introduces more
scatter. Samples mounted in a hot press may have been
further aged by the mounting process. Most published
profiles are for the midthickness of a plate, but it is
apparent from Fig. 8 that variations through thickness
will be expected. Only occasional studies map the whole
cross-section, due to the effort involved.
Most friction stir welds in heat-treatable alloys,
welded in the peak aged or overaged conditions (T6/
T7 tempers), exhibit a characteristic hardness profile, as
typified in Fig. 23. The significant effect of post-weld
natural aging for a 7075T6 weld is illustrated in
Fig. 24.178 In the HAZ, softening is observed, with a
rapid drop in hardness as the TMAZ is approached. The

66

International Materials Reviews

2009

VOL

54

NO

greatest recovery in strength is observed in the nugget.


Both coarsening and dissolution lead to a drop in
hardness, but strength recovery only occurs following
dissolution. The hardness profiles are therefore consistent with precipitate coarsening being dominant in the
HAZ (lower peak temperatures) and dissolution in the
nugget (peak temperatures above the solvus of the initial
precipitates), followed by natural aging. The nugget
strength may also be augmented by its deformation
substructure, as discussed above. For naturally aged
tempers (T3/T4) there is the additional possibility that
coarsening causes a strength increase in the HAZ, in the
region where the peak temperature is comparable with
conventional aging temperatures (150200uC), and the
initial GP zones can evolve to a more hardening state. It
is also important to recognise that, in all heat-treatable
alloys, precipitate evolution is not limited to the bulk of
the grains. For example, grain boundaries may exhibit
widening of precipitate free zones.182,191 This will clearly
not be evident in hardness profiles, but may be
significant in relation to ductility, toughness, fatigue or
corrosion.
Faster welding speeds generally lead to colder welds,
though there is a more complex relationship between
thermal history and speed than in fusion welds, due to
the dependence of the heat input itself on the welding
speed. Higher speed welds tend to have a narrower
HAZ,233,234 and the nugget hardness is also often
increased due to the reduced deformation temperature
and increased strain rate. Figure 25 shows hardness data
for 6 mm 7075T7351 alloy for plates welded at
different times during a five year period in which tools
and practices evolved significantly: from a traditional
threaded pin tool (1995), a traditional tool with
accelerated cooling (1997), to an advanced (Triflute)
tool allowing a substantial increase in welding speed
(2005). A traditional threaded pin tool is now recognised
as being of low efficiency. Then the application of a high
conductivity toolbed, which extracted heat far more
rapidly than a conventional steel bedplate, was applied.
More recently, this has been allied to the use of a high
efficiency tool (in this case a Triflute tool, see Fig. 2). It
is clear that progressive reductions in heat have
increased the hardness of the nugget, increased the
minimum hardness measured at the HAZ/TMAZ
boundary, and reduced the overall width of the HAZ.
Since strength recovery is commonly observed in the
nugget due to natural aging, it might be expected that

Threadgill et al.

25 Progressive improvements in hardness of 7xxx alloy


with onset of more efcient welding methods: in 1995
a traditional threaded pin tool was used; in 1997 same
parameters were applied on a high conductivity tool
bed; in 2000 a Triute tool was used

post-weld heat treatment would be employed to promote


artificial aging. However this is not widely practised on
aluminium fabrications, for which a complete resolution heat treatment is not often feasible, or
economic. Low temperature heat treatment is practised
in some companies to stabilise the microstructure in
2xxx and 7xxx welds. In T3/T4 tempers there is little
benefit in elevated temperature post-weld heat treatment. Natural aging potentially restores the weld to the
initial strength, if complete dissolution took place in the
weld nugget, however the HAZ region will remain softer
even after very long periods.
Precipitate evolution in heat-treatable alloys

The detail of precipitation behaviour in friction stir


welds is of course more complex than the simple outline
above, as revealed by transmission electron microscopy.
Early studies6,187 on welds in alloys 7075 and 2014
identified that the weld nugget contained overaged
precipitates. However, these studies and others178,191
demonstrate that the hardness of the weld nugget

Friction stir welding of aluminium alloys

commonly increases during subsequent aging, indicating


that the nugget contains sufficient solute capable of
sustaining subsequent metastable precipitation and age
hardening. Strangwood et al.187 have postulated, based
upon TEM observations of precipitate sizes and volume
fractions, that both aging and re-solution of precipitates
occur in the nugget during welding of alloys 7075 and
2014. This is further supported by temperature measurements made during welding, which demonstrate that the
theoretical dissolution temperatures for hardening precipitates in both 7075 and 2014 were exceeded. Su
et al.182 have argued that the temperatures achieved
during the welding thermal cycle for alloy 7050T651
were sufficiently high to take all of the strengthening
precipitates in the nugget into solution. The cooling rate
was not sufficiently rapid to prevent reprecipitation, but
this was favoured by heterogeneous nucleation at
dislocations and constituent particles, resulting in only
large precipitates forming. It has been argued that the
observed hardness peak in the nugget is due to a number
of factors: a decrease in grain size, solid solution
strengthening and comminution of constituent particles.178,187 With regard to comminution, it has been
postulated that some dissolution of constituent particles
may occur as a result of the thermomechanical processing achieved in the stirred region.178
Precipitation studies in the HAZ of alloys 7075 and
2014 have shown full precipitation of the equilibrium
phases,6,187,191 in common with arc welding.235
Strangwood et al.187 and Svensson and Karlsson188
have shown that precipitation is favoured at matrix/
constituent particle interfaces. Full strength may be
recovered only by a full solution anneal;187 a post-weld
aging heat treatment is not applicable.6,187 It is worth
noting that some aging of the HAZ has been observed in
alloys 2014AT651 and 7075T651,178 suggesting that
some reversion may take place, but this was within the
first few hours after welding, after which no appreciable
aging was observed.
Figure 26 shows 7050 welds made in 6?35 mm 7050
T7451 alloy. This temper is overaged, designed to give a

a 7050 parent material: well characterised intragranular precipitates are g9 [Mg(Zn,Cu,Al)2] and intergranular precipitates
are g (MgZn2) and/or Mg3Zn3Al2; b area in HAZ, where same precipitates are found but thermal cycle has led to 56
increase in size; c area from recrystallised nugget, where some grains had significant dislocation densities: dislocations
are pinned by Al3Zr dispersoids or Al7Cu2Fe inclusions
26 Comparison of microstructures in 7050 FSW191 (TEM bright eld images: welds made at 396 rev min21,
102 mm min21)

International Materials Reviews

2009

VOL

54

NO

67

Threadgill et al.

Friction stir welding of aluminium alloys

27 Small angle X-ray scattering maps of volume fraction


and size of precipitates in friction stir welds of T3 and
T79 7449 alloy produced at low welding speeds:230
FSW tool shoulder diameter 23 mm

good compromise between mechanical strength and


corrosion performance. More recent TEM studies have
systematically investigated the precipitation state in the
HAZ, TMAZ and weld nugget, as a function of welding
conditions (and also peak temperature of the thermal
cycle, inferred from thermal modelling).236 While TEM
studies reveal precipitation behaviour in detail, they
have limited quantitative capability, and are restricted
practically to a few locations per weld. A powerful technique for obtaining quantitative data covering whole
welds is small angle X-ray scattering (SAXS) using
synchrotron X-rays. This is an expensive technique and
not available routinely, but the studies conducted on
FSW offer valuable insight into the process.230,232,236
Figure 27 shows representative SAXS maps for precipitate volume fraction and size of g precipitates in 7449
FSW. This technique was used to quantify and explain
the differences in final precipitate distributions between
T3 and T79 tempers, each welded at high and low speed.
The SAXS data were supported by selective TEM
observations and provide robust evidence for the interpretation of hardness profiles. In particular, the relative
extent of coarsening in the HAZ, and dissolution in the
nugget are clearly evident. The higher precipitate
fraction in the plate in the T79 condition compared to
T3 is also evident.
Svensson and Karlsson188 and Svensson et al.237
studied the complex precipitation sequence in various
regions of a friction stir weld in 6082, which is one of the
most common alloys welded by this process. The normal
hardening precipitate in this alloy is b99 (Mg5Si6), but
this can dissolve easily at temperatures of 200250uC,
which are easily reached in the HAZ, and another
precipitate, b9 (Mg1?7Si), forms on dispersoids in the
matrix very easily at ,300uC, a typical temperature in
the HAZ. This is less effective as a hardening precipitate.
However, in the nugget region, temperatures are much
higher, allowing dissolution of all precipitates, and b9

68

International Materials Reviews

2009

VOL

54

NO

a in nugget zone dispersoids are free from precipitates;


b in low hardness zone, needle-like precipitates of b9Mg1?7Si are found on dispersoids237
28 Images (TEM) of friction stir welds in 6082T6 alloy

does not form easily, as the weld will cool very rapidly
through the temperature range where such precipitation
can occur. Figure 28 shows typical examples of HAZ
and nugget microstructures.
Fonda et al.196 have studied the precipitation
sequence in an underaged 2195 alloy of 25 mm
thickness. Figure 29 shows examples of the precipitate
structure in the HAZ, TMAZ and nugget for 2195, and
clearly exemplifies the significant changes which occur.
Evidence is shown for the coarsening of the T1 and h9
precipitates. In the TMAZ, these precipitates are
gradually replaced by a d9 phase (Al3Li), with none of
the initial phases being present in the hottest part of the
unrecrystallised TMAZ. GuinierPreston zones were
also detected at this point. Rod shaped precipitates,
identified as TB phase (Al7Cu4Li), were detected in the
nugget region. These often nucleate on b9 (Al3Zr) which
is difficult to distinguish from the d9 phase.
Generally, the trends in precipitation observed in 2xxx
and 7xxx series alloys have been matched in 6xxx
series alloys. Coarsening of b9 precipitates and solution
of needle precipitates was observed in the HAZ of
welds in alloy 6063.142,192,204,228 However, studies of

Threadgill et al.

Friction stir welding of aluminium alloys

29 Transmission electron diffraction patterns and micrographs from HAZ, TMAZ and nugget region of friction stir weld
in 2195 alloy196

precipitation behaviour in the weld nugget of 6xxx series


alloys have produced some conflicting results. Murr
et al.8 published TEM micrographs of the weld nugget in
alloy 6061 showing large precipitates present, but their
identity was not established. However, Lienert and
Grylls189 reported that only second phase particles,
presumably constituent particles, were present in the
weld nugget of the same alloy; the hardening precipitate
b99 was noted to be absent from the microstructure,
implying that dissolution and not overaging had
occurred during welding of this particular alloy. Temperature measurements taken during welding, which
indicated that the precipitate dissolution temperature
was exceeded, were reported to support this observation.
A similar result has been reported in a series of studies
on alloy 6063, welded in both the T4 and T5 conditions.142,192,204,228 Evidently, welding conditions have
an effect on the thermal cycle experienced by the nugget
and this may explain the observations. As observed in
other precipitation hardenable aluminium alloy systems,
aging studies have shown that joint properties may be
improved by post-weld heat treatment.204
Modelling of microstructure and hardness evolution in
heat-treatable alloys

The evolution of microstructure and hardness in heattreatable alloy FSW has been modelled in most detail,
adapting methods developed for arc welding.238244 For
the HAZ, the problem is purely thermal; for the TMAZ
and nugget there is the potential added complexity of
coupling between the deformation microstructure and

precipitation. These microstructure models fall into two


categories:
(i) semi-empirical (with some physical basis), based
on isothermal heat treatments and indirect
calibration via hardness measurement, and able
to predict hardness profiles across welds
(ii) physically based, using detailed thermodynamics
and kinetics of phase transformations, calibrated
on direct measurement of microstructural features,
and able to predict hardness and strength, with the
potential for extension to ductility, fracture toughness, fatigue and corrosion properties.
The semi-empirical methodology has been applied to
FSW in 2000, 6000 and 7000 series alloys.51,52,121,168,245249
This method uses the kinetic strength concept, outlined
in the section on Deformation microstructure in weld
nugget above. Isothermal softening experiments are
conducted, and the hardness scaled to the residual
volume fraction of hardening precipitates. A simple
kinetic model for dissolution is used to calibrate the
temperature dependent dissolution time t*(T). Thermal
cycles are converted to equivalent isothermal holds, and
the final precipitate fraction and hardness predicted,
including the extent of subsequent natural aging. It is
currently limited to artificially aged tempers (T5, T6, or
T7), for which dissolution is the dominant mechanism.
Figure 30 shows predicted and measured hardness
profiles, including predicted curves immediately after
welding and after natural aging.52
More sophisticated approaches to the evolution of
precipitation in heat-treatable aluminium alloys have

International Materials Reviews

2009

VOL

54

NO

69

Threadgill et al.

Friction stir welding of aluminium alloys

30 Semi-empirical model predictions and measured hardness proles in 20 mm thick friction stir weld in 7449TAF, at
six different depths through thickness (after Colegrove et al.)248

been developed and applied to FSW.242244,250,251 Both


these and the simple semi-empirical models have been
incorporated into integrated modelling platforms for
FSW, spanning heat generation, thermal modelling and
prediction of microstructure and hardness profiles for
heat-treatable aerospace alloys.248
In these analyses, the evolution of the full size
distribution of precipitates is modelled, capturing the
competition between dissolution, coarsening and transformation from one phase to another. Extensive use is
made of direct measurement of volume fractions and
particle radii by SAXS and electron microscopy (TEM
or FEG SEM) for calibration and validation of the
model. The key ingredients of the physically based
methodology are:

(i) thermodynamic calculation of phase stability for


both metastable and equilibrium precipitates,
employing thermodynamic database software
(ii) classical isothermal nucleation, growth and coarsening theory, applied to thermal cycles.
More than one population of precipitates may be considered simultaneously, with the competition between
phases and evolution of each phase determined by the
instantaneous microstructural state and temperature.
Figure 31 shows an example of the predicted evolution of various phases for different locations in a FSW
of alloy AA7449T7.248
The models are complex, requiring expert users, and
have a significant computational penalty, but the
potential benefits are large. For example, FEG SEM is

31 Predicted evolution of a precipitate volume fraction and b equivalent radius in AA7449T7 friction stir weld, for three
positions typical of nugget, TMAZ and HAZ (after Colegrove et al.)248

70

International Materials Reviews

2009

VOL

54

NO

Threadgill et al.

Friction stir welding of aluminium alloys

a Mg trace for joining 5083H321 (light) to 6061T6 (dark); b Cu trace for joining 2219T87 (light) to 7075T6 (dark)
32 X-ray maps from nugget regions in dissimilar welds (100 mm markers)

able to provide independent data for grain bulk and


grain boundaries. This opens up the potential for
modelling the effect of dislocation structures on
precipitation within the TMAZ and nugget, including
quench sensitivity effects (i.e. precipitation of nonhardening phases during the cooling part of the thermal
cycle). Strength and hardness predictions can also be
made at a more detailed level than in the semi-empirical
approach, using the predicted volume fraction and
average radius.252,253 In principle, the detailed description of the precipitate state (including distinctions
between grain interiors and boundaries) can be used to
predict more complex but industrially critical properties,
such as ductility, fracture toughness, fatigue and
corrosion. Some recent results for fracture toughness
modelling are presented by Derry and Robson.254
Microstructural models of this type require careful
calibration to data from thermodynamic computation
and high resolution microscopy. They are primarily used
to develop scientific understanding, but, suitably packaged, can offer industrial users with a tool to reduce the
number of experimental trials in, for example, a new
joint design, for previously calibrated alloys. But there
remains the prospect in future of using such tools for
alloy development, in which the friction stir weldability
of a new alloy variant is considered earlier in the
development programme, alongside the core mechanical
and corrosion properties.

between the phases are clearly evident. Hardness data


show the expected behaviour in the HAZ and unrecrystallised TMAZ regions. However, in the nugget different
behaviour is often observed. The hardness in the nugget
can oscillate between the hardness levels expected for
each of the two alloys as seen in Fig. 33a, although this
is not always the case: in Fig. 33b the response is simply
a composite of the responses of the two constituent
alloys (see Fig. 22), with the 6082 hardness in the nugget
immediately after welding being much less than after
aging. In common with welds in a single alloy, the
nugget region is not generally the origin of tensile
failure, so in dissimilar alloys failure will be controlled

Dissimilar welds
In agreement with the distinct macroscale separation
discussed in the section on Mixing across dissimilar
welds, close examination of weld cross-sections commonly shows alternating bands of each alloy, often only
micrometres wide. Larsson et al.49 have described EDS
scans of Mg across two boundaries between 5083 and
6082 regions. The transition between the two Mg levels
is located to within a few micrometres, suggesting only
limited diffusion occurs. No evidence of regions with an
intermediate composition can be seen. Many other
examples exist which show the sharp transition, for
example that shown in Fig. 32, which shows X-ray maps
from 2219T87/7075T6 and 5083H321/6082T6 dissimilar welds. In both cases the sharp boundaries

33 Hardness traverses across 6mm thickness dissimilar


welds:52 (a) 7075T7351/2219T6 and 2219/7075; (b)
6082T6/5083H32152

International Materials Reviews

2009

VOL

54

NO

71

Threadgill et al.

Friction stir welding of aluminium alloys

by the HAZ/TMAZ of lowest strength in the two alloys


concerned.
In summary, both heat-treatable and non-heat-treatable alloys have marked variations in microstructure
and hardness across FSWs. In the former the growth,
dissolution and reprecipitation of strengthening precipitates across the welds must be accounted for, while loss
of work hardening is important in the latter. Models
capable of capturing these effects have been developed.
For both alloy types the complex grain structure
variations can be explained in terms of grain growth,
recovery and recrystallisation driven primarily by the
reduction in stored energy as a function of the local peak
temperature attained.

Mechanical properties
There is an ever increasing volume of data in the
literature on mechanical properties of friction stir
welds.255 It is thus beyond the scope of the present
review to summarise all aspects. An ISO standard (ISO
25239: Friction stir welding aluminium) on FSW will
be published in 2009; this uses other ISO standards to
define standard mechanical tests. When considering
specimen design, as for all welding processes, the
dimension of the initial test specimens should have some
relationship to the final structure, in particular to ensure
that the heat sink is representative, and that a steady
state is reached (at least approximately) and that the
residual stresses are comparable. Finally it should be
noted that hardness was discussed above as a useful
means of delineating microstructural changes across
friction stir welds.

Tensile properties
It is often stated that the tensile properties of friction stir
welds generally equal or exceed those reported for fusion
welds. Although this is often the case, some qualification
of this statement is in order:
(i) tensile properties of fusion welds made with a
filler are often determined as much by the filler
as by the welding process: in general when fusion
welding aluminium alloys, the filler wire is not
the same composition as the parent material, and
therefore may not have the same mechanical
properties
(ii) alloys such as 2xxx and 7xxx are designed to
have high strength, and therefore the strength of
welds is of particular importance. Unfortunately,
these alloys are generally difficult, and sometimes impossible, to weld by fusion processes;
thus, comparative data from high quality fusion
welds is scarce, or non-existent, and comparisons
with FSW are not always straightforward
(iii) when comparing tensile data on different types
of weld, care should be taken to establish how
the measurements have been made, e.g. removing
the overfill in fusion welds may affect the
properties, and not all authors state whether or
not this has been done; determination of yield
stress is dependent on the technique and equipment used (again not always stated).
When assessing tensile data, it should be remembered
that, as shown above, the microstructure across a
friction stir weld is typically highly non-uniform. As a
result yield strength, tensile strength and ductility may

72

International Materials Reviews

2009

VOL

54

NO

change considerably over very short distances. Consequently very different results can be obtained according to whether the welds have been tested longitudinal or
transverse to the weld. The stressstrain response will
vary even for cross-weld tests according to the width of
sample, since this will determine the retained residual
stresses (see below), and the length of the testpieces since
this will determine the average ductility/overall elongation, 0?2% yield stress, etc. There are many reports of
low elongation in cross-weld tensile tests of welds;
however, in many cases this is not due to low ductility,
as confirmed by the significant reduction in area.
Instead, the strain will have been concentrated in a very
small part of the gauge length where a locally softer
microstructure may have formed. Studies of deformation by Mahoney et al.6 on 6?35 mm 7075T7541 and
Liu et al.256 on 5 mm 1050H24, 6061T6 and 2017
T351 have demonstrated the variability in strain across
transverse tensile samples. Consequently, overall elongation measurements made on cross-weld samples tend to
be unrepresentative of any region of the weld and serve
only to identify the likely failure location under static
loading.
Two strategies have been developed for extracting
more representative data to map the properties across
friction stir welds, the former more suited to extracting
longitudinal properties, the latter capable of extracting
transverse properties:
(i) the excision of matchstick style microtensile test
samples: in this manner samples can be removed
parallel to the welding direction x that are
representative of parent, TMAZ, HAZ or weld
nugget
(ii) the cutting out of thin cross-weld testpieces: here,
the microstructure varies along the length y of the
testpiece, but not through the thickness z or
width x, provided the testpiece is sufficiently thin.
In such a case deformation behaviour will vary as
a function of y position. This can be monitored
by full field strain mapping using laser speckle
interferometry or digital image correlation, for
example.
Microtensile testing has provided the bulk of the data
delineating the variation in mechanical properties across
the various microstructural zones of friction stir welds
(Fig. 34a and b). Such studies include those by von
Strombeck and co-workers,257259 Allehaux et al.260
using 10 mm thick 7349T6 alloy and Denquin et al.261
using 6 mm thick 6056 alloy. In the last case, minimum
ductility was reported in the centre of the nugget region;
the stressstrain response parallel to the welding
direction was measured as a function of lateral distance
from the weld, e.g. Fig. 34b.
Cross-weld testing can provide useful insights if the
strain is measured as a function of position through the
microstructural zones. Initially, this was carried out by
monitoring closely spaced parallel lines6 or Vickers
indents,204 laser extensometry262 or a small number of
strain gauges.263 Mahoney et al.6 recorded the distribution of strain at failure for 7075T651 alloy FSW which
showed very close agreement with the hardness variation
characteristic of FSW for the alloy. The peak elongation
(,15%) corresponded to the HAZ with the strain in the
weld nugget close to that of the parent in accord with
their similar hardness. A more sophisticated approach is

Threadgill et al.

Friction stir welding of aluminium alloys

34 a hardness proles for Al2024 friction stir weld at three depths,259 b corresponding longitudinal tensile performance
as determined by microtensile specimens,259 c variation in hardness (bold circles) and 0?2% proof stress (open circles) as determined from cross-weld tensile test180 monitored by electronic speckle pattern interferometry for FSW
AA5083 welded at 200 mm min21 and d corresponding evolution of tensile strain with position across weld as crossweld load is raised180 (HAZ boundaries marked by dashed lines)

to measure the strains by digital image correlation264 or


by electronic speckle pattern interferometry.180,265 The
variation in 0?2% proof stress across an AA5083 FSW
joint are shown in Fig. 34c. It is clear from Fig. 34d that
failure occurs just outside the tool shoulder because this
is the softest region of the weld zone. In this case the
performance perpendicular to the weld direction is
measured as a function of lateral position from the
weld, e.g. Fig. 34d.
When collecting tensile strength data on the basis of
cross-weld testing it should be remembered that friction
stir welds typically have significant residual stresses (see
below). Test samples cut from larger plates may retain a
significant proportion of the weld stresses which may
compromise tensile strength measurements. Only when
the cross-weld sample width is less than the size of the
tensile zone (approximately the width of the HAZ) is the
residual stress negligibly small.266
As a result of an international collaborative effort
(Eurostir),267 data for a variety of alloys have been
published: the performance of different alloy types and
tempers from this and numerous other sources (where
indicated) can be grouped together. Their generic
hardness responses are shown schematically in Fig. 23
and their mechanical performance is as follows.
For heat-treatable alloys (e.g. 2xxx, 6xxx or 7xxx
alloys in the T6 or T7 condition), irrespective of the heat
treatment condition, cross-weld tensile tests normally
fail at the side of the nugget, at or close to the HAZ/
TMAZ boundary.268 The failure mechanism is a ductile
shear failure, showing 45u facets. In thicker samples, the
faceting may be more complex, but the mechanism is the
same. The elongation is almost invariably less than
found in the parent material, due entirely to concentration of strain in softer regions. The local ductility at
failure can be estimated from the reduction in area, and

substantial necking, indicating good ductility, is


common.
Tensile strength is typically about the same as found
in the parent material in the annealed condition,
although significant improvements can be made by
minimising the thermal cycle. Joint efficiencies exceeding
90% have been reported in 7xxx alloys. Liu et al.269
reported joint efficiencies as high as 82% for 2017T351.
Sato and Kokawa204 report an inverted top hat profile
for yield strength, the trough in the nugget and HAZ
being 50% of the parent T5 condition for 6063Al. Upon
post-weld aging 95% of parent strength in the weld
region was recovered, although the fracture location
remained unchanged.
Failures can occur on advancing and retreating sides,
although for a series of welds they will usually all fail on
one side or all fail on the other.
For work hardened non-heat-treatable alloys (e.g.
5xxx alloys in the H1xx, H2xx or H3xx conditions),
failure of cross-weld tensile specimens normally occurs
in the centre of the weld, where the hardness is at a
minimum (Fig. 34). For example, in Fig. 15, the
AA5454 joint has failed at a kissing bond.156 Despite
this, the failure stress and elongation is only some 10%
less than the parent. For AA1050H24 failure was on
the advancing or retreating side with the distance from
the weld centre decreasing with decreasing pitch in
correspondence with a narrowing troughs in hardness.268 The failure stress observed is typically close to
the annealed strength of the material, although higher
values can be obtained if the heat input is minimised.
Elongation values are normally a little below the parent
value, but the reduction is less than in heat-treatable
alloys. Failures are fully ductile, with extensive necking.
For annealed non-heat-treatable alloys (e.g 5xxx in the
O condition), failure of cross-weld tensile alloys can

International Materials Reviews

2009

VOL

54

NO

73

Threadgill et al.

Friction stir welding of aluminium alloys

occur anywhere on the sample, although they usually


occur away from the weld in the parent material.
Elongation is therefore typically the same as the parent
material, and the failure mode is invariably very ductile.
Thus joint efficiencies of 100% can be obtained.
One advantage often quoted for tensile properties in
friction stir welds is the very consistent performance
from weld to weld. This is perhaps well illustrated by
data from Lockheed Martin,270 where the analysis of a
large number of welds in a 2xxx alloy showed that FSW
gave rise to a small increase in average tensile strength,
but a very much reduced scatter band. As design of the
component in question was based on minimum strength
which can be guaranteed, the higher repeatability of
friction stir welds allowed an extra 20% strength to be
used in the design, even though the average strength of
the friction stir welds was not much more than that of
fusion welds. It should be noted that this is due to the
low variability in weld properties rather than the
incidence of defects. Typical data are shown in Table 3.
Ideally, the loss of parent material strength in an FSW
would be negligible. At present, parent material properties can only be achieved with annealed alloys which
cannot be softened by further heating during welding.
Since loss of strength and hardness is usually related to
either overaging in precipitation hardened alloys, or
annealing in work hardened alloys, minimising the heat
input should offer a way of improving properties.
However, this approach is limited by the fact that
the material being welded must be hot enough to flow,
and in aluminium alloys this temperature will cause
softening.
Heat input can be minimised by several methods for
example:
(i) use of more efficient tool designs that require
less energy to push them through the weld
(ii) use of higher welding speeds and/or lower
rotation speeds
(iii) use of artificial cooling (water sprays, welding
underwater, etc.)
(iv) active tool cooling.

Residual stresses
As welded residual stresses

Residual stresses in welds are of great significance in


determining weld performance, in particular fatigue and
fracture toughness. In fusion welding, residual stress
levels are often at, or very close to, parent material or
weld metal yield strength. In solid state welds the
residual stresses can be substantially lower, although this
is not necessarily so. In comparing residual stress levels
in friction stir welds, great care must be taken in
interpretation of the results, as there are several
techniques which are commonly used. Determination
Table 3 Tensile data for variable polarity plasma arc
(VPPA) and FSW joints showing benet of FSW
Ultimate tensile strength, MPa
VPPA
Thickness,
mm
Form
5.1
8.19.8

74

FSW

Average Minimum Average Minimum

Plate
358
Extrusion 323

International Materials Reviews

310
200

2009

392
394

VOL

378
369

54

NO

35 Longitudinal residual stress distribution normalised


by pin shoulder diameter for friction stir welds in
7449,326 2199,301 6082,54 2024,324 and 5083 alloy54

of residual stresses is a complex area, and an authoritative summary of the methods which can be used is
beyond the scope of this review. Readers are referred to
other texts for more detailed information on the origins
of weld residual stresses,271,272 as well as the measurement of residual stresses by destructive (e.g. hole
drilling,273 contour method)274,275 and non-destructive
(e.g. neutron diffraction,273,276,277 synchrotron X-ray
diffraction277,278 and magnetic)275,279 techniques.
Several authors have determined residual stresses
non-destructively using synchrotron X-ray diffraction.180,280284 Although different materials were examined, there is broad agreement in the results, in that the
longitudinal stresses tend to show the largest variation,
being most tensile in the HAZ, lower in the nugget and
compressive in the parent plate (Fig. 35).
Similar trends have been recorded by Staron et al.285
using neutron diffraction. Destructive methods such as
the contour method,286,287 the crack compliance
method288 and incremental centre hole drilling289 have
also been applied.
The characteristic magnitude and profile of the
longitudinal stresses across a friction stir weld are
shown in Fig. 35 for a range of alloys. The longitudinal
stresses are typically much greater than the transverse.
As is clear from the figure, the stresses tend to be tensile
over a region extending just beyond the diameter of the
tool shoulder. The tensile region tends to encompass the
nugget and TMAZ and reflects the extent of the hot
region beneath the shoulder. The peak stresses are often
found just inside or just beyond the shoulder radius.
Often the peak stress lies within the HAZ despite the
lower hardness often found there. Lower level compressive residual stresses are typically found in the parent
plate beyond the HAZ. The depth of the tensile plateau
below the tensile peaks and the presence of a subsidiary
peak on the weld centreline appear to be alloys specific.
It should also be noted that the breadth of the tensile
region and the magnitude of the stresses vary greatly
according to the processing conditions. For example,

Threadgill et al.

Friction stir welding of aluminium alloys

36 Longitudinal and transverse residual stress variation: above, plate cross-section in 20 mm thick AA7449/7449 friction
stir weld (tool diameter 34 mm);326 below, at midthickness for dissimilar 6082/5083 weld (tool diameter 18 mm) in
3 mm thick plate54 (weld started at white spot and nished at black spot). In both cases advancing side is on right

much lower stresses than those plotted in Fig. 35 are


reported for 2024,290 5083,180 and 6013.281
Although almost all the attention to date has focused
on the variation of longitudinal stresses lateral to the
weld line measured midthickness and midweld length,
where the situation can be described as steady state (e.g.
Fig. 35), in practice the stress field varies both through
thickness and along the welding direction. The variation
through thickness for 20 mm 7449 plate is shown in the
upper part of Fig. 36. It is clear that the largest stresses
are found near the surface (in this case directly under the
edge of the shoulder). Further the stress field profile
broadly follows that of the HAZ and the hardness
variation being narrower at the base where the heat
input is less. For plates less than 6 mm thick the stress
field is essentially uniform with depth through thickness.282 The lower part of Fig. 36 shows the inplane
variation in residual stress for a 3 mm thick 6082/5083
weld. It is clear that, in the case of the 100 mm long weld
shown, the largest stresses are not found midlength but
continue to rise towards the end position. This is in
contrast to TIG welding similarly sized plates where the
stresses the fell towards the end of the plate as the plate
warmed up and thus the mismatch between weld and
plate decreased.277,291 Also, in contrast to TIG welding,291 the stresses around the start and end positions are
not particularly high. Both these observations may be
the result of lower heat input associated with FSW. The
stresses are not symmetrical across the weld line in
Fig. 36 because the stresses for the 6082 are lower in the
nugget than for the 5083,54 due to dissolution of the

hardening precipitates there. The transverse stresses are


generally much lower than those in the longitudinal
direction with the most significant tensile stresses (about
50100 MPa) found around the exit hole.
If the stress is measured, as is often the case, on a
cross-weld sample extracted from a larger welded plate,
account should be taken of the possibility of stress
relaxation. In essence a weld length approximately eight
times the diameter of the tool must be retained if 90% of
the residual stresses are to be retained on cutting out the
testpiece.266,286 This criterion is often not fulfilled, which
may explain the low stresses observed in some
studies.281,286,292
To understand how the residual stresses arise, finite
element modelling (FEM) provides a picture of how the
stresses evolve during welding. In this respect it is
important to note that current FEMs for predicting
residual stresses tend to ignore the mechanical stirring
effect of the tool and regard it simply as a heat
source.293296 Such models involve a decoupled heat
transfer and a subsequent thermomechanical analysis.
An attempt has been made to incorporate the mechanical effect of the tool using FEM, representing the down
force and torque loading elastically, without incorporating large plastic strains.128 However, to take proper
account of the material flow around the pin an arbitrary
LagrangianEulerian formulation must be applied to
avoid excessively severe distortions of the mesh.152 Fluid
dynamics based models are now beginning to emerge
capable of predicting residual stresses, which explicitly
take into account fluid flow.119,297,298 Unlike purely

International Materials Reviews

2009

VOL

54

NO

75

Threadgill et al.

Friction stir welding of aluminium alloys

37 Measured and predicted effect of traverse speed and rotation speed: experimental data are for 2199,301 predictions
for 7050297

thermal models, these are capable of predicting differences in stress between the advancing and retreating
sides of the weld. Typically very little difference is either
predicted or observed in practice; however, Fig. 37 does
show slightly high measured and predicted stresses on
the advancing side and this observation is supported by
the work of others.299
The evolution of longitudinal stress as the tool passes
is depicted in Fig. 38. It demonstrates that ahead of the
tool the compressive stress caused by the expanding hot
material impinges on the compressive yield stress locus,
causing local plastic straining. Just behind the tool
longitudinal tensile stresses begin to generate as the weld
material cools. Initially stress development near to the
weld line is limited by the low tensile yield stress
(Fig. 38c). This local tensile plastic straining at the weld
line results in the characteristic M shape typically
observed in the welded plate (Fig. 38d), as the hot region
plastically deformed in compression, to a greater width,
ahead of the tool becomes stressed in tension as it cools
behind the tool. As the tool travels forwards and the
temperature falls, the tensile stress level builds up at a

76

International Materials Reviews

2009

VOL

54

NO

rate slower than that at which the yield stress rises so


that a point is reached very soon after the tool has
passed when no further yielding occurs and the
increasing misfit is then accommodated elastically. As
the tensile stresses develop near the weld line these are
balanced by a compressive stress towards the edges of
the plate. Lombard et al.300 have noted that for AA 5083
the width of the tensile region increases and the
maximum tensile stress decreases with increasing heat
input. In agreement with experiment,180,284,300,301 the
residual stress distribution has been found by modelling294,295 to be dependent on the welding traverse and
rotational speeds (Fig. 37). Despite the fact that the
parametric study is for AA2199,301 and the modelling is
for AA7050,297 the responses in Fig. 37 show a high
level of agreement. In both cases the M profile
characteristic of FSW shows a small secondary peak
centred on the weld line, the peak situated just outside
the tool shoulder is slightly higher on the advancing side
and the width of the tensile region increases, and the
magnitude of the tensile region decreases, as the traverse
speed is lowered and the rotation speed is increased. This

Threadgill et al.

Friction stir welding of aluminium alloys

a as heat source approaches and compressive stresses form; b directly through tool centre, showing resultant reduction
in thermal strain by compressive yielding; c 8 mm behind centre of pin and at edge of shoulder, as heat source
retreats, material begins to cool and tensile yielding occurs; d final stress state after removal of tensioning loads
38 Effect of external mechanical tensioning (given as percentage of parent plate yield stress) on predicted longitudinal
residual stress proles for AA2024T6 plate friction stir welded at 770 rev min21 and 195 mm min21 (tensile and compressive yield loci are shown as dashed lines)296

is believed to be because the high heat input associated


with low traverse speeds and high rotation speeds (see
Fig. 5) leads to more extensive softening in the weld
region. This in turn results in overall reduction in
magnitude of the longitudinal residual stress, but a
wider tensile stress zone (Fig. 37). The residual stress
field is governed primarily by the as-welded yield
strength distribution in the stir zone, HAZ and
TMAZ. On the other hand, the natural aging process,
while strongly affecting the final weld strength, shows
minimal influence on the residual stress.54,297
Residual stress control

Residual stresses arise from the accumulation of misfit


between the weld region and the remaining plate. There
are a number of means by which the misfit, and hence
the residual stresses, can be manipulated, e.g.
(i) thermal tensioning
(ii) mechanical tensioning
(iii) subsequent processing treatments.
One of the earliest reported applications of thermal
tempering was by Greene and Holzbaur,302 who in 1946
used superimposed temperature gradients to achieve
reduced residual stresses in longitudinal butt welds in
ship hull structures. Local induction heating has also
been investigated for residual stress improvement.303
Michaleris and Sun304 and Dull et al.305 have applied
thermal tensioning to reduce buckling distortion,
whereas Dong et al.306 developed an in-process thermal
stretching technique for effective mitigation of residual
stresses and distortion on repair welding of aluminium

panels. Barber et al.,307 van der Aa et al.308 and Williams


and co-workers285,309311 have applied local cooling to
FSW, using either solid or liquid CO2 trailing the heat
source, as a means of creating dynamically controlled
low residual stress and distortion free welds; others have
used water jets.312
Several mechanical tensioning systems have been
proposed. Yang et al.313,314 have mechanically compressed the weld on cooling using a pair of rollers
positioned on either side of the weld line, reducing both
residual stress and buckling distortion. Williams and coworkers79,296,315 have shown that the application of
global, or far field, mechanical tensioning externally
during the welding process can greatly reduce the tensile
residual stresses in FSW joints. In global mechanical
tensioning a load is applied uniformly along opposite
ends of the plates prior to clamping the parts for
welding, so that a uniform tensile stress is maintained in
the two butted plates parallel to the weld line. The
clamping and tensioning loads are then released after the
FSW tool has traversed along the join line forming a
weld. Perhaps counter intuitively, Williams et al.315
found that high levels of mechanical tensioning parallel
to the welding direction can actually reverse the state of
stress, so that compressive longitudinal residual stresses
are found in the weld region.
Global mechanical tensioning operates by reducing
the compressive bow wave ahead of the travelling heat
source and increasing the tensile plastic strain developed
in the hot zone trailing the weld. Figure 38 indicates that
for low levels of tensioning (,40%) a reduction in the

International Materials Reviews

2009

VOL

54

NO

77

Threadgill et al.

Friction stir welding of aluminium alloys

39 a comparison of measured (left)266 and predicted (right) longitudinal stress proles for AA7449W51 welded plates as
function of tensioning level (0, 5, 10, 20, 30% of parent alloy yield stress): dotted prole represents predicted untensioned (0%) case for which there were no measured results; b residual stress at midthickness near weld line as function of applied global tensioning level for various alloys290,326

compressive plastic strain field ahead of the tool, as the


hot material expands, is mainly responsible for reducing
the final residual stresses. At higher tensioning loads,
little or no compressive misfit develops ahead of the tool.
Instead, larger tensile plastic straining of the softened
hot material after the tool has passed causes a tensile
misfit, or over tensioning, once the tensioning forces
are removed. This results in the compressive longitudinal stresses seen along the weld line (Fig. 39b).
Taken together, these effects give rise to the observed
approximately linear reduction in longitudinal weld
stresses with tensioning level and zero residual stresses
can be engineered at tensioning levels of about 4050%,
depending on the material and welding conditions. The
results indicate that while the level of residual stresses
present in the untensioned case is a function of the alloy,
the rate of residual stress reduction brought about by
mechanical tensioning is essentially alloy independent
(Fig. 39b). In all cases studied it is essentially linear with
respect to the tensioning load, so that the tensioning
required to reduce the weld stresses to zero can be
calculated directly from the stresses present in the
untensioned case. For thin plates a guideline rule is that
1 MPa of tensioning reduces the tensile stress by
approximately 1 MPa. Global tensioning was found to
be less effective at greater depths in thick plates.
Furthermore a reduction of the bending distortion and
an increase in angular distortion was observed with
increased tensioning, while no effects on the weld
microstructure and hardness were observed.
Besides post-weld heat treatment,316 there are a
number of post-weld treatments by which the misfit
introduced during FSW can be reduced to lower the
residual stress level, or replaced with another one
leading to beneficial residual stresses. Burnishing,317,318
laser and conventional shot peening263,319321 have been
used to introduce new misfits and thus modify the near
surface state of friction stir welds introducing
compressive stresses that are beneficial to the fatigue
and stress corrosion behaviour.319,322
It is also possible to apply the roller tensioning
method described above after welding.323 Whereas
during welding two rollers are passed on either side of
the weld, a single roller approximately equal in width to

78

International Materials Reviews

2009

VOL

54

NO

the FSW tool shoulder can be rolled along the weld line
once the weld has cooled. Recent stress measurements
suggest that this approach is much more effective than
when applied during welding (Fig. 40) with loads in
excess of 15 kN leading to compressive weld stresses for
2199.324 This compares with little effect during welding
using two rollers and a combined down force of 75 kN.
By contrast post-weld mechanical tensioning is much
less effective than that applied during welding.296
Distortion

There is of course a strong link between residual stresses


and distortion in welds. In most cases, FSW of
aluminium produces low distortion levels compared
with arc welding. However, significant distortion can
occur in friction stir welds, in particular in thin gauge
welds where the design imparts an asymmetry in
restraint or heat sink. Preliminary studies show that
thermal and mechanical tensioning methods used to
control stress are also effective at reducing longitudinal
distortion.79,325,326
In summary, residual stresses and distortion in
friction stir welds in aluminium alloys can be engineered
to be considerably less than those typical of fusion
welds. The characteristic tensile longitudinal stress tends
to be of lower magnitude but broader in extent the

40 Longitudinal midthickness residual stress proles as


function of post-weld roller tensioning load for
AA2199 friction stir weld:324 roller is shown inset, its
footprint is indicated by horizontal solid line

Threadgill et al.

Friction stir welding of aluminium alloys

a initial residual stress before crack (solid line) and contribution of crack to stress intensity factor (dashed line) as function of crack growth for transverse crack; b corresponding residual stress redistribution ahead of crack tip
41 Residual stresses in 3?2 mm thick 2024 friction stir weld testpiece, measured by crack compliance technique (horizontal dashed line indicates compressive yield strength of material)299

greater the heat input. As a solid state welding process


there is considerable scope for manipulating the state of
stress during welding. Thermal and mechanical tensioning have been found to be successful in reducing and
even reversing the state of stress. Roller tensioning after
welding has also been found to be successful in lowering
the tensile residual stresses.

Fatigue
As with other mechanical property data, care needs to be
taken to ensure that full information on the test
procedure is available. In particular, in some fatigue
tests, the surface of friction stir welds is dressed to remove
flash and the surface markings. Occasionally even more
material is removed, and this is acceptable if the weld in
question is machined in the same way in service. When
comparing data, R values should be checked to ensure
they are the same especially in cases where residual
stresses may exist in the testpiece, but these are not always
reported. Examination of the literature has shown that
the following generalisations can be made:
(i) in simple SN tests on cross-weld samples, the
fatigue performance of friction stir butt welds is
typically less good than that of the parent
material tested under the same conditions.327,328
Many studies have found that after milling the
top surface, the fatigue performance of 2014,
6013 and 7475 FSW joints approached that of
the parent alloys,329331 yet in other studies332
the properties remain significantly below parent
material benchmarks
(ii) the fatigue performance of friction stir butt
welds generally comfortably exceeds that of
comparable fusion welds,93,226,330,331,333,334 a
trend reported for many alloy grades
(iii) failure is normally (but not always) associated
with an initiation event at the geometric stress
concentration at the side of the weld on the
upper surface;335 where this has been machined
away, failure normally initiates in the region of
lowest strength. For many alloy groups, these
two locations are very close together

(iv) residual stresses can play a significant role in the


fatigue behaviour191,336338 and vary according
to crack test geometry.339 In this respect it
should be noted that the dimensions of most
crack test specimens lie between the limits for
negligible and complete relaxation of the
residual stresses introduced by FSW.266
In general, for machined welds, for which surface finish
is not an important influence, three factors appear to
play a dominant role in the fatigue performance:
residual stress, microstructure and defects. As a consequence, without a detailed picture of all three and an
understanding of their interactions, it is difficult to
discern simple trends as a function of FSW processing
conditions from the literature. Indeed, because of the
overriding and interacting importance of these three
poorly reported factors much of the data appears
inconsistent and contradictory.
Dalle Donne et al.338 found that residual stress was by
far the most important effect in their compact tension
fatigue study in which fatigue crack propagation varied
strongly as a function of R (maximum/minimum stress
ratio). In agreement with others340,341 they found that
once residual stress was taken into account very little
variation in fatigue performance was observed as a
function of FSW process parameters.
Naturally, as the crack grows the residual stresses
redistribute. This can be especially important in
considering residual stress effects on crack growth
across the weld line (i.e. transverse) because the
longitudinal stresses are typically the largest (see above).
A number of researchers have calculated the effect of
residual stress redistribution on the stress intensity
factor for cracks growing through FSW joints.299,342
Figure 41 shows the initial residual stress field in an
FSW testpiece, as determined by the crack compliance
technique, along with the contribution to the stress
intensity arising from the residual stresses.299 It is clear
that the contribution can be very significant.
In general the refined microstructure characteristic of
FSWs leads to improved fatigue properties compared
with fusion welds.226,331,343

International Materials Reviews

2009

VOL

54

NO

79

Threadgill et al.

Friction stir welding of aluminium alloys

a in as-welded condition; b after 2% stretching (filled and open symbols correspond to different samples)
42 Crack growth data for FSW in 2024T351 for cracks growing parallel to weld in CT samples:341 plots are for cracks
located at various distances from weld line propagating parallel to weld

Unsurprisingly, the presence of certain defects,


especially root flaws such as zig zag kissing bonds
containing oxide defects, can compromise fatigue
performance.161,343347 James et al.348,349 comment on
planar onion ring defects associated with the weld
nugget that can accelerate the link-up of cracks in 5083
and 5383. Dickerson and Przydatek 350 suggest that root
flaws up to 0?35 mm deep are tolerable without a
significant loss in performance when compared to
nominally flaw free welds. Widener et al.351 found that
for FSW 2024 once defects were eliminated by careful
process control similar fatigue lives were obtained across
a wide range of process parameters.
Effect of alloy system

The plethora of fatigue studies can be broken down by


alloy system.
Extensive research has been carried out on 2xxx
alloys.343,346,351353 For example, Fig. 42 compares compact tensile crack propagation tests through welded and
parent (TL: transverselongitudinal section) 2024T351
material (cracks parallel to the weld direction). The
lowest threshold DK values were found for cracks
propagating in the HAZ, which corresponds to a region
of minimum hardness. By contrast, cracks propagating
through the weld grew at rates 10 times slower than
those of the parent plate. The largest threshold DK
values corresponded to the TMAZ, which was some 15
times that for the parent plate. All the data appear to
converge at large DK. To investigate the origin of these
differences, the residual stresses were relieved by over
90% by a 2% stretch normal to the weld direction, which
was insufficient to appreciably affect the microstructures. The fact that the fatigue crack growth rates for
weld line and TMAZ material subsequently coincide
with those for the parent plate was taken as evidence that
the differences in crack growth rate were primarily due to
residual stress. After stretching, the crack growth rate in

80

International Materials Reviews

2009

VOL

54

NO

the HAZ material remains inferior to parent and weld


metal, which may be a microstructural effect. For
thumbnail cracks growing across the welds, the opposite
effect was observed with the slowest rates corresponding
to the HAZ and the fastest in the TMAZ. This is
consistent with the hypothesis that the residual stresses
control fatigue crack growth resistance since the longitudinal stresses in 2024 are normally significantly tensile,
with the TMAZ having the largest tensile stresses and the
HAZ the compressive or low tensile stresses (see Fig. 35).
Milan et al.354 found that under transverse cracking
compressive residual stresses increase fatigue resistance in
2024 until the more brittle weld region is reached.
A number of studies have been undertaken on 5xxx
series alloys.336,337,355359 James et al.348,349 found that
lower welding power (but sufficient to give adequate
plasticisation for a good weld bond to be made) gives the
highest observed fatigue lives. However they found little
correlation between the static and dynamic performance
indicators, suggesting that crack path effects associated
with onion skin defects are the primary cause of fatigue
crack initiation and growth.
Most of the work concerning the fatigue properties of
6xxx series alloys focuses on 6055,357 6056,316,360
6061334,336,361 and 6082.316,333,334,340,357,362,363 It has
been found that fatigue crack growth rates for compact
tension specimens in the dynamically recrystallised zone
are lower than for the parent metal especially at low DK.
As in other systems this has been ascribed to the
beneficial effects of compressive residual stresses in
competition with detrimental grain refinement that
brings about intergranular failure.361
Some fatigue crack growth data have been reported
for 7xxx alloys.191,339,353 Jata et al.191 carried out
eccentrically (or extended) loaded single edge tension
tests on 7050T7451 with the crack running through
different regions parallel to the weld direction. They
found that, at a load ratio of 0?33, fatigue crack

Threadgill et al.

Friction stir welding of aluminium alloys

lifetime of the specimen is the time necessary for one of


these cracks to propagate up to a critical length. Fatigue
efficiency of some 15% of the static strength was
estimated, the cracks failing under a mixture of mode I
and mode II.
Fatigue extension strategies

43 Comparison of fatigue crack growth rates in laboratory air between weld nugget and HAZ for friction stir
welds in 7050T7451 plate at load ratios R50?33 and
0?7: specimens were in as FSWzT6 condition191

propagation in the weld nugget region is inferior, while


the growth rate in the HAZ is superior to that of the
parent (Fig. 43). On the basis of crack closure,
fractography, microstructure and residual stresses they
concluded that compressive residual stresses dominate
fatigue crack growth in the HAZ, whereas for the fine
(15 mm) grain dynamically recrystallised weld nugget
region, microstructure and intergranular failure
mechanism dominate. This observation is corroborated
by a tendency for differences in fatigue crack growth
rates to become smaller at a load ratio of 0?7 (Fig. 43).
Kumagai et al.364 found that for 7050 the fatigue
strength was close to that of the base metal with fracture
in the softer HAZ.
Lap welds

There is much less information available on lap welds,


and comparison with welds made by other processes is
not possible at present. However, recent work by
Ericsson and Sandstrom365 has shown that improved
tool designs which increase the volume of disturbed
metal at the interface, and which minimise the plate
thinning and hooking defects associated with inadequately optimised lap welding can lead to significant
improvements. Similarly, recent work by Thomas
et al.366 has reported promising results for lap welds in
5083H111 made using an advanced tool design.
Friction stir welded lap joints in 2024T351,367 and
7075T7351,368 have shown improved fatigue performance over mechanically fastened joints of similar
geometry when tested under similar conditions.
However, Shepherd has indicated that in 2024T351
lap welds, an improvement over bolted joints is only
obtained after weld surface removal.369 Christner et al.92
have published information on fatigue of lap welded
joints in thin gauge 2024T351, and found that fatigue
properties equalled or exceeded those of bolted joints.
There is substantial evidence that more complex tool
motions may also be useful in improving lap weld
quality.370 As discussed above, single lap joints inevitably contain two crack-like regions, which can be
straight or deflected (termed hooking, see Fig. 18).
Fersini371 has considered these regions as cracks in
order to estimate the fatigue lifetime using FEM. The

Until recently, little attention has been paid to fatigue


improvement techniques. Most of the attention has been
focused on the use of mechanical surface treatments
designed to place the near surface region into a state of
residual compression. Laser and shot peening have been
examined.263,319321 Typically shot peening introduces
compressive residual stresses to a depth of around
200 mm, but laser peening can introduce residual stresses
to a depth in excess of 1 mm. In some cases shot peening
has been found to be effective,319,321 whereas in others263
only laser peening provided a significant increase
(,120% compared with 10% for shot peening) in fatigue
life. Low plasticity burnishing317,322,372 can also introduce deep (.1 mm) compressive residual stresses.
Jayaraman et al.322 report increases of 80% in the high
cycle fatigue endurance of aluminium alloy friction stir
welds.
In summary, although good fatigue results are
consistently obtained with FSW, it is premature for
design codes and joint classifications to be relaxed for
friction stir welds, although this may in due course be
justified. Lomolino et al. have collected and statistically
analysed fatigue data on FSW for a range of alloys to
derive a first set of reference fatigue curves.373

Fracture toughness
Fracture toughness is not normally a problem in
aluminium alloys, but there are nevertheless areas where
it is important, in particular at very low temperatures
such as those encountered in cryogenic structures, or in
exterior surfaces of airframes. Fracture toughness has
been studied using a number of testing configurations
for alloys including 2014,374,375 2024,257,376 2139,116,377
2195,378 5083,374,379 6061,257,258 7075,374 and 7449.254,380
Mochizuki et al.379 found that for 5083O, in contrast
to hardness and static strength, the Charpy impact
energy and critical crack tip opening displacement
(CTOD) in the friction stir weld are much higher than
those corresponding to the parent metal or the HAZ.
This was ascribed to the fact that the fine grained
microstructure in the stir zone helps to increase ductile
crack initiation and propagation resistance.
Dawes et al.374 investigated the R curve behaviour of
2014A, 5083 and 7075; typical data are shown in Fig. 44.
This shows that the fracture toughness in the nugget and
the HAZ/TMAZ region exceeded that of the parent
material, presumably because of the very fine grain
structures. This effect was found for all alloys tested,
although the magnitude of the difference varied between
alloy types. von Strombeck et al.259 obtained similar
results on 2024T351, 5005H14 and 6061T6 alloys
using a CTOD parameter to measure toughness rather
than the J parameter used by Dawes. Only 2024 joints
exhibited similar or slightly lower fracture toughness
than the parent alloy. This behaviour was attributed to
changes in the characteristics of the inclusion and
precipitates population. Supporting these observations,
Brinkmann et al.258 have reported that the fracture
toughness of the nugget in 3 mm 6061T6 friction stir

International Materials Reviews

2009

VOL

54

NO

81

Threadgill et al.

Friction stir welding of aluminium alloys

44 Toughness data for 2014T6 friction stir welds381

welds (and repair welds) exceeds that of the parent


material by a considerable margin.
A number of studies on the fracture toughness of
2195T8 have given broadly similar results.377,378,382
Kroninger and Reynolds in particular studied R curve
behaviour in 2195.378 The observed improvement in R
curve behaviour for weld line and nugget seen in Fig. 45
is ascribed primarily to microstructural origins: these
regions being solution treated and naturally aged by the
tool giving greater toughness than the overaged HAZ
and peak aged parent. The poor response of the VPPA
weld is ascribed to brittle solidification products absent
in the FSW.
More recently Kristensen et al.267 published limited
data on 6 mm thick 2024T3, 5083H111 and 6082T6
alloys. Conventional CTOD tests at room temperature
showed that the fracture toughness of the nugget in all
cases equalled (for 2024T3), or exceeded, the parent
material values, in contrast to the R curve behaviour
reported by von Strombeck et al.259 Mochizuki et al.379
studied the toughness of welds in 25 mm thick 5083O
at ambient temperatures and at 196uC. At both
temperatures, the toughness of the stir zone exceeded
the toughness of the parent plate by a substantial
margin. HAZ toughness was slightly above that of the
parent material toughness.
A good example of the importance of microstructure
for toughness is given by the work of Derry and
Robson254,380 on FSWs in 7449TAF. Failure is
predominantly intergranular, with a low level of
transgranular failure through microvoid coalescence.
This is because the metastable g9 precipitates are
semicoherent with the matrix and therefore allow
dislocations to pass through them, resulting in inhomogeneous slip and stress concentrations where slip bands
meet the grain boundaries. Grain boundary failure is
exacerbated by low strength precipitate free zones at the
boundaries. In the HAZ, precipitate coarsening and
transformation to incoherent equilibrium g precipitates
means that dislocations are held up, leading to more
homogeneous deformation. Thus failure occurs by
conventional nucleation and growth of voids at coarse
intermetallics rather than at grain boundaries, and
consequently toughness is higher. In the weld nugget,
temperatures are sufficient to cause precipitate dissolution (Fig. 27) and subsequently natural aging on
cooling. The resulting fine dispersion of GunierPreston
zones and g9 precipitates leads to intergranular failure in
the nugget by the same mechanism as in the parent

82

International Materials Reviews

2009

VOL

54

NO

45 Comparison of R curves for cracks growing through


parent plate, weld centreline, weld nugget and tool
shoulder regions of 2195 friction stir weld, and for
VPPA weld centreline378

material. In fact, the situation is worsened by the


increase in grain boundary area and the coarsened grain
boundary particles mean failure is wholly intergranular.
As a consequence the HAZ is eight times tougher than
the parent and the nugget half as tough, as determined
by Kahn tear testing. These effects can be predicted for
the HAZ in terms of simple hardness383 models simply
by assuming toughness to be inversely proportional to
hardness.
In summary, all the results confirm that microstructural factors play a determining role in fracture
toughness. Typically the FSW nugget zone shows a
higher toughness than the parent alloys. This is in
contrast to the fatigue behaviour where residual stresses
dominate, presumably because the plastic strain washes
out any stored residual stresses. It is noteworthy that as
well as the expected correlation between increasing
hardness and strength and decreasing fracture toughness
across FSW in three alloys systems, Dawes has also
found remarkably good correlation between Charpy
data and more sophisticated fracture mechanics J
integral test methods374 (Fig. 46).

Corrosion
It is well established that microstructure is an important
factor in determining the corrosion behaviour of
aluminium alloys.384 A great deal of attention has been
focused on the Cu containing 2xxx and 7xxx series
(e.g. 2024,48,385390 7010, 7050, 7075),385,388,391393 which
show that the nugget becomes sensitised (Table 4). The
severe themomechanical processing refines the grain
structure and alters the precipitation distribution and
chemistry, particularly near the grain boundaries
(Tables 4 and 5). The relationship between microstructure and corrosion for 2024 is summarised in the form
of a timetemperaturecorrosion map384,410 in Fig. 47.
From this it is clear that the maximum thermal
excursion in the FSW weld region is above the knee in
the timetemperaturecorrosion curve and the cooling
rate sufficiently slow for pitting and intergranular
corrosion to occur.

Threadgill et al.

46 Charpy energytoughness correlations for friction stir


welds in various alloys381

Localised corrosion

Pitting, the removal of metal at localised sites, is a


common corrosion mechanism in FSW Al alloys. The
tendency for pitting is characterised by the pitting
potential, namely, the potential at which the protective
anodic film breaks down. Connolly et al.395 and Paglia
et al.407 have used micro-electronic cells to map this
spatially. The former team used a micro-electrochemical
test set-up (droplet cell) comprising a three electrode cell
within a fine (10 mm to 1 mm) pipette mounted on an
optical microscope for precise positioning to delineate
the variation in pitting potential as a function of
position across the microstructural zones. For 5456 the
breakdown potential for parent, HAZ and TMAZ were
similar; however, the nugget was significantly lower in
accordance with the higher level of corrosion attack
found there. This occurs due to preferential attack of

Friction stir welding of aluminium alloys

47 Timetemperaturecorrosion plot for 2024 friction stir


weld based on interrupted quench experiments, showing predominant corrosion mechanisms determined
on basis of accelerated corrosion tests on sheet:410
timetemperature band added by Lumsden384 to indicate conditions typical of most sensitised FSW HAZ
region

b-phase precipitates. Interestingly the advancing side of


the nugget was found to corrode preferentially. For the
2024 welds discussed above the pitting potential
(Fig. 48) was found to vary with increasing heat
input,396 again in accordance with Fig. 47. At low heat
inputs the whole TMAZ is most affected, but with
increasing heat input the HAZ becomes significantly
more affected than either the parent or nugget. Gel
visualisation comprising agar, NaCl and universal
indicator can also be used to highlight oxygen reduction
(blue/green) and hydrolysis of metal ions (yellow).395 It
can be seen in Fig. 49 that this method is in good
agreement with the pitting potential results, showing

Table 4 Summary of corrosion observations for FSW Al alloys


Alloy Corroding zone

Mechanism

Test

Ref.

2024 TMAZ
Nugget
Nugget
HAZ/parent
HAZ
Nugget and HAZ

Exfoliation
Pitting/blistering
Intergranular/pitting (150 mm)
Pitting to 150 mm
Intergranular
Intergranular attack (low rotation spends
in nugget/high speeds predominantly HAZ))
Passive Pitting

ASTM G34

48

Immersion (NaClzH2O2)

48

Immersion (NaCl)
Gel visualisation and Immersion
(NaCLzH2O2)
Poarisation curves and ellectrochemical
impedance spectroscopy in NaCl
Immersion (NaCl)

385, 394
395, 396

Immersion (NaClzH2O2)
Immersion (NaCl)

116
386, 398

EXCO (KCl, KNO3, HNO3)


Immersion (phosphoric acid)
Immersion (NaClz acetic acid)
Immersion (NaCl)
Immersion (NaCl)
Immersion tests in NaCl solution

400
395
401
402
385
213, 392, 403,
404, 405
391
392, 404, 406
407
408
409
401

Nugget/HAZ Parent
2219 Parent
2139 Parent
2195 No preference
5083
5456
6082
6013
7010
7050

Parent
Nugget
Parent
HAZ
Nugget, TMAZ/HAZ

Slight pitting at bottom of nugget, overall


better than parent
Pitting z intergranular
Even pitting across sample or weld better
than parent
Welds show lower pitting tendency
Preferential to advancing side
Even pitting across sample
Intergranular corrosion
Intergranular corrosion
Intergranular corrosion

7075 HAZ
Intergranular
HAZ
HAZ
TMAZ/HAZ boundary
7108 TMAZ
Localised intergranular corrosion
7150 TMAZ

Immersion (NaClKNO3HNO3)
Immersion
Pitting potential cell
Salt spray
ASTM G34 EXCO
Immersion

International Materials Reviews

2009

397
398, 399

VOL

54

NO

83

Threadgill et al.

Friction stir welding of aluminium alloys

48 Pitting breakdown potential measured using droplet


cell with NaCl across ve 2024 FSWs described in
Fig. 49396

increased corrosion in the HAZ as the heat input is


increased, the nugget becoming less affected.
There is a general consensus that for copper containing alloys (e.g. 2024, 7xxx) intergranular attack (see
Table 4) is encouraged by the precipitation of a network

of Cu rich intermetallics at grain boundaries (Fig. 50b),


which lead to locally depleted zones that are vulnerable
to localised attack,391,394,395,411 This vulnerability has
been linked to the lowering of the pitting potential
(Fig. 48). For 7xxx and 2xxx alloys intergranular
corrosion is thus generally preferentially observed in
the TMAZ/nugget and HAZ depending on the heat
input (Table 4). Figure 50a shows a scanning electron
micrograph of a 2024 weld following immersion in Exco
test solution for 8 h, revealing a severely corroded,
narrow band in the HAZ. On the right side of the
corrosion band, i.e. in the parent alloy, little development of corrosion is observed. Conversely, on the left
side of the corrosion band, i.e. within the TMAZ, a
number of randomly distributed pits, of dimension 50
300 mm had developed. Because 7075 contains less Cu
than 7050, the enrichment of Mg and depletion of Cu is
less, leading to lower susceptibility.412 For other alloy
systems, for example 2195, 2219, 5983, 5456, 6061, 6081,
the proclivity for the weld region to corrode is much
reduced being more similar to that of the parent

49 Macrographs (left) and corresponding gel visualisations (right) of corrosion attack on cross-sections of 2024 friction
stir welds after immersion in 0?1M NaCl for 24 h: welds were produced at different speeds and are arranged from top
to bottom in order of decreasing heat input396
Table 5 Typical relationship between microstructure, signicant localised corrosion mechanisms and main investigation
techniques for high strength aluminium alloy friction stir welds414
Zone

Microstructure

Form of localised corrosion Investigation technique

Parent

Strengthening ppts

Pitting

HAZ

Small grain boundary phases


Narrow ppt free zones
Coarse intragranular ppts

Intergranular corrosion
Other corrosion phenomena
Intergranular corrosion

Coarse grain boundary phases


Wide ppt free zones

Intersubgranular corrosion
Pitting

TMAZ

Intergranular corrosion
Intersubgranular corrosion
Pitting
Nugget General absence of intragranular ppts Pitting
and ppt free zones
Intergranular corrosion
For some alloys, presence of
intragranular ppts and ppt free zones

84

Variable dimensions of the ppts and


grain boundary phases

International Materials Reviews

2009

VOL

54

NO

Polarisation techniques, spray tests, droplet


cell methods, other methods

Corrosion immersion tests: appropriate for


the investigation of intergranular corrosion if
combined with microstructure investigations
(optical or SEM in back scattered mode)
Polarisation techniques: not appropriate for a
clear discrimination of the extent of the
intergranular corrosion
As for HAZ

Immersion, polarisation tests with microstructural


investigations (optical or SEM in back scattered
mode)
Conventional methods appropriate for pitting
corrosion

Threadgill et al.

Friction stir welding of aluminium alloys

a general view following corrosion testing in Exco solution for 8 h; b network of CuMgAl2; c close-up of intergranular
corrosion
50 Scanning electron micrographs of intergranular attack on 2024 friction stir weld394

(Table 4). Hu and Melekis398 ascribe the better pitting


resistance of 2219 compared with 2195 to the lower Cu
concentration. Generally the welded zones for 2219 and
2195 show more non-uniform pitting, but better corrosion behaviour than the parent alloys.
Exfoliation occurs where corrosion products having a
larger volume than the metal they consume produce a
wedge-like action.384 It is not common in the weld
nugget because large flat grains are much more susceptible than the fine grains typical of FSW. Exfoliation
corrosion is usually tested by exposure to an oxidising
acidic chloride (Exco) solution using ASTM G34.
In summary, corrosion of FSWs in 2xxx and 7xxx
alloys continues to be a challenge; those interested are
directed to more comprehensive reviews of the subject.413,414 It should be borne in mind that FSWs will
naturally be associated with residual stresses; these may
well play a role in corrosion studies. Finally, some work
has been undertaken on the corrosion performance of
dissimilar systems.388,415418 In such cases there is clearly
an opportunity for galvanic corrosion in the nugget, for
example 2024 is anodic with respect to 7010.388
Stress corrosion

The location for failure by stress corrosion cracking or


corrosion fatigue is often the HAZ of 7075 and 2049
alloys or the TMAZ/HAZ for 7050 and 5454 (Table 4).
If the weld is not particularly susceptible, as for 2195
and 2219 (Table 4), failure typically occurs in the softest
region of the weld. Intergranular failures are usually
associated with the sensitisation of the microstructure.
Increased pitting corrosion may act as an initial stage for
intergranular corrosion.414 In constant strain rate tests,
FSWs in 7075 have been found to exhibit much better
environmental cracking resistance than 7050,404,419 failure occurring in the HAZ in accordance with the
increased susceptibility there (Table 4).
Strategies to reduce corrosion susceptibility

Short term post-weld heat treatments (artificial aging)


(PWHT/PWAA), with thermal exposures similar to that

during welding, may be an efficient way to rehomogenise


the sensitised microstructure and thereby increase the
corrosion resistance of the welds.414 Such treatments may
also reduce the level of residual stresses, thereby lowering
susceptibility to stress corrosion. Lumsden et al.419 found
that 7050 is very sensitive to stress corrosion, exhibiting a
ductility just 13% of the in air ductility when slow strain
rate tested (according to ASTM G129)384 in NaCl
solution. Aging for one week at 100uC (equivalent to 10
years natural aging) restored the ductility to 80% of the in
air value. Alloy 7075 was found to be less sensitive to SCC
(73% of strain to failure in air), with a T73 post-weld
temper restoring this to 85%. Widener et al.420 found
joining 7075 material originally in the T73 condition
followed by PWAA to be preferable (in terms of higher
tensile and yield strengths and better exfoliation corrosion resistance) to welding in the T6 temper condition
followed by aging to T73. Retrogression and re-aging
treatments were not found to improve joint properties of
T6 material due to the severity of the overaging in the
HAZ caused by the welding process. Merati et al.,421
however, did have success with a localised retrogression
and re-aging treatment to 7475T73. Overall, 4 h at
190uC was found both to stabilise the microstructure and
to enhance the corrosion resistance for 7075, with only a
slight reduction in tensile strength and the added
advantage of annealing out residual stresses.420 Less
precise treatments have been applied using local heating
of joints with torch flames in 7075 and 2219.414
The effect of starting temper of the parent material
and PWHTs were investigated for FSW joints in
2024.422 It was found that the exfoliation resistance of
FSW Al 2024 joints may be restored through PWHT to
the T81 temper or when initially welded in the T81
temper, followed by naturally aging.
Other approaches to reduce the sensitivity to corrosion include modifications to the tool design; for
example, by using a tool with a scroll shoulder instead
of a threaded pin/flat shoulder tool when welding 7050,
the sensitivity to SCC can be eliminated.423 Other

International Materials Reviews

2009

VOL

54

NO

85

Threadgill et al.

Friction stir welding of aluminium alloys

studies have looked at the minimisation of corrosion


through optimisation of the welding procedure for 6xxx
alloys.424
The effect of chromate, molybdate and cerium nitrate
inhibitor additions to NaCl solution were examined for
dissimilar 7075/6056 FSW joints.425 The chromates were
better in terms of inhibition efficiency and inhibit all
regions of the weld area to a similar extent. Though less
effective, from an environmental viewpoint, molybdate
and cerate may offer advantages.
Approximately 30 mm micro-arc oxidation coatings
have been successfully applied to 2219 and 7018 FSW
joints subjected to salt spray corrosion.426
Other surface treatments shown to be successful
include low plasticity burnishing to retard corrosion
fatigue of 2219 FSW joints.372 An 80% increase in
corrosion fatigue performance has been achieved for the
same system by LPB,322,427 completely mitigating pitting
corrosion damage, with comparable fatigue perfrmance
regardless of salt fog exposure. Laser surface melting has
also been found to be effective in reducing corrosion of
2024.395

Comparison with other joining processes


A reasonable volume of data (e.g. Table 3) exists to
compare the properties of friction stir welds with other
processes. Consequently, it is now well established that
the mechanical properties associated with FSW joints
are generally better than those obtainable in arc welds.
However, mechanical properties are only part of the
picture: process economics are also of significance, as is
the quality of the weld which can be reliably obtained.
For example, Hori et al.330 have described work on a
Japanese alloy (6N01) similar to 6005. Comparison of
tensile properties with MIG and laser welding showed
no improvement in tensile properties over laser or MIG
(in fact FSW performed slightly worse), but a significant
improvement in fatigue performance was reported.
Furthermore, the tolerance on misfit, the remarkably
low weld to weld variability and freedom from defects
made the FSW process more attractive than its
competition. A further example can be found in a recent
study by Gesto et al.,84 who compared FSW and
GMAW of 6082T6 for a marine application. They
concluded that FSW is superior in terms of properties
and weld quality, but at the price of higher capital costs.
In summary, in many cases improved mechanical
properties are only part of a complex decision making
process for selecting FSW. For example, the low defect
incidence and low repair rates have been instrumental in
process selection, especially for critical applications such
as rocket fuel tanks.90,91 The reverse argument is also
true. In a number of cases companies have judged the
potential advantages of the process in terms of mechanical properties and weld quality to be outweighed by the
capital costs and other process conversion costs,
particularly in situations where an expensive machine
could not be kept fully occupied, or in areas where low
cost is the highest priority.

Concluding remarks
Friction stir welding of aluminium is now a mature and
robust process, which is becoming increasingly well
established in the fabrication of critical components. It is
true to say that FSW has extended the use of welding in

86

International Materials Reviews

2009

VOL

54

NO

certain materials and applications, in particular in the


welding of 2xxx and 7xxx alloys for the aerospace
industry. The qualities making the process attractive
include reduced cost, minimal repair requirement, good
properties and total automation leading to a high level
of consistency.
At the present time, FSW can compete with other
welding processes for quality of welds and performance.
It should be noted that FSW is still relatively new, and
has been in commercial production for less than 15
years. Nevertheless, progress has been significant, and
further improvements and developments can be
expected.
As with fusion welding, FSW is basically a thermal
process. Temperatures reached (typically around 500uC)
are sufficient to cause major microstructural changes in
precipitation hardened or work hardened alloys. Unlike
fusion processes, FSW also involves extremely high
shear strains and strain rates, which will have a
profound influence on the development of microstructures. The debate on the relative importance of recovery
and recrystallisation continues to be lively, and further
work is required in this area to gain a full understanding
of the complex processes and their interactions which
determine microstructures.
Friction stir welding is already one of the most energy
efficient processes available, although improved process
developments (in particular better tool designs) will no
doubt further reduce the energy required to make the
weld. Now industry has a welding process that can
provide high quality and defect free welds in the high
strength 2xxx and 7xxx alloys, development of improved
alloys which can be welded without loss of properties is
required, and this is a major challenge for the aluminium
producers.
Although the process is asymmetrical in terms of
material movement and heat generation, there is limited
evidence that this has an adverse effect on the properties
of the weld. Aluminium is an excellent conductor of
heat, so thermal gradients will rapidly disappear. The
significant heat transport in the rapidly moving material
will again help minimise differences. Although many
welds will have slight differences in properties on each
side, this is seldom significant, and probably influenced
as much by the thermodynamics and kinetics of the
alloys as by the asymmetry of the process.
In a similar vein, there was considerable interest in the
early days of FSW in the so called onion rings in the
weld. Several suggestions have been made regarding
the origin of these, and evidence has been found of
changes in grain size, texture, precipitation density,
composition, etc., the exact mechanism varying from
alloy to alloy. To date, little evidence has emerged to
suggest they are harmful, and indeed their impact on
weld performance seems to be very small.
The present review has demonstrated the extensive
research effort that continues to progress the understanding of FSW of aluminium alloys and its influence
on their microstructure and properties. It identifies a
number of areas that are worthwhile for further study.
From an engineering perspective, there is a need to
investigate the occurrence and significance of flaws in
friction stir welds. In particular, the influence of tool
design on flaw occurrence and the development of nondestructive testing techniques to identify flaws in both

Threadgill et al.

lap and butt welds would be beneficial. Metal flow


modelling may have a role to play here, though
capturing this aspect of the thermomechanical behaviour remains a significant challenge. Furthermore, the
development of NDT techniques capable of mapping
and locating the defects that arise in FSW is an area
requiring further work if they are to be utilised in safety
critical applications.
The review has attempted to outline the process
microstructureproperty relationships that need to be
considered when friction stir welding aluminium alloys.
While it is clear that FSW of aluminium alloys has
reached a level of maturity that enables commercial
exploitation of the technique, much remains to be done
before it is possible intelligently to optimise the process
to tailor the microstructure and residual stress for
particular service environments, whether to provide
improved tensile, fatigue, creep or corrosion resistance.
Development of greater understanding of friction stir
processing on sheet forming capabilities would also be of
interest.
Modelling of heat generation and thermal history is
reasonably mature, particularly if the heat input is
measured independently from the machine itself.
Microstructure evolution in both heat-treatable and
non-heat-treatable alloys has been modelled at various
levels of complexity, enabling prediction of hardness
profiles after a degree of calibration to specific alloys
and tempers. Some progress has been made to package
these models for industrial use by the non-expert, and
for predicting more difficult properties such as toughness. Future microstructure modelling challenges
include the ability to consider FS weldability in alloy
development programmes, and improved understanding
of dissimilar alloy FSW, particularly for joining
aluminium to other alloys such as steel for automotive
applications.
Comparisons with other processes are of course
inevitable. The main competitors to FSW are MIG
and TIG welding, and laser welding. In terms of heat
input, FSW typically introduces more heat per unit
length than laser welding, but less than MIG or TIG.
The fact that FSW is basically a machine tool process,
closely related to milling, gives it an advantage of full
automation, but at the cost of the flexibility which can
be achieved with MIG and TIG. The absence of a filler
in FSW is an advantage in many cases, but can also be a
disadvantage since it prevents the process being used for
fillet welds. The absence of filler also requires a higher
standard of preparation in terms of weld gap and fit-up
than fusion processes. Friction stir welding has a much
greater range of thickness than any fusion process (with
the exception of electron beam welding). The process has
been used to make single pass welds in thicknesses from
about 0?5 to 100 mm. Static mechanical properties in
aluminium alloys are generally related to the heat input,
and so FSW generally equals or exceeds the performance of MIG welds. Dynamic properties are generally
better than fusion welds, irrespective of whether the weld
surfaces are dressed. Finally, as FSW is a solid state
process, it can cope with any aluminium alloy, whereas
many high strength aluminium alloys are challenging or
impossible to weld by fusion processes. Friction stir
welding therefore joins the armoury of welding processes, but it will not displace other established processes

Friction stir welding of aluminium alloys

in all applications. It has, however, succeeded in making


the welding of high strength alloys a reality.
In the future, it seems likely that FSW will continue to
displace MIG in applications requiring long, essentially
straight welds and is well suited to the joining of
aluminium extrusions and panels. It is expected to
encroach more gradually on MIG for lower volume and
more complex joints. However, MIG is likely to be
retained in cases requiring a filler or where a manual/
portable process is preferred. The barriers to growth of
FSW for Al are primarily high cost of bespoke
equipment (almost all machines built so far have been
one-off designs (e.g. Fig. 2a), but low cost milling
machines can be modified), under utilisation of expensive equipment, licence fees, lack of familiarity with the
process by customers, plus poor availability of skilled
practitioners with industrial experience, regulatory
authorities, etc. All these issues are gradually being
eroded, making FSW increasingly attractive. In addition
FSW is becoming indispensible where it is necessary to
join different aluminium alloys. Finally, it is worth
pointing to the steeply increasing volume of work
looking at FSW of a wide range of systems beyond
aluminium alloys. This suggests that commercial applications of FSW are set to expand widely in coming
years.

Acknowledgements
The authors are deeply indebted to many colleagues
within the global FSW community for numerous direct
and indirect contributions to this work, for making
available micrographs and figures and for many useful
discussions over the past few years. PLT and AJL were
afforded time by TWI to write this review. PJW is
grateful to Dr Altenkirch and Dr Steuwer for assistance
in providing some of the data presented.

References
1. W. M. Thomas, E. D. Nicholas, J. C. Needham, M. G. Murch,
P. Temple-Smith and C. J. Dawes: Friction stir butt welding, GB
patent no. 9125978?8, 1991.
2. W. M. Thomas, E. D. Nicholas, J. C. Needham, M. G. Murch,
P. Temple-Smith and C. J. Dawes: Improvements relating to
friction welding, US patent no. 5 460 317; EPS 0 616 490, 1991.
3. I. J. Smith and D. D. R. Lord: Proc. 7th Int. Symp. on Friction
stir welding, Awaji Island, Japan, May 2008, TWI, Paper
no. 2007-01-1707.
4. P. L. Threadgill: Sci. Technol. Weld. Join., 2007, 12, 357360.
5. W. Tang, X. Guo, J. C. McClure and L. E. Murr: J. Mater.
Process. Manuf. Sci., 1999, 7, 163172.
6. M. W. Mahoney, C. G. Rhodes, J. G. Flintoff, R. A. Spurling and
W. H. Bingel: Metall. Mater. Trans. A, 1998, 29A, 19551964.
7. A. P. Reynolds, W. D. Lockwood and T. U. Seidel: Mater. Sci.
Forum, 2000, 331337, 17191724.
8. L. E. Murr, G. Liu and J. C. McClure: J. Mater. Sci., 1998, 33,
12431251.
9. G. J. Bendszak, T. H. North and C. B. Smith: Proc. 2nd Int.
Symp. on Friction stir welding, Gothenburg, Sweden, June 2000,
TWI.
10. K. A. A. Hassan, P. B. Prangnell, A. F. Norman, D. A. Price and
S. W. Williams: Sci. Technol. Weld. Join., 2003, 8, 257268.
11. Y. K. Yang, H. Dong and S. Kou: Weld. J., 2008, 87, 202s211s
12. A. Gerlich, M. Yamamoto and T. H. North: Sci. Technol. Weld.
Join., 2007, 12, 472480.
13. J. H. Yan, M. A. Sutton and A. P. Reynolds: Sci. Technol. Weld.
Join., 2005, 10, 725736.
14. P. A. Colegrove, H. R. Shercliff and R. Zettler: Sci. Technol.
Weld. Join., 2007, 12, 284297.

International Materials Reviews

2009

VOL

54

NO

87

Threadgill et al.

Friction stir welding of aluminium alloys

15. C.-G. Andersson, R. E. Andrews, B. G. I. Dance, M. J. Russell,


E. J. Olden and R. M. Sanderson: Proc. 2nd Int. Symp. on
Friction stir welding, Gothenburg, Sweden, June 2000, TWI.
16. L. Cedeqvist and R. E. Andrews: Proc. 4th Int. Symp. on Friction
stir welding, Park City, UT, USA, May 2003, TWI, 14001430.
17. K. Savolainen, J. Mononen, T. Saukkonen, H. Hanninen and
J. Koivula: Proc. 5th Int. Symp. on Friction stir welding, Metz,
France, September 2004, TWI, 19.
18. W. B. Lee and S. B. Jung: Mater. Lett., 2004, 58, 10411046.
19. H. S. Park, T. Kimura, T. Murakami, Y. Nagano, K. Nakata and
M. Ushio: Mater. Sci. Eng. A, 2004, A371, 160169.
20. T. Sakthivel and J. Mukhopadhyay: J. Mater. Sci., 2007, 42,
81268129.
21. G. M. Xie, Z. Y. Ma and L. Geng: Scr. Mater., 2007, 57, 7376.
22. P. Volovitch, J.-E. Masse, T. Baudin, B. da Costa, J. C. Goussain,
W. Saikaly and L. Barrallier: Proc. 5th Int. Symp. on Friction stir
welding, Metz, France, September 2004, TWI.
23. J. A. Esparza, W. C. Davis, E. A. Trillo and L. E. Murr: Mater.
Sci. Lett., 2002, 21, 917920.
24. R. Johnson: Mater. Sci. Forum, 2003, 419422, 365370.
25. P. L. Threadgill and R. Johnson: Proc. Symp. on Magnesium
technology, San Diego, CA, USA, March 2003, TMS.
26. W. B. Lee, J. W. Kim, Y. M. Yeon and S. B. Jung: Mater. Trans.,
2003, 44, 917923.
27. S. H. C. Park, Y. S. Sato and H. Kokawa: Scr. Mater., 2003, 49,
161166.
28. S. H. C. Park, Y. S. Sato, H. Kokawa and T. Tsukeda: in Trends
in welding research, (ed. S. A. David et al.), 267272; 2003,
Materials Park, OH, ASM International.
29. N. Afrin, D. L. Chen, X. Cao and M. Jahazi: Mater. Sci. Eng. A,
2008, A472, 179186.
30. W. B. Lee, C. Y. Lee, W. S. Chang, Y. M. Yeon and S. B. Jung:
Mater. Lett., 2005, 59, 33153318.
31. A. J. Ramirez and M. C. Juhas: Mater. Sci. Forum, 2003, 426432,
29993004.
32. A. P. Reynolds, E. Hood and W. Tang: Scr. Mater., 2005, 52,
491494.
33. B. P. Wynne, P. S. Davies, M. J. Thomas, B. S. Ng and P. L.
Threadgill: Proc. 7th Int. Symp. on Friction stir welding, Awaji
Island, Japan, May 2008, TWI.
34. M. J. Russell, P. L. Threadgill, M. J. Thomas and B. P. Wynne:
Proc. 11th Int. Conf. on Titanium, Kyoto, Japan, June 2007,
Japan Institute of Metals, 10951098.
35. W. M. Thomas, P. L. Threadgill and E. D. Nicholas: Sci. Technol.
Weld. Join., 1999, 4, 365372.
36. R. Johnson and P. L. Threadgill: Proc. 6th Int. Conf. on Trends
in welding research, Pine Mountain, GA, USA, June 2002, TWI.
37. T. J. Lienert, W. L. Stellwag, B. B. Grimmett and R. W. Warke:
Weld. J., 2003, 82, 1S9S.
38. S. H. C. Park, Y. S. Sato, H. Kokawa, K. Okamoto, S. Hirano
and M. Inagaki: Scr. Mater., 2003, 49, 11751180.
39. A. P. Reynolds, W. Tang, T. GnaupelHerold and H. Prask: Scr.
Mater., 2003, 48, 12891294.
40. A. P. Reynolds, W. Tang, M. Posada and J. DeLoach: Sci.
Technol. Weld. Join., 2003, 8, 455460.
41. H. Fujii, L. Cui, N. Tsuji, M. Maeda, K. Nakata and K. Nogi:
Mater. Sci. Eng. A, 2006, A429, 5057.
42. L. Cui, H. Fujii, N. Tsuji and K. Nogi: Scr. Mater., 2007, 56, 637
640.
43. S. J. Barnes, A. Steuwer, S. Mahawish, R. Johnson and P. J.
Withers: Mater. Sci. Eng. A, 2008, A492, 3544.
44. H. J. Jun and R. Ayer: Proc. 6th Int. Symp. on Friction stir
welding, Saint-Sauveur, Montreal, Canada, October 2006, TWI.
45. H. J. Jun, R. Ayer, T. Neeraj and R. Steel: Mater. Sci. Forum,
2007, 539543, 37633768.
46. F. X. Ye, H. Fujii, T. Tsumura and K. Nakata: J. Mater. Sci.,
2006, 41, 53765379.
47. H. Fujii, H. Kato, K. Nakata and K. Nogi: Proc. 6th Int. Symp.
on Friction stir welding, Saint-Sauveur, Montreal, Canada,
October 2006, TWI.
48. G. Biallas, R. Braun, C. Dalle Donne, G. Staniek and W. A.
Kaysser: Proc. Conf. 1st Int. Symp. on ;Friction stir welding;,
Thousand Oaks, CA, USA, June 1999, TWI.
49. H. Larsson, L. Karlsson, S. Stoltz and E.-L. Bergqvist: Proc. 2nd
Int. Symp. on Friction stir welding, Gothenburg, Sweden, June
2000, TWI.
50. S. Tanaka and M. Kumagai: Proc. 3rd Int. Symp. on Friction stir
welding, Kobe, Japan, September 2001, TWI.
51. H. R. Shercliff, M. J. Russell, A. D. Taylor and T. L. Dickerson:
Mecan. Indust., 2005, 6, 2535.

88

International Materials Reviews

2009

VOL

54

NO

52. M. J. Peel, A. Steuwer and P. J. Withers: Metall. Mater. Trans. A,


2006, 37A, 21952206.
53. M. J. Peel, A. Steuwer, P. J. Withers, T. Dickerson, Q. Shi and
H. Shercliff: Metall. Mater. Trans. A, 2006, 37A, 21832193.
54. A. Steuwer, M. J. Peel and P. J. Withers: Mater. Sci. Eng. A, 2006,
A441, 187196.
55. F. Palm: Proc. Conf. Materials Week 99, Munich, Germany,
September 1999, DGM.
56. W. B. Lee, Y. M. Yeon and S. B. Jung: Scr. Mater., 2003, 49, 423
428.
57. Y. Li, L. E. Murr and J. C. McClure: Mater. Sci. Eng. A, 1999,
A271, 213223.
58. Y. Li, L. E. Murr and J. C. McClure: Scr. Mater., 1999, 40, 1041
1046.
59. S. Lim, S. Kim and C. G. Lee: Metall. Mater. Trans. A, 2004, 35A,
28372843.
60. O. T. Midling: Proc. 4th Int. Conf. on Aluminium alloys,
Atlanta, GA, USA, September 1994, Georgia Institute of
Technology, 451458.
61. J. H. Ouyang and R. Kovacevic: J. Mater. Eng. Perform., 2002,
11, 5163.
62. A. Gerlich, P. Su and T. H. North: Sci. Technol. Weld. Join., 2005,
10, 647652.
63. S. A. Khodir and T. Shibayanagi: Mater. Trans., 2007, 48, 2501
2505.
64. Y. S. Sato, S. H. C. Park, M. Michiuchi and H. Kokawa: Scr.
Mater., 2004, 50, 12331236.
65. A. C. Somasekharan and L. E. Murr: Mater. Charact., 2004, 52,
4964.
66. J. A. Wert: Scr. Mater., 2003, 49, 607612.
67. C. M. Chen and R. Kovacevic: Int. J. Mach. Tools Manuf., 2004,
44, 12051214.
68. K. Kimapong and T. Watanabe: Weld. J., 2004, 83, 277S282S.
69. H. Uzun, C. D. Donne, A. Argagnotto, T. Ghidini and
C. Gambaro: Mater. Design, 2005, 26, 4146.
70. W. B. Lee and S. B. Jung: Mater. Res. Innov., 2004, 8, 9396.
71. L. E. Murr, Y. Li, R. D. Flores, E. A. Trillo and J. C. McClure:
Mater. Res. Innov., 1998, 2, 150163.
72. L. E. Murr, R. D. Flores, O. V. Flores, J. C. McClure, G. Liu and
D. Brown: Mater. Res. Innov., 1998, 1, 211223.
73. L. E. Murr, Y. Li, E. A. Trillo, R. D. Flores and J. C. McClure:
J. Mater. Process. Manuf. Sci., 1999, 7, 145161.
74. R. S. Mishra and Z. Y. Ma: Mater. Sci. Eng. R, 2005, R50, 178.
75. R. Nandan, T. DebRoy and H. K. D. H. Bhadeshia: Prog. Mater.
Sci., 2008, 53, 9801023.
76. R. S. Mishra and M. W. Mahoney (eds.): in ASM specialty
handbook: friction stir welding and processing; 2007, Materials
Park, OH, ASM International.
77. H. R. Shercliff and P. A. Colegrove: in Mathematical modelling
of weld phenomena 6, (ed. H. Cerjak), 927974; 2002, London,
The Institute of Metals.
78. H. R. Shercliff and P. A. Colegrove: in Friction stir welding and
processing, (ed. R. S. Mishra et al.), 187217; 2007, Materials
Park, OH, ASM International.
79. D. A. Price, S. W. Williams, A. Wescott, C. J. C. Harrison,
A. Rezai, A. Steuwer, M. Peel, P. Staron and M. Kocak: Sci.
Technol. Weld. Join., 2007, 12, 620633.
80. M. M. Z. Ahmed, B. P. Wynne, W. M. Rainforth and P. L.
Threadgill: Scr. Mater., 2008, 59, 507510.
81. P. L. Threadgill and M. E. Nunn: A review of friction stir
welding: part 1 process overview , TWI members report no. 760/
2003, TWI, Abington, UK, 2003.
82. Y. Uematsu, K. Tokaji, Y. Tozaki, T. Kurita and S. Murata: Int.
J. Fatig., 2008, 30, 19561966.
83. O. T. Midling, J. S. Kvale and O. Dahl: Proc. 1st Int. Symp. on
Friction stir welding, Thousand Oaks, CA, USA, June 1999,
TWI.
84. D. Gesto, V. Pintos, J. Vazquez, J. Rasilla and S. Barreras: Proc.
7th Int. Symp. on Friction stir welding, Awaji Island, Japan,
May 2008, TWI.
85. Super high-speed passenger cargo TSL successfully completed its
sea trial, http://www.mes.co.jp/english/press/2005/20051014.html
2005 (accessed September 2008).
86. J. Ray: Delta 4 fleet goes from: Medium to Heavy, in
Spaceflight Now, 2002.
87. M. R. Johnsen: Weld. J., 1999, 78, 3539.
88. D. J. Waldron and R. W. Roberts: Proc. Conf. on Aerospace
automated fastening, Long Beach, CA, USA, September 1998,
SAE, 1517.

Threadgill et al.

89. J. Ding, R. Carter, K. Lawless, A. Nunes, C. Russell, M. Suites


and J. Schneider. A decade of friction stir welding R and D at
NASAs Marshall Space Flight Center and a glance into the
future,
http://ntrs.nasa.gov/archive/nasa/casi.ntrs.nasa.gov/
20080009619_2008009118.pdf 2005 (accessed September 2008).
90. C. Dawes: Proc. AIAA Int. Air and Space Symp., Dayton, OH,
USA, July 2003, AIAA, AIAA20032769.
91. Z. S. Loftus, W. J. Arbegast and P. J. Hartley: Proc. 5th Int. Conf.
on Trends in welding research, Pine Mountain, GA, USA, June
1998, ASM International, 580.
92. B. Christner, J. McCoury and S. Higgins: Proc. 4th Int. Symp. on
Friction stir welding, Park City, UT, USA, May 2003, TWI.
93. T. Kawasaki, T. Makino, S. Todori, H. Takai, M. Ezumi and
Y. Ina: Proc. 2nd Int. Symp. on Friction stir welding,
Gothenburg, Sweden, June 2000, TWI.
94. D. Otsuka and Y. Sakai: Proc. 7th Int. Symp. on Friction stir
welding, Awaji Island, Japan, May 2008, TWI.
95. M. M. Shahri and R. Sandstrom: Proc. 7th Int. Symp. on
Friction stir welding, Awaji Island, Japan, May 2008, TWI.
96. Anon: Mach. Design, 2003, 75, S2.
97. J. C. Bassett and S. S. Birley: Proc 2nd Int. Symp. on Friction stir
welding, Gothenburg, Sweden, June 2000, TWI.
98. G. Campbell and T. Stotler: Weld. J., 1999, 78, 4547.
99. R. S. Mishra and M. W. Mahoney: in Superplasticity in advanced
materials: ICSAM2000, 507512; 2001, Zurich-Uetikon, Trans
Tech Publications Ltd.
100. J. Q. Su, T. W. Nelson and C. J. Sterling: Scr. Mater., 2005, 52,
135140.
101. R. S. Mishra, M. W. Mahoney, S. X. McFadden, N. A. Mara and
A. K. Mukherjee: Scr. Mater., 1999, 42, 163168.
102. P. B. Berbon, W. H. Bingel, R. S. Mishra, C. C. Bampton and
M. W. Mahoney: Scr. Mater., 2001, 44, 6166.
103. J. C. Bersaas, A. Oosterkamp and L. D. Oosterkamp: Friction
stir welding method and apparatus, Patent no. WO/2001/028732,
2001.
104. P. Su, A. Gerlich, T. H. North and G. J. Bendzsak: Sci. Technol.
Weld. Join., 2006, 11, 6171.
105. W. S. Chang, C. K. Chun, H. J. Kim, H. J. Cho and T. K. Kim:
Proc. IWJC-Korea 2007, Seoul, South Korea, May 2007,
Genicom Co., Ltd, 435438.
106. M. Mahoney, R. S. Mishra, T. Nelson, J. Flintoff, R. Islamgaliev
and Y. Hovansky: in Friction stir welding and processing, (ed.
K. V. Jata et al.), 183194; 2001, Warrendale, PA, TMS.
107. I. Charit, R. S. Mishra and M. W. Mahoney: Scr. Mater., 2002,
47, 631636.
108. I. Charit, R. S. Mishra and K. V. Jata: in Friction stir welding
and processing, (ed. K. V. Jata et al.), 225234; 2001,
Warrendale, PA, TMS.
109. I. Shigematsu, N. Saito, T. Komaya, T. Tamaki, G. Yamauchi
and M. Nakamura: in Friction stir welding and processing, (ed.
K. V. Jata et al.), 217224; 2001, Warrendale, PA, TMS.
110. J. Zheng, R. S. Mishra, P. B. Berbon and M. W. Mahoney: in
Friction stir welding and processing, (ed. K. V. Jata et al.), 235
242; 2001, Warrendale, PA, TMS.
111. T. U. Seidel and A. P. Reynolds: Sci. Technol. Weld. Join., 2003,
8, 175183.
112. P. A. Colegrove and H. R. Shercliff: Sci. Technol. Weld. Join.,
2004, 9, 483492.
113. A. P. Reynolds: Sci. Technol. Weld. Join., 2000, 5, 120124.
114. W. M. Thomas, E. D. Nicholas and S. D. Smith: in Aluminum
2001, (ed. S. K. Das et al.), 213224; 2001, Warrendale, PA,
TMS.
115. P. A. Colegrove and H. R. Shercliff: Sci. Technol. Weld. Join.,
2004, 9, 352361.
116. D. Allehaux and F. Marie: Mater. Sci. Forum, 2006, 519521,
11311138.
117. C. B. Fuller: in Friction stir welding and processing, (ed. R. S.
Mishra et al.), 735; 2007, Materials Park, OH, ASM
International.
118. L. Dubourg and P. Dacheux: Proc. 6th Int. Symp. on Friction stir
welding, Saint-Sauveur, Mont., Canada, October 2006, TMS.
119. A. Bastier, M. H. Maitournam, K. D. Van and F. Roger: Sci.
Technol. Weld. Join., 2006, 11, 278288.
120. K. Colligan: Weld. J., 1999, 78, 229S237S.
121. O. Frigaard, O. Grong and O. T. Midling: Metall. Mater. Trans.
A, 2001, 32A, 11891200.
122. M. J. Russell, H. R. Shercliff and P. L. Threadgill: in Aluminum
2001, (ed. S. K. Das et al.), 2001, Warrendale, PA, TMS.
123. Y. J. Chao, X. H. Qi and W. Tang: Trans. ASME J. Manuf. Sci.
Eng., 2003, 125, 138145.

Friction stir welding of aluminium alloys

124. M. Song and R. Kovacevic: Proc. Inst. Mech. Eng. B, 2003, 217B,
7385.
125. A. P. Reynolds, Z. Khandkar, T. Long, W. Tang and J. Khan:
Mater. Sci. Forum, 2003, 426432, 29592964.
126. H. Schmidt, J. Hattel and J. Wert: Model. Simul. Mater. Sci. Eng.,
2004, 12, 143157.
127. R. Kovacevic, V. Soundararajan and S. Zekovic: Int. J. Mach.
Tools Manuf., 2005, 45, 15771587.
128. Q.-Y. Shi, T. L. Dickerson and H. R. Shercliff: Proc. 4th Int.
Symp. on Friction stir welding, Park City, UT, USA, May 2003,
TWI.
129. A. Simar, T. Pardoen and B. Meester: Proc. 5th Int. Symp. on
Friction stir welding, Metz, France, September 2004, TWI.
130. P. A. Colegrove and H. R. Shercliff: Sci. Technol. Weld. Join.,
2006, 11, 429441.
131. V. I. Vill: Friction welding of metals; 1959, Mashgiz, Leningrad.
132. R. E. Andrews and K. A. Beamish: Characterisation of and
guidelines for rotary friction welding of common metallic
engineering materials, TWI members report no. 824, TWI,
Abington, UK, 2005.
133. Y. G. Kim, H. Fujii, T. Tsumura, T. Komazaki and K. Nakata:
Mater. Sci. Eng. A, 2006, A415, 250254.
134. P. L. Threadgill: TWI Bull., 1997, 28, 3033.
135. P. L. Threadgill and A. J. Leonard: Macro and microstructural
features of friction stir welds in various materials, TWI members
report no. 693/1999, TWI, Abington, UK, 1999.
136. Specification for friction stir welding of aluminum alloys for
aerospace applications, Standard D17?3:200X, American
Welding Society, Miami, FL, USA, 2006.
137. A. Gerlich, P. Su, M. Yamamoto and T. H. North: Sci. Technol.
Weld. Join., 2008, 13, 254264.
138. A. D. Gingell and T. G. Gooch: Review of factors influencing
porosity in aluminium arc welds, TWI members report no. 625/
1997, TWI, Abington, UK, 1997.
139. M. F. Gittos and M. H. Scott: TWI Bull., 1987, 28, 259263.
140. A. J. Leonard and S. A. Lockyer: Proc. 4th Int. Symp. on
Friction stir welding, Park City, UT, USA, May 2003, TWI.
141. T. Hashimoto, S. Jyogan, K. Nakata, Y. G. Kiu and M. Ushio:
Proc. 1st Int. Symp. on Friction stir welding, Thousand Oaks,
CA, USA, June 1999, TWI.
142. Y. S. Sato, M. Urata and H. Kokawa: Metall. Mater. Trans. A,
2002, 33A, 625635.
143. B. K. Christner and G. D. Sylva: Proc. Conf. ICAWT 96,
Columbus, OH, USA, November 1996, EWI, 359368.
144. S. K. Chimbli, D. J. Medlin and W. J. Arbegast: in Friction stir
welding and processing IV, (ed. R. S. Mishra et al.), 135142;
2007, Warrendale, PA, TMS.
145. V. Balasubramanian: Mater. Sci. Eng. A, 2008, A480, 397403.
146. K. Kumar, S. V. Kailas and T. S. Srivatsan: Mater. Manuf.
Process., 2008, 23, 189195.
147. S. T. Wei, C. Y. Hao and J. C. Chen: Mater. Sci. Eng. A, 2007,
A452, 170177.
148. Y. H. Zhao, S. B. Lin, L. Wu and F. X. Qu: Mater. Lett., 2005,
59, 29482952.
149. R. Crawford, G. E. Cook, A. M. Strauss, D. A. Hartman and
M. A. Stremler: Sci. Technol. Weld. Join., 2006, 11, 657665.
150. O. Lorrain, V. Favier, H. Zahrouni and M. E. Hadrouz: Proc. 7th
Int. Symp. on Friction stir welding, Awaji Island, Japan, 2008,
TWI.
151. H. Schmidt and J. Hattel: in Friction stir welding processing III,
(ed. K. V. Jata et al.), 225232; 2005, Warrendale, PA, TMS.
152. H. Schmidt and J. Hattel: Model. Simul. Mater. Sci. Eng., 2005,
13, 7793.
153. H. Jin, S. Saimoto, M. Ball and P. L. Threadgill: Mater. Sci.
Technol., 2001, 17, 16051614.
154. H. Okamura, K. Aota, M. Sakamoto, M. Ezumi and K. Ikeuchi:
Weld. Int., 2002, 16, 266275.
155. F. Palm, H. Steiger and U. Hennebohle: Proc. 4th Int. Symp. on
Friction stir welding, Park City, UT, USA, May 2003, TWI.
156. T. Jene, G. Dobmann, G. Wagner and D. Eifler: Proc. 6th Int.
Symp. on Friction stir welding, Saint-Sauveur, Montreal,
Canada, October 2006, TWI.
157. K. Savolainen, T. Saukkonen, J. Mononen and H. Hanninen:
Proc. 7th Int. Symp. on Friction stir welding, Awaji Island,
Japan, May 2008, TWI.
158. Y. S. Sato, F. Yamashita, Y. Sugiura, S. H. C. Park and
H. Kokawa: Scr. Mater., 2004, 50, 365369.
159. H. Okamura, K. Aota, M. Sakamoto, M. Ezumi and K. Ikeuchi:
J. Jpn Weld. Soc., 2001, 19, 446456.

International Materials Reviews

2009

VOL

54

NO

89

Threadgill et al.

Friction stir welding of aluminium alloys

160. T. L. Dickerson, H. R. Shercliff and H. Schmidt: Proc. 4th Int.


Symp. on Friction stir welding, Park City, UT, USA, May 2003,
TWI.
161. C. Z. Zhou, X. Q. Yang and G. H. Luan: Scr. Mater., 2006, 54,
15151520.
162. J. Pryzdatek: Proc 1st Int. Symp. on Friction stir welding,
Thousand Oaks, CA, USA, June 1999, TWI.
163. A. J. Leonard: Unpublished work, TWI Ltd, 2001.
164. C. J. Goodfellow and A. J. Leonard: Unpublished work, TWI and
Birmingham University, 2003.
165. R. Johnson: Further assessment of the friction stir welding of
magnesium alloys, TWI members report no. 766/2003, TWI,
Abington, UK, 2003.
166. L. Cederqvist and A. R. Reynolds: Weld. J., 2001, 80, 281S287S.
167. W. M. Thomas, D. G. Staines, I. M. Norris and R. de Frias: Proc.
FSW Semin., Porto, Portugal, December 2002, IST.
168. P. A. Colegrove, T. Hyoe and H. R. Shercliff: Proc. 5th Int. Symp.
on Friction stir welding, Metz, France, September 2004, TWI.
169. D. Levesque, C. Mandache, L. Dubourg and P. Gougeon: Proc.
7th Int. Symp. on Friction stir welding, Awaji Island, Japan,
May 2008, TWI.
170. C. R. Bird: Proc. 4th Int. Symp. on Friction stir welding, Park
City, Utah, USA, May 2003, TWI.
171. C. R. Bird: Proc. 5th Int. Symp. on Friction stir welding, Metz,
France, September 2004, TWI.
172. T. Vugrin, G. Staniek, W. Hillger and C. Dalle Donne: Proc. 5th
Int. Symp. on Friction stir welding, Metz, France, September
2004, TWI.
173. M. Moles and A. Lamarre: Proc. 4th Int. Symp. on Friction stir
welding, Park City, UT, USA, May 2003, TWI.
174. N. Goldfine, D. Grundy, V. Zilberstein and D. G. Kinchen: Proc.
6th Int. Conf. on Trends in welding research, Pine Mountain,
GA, April 2002, ASM International.
175. N. Oiwa, S. Iwaki, T. Okada, N. Eguchi, S. Tanaka and
K. Namba: Proc. Int. Conf. on Aluminium connections,
Cleveland, OH, USA, June 2004, Lincoln Electric Company.
176. A. Lamarre, O. Dupuis and M. Moles: Proc. WCNDT 2004,
Montreal, Canada, AugustSeptember 2004, Paper 84.
177. A. Steuwer, M. Dumont, J. Altenkirch, S. Birosca, A. Deschamps,
P. B. Prangnell and P. J. Withers: Friction stir welding AlLi
AA2199: I microstructural aspects, In preparation, 2008.
178. A. J. Leonard: Proc. 2nd Int. Symp. on Friction stir welding,
Gothenburg, Sweden, June 2000, TWI.
179. J. Karlsson, B. Karlsson, H. Larsson, L. Karlsson and L.-E.
Svensson: Proc. 7th Int. Conf. on Joints in aluminium,
Cambridge, UK, April 1998, TWI, 221230.
180. M. Peel, A. Steuwer, M. Preuss and P. J. Withers: Acta Mater.,
2003, 51, 47914801.
181. H. W. Hayden, S. A. David, S. S. Babu, S. Spooner, J. M. Vitek
and P. J. Hartley: Prof. Conf. 3rd Int. Forum on Aluminium
ships, Haugesund, Norway, May 1998.
182. J. Q. Su, T. W. Nelson, R. Mishra and M. Mahoney: Acta Mater.,
2003, 51, 713729.
183. A. F. Norman, I. Brough and P. B. Prangnell: Mater. Sci. Forum,
2000, 331333, 17131718.
184. M. M. Attallah, C. L. Davis and M. Strangwood: Sci. Technol.
Weld. Join., 2007, 12, 361369.
185. K. N. Krishnan: Mater. Sci. Eng. A, 2002, A327, 246251.
186. T. U. Seidel and A. P. Reynolds: Metall. Mater. Trans. A, 2001,
32A, 28792884.
187. M. Strangwood, J. E. Berry, D. P. Cleugh, A. J. Leonard and P. L.
Threadgill: Proc. 1st Int. Symp. on Friction stir welding,
Thousand Oaks, CA, USA, June 1999, TWI.
188. L.-E. Svensson and L. Karlsson: Proc. 1st Int. Symp. on Friction
stir welding, Thousand Oaks, CA, USA, June 1999, TWI.
189. T. J. Lienert and R. J. Grylls: Proc. 1st Int. Symp. on Friction stir
welding, Thousand Oaks, CA, USA, June 1999, TWI.
190. C. G. Rhodes, M. W. Mahoney, W. H. Bingel and M. Calabrese:
Scr. Mater., 2003, 48, 14511455.
191. K. V. Jata, K. K. Sankaran and J. J. Ruschau: Metall. Mater.
Trans. A, 2000, 31A, 21812192.
192. Y. S. Sato, H. Kokawa, M. Enomoto and S. Jogan: Metall.
Mater. Trans. A, 1999, 30A, 24292437.
193. K. A. A. Hassan, B. P. Wynne and P. B. Prangnell: Proc. 4th Int.
Symp. on Friction stir welding, Park City, UT, May 2003, TWI.
194. H. S. Yang: Proc. Conf. 6th Int. Conf. on Aluminium alloys,
Toyohashi, Japan, July 1998, The Japan Institute of Light Metals,
14831488.

90

International Materials Reviews

2009

VOL

54

NO

195. . Frigaard, . Grong, J. Hjelen, S. Gulbrandsen-Dahl and O. T.


Midling: Proc. 1st Int. Symp. on Friction stir welding, Thousand
Oaks, CA, USA, June 1999, TWI.
196. R. W. Fonda, J. F. Bingert and K. J. Colligan: Proc. 5th Int.
Symp. on Friction stir welding, Metz, France, September 2004,
TWI.
197. R. S. Mishra, R. K. Islamgaliev, T. W. Nelson, Y. Hovansky and
M. W. Mahoney: in Friction stir welding and processing, (ed.
K. V. Jata et al.), 205216; 2001, Warrendale, PA, TMS.
198. K. V. Jata and S. L. Semiatin: Scr. Mater., 2000, 43, 743749.
199. P. B. Prangnell and C. P. Heason: Acta Mater., 2005, 53, 3179
3192.
200. Y. S. Sato, H. Watanabe, S. H. C. Park and H. Kokawa: Proc. 5th
Int. Symp. on Friction stir welding, Metz, France, September
2004, TWI.
201. R. S. Mishra, R. K. Islamgaliev, T. W. Nelson, Y. Hovansky and
M. W. Mahoney: in Friction stir welding and processing, (ed.
K. V. Jata et al.), 205216; 2001, Warrendale, PA, TMS.
202. E. Litwinski: Proc. 3rd Int. Symp. on Friction stir welding,
Kobe, Japan, September 2001, TWI.
203. M. M. Attallah and H. G. Salem Hassan: Proc. 5th Int. Symp. on
Friction stir welding, Metz, France, September 2004, TWI.
204. Y. S. Sato and H. Kokawa: Metall. Mater. Trans. A, 2001, 32A,
30233031.
205. K. N. Krishnan: J. Mater. Sci., 2002, 37, 473480.
206. M. Karlsen, S. Tangen, J. Hjelen, . Frigaard and . Grong:
Proc. 3rd Int. Symp. on Friction stir welding, Kobe, Japan,
September 2001, TWI.
207. K. A. A. Hassan, A. F. Norman, D. A. Price and P. B. Prangnell:
Acta Mater., 2003, 51, 19231936.
208. F. J. Humphreys and M. Hatherly: Recrystallisation and related
annealing phenomena; 1995, Oxford, Pergamon.
209. F. J. Humphreys: Acta Mater., 1997, 45, 50315039.
210. F. J. Humphreys: Acta Mater., 1997, 45, 42314240.
211. M. F. Gittos and K. Bridges: A study of arc and friction stir
welding of two Al alloys containing a low level scandium
addition, TWI members report no. 776/2003, TWI, Abington,
UK, 2003.
212. B. Huneau, X. Sauvage, S. Marya and A. Poitou: in Friction stir
welding processing III, (ed. K. V. Jata et al.), 253260; 2005,
Warrendale, PA, TMS.
213. C. S. Paglia, K. V. Jata and R. G. Buchheit: Mater. Sci. Eng. A,
2006, A424, 196204.
214. Y. S. Sato, M. Urata, H. Kokawa and K. Ikeda: Scr. Mater.,
2002, 47, 869873.
215. I. Charit and R. S. Mishra: Acta Mater., 2005, 53, 42114223.
216. C. J. Hsu, C. Y. Chang, P. W. Kao, N. J. Ho and C. P. Chang:
Acta Mater., 2006, 54, 52415249.
217. Y. S. Sato, H. Kokawa, K. Ikeda, M. Enomoto, S. Jogan and
T. Hashimoto: Metall. Mater. Trans. A, 2001, 32A, 941948.
218. D. P. Field, T. W. Nelson, Y. Hovanski and K. V. Jata: Metall.
Mater. Trans. A, 2001, 32A, 28692877.
219. T. W. Nelson, B. Hunsaker and D. P. Field: Proc. 1st Int. Symp.
on Friction stir welding, Thousand Oaks, CA, USA, June 1999,
TWI.
220. O. V. Flores, C. Kennedy, L. E. Murr, D. Brown, S. Pappu, B. M.
Nowak and J. C. McClure: Scr. Mater., 1998, 38, 703708.
221. H. Larsson, L. Karlsson and L.-E. Svensson: Proc. 6th Int. Conf.
on Aluminium alloys, Toyohashi, Japan, July 1998, The Japan
Institute of Light Metals, 14711476.
222. C. Genevois, A. Deschamps and A. Denquin: Proc. 5th Int. Symp.
on Friction stir welding, Metz, France, September 2004, TWI.
223. M. M. Attallah, C. L. Davis and M. Strangwood: J. Mater. Sci.,
2007, 42, 72997306.
224. R. W. Fonda and J. F. Bingert: in Friction stir welding and
processing II, (ed. K. V. Jata et al.), 191198; 2003, Warrendale,
PA, TMS.
225. C. G. Rhodes, M. W. Mahoney, W. H. Bingel, R. A. Spurling and
C. C. Bampton: Scr. Mater., 1997, 36, 6975.
226. M. Kumagai and S. Tanaka: Proc. 1st Int. Symp. on Friction stir
welding, Thousand Oaks, CA, USA, June 1999, TWI.
227. Y. S. Sato, H. Kokawa, M. Enomoto, S. Jogan and T. Hashimoto:
Metall. Mater. Trans. A, 1999, 30A, 31253130.
228. Y. S. Sato, S. H. C. Park and H. Kokawa: Metall. Mater. Trans.
A, 2001, 32A, 30333042.
229. J. D. Robson, A. Sullivan, H. R. Shercliff and G. McShane: Proc.
5th Int. Symp. on Friction stir welding, Metz, France, September
2004, TWI.
230. M. Dumont, A. Steuwer, A. Deschamps, M. Peel and P. J.
Withers: Acta Mater., 2006, 54, 47934801.

Threadgill et al.

231. R. W. Fonda, J. F. Bingert and K. J. Colligan: Scr. Mater., 2004,


51, 243248.
232. C. Genevois, A. Deschamps, A. Denquin and B. Doisneaucottignies: Acta Mater., 2005, 53, 24472458.
233. C. J. Dawes: TWI Bull., 2000, 41, 5155.
234. K. Lindner, Z. Khandkar, J. Khan, W. Tang and A. P. Reynolds:
Proc. 4th Int. Symp. on Friction stir welding, Park City, UT,
USA, May 2003, TWI.
235. R. Y. Hwang and C. P. Chou: Scr. Mater., 1997, 38, 215221.
236. A. Sullivan and J. D. Robson: Mater. Sci. Eng. A, 2008, A478,
351360.
237. L. E. Svensson, L. Karlsson, H. Larsson, B. Karlsson, M. Fazzini
and J. Karlsson: Sci. Technol. Weld. Join., 2000, 5, 285296.
238. O. R. Myhr and . Grong: Acta Metall. Mater, 1991, 39, 2703
2711.
239. O. R. Myhr and . Grong: Acta Metall. Mater, 1991, 39, 2693
2702.
240. O. R. Myhr, O. Grong, S. Klokkehaug, H. G. Fjoer and A. O.
Kluken: Sci. Technol. Weld. Join., 1997, 2, 245253.
241. . Grong and H. R. Shercliff: Prog. Mater. Sci., 2002, 47, 163
282.
242. O. Grong and O. R. Myhr: Acta Mater., 2000, 48, 445452.
243. O. R. Myhr and O. Grong: Acta Mater., 2000, 48, 16051615.
244. M. Nicolas and A. Deschamps: Acta Mater., 2003, 2003, 6077
6094.
245. T. Hyoe, P. A. Colegrove and H. R. Shercliff: in Friction stir
welding and processing II, (ed. K. V. Jata et al.), 3342; 2003,
Warrendale, PA, TMS.
246. J. D. Robson, A. Sullivan, G. McShane and H. R. Shercliff: Proc.
5th Int. Symp. on Friction stir welding, Metz, France, September
2004, TWI.
247. C. Gallais, A. Denquin, A. Pic, A. Simar and T. Pardoen: Proc.
5th Int. Symp. on Friction stir welding, Metz, France, September
2004, TWI.
248. P. A. Colegrove, H. R. Shercliff, J. D. Robson, N. Kamp,
A. Sullivan and S. W. Williams: in Mathematical modelling of
weld phenomena 8, (ed. H. Cerjak et al.), 2008, London, Maney
Publishing.
249. A. Sullivan, N. Kamp and J. D. Robson: Mater. Sci. Forum, 2006,
519521, 11811186.
250. N. Kamp, A. Sullivan, R. Tomasi and J. D. Robson: Acta Mater.,
2006, 54, 20032014.
251. N. Kamp, A. Sullivan and J. D. Robson: Mater. Sci. Eng. A,
2007, A466, 246255.
252. A. Deschamps and Y. Brechet: Acta Mater., 1998, 47, 293305.
253. A. Deschamps, F. Livet and Y. Brechet: Acta Mater., 1998, 47,
281292.
254. C. G. Derry and J. D. Robson: Mater. Sci. Eng. A, 2008, A490,
328334.
255. M. W. Mahoney: in Friction stir welding and processing, (ed.
R. S. Mishra et al.), 187217; 2007, Materials Park, OH, ASM
International.
256. H. J. Liu, H. Fujii, M. Maeda and K. Nogi: Proc. 4th Int. Symp.
on Friction stir welding, Park City, UT, USA, May 2003, TWI.
257. A. von Strombeck, G. C
am, J. F. dos Santos, V. Venzke and
M. Kocak: in Aluminum 2001, (ed. S. K. Das et al.), 249264;
2001, Warrendale, PA, TMS.
258. S. Brinkmann, A. von Strombeck, C. Schilling, J. F. dos Santos,
D. Lohwasser and M. Kocak: Proc. 2nd Int. Symp. on Friction
stir welding, Gothenburg, Sweden, June 2000, TWI.
259. A. von Strombeck, J. F. dos Santos, F. Torster, P. Laureano and
M. Kocak: Proc. 1st Int. Symp. on Friction stir welding,
Thousand Oaks, CA, USA, June 1999, TWI.
260. D. Allehaux, G. Petit, M.H. Campagnac, G. Lapasset and
A. Denquin: Proc. 4th Int. Symp. on Friction stir welding, Park
City, UT, USA, May 2003, TWI.
261. A. Denquin, D. Allehaux, H.-H. Campagnac and G. Lapasset:
Proc. 3rd Int. Symp. on Friction stir welding, Kobe, Japan,
September 2001, TWI.
262. G. Biallas and C. D. Donne: Materialprufung, 2000, 42, 236239.
263. O. Hatamleh, J. Lyons and R. Forman: Int. J. Fatig., 2007, 29,
421434.
264. M. A. Sutton, B. C. Yang, A. P. Reynolds and J. H. Yan: Mater.
Sci. Eng. A, 2004, A364, 6674.
265. M. J. Peel, M. Preuss, A. Steuwer, M. Turski and P. J. Withers: in
Trends in welding research, proceedings, (ed. S. A. David et al.),
273278; 2003, Materials Park, OH, ASM International.
266. J. Altenkirch, A. Steuwer, M. J. Peel, D. G. Richards and P. J.
Withers: Mater. Sci. Eng. A, 2008, A488, 1624.

Friction stir welding of aluminium alloys

267. J. K. Kristensen, C. Dalle Donne, T. Ghidini, J. Mononen,


A. Norman, A. Pietras, M. J. Russell and S. Slater: Proc. 5th Int.
Symp. on Friction stir welding, Metz, France, September 2004,
TWI.
268. H. J. Liu, H. Fujii, M. Maeda and K. Nogi: Mater. Sci. Technol.,
2004, 20, 103105.
269. H. J. Liu, H. Fujii, M. Maeda and K. Nogi: J. Mater. Process.
Technol., 2003, 142, 692696.
270. Lockheed Martin: presented at Aeromat 99, Dayton, OH, USA,
June 1999, ASM.
271. K. Masubuchi: in Encyclopedia of materials: science and
technology, (ed. K. H. J. Buschow et al.), 81218126; 2003,
Oxford, Elsevier.
272. P. J. Withers and H. K. D. H. Bhadeshia: Mater. Sci. Technol.,
2001, 17, 366375.
273. J. Lu (ed.): in Handbook of measurement of residual stresses;
1996, Lilburn, GA, Fairmont Press.
274. M. B. Prime: J. Mater. Eng. Technol., 2001, 123, 162168.
275. P. J. Withers, M. Turski, L. Edwards, P. J. Bouchard and D. J.
Buttle: Int. J. Press. Vess. Pip., 2008, 85, 118127.
276. P. J. Withers: C. R. Phys., 2007, 8, 806820.
277. P. J. Withers and P. J. Webster: Strain, 2001, 37, 1931.
278. P. J. Withers, M. Preuss, P. J. Webster, D. J. Hughes and A. M.
Korsunsky: Mater. Sci. Forum, 2002, 404407, 110.
279. D. J. Buttle and C. Scruby: in Encyclopedia of materials: science
and technology, (ed. K. H. J. Buschow et al.), 81738180; 2001,
Oxford, Elsevier.
280. L. D. Oosterkamp, P. J. Webster, P. A. Browne, G. B. M.
Vaughan and P. J. Withers: Proc. 5th Eur. Conf. on Residual
stresses, Mater. Sci. Forum, 2000, 347349, 687693.
281. C. Dalle Donne: Proc. 3rd Int. Symp. on Friction stir welding,
Kobe, Japan, September 2001, TWI.
282. P. J. Webster, L. D. Oosterkamp, P. A. Browne, D. J. Hughes,
W. P. Kang, P. J. Withers and G. B. M. Vaughan: J. Strain Anal.
Eng. Design, 2001, 36, 6170.
283. M. N. James, D. J. Hughes, D. G. Hattingh, G. R. Bradley,
G. Mills and P. J. Webster: Fatig. Fract. Eng. Mater. Struct., 2004,
27, 187202.
284. X. L. Wang, Z. Feng, S. A. David, S. Spooner and C. R.
Hubbard: Proc. 6th Int. Conf. on Residual stresses, Oxford, UK,
July 2000, IOM.
285. P. Staron, M. Kocak and S. Williams: in Trends in welding
research, proceedings, (ed. S. A. David et al.), 253256; 2003,
Materials Park, OH, ASM International.
286. M. B. Prime, T. GnaupelHerold, J. A. Baumann, R. J. Lederich,
D. M. Bowden and R. J. Sebring: Acta Mater., 2006, 54, 4013
4021.
287. M. Mahoney, C. Fuller, A. DeWald and M. R. Hill: Proc. 6th Int.
Symp. on Friction stir welding, Saint-Sauveur, Montreal,
Canada, October 2006, TWI.
288. T. Ghidini and C. Dalle Donne: Fatig. Fract. Eng. Mater. Struct.,
2007, 30, 214222.
289. D. Stefanescu, C. E. Truman, D. J. Smith and P. S. Whitehead:
Experim. Mech., 2006, 46, 417427.
290. P. Staron, M. Kocak, S. Williams and A. Wescott: Physica B,
2004, 350B, E491E493.
291. R. A. Owen, R. V. Preston, P. J. Withers, H. R. Shercliff and P. J.
Webster: Mater. Sci. Eng. A, 2003, A346, 159167.
292. M. A. Sutton, A. P. Reynolds, D. Q. Wang and C. R. Hubbard:
J. Eng. Mater. Technol., 2002, 124, 215221.
293. Y. J. Chao and X. H. Qi: J. Mater. Process. Manuf. Sci., 1999, 7,
215233.
294. C. M. Chen and R. Kovacevic: Int. J. Mach. Tools Manuf., 2003,
43, 13191326.
295. Z. Feng, X. L. Wang, S. A. David and P. S. Sklad: Sci. Technol.
Weld. Join., 2007, 12, 348356.
296. D. G. Richards, P. B. Prangnell, S. W. Williams and P. J. Withers:
Mater. Sci. Eng., 2008, 489, 351362.
297. A. Bastier, M. H. Maitournam, F. Roger and K. D. Van: J. Mater.
Process. Technol., 2008, 200, 2537.
298. T. D. Vuyst, V. Madhavan, B. Ducoeur, A. Simar, B. D. Meester
and L. DAlvise: Proc. 7th Int. Symp. on Friction stir welding,
Awaji Island, Japan, May 2008, TWI.
299. M. T. Milan, W. W. Bose, J. R. Tarpani, A. M. S. Malafaia, C. P.
O. Silva, B. C. Pellizer and L. E. Pereira: J. Mater. Eng. Perform.,
2007, 16, 8692.
300. H. Lombard, D. G. Hattingh, A. Steuwer and M. N. James:
Mater. Sci. Eng. A, 2009, A501, 119124.

International Materials Reviews

2009

VOL

54

NO

91

Threadgill et al.

Friction stir welding of aluminium alloys

301. J. Altenkirch, A. Steuwer, M. Poad and P. J. Withers: Residual


stresses in 2199 as a function of welding conditions, to be
submitted.
302. T. W. Greene and A. A. Holzbaur: Weld. J. Res. Suppl., 1946, 11,
171s185s.
303. P. A. McGuire and J. J. Groom: Computational analysis and
experimental evaluation for residual stresses from induction
heating, RP1394-4, Battelle Memorial Institute, 1979.
304. P. Michaleris and X. Sun: Weld. J., 1997, 76, 451s457s.
305. R. M. Dull, J. R. Dydo and J. J. Russell: Proc. 82nd Annual AWS
Convention, Cleveland, OH, USA, May 2001, AWS, 9596.
306. P. Dong, J. K. Hong and P. Rogers: Weld. J., 1998, 77, 439s445s.
307. T. E. Barber, F. W. Brust, H. W. Mishler and M. F. Kanninen:
Controlling residual stresses by heat sink welding, EPRI report
no. NP2159LD, Electric Power Research Institute, Palo Alto,
CA, USA, 1981.
308. E. M. van der Aa, M. J. M. Hermans, I. M. Richardson, N. M.
van der Pers and R. Delhez: Mater. Sci. Forum, 2006, 524525,
479484.
309. J. Gabzdyl, M. Cole, S. W. Williams and D. Price: Proc. ICALEO
2001, Jacksonville, FL, USA, October 2001, Laser Institute of
America, Vols. 92 and 93, 12361245.
310. J. T. Gabzdyl: Thermal welding, US patent no. 20010054639,
2001.
311. D. G. Richards, P. B. Prangnell, P. J. Withers, S. W. Williams and
S. Morgan: Proc. 7th Int. Symp. on Friction stir welding, Awaji
Island, Japan, May 2008, TWI.
312. G. Luan, G. Li, C. Li and C. Dong: Proc. 7th Int. Symp. on
Friction stir welding, Awaji Island, Japan, May 2008, TWI.
313. Y. P. Yang, P. Dong, X. Tian and Z. Zhang: Proc. 5th Int. Conf.
on Trends in welding research, (ed. S. A. David et al.), 700705;
1998, Materials Park, OH, ASM International.
314. Y. P. Yang, P. Dong, J. Zhang and X. Tian: Weld. J. Res. Suppl.,
2000, 79, 9s17s.
315. S. W. Williams, D. A. Price, A. Wescott, C. J. C. Harrison,
P. Staron and M. Kocak: Proc. Conf. on Welding and brazing of
aerospace structures modern applications and materials for new
and in-service, Berlin, Germany, May 2004, DVS Berichte, 95
101.
316. A. L. Lafly, C. D. Donne, G. Biallas, D. Allehaux and F. Marie:
Mater. Sci. Forum, 2006, 519521, 10891094.
317. P. Prevey and M. Mahoney: Mater. Sci. Forum, 2003, 426432,
29332939.
318. P. S. Prevey, D. J. Hornbach and N. Jayaraman: Mater. Sci.
Forum, 2007, 539547, 38073813.
319. A. Ali, X. An, C. A. Rodopoulos, M. W. Brown, P. OHara,
A. Levers and S. Gardiner: Int.l J. Fatig., 2007, 29, 15311545.
320. O. Hatamleh, J. Lyons and R. Forman: Fatig. Fract. Eng. Mater.
Struct., 2007, 30, 115130.
321. O. Hatamleh, I. V. Rivero and J. Lyons: J. Mater. Eng. Perform.,
2007, 16, 549553.
322. N. Jayaraman, P. Prevey and M. Mahoney: in Friction stir
welding and processing II, (ed. K. V. Jata et al.), 259269; 2003,
Warrendale, PA, TMS.
323. S. W. Wen, S. W. Williams, S. A. Morgan, A. Wescott and
M. Poad: submitted to Sci. Technol. Weld. Join.
324. J. Altenkirch, A. Steuwer, P. J. Withers, S. W. Williams, M. Poad
and S. Wen: Sci. Technol. Weld. Join., 2009, 14, 185192.
325. S. W. Williams, D. A. Price, W. Wescott, A. Steuwer, M. Peel,
J. Altenkirch, P. J. Withers and M. Poad: Proc. 6th Int. Symp. on
Friction stir welding, Saint-Sauveur, Montreal, Canada, October
2006, TWI.
326. J. Altenkirch, A. Steuwer, M. J. Peel, P. J. Withers, S. Williams
and M. Poad: Metall. Mater. Trans. A, 2008, 39A, 32463259.
327. J. Z. Zhang, R. Pedwell and H. Davies: Proc. 2nd Int. Symp. on
Friction stir welding, Gothenburg, Sweden, June 2000, TWI.
328. G. E. Shepherd: Proc. 2nd Int. Symp. on Friction stir welding,
Gothenburg, Sweden, June 2000, TWI.
329. L. Magnusson and L. Kallman: Proc. 2nd Int. Symp. on Friction
stir welding, Gothenburg, Sweden, June 2000, TWI.
330. H. Hori, S. Makita and H. Hino: Proc. 1st Int. Symp. on Friction
stir welding, Thousand Oaks, CA, June 1999, TWI.
331. G. Bussu and P. E. Irving: Proc. 1st Int. Symp. on Friction stir
welding, Thousand Oaks, CA, June 1999, TWI.
332. D. Lohwasser: Proc. 4th Int. Symp. on Friction stir welding,
Park City, UT, USA, May 2003, TWI.
333. M. Ericsson and R. Sandstrom: Int. J. Fatig., 2003, 25, 1379
1387.

92

International Materials Reviews

2009

VOL

54

NO

334. P. Moreira, A. M. P. de Jesus, A. S. Ribeiro and P. de Castro: in


Advances in fracture and damage mechanics VI, (ed. J. Alfaiate
et al.), 209212; 2007, Zurich, Trans Tech Publications.
335. M. Ericsson, R. Sandstrom and J. Hagstrom: Proc. 2nd Int.
Symp. on Friction stir welding, Gothenburg, Sweden, June 2000,
TWI.
336. S. Kim, C. G. Lee and S.-J. Kim: Mater. Sci. Eng. A, 2008, A478,
5664.
337. H. Lombard, D. G. Hattingh, A. Steuwer and M. N. James: Eng.
Fract. Mech., 2008, 75, 341354.
338. C. Dalle Donne, G. Biallas, T. Ghidini and G. Raimbeaux: Proc.
2nd Int. Symp. on Friction stir welding, Gothenburg, Sweden,
June 2000, TWI.
339. R. John and K. V. Jata: in Friction stir welding and processing,
(ed. K. V. Jata et al.), 5769; 2001, Warrendale, PA, TMS.
340. A. Cirello, G. Buffa, L. Fratini and S. Pasta: Proc. Inst. Mech.
Eng. B, 2006, 220B, 805811.
341. G. Bussu and P. E. Irving: Int. J. Fatig., 2003, 25, 7788.
342. J. M. L. Tan, M. E. Fitzpatrick and L. Edwards: Eng. Fract.
Mech., 2007, 74, 20302054.
343. S. S. Di, X. Q. Yang, G. H. Luan and B. Jian: Mater. Sci. Eng. A,
2006, A435, 389395.
344. C. Z. Zhou, X. Q. Yang and G. H. Luan: J. Mater. Sci., 2006, 41,
27712777.
345. A. Oosterkamp, L. D. Oosterkamp and A. Nordeide: Weld. J.,
2004, 83, 225S231S.
346. A. Ali, M. W. Brown and C. A. Rodopoulos: Proc. 6th Int. Symp.
on Friction stir welding, Saint-Sauveur, Montreal, Canada,
October 2006, TWI.
347. H. B. Chen, K. Yan, T. Lin, S. B. Chen, C. Y. Jiang and Y. Zhao:
Mater. Sci. Eng. A, 2006, A433, 6469.
348. M. N. James, D. G. Hattingh and G. R. Bradley: Int. J. Fatig.,
2003, 25, 13891398.
349. M. N. James, G. R. Bradley, H. Lombard and D. G. Hattingh:
Fatig. Fract. Eng. Mater. Struct., 2005, 28, 245256.
350. T. L. Dickerson and J. Przydatek: Int. J. Fatig., 2003, 25, 1399
1409.
351. C. A. Widener, B. M. Tweedy and D. A. Burford: Proc. 7th Int.
Symp. on Friction stir welding, Awaji Island, Japan, May 2008,
TWI.
352. C. Dalle Donne and G. Biallas: Proc. Eur. Conf. on Spacecraft
structures, materials and mechanical testing, Noordwijk, The
Netherlands, NovemberDecember 1999, ESA, 309314.
353. S. G. Russell, M. Tester, E. Nichols, A. Cleaver and J. Maynor: in
Friction stir welding and processing, (ed. K. V. Jata et al.), 93
104; 2001, Warrendale, PA, TMS.
354. M. T. Milan, W. W. B. Filho, C. O. F. T. Ruckert and J. R.
Tarpani: Fatig. Fract. Eng. Mater. Struct., 2008, 31, 526538.
355. C. Z. Zhou, X. Q. Yang and G. H. Luan: Scr. Mater., 2005, 53,
11871191.
356. H. Q. Qu, M. Tsujikawa, S. W. Chung, S. Oki and K. Higashi: in
Recrystallization and grain growth III, (ed. S. J. L. Kang et al.),
793796; 2007, Zurich, Trans Tech Publications.
357. M. Ranes, A. O. Kluken and O. T. Midling: Proc. 4th Int. Conf.
on Trends in welding research, Gatlinburg, TN, USA, June 1995,
ASM International.
358. S. Hong, S. Kim, C. G. Lee and S. J. Kim: J. Mater. Sci., 2007, 42,
98889893.
359. P. S. Pao, R. W. Fonda, H. N. Jones, C. R. Feng, B. J. Connolly
and A. J. Davenport: in Friction stir welding processing III, (ed.
K. V. Jata et al.), 2734; 2005, Warrendale, PA, TMS.
360. P. Cavaliere, G. Campanile, F. Panella and A. Squillace: J. Mater.
Process. Technol., 2006, 180, 263270.
361. S. J. Hong, S. S. Kim, C. G. Lee and S. J. Kim: in Advances in
nanomaterials and processing, (ed. B. T. Ahn et al.), 13211324;
2007, Zurich, Trans Tech Publications.
362. M. Ericsson and R. Sandstrom: Steel Res. Int., 2006, 77, 450455.
363. P. J. Haagensen, O. T. Midling and M. Ranes: in Computer
methods and experimental measurements for surface treatment
effects II, (ed. M. H. Aliabadi et al.), 225237; 1995,
Southampton, Wit Pr/Computational Mechanics.
364. M. Kumagai, S. Tanaka, H. Hatta, H. Yoshida and H. Sato:
Proc. 3rd Int. Symp. on Friction stir welding, Kobe, Japan,
September 2001, TWI.
365. M. Ericsson and R. Sandstrom: Proc. 5th Int. Symp. on Friction
stir welding, Metz, France, September 2004, TWI.
366. W. M. Thomas, M. F. Gittos, P. Tubby, D. G. Staines and
S. Lockyer: Friction skew stir lap welding of 5083H111preliminary fatigue studies, TWI members report, TWI,
Abington, UK, to be published.

Threadgill et al.

367. R. Pedwell, H. Davies and A. Jefferson: Proc. 1st Int. Symp. on


Friction stir welding, Thousand Oaks, CA, June 1999, TWI.
368. R. Talwar, D. Bolser, R. Lederich and J. Baumann: Proc. 2nd Int.
Symp. on Friction stir welding, Gothenburg, Sweden, June 2000,
TWI.
369. G. E. Shepherd: Proc. 4th Int. Symp. on Friction stir welding,
Park City, UT, USA, May 2003, TWI.
370. W. M. Thomas, I. M. Norris, D. G. Staines and E. R. Watts:
Proc. SME Summit, Oconomowoc, WI, USA, August 2005, SME.
371. D. Fersini and A. Pirondi: Eng. Fract. Mech., 2008, 75, 790803.
372. D. Hornbach, M. Mahoney, P. Prevey, D. Waldron and
J. Cammett: in Trends in welding research, proceedings, (ed.
S. A. David et al.), 302306; 2003, Materials Park, OH, ASM
International.
373. S. Lomolino, R. Tovo and J. dos Santos: Int. J. Fatig., 2005, 27,
305316.
374. M. G. Dawes, T. L. Dickerson, S. A. Karger and J. Przydatek:
Proc. 2nd Int. Symp. on Friction stir welding, Gothenburg,
Sweden, June 2000, TWI.
375. M. K. Kulekci, F. Mendi, I. Sevim and O. Basturk: Metalurgija,
2005, 44, 209213.
376. M. A. Sutton, A. P. Reynolds, B. C. Yang and R. Taylor: Eng.
Fract. Mech., 2003, 70, 22152234.
377. W. J. Arbegast, K. S. Baker and P. J. Hartley: Proc. 5th Int. Conf.
on Trends in welding research, Pine Mountain, GA, USA, June
1998, ASM International.
378. H. R. Kroninger and A. P. Reynolds: Fatig. Fract. Eng. Mater.
Struct., 2002, 25, 283290.
379. M. Mochizuki, M. Inuzuka, H. Nishida, K. Nakata and
M. Toyoda: Sci. Technol. Weld. Join., 2006, 11, 366370.
380. C. G. Derry and J. D. Robson: Proc. 7th Int. Symp. on Friction
stir welding, Awaji Island, Japan, May 2008, TWI.
381. M. G. Dawes, S. A. Karger and J. Przydatek: Fracture toughness
of friction stir welds in 2014A, 7075 and 5083 aluminium alloys,
TWI report no. 705/2000, TWI, Abington, UK, 2000.
382. D. G. Kinchen, Z. Li and G. P. Adams: Proc. 1st Int. Symp. on
Friction stir welding, Thousand Oaks, CA, June 1999, TWI.
383. D. Dumont, A. Deschamps and Y. Brechet: Acta Mater., 2004,
52, 25292540.
384. J. Lumsden: in Friction Stir Welding and Processing, (ed. R. S.
Mishra et al.), 187217; 2007, Materials Park, OH, ASM
International.
385. F. Hannour, A. J. Davenport and M. Strangwood: Proc. 2nd Int.
Symp. on Friction stir welding, Gothenburg, Sweden, June 2000,
TWI.
386. J. Corral, E. A. Trillo, Y. Li and L. E. Murr: J. Mater. Sci. Lett.,
2000, 19, 21172122.
387. D. P. P. Booth, M. J. Starink and I. Sinclair: Mater. Sci. Technol.,
2007, 23, 276284.
388. A. J. Davenport, M. Jariyaboon, C. Padovani, N. Tareelap, B. J.
Connolly, S. Williams and E. Siggs: Mater. Sci. Forum, 2006, 699,
519521.
389. C. Dalle Donne, R. Braun, G. Staniek, A. Jung and W. A.
Kaysser: Mater. Werkst., 1998, 29, 609617.
390. D. B. Mitton, A. Squillace, A. De Fenzo, C. Padovani,
T. Monetta, G. Giorleo, P. Cozzolino and F. Bellucci: Corros.
Rev., 2007, 25, 449459.
391. J. B. Lumsden, M. W. Mahoney, G. Pollock and C. G. Rhodes:
Corrosion, 1999, 55, 11271135.
392. R. G. Buchheit and C. S. Paglia: in Corrosion and protection of
light metal alloys, (ed. R. G. Buchheit et al.), 94103; 2004,
Pennington, NJ, Electrochemical Society.
393. C. S. Paglia, K. V. Jata and R. G. Buchheit: Mater. Corros., 2007,
58, 737750.
394. X. Zhou, Y. Younes, D. Wadeson, T. Hashimoto and G. E.
Thompson: Adv. Mater. Res., 2008 38, 298305.
395. B. J. Connolly, A. J. Davenport, M. Jariyaboon, C. Padovani,
R. Ambat, S. W. Williams, D. A. Price, A. Wescott, C. J.
Goodfellow and C.-M. Lee: Proc. 5th Int. Symp. on Friction stir
welding, Metz, France, September 2004, TWI.
396. M. Jariyaboon, A. J. Davenport, R. Ambat, B. J. Connolly, S. W.
Williams and D. A. Price: Corros. Sci., 2007, 49, 877909.

Friction stir welding of aluminium alloys

397. A. Squillace, A. de Fenzo, G. Giorleo and F. Bellucci: J. Mater.


Process. Technol., 2004, 152, 97105.
398. W. Hu and E. I. Meletis: Mater. Sci. Forum, 2000, 331337, 1683
1688.
399. C. S. Paglia and R. G. Buchheit: Mater. Sci. Eng. A, 2006, A429,
107114.
400. F. Zucchi, G. Trabanelli and V. Grassi: Mater. Corros., 2001, 52,
853859.
401. S. Maggiolino and C. Schmid: J. Mater. Process. Technol., 2008,
197, 237240.
402. R. Braun, C. D. Donne and G. Staniek: Mater. Werkst., 2000, 31,
10171026.
403. J. B. Lumsden, M. W. Mahoney, C. G. Rhodes and G. A.
Pollock: Corrosion, 2003, 59, 212219.
404. C. S. Paglia, L. M. Ungaro, B. C. Pitts, M. C. Carroll, A. P.
Reynolds and R. G. Buchheit: in Friction stir welding and
processing II, (ed. K. V. Jata et al.), 6575; 2003, Warrendale,
PA, TMS.
405. K. K. Sankaran, H. L. Smith and K. Jata: in Trends in welding
research, proceedings, (ed. S. A. David et al.), 284286; 2003,
Materials Park, OH, ASM International.
406. D. P. Field, T. W. Nelson, Y. Hovanski and D. F. Bahr: in
Friction stir welding and processing, (ed. K. V. Jata et al.), 83
91; 2001, Warrendale, PA, TMS.
407. C. S. Paglia, M. C. Carroll, B. C. Pitts, T. Reynolds and R. G.
Buchheit: Mater. Sci. Forum, 2002, 396402, 16771684.
408. X. Y. Yun, Y. Motohashi, T. Ito, T. Asano and S. Hirano: J. Jpn
Inst. Met., 2006, 70, 96105.
409. D. A. Wadeson, X. Zhou, G. E. Thompson, P. Skeldon, L. D.
Oosterkamp and G. Scamans: Corros. Sci., 2006, 48, 887897.
410. L. A. Willey: in Aluminium: properties, physical metallurgy and
phase diagrams, (ed. K. R. van Horn), 140158; 1967, Cleveland,
OH, ASM.
411. J. R. Galvele and S. M. de Micheli: Corros. Sci., 1970, 10, 795
807.
412. J. T. Staley, S. C. Byrne, E. L. Colvin and K. P. Palmer: Mater.
Sci. Forum, 1996, 217, 15871592.
413. C. M. Lee: Corrosion, submitted.
414. C. S. Paglia and R. G. Buchheit: Scr. Mater., 2008, 58, 383387.
415. R. Cook, T. Handboy, S. L. Fox and W. Arbegast: in Friction
stir welding processing III, (ed. K. V. Jata et al.), 3542; 2005,
Warrendale, PA, TMS.
416. P. B. Srinivasan, W. Dietzel, R. Zettler, J. F. dos Santos and
V. Sivan: Mater. Sci. Eng. A, 2005, A392, 292300.
417. C. A. Widener, J. E. Talia, B. M. Tweedy and D. A. Burford: in
Friction stir welding and processing IV, (ed. R. S. Mishra et al.),
449458; 2007, Warrendale, PA, TMS.
418. M. Jariyaboon, A. J. Davenport, R. Ambat, B. J. Connolly, S. W.
Williams and D. A. Price: Corros. Eng. Sci. Technol., 2006, 41,
135142.
419. J. Lumsden, G. Pollock and M. Mahoney: Mater. Sci. Forum,
2003, 426432, 28672872.
420. C. A. Widener, D. A. Burford, B. Kumar, J. E. Talia and
B. Tweedy: Mater. Sci. Forum, 2007, 539543, 37813788.
421. A. Merati, K. Sarda, D. Raizenne and C. D. Donne: in Friction
stir welding and processing II, (ed. K. V. Jata et al.), 7790; 2003,
Warrendale, PA, TMS.
422. C. A. Widener, J. E. Talia, B. M. Tweedy and D. A. Burford: in
Friction stir welding and processing IV, (ed. R. S. Mishra et al.),
459468; 2007, Warrendale, PA, TMS.
423. J. Lumsden, G. Pollock and M. Mahoney: in Friction stir welding
processing III, (ed. K. V. Jata et al.), 1925; 2005, Warrendale,
PA, TMS.
424. C. Padovani, L. Fratini, A. Squillace and F. Bellucci: Corros.
Rev., 2007, 25, 475489.
425. P. B. Srinivasan, W. Dietzel, R. Zettler, J. F. dos Santos and
V. Sivan: Corros. Eng. Sci. Technol., 2007, 42, 161167.
426. K. P. Rao, G. D. J. Ram and B. E. Stucker: Scr. Mater., 2008, 58,
9981001.
427. P. Prevey, D. Hornbach, P. Mason and M. Mahoney: in Surface
engineering: coating and heat treatments, proceedings, (ed.
O. Popoola et al.), 131137; 2003, Materials Park, OH, ASM
International.

International Materials Reviews

2009

VOL

54

NO

93

S-ar putea să vă placă și