You are on page 1of 9

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/228849680

Investigation of microstructure, surface and


subsurface characteristics in titanium alloy
friction stir welds of varied thicknesses
ARTICLE in SCIENCE AND TECHNOLOGY OF WELDING & JOINING JULY 2009
Impact Factor: 1.38 DOI: 10.1179/136217109X425838

CITATIONS

15

2 AUTHORS:
P. Edwards

M. Ramulu

Tesla Motors

University of Washington Seattle

49 PUBLICATIONS 151 CITATIONS

182 PUBLICATIONS 2,639 CITATIONS

SEE PROFILE

SEE PROFILE

Available from: P. Edwards


Retrieved on: 27 August 2015

Investigation of microstructure, surface and


subsurface characteristics in titanium alloy
friction stir welds of varied thicknesses
P. D. Edwards1,2 and M. Ramulu*1
Friction stir welding of titanium alloy (Ti6Al4V) was demonstrated on 3, 6, 9 and 12 mm
thickness square groove butt joints. Complete microstructural and microhardness evaluations
were conducted in addition to surface and subsurface examinations for each case. The 3 mm
welds exhibited an extremely fine grained microstructure with evidence of processing
temperatures below the beta transus temperature of the alloy. The 6, 9 and 12 mm samples
possessed larger grains formed by a slower cooling rate from above the beta transus
temperatures. The thick section weld exhibited a nearly uniform microhardness, while the thinner
welds showed a slight, 6%, increase in hardness compared with the parent material.
Keywords: Friction stir welding, Titanium, Microstructure, Microhardness

Introduction
Ti6Al4V is the most used of all the titanium alloys. It
accounts for more than 50% of worldwide titanium
usage.1 Furthermore, of all the titanium alloys, Ti6Al
4V is considered to be the most highly weldable via
standard fusion welding processes such as arc, laser or
electron beam. These joining processes produce large
grained, martensitic, cast type microstructures that
result in adverse effects on the mechanical properties
of the joints.1 Solid state joining processes such as
friction stir welding (FSW) are able to retain the
microstructural integrity of the parent material in the
welded joint, thereby producing mechanical properties
that are more comparable with those of the parent
material. Friction stir welding has been successfully
applied to the joining of metals such as aluminium,
magnesium, copper and even steel.2,3 Only recently has
FSW been applied to high strength, high temperature
materials such as steel and titanium mainly due to the
difficulties associated identifying tooling materials that
can withstand the temperatures and loads involved with
welding such materials.
Mishra et al.2,3 and Nandan et al.4 have recently
provided excellent reviews of the FSW process, weld
characterisation, modelling and its applications for a
variety of materials including aluminium and titanium.
However, because the application of FSW to titanium
alloys is relatively new and information available is
scarce, there still exists an opportunity for process
development and joint characterisation. The purpose of
this paper is to review the developments in titanium

University of Washington, Seattle, WA, USA


The Boeing Company, Seattle, WA, USA

*Corresponding author, email ramulum@u.washington.edu

2009 Institute of Materials, Minerals and Mining


Published by Maney on behalf of the Institute
Received 17 January 2009; accepted 10 February 2009
DOI 10.1179/136217109X425838

FSW to date and present the results of a current study


intended to examine the process conditions required to
produce high quality square groove butt joint configuration welds in a wide range of material thicknesses
along with the surface and subsurface characteristics of
those joints.

Brief review of titanium friction stir


welds
There have been several papers published on the FSW of
titanium.517 Titanium alloys friction stir welded include
CP-Ti, Ti 15-3-3-3, Ti 17, Ti 6-2-4-2, Ti 17 and beta 21S.
Welded thicknesses reported range from 1?2 to 12 mm.
However, process conditions and tooling information
available were rather limited. Since the most widely used
titanium alloy is Ti6Al4V, it was of particular interest.
From the limited literature available, the thicknesses
welded and processing conditions used in Ti6Al4V are
summarised in Table 1. In most cases, welds were
performed in position control using a simply shaped
tungsten tool and some employed the use of thermal
management.5,6 Other tool materials such as TiC,7
PCBN8 and Mo based9 alloy were used, but these were
far less successful than the tungsten based tools. Tool
dimensions used ranged from 1519 mm shoulders and
38 mm diameter pins.9,10 In some cases, the pin tips
were tapered from 5 mm at the shoulder to 3 mm at the
tip.9
Characterisation of Ti6Al4V FSWs typically
included macro- and microstructural evaluations along
with microhardness and tensile testing. Macrodefects
observed in welds made under non-optimal conditions
include undercut, root voids and lack of penetration.9
Microstructural evaluations typically identified a refined
weld nugget microstructure compared with the parent
material with evidence of exceeding the beta transus

Science and Technology of Welding and Joining

2009

VOL

14

NO

476

Edwards and Ramulu

Characteristics in Ti alloy friction stir welds

a optical micrograph of standard Ti6Al4V microstructure; b scanning electron micrograph of standard Ti6Al4V
microstructure
1 Standard Ti6Al4V micrographs

temperature during processing.5,6,9,10,13 Microstructural


evaluations of the heat affected zones (HAZ) and
thermomechanically affected zones (TMAZ) indicated
that the peak temperatures in these portions of the weld
are below the beta transus temperature10 and the TMAZ
is relatively small compared with aluminium FSWs.13
Microhardness evaluations showed that the hardness of
the weld nugget is greater than the surrounding parent
material.9,10 Reported hardness measurements of the
HAZ varied. Some reported softening in the HAZ,9
while others found hardening10 in the HAZ. It is
generally expected that the weld nugget will be harder
than the parent material due to the refined grain size,
while the HAZ will be softer due to the enlarged grains
caused by exposure to elevated temperatures.9 Tensile
testing typically resulted in lower elongations and
strengths with failure occurring in the HAZ.9,10 Other
tensile property studies14,15 observed lower to equivalent
elongations and strengths of the weld specimen compared with the parent material with failure occurring in
either the base metal or weld depending on the quality of
the joints. Tensile testing of the weld nugget alone
showed increased strength and elongation compared
with the parent material.9 Again, this is assumed to be
due to the microstructural differences among the HAZ,
weld nugget and base material.9

Fig. 1. All materials were machined along the abutting


edges and chemically etched before welding. Test
materials were welded then evaluated visually and
microstructurally to determine the joint quality. Test
parts were welded using the facilities at the Edison
Welding Institute (EWI). This development effort was a
collaborative effort among EWI, the Boeing Company
and the University of Washington.

Friction stir welding


The tooling design configurations, tooling materials,
process conditions and thermal management techniques
utilised in the work by Sanders11,12 were the baseline
used in this study. A tapered, tungsten lanthanide, pin
tool with a relatively small shoulder diameter was used
through out this study. Thermal management of the pin
tool was also used to control the high temperatures
produced during the Ti FSW process.
In general, spindle speed, travel speed, shoulder
engagement, penetration ligament tool tilt, coolant flow
through the weld tool and the coolant flow through the
backing anvil are the process parameters that dictate the
quality of the joint. Tool tilt, penetration ligament and
shoulder engagement are defined in Fig. 2. The process
parameters and basic tooling configuration dimensions
used during weld development effort are summarised in
Table 2.

Microstructural, surface and subsurface


characterisation

Experimental procedure
Material
The material used this investigation is the Ti6Al4V
titanium alloy possessing a average grain size that is
typically on the order of 810 mm. The thicknesses used
are 3, 6, 9 and 12 mm. The chemical composition for
this material is 6Al0?8C0?15H0?4Fe0?05N0?2O
9Ti4V and typical representative microstructures at low
(Fig. 1a) and high (Fig. 1b) magnifications are shown in

Microstructural evaluations were carried out in order to


examine the grain structure of the weld nugget, the HAZ
and the base metal after the FSW process. To
accomplish this, the samples were first cut, via abrasive
water jet, from a titanium plate that had been friction
stir welded. The sample included a cross-section of the
weld nugget oriented perpendicularly to the joining line.
The samples were then polished on 5, 1 and finally

Table 1 Summary of process conditions used for FSW of Ti6Al4V


Thickness, mm

Spindle speed, rev min21

Feedrate, mm min21

Ref. no.

1?5
2
2
6

325
200
300, 400, 500
275

100
100
60
100

11
5, 6
9
10

Science and Technology of Welding and Joining

2009

VOL

14

NO

477

Edwards and Ramulu

Characteristics in Ti alloy friction stir welds

Table 3 Optimal FSW processing parameters for given


tool design and thickness

2 Schematic dening tool tilt, shoulder engagement and


penetration ligament

0?3 mm alumina oxide polishing wheels. The sample was


then etched to expose the microstructure. A JEOL JSM840A scanning electron microscope (SEM) was used in
conjunction with a JEOL DSG Plus digital scan imaging
system to digitally capture the SEM images. First, a
series of low magnification (610) optical macrographs
were taken to encompass the whole sample. This was
followed by a series of higher magnification (650 and
6500) optical micrographs. These were taken at
approximately the mid-thickness of each weld. Scanning
electron images were captured at 62000 magnification
under a 10 kV accelerating voltage.
Surface topography was investigated using optical
microscopy and surface profilometry. Surface profiles of
the weldments and average roughness parameters were
obtained using MarSurf XR 20 surface analyser. This
XR 20 is equipped with a stylus based, 2?54 mm radius
conical diamond tip with 5 nm resolution. The unit has
high sensitivity capability to capture exceptionally small
changes in surface roughness as it traverses across a
profile. The diamond tip is necessary to avoid damage
when measuring hard metal alloys with rough surfaces,
such as the sharp peaks of notches present on the
titanium FSW coupons. Surfanlyser is that it automatically stores all of the data to disk and then it computes
the standard roughness parameters. Longitudinal passes
along specimens centreline were obtained with a cutoff
length of 0?8 mm and a traverse length of 5?6 mm. In
order to quantify the surface topology, the surface
roughness parameters Ra, the arithmetic average roughness height Ry, the maximum peak to valley height and
Rz, the 10 point height as calculated by measuring the
distance between the five highest peaks and the five
lowest valleys in the sampling length were calculated
from the recorded surface profile.

Joint thickness,
mm

Spindle speed,
rev min21

Feedrate,
mm min21

3
6
9
12

300
280
270
170

75
100
65
65

Microhardness tests were utilised to assess the surface


and subsurface characteristics of both the base material
and the weld nugget. These samples were prepared in the
same manner as discussed for the microstructural
evaluations. The samples were tested with a 100 g load
applied for 15 s using an automated Vickers hardness
tester. The size of the resulting indent was then measured to determine the microhardness of the material
at that point. For each weld cross-section, a twodimensional microhardness grid was laid out over the
specimen. The microhardness indent spacing was
y500 mm in the through thickness and transverse directions. Depending on the weld thickness, this resulted in
10006000 indents per sample. The first microhardness
point for a given traverse was started at the left edge of
the sample, in the base material, and then moved to the
HAZ, the weld nugget, back into the HAZ and finally
into the base material on the opposite side of the weld
nugget again.

Results
Process parameters for welding 3, 6, 9 and 12 mm Ti
6Al4V were identified and evaluated. The optimal
processing conditions for each thickness based on this
development effort are given in Table 3. Macrographs
for each of these optimal welding conditions in each
thickness are shown in Fig. 3.
In addition to macrographic examinations, each of
the welds made under the best set of process conditions
identified was prepared for microstructural examination
via optical microscopy and SEM as described previously. Low magnification (650) optical photographs
of the weld nuggetbase metal transition areas, which
include HAZ and TMAZ were taken for each weld
thickness along with higher magnification (6500)
optical photographs of the microstructure in the centre
of the welds. Representative optical micrographs for
the 6 and 9 mm weld cases are shown in Fig. 4.
Representative SEM images from the centre of the 3
and 12 mm welds are given in Fig. 5. Scanning electron
microscope images (62000) which show the grain size
distribution through the thickness of the 3 and 12 mm
welds are shown in Fig. 6. Finally, a 62000 SEM image
of the HAZ in a 9 mm weld sample is shown in Fig. 7.
The surface texture of the friction stir welded samples
was characterised with optical macrographs and with
surface roughness profiles. As the FSW tool shoulder

Table 2 Summary of Ti6Al4V FSW parameters tested


Joint thickness, mm

Spindle speed, rev min21

Feedrate, mm min21

Shoulder diameter, mm

Pin tip diameter, mm

3
6
9
12

300
250320
250285
140190

50130
45100
65100
4075

20
25
25
30

8
10
10
10

Science and Technology of Welding and Joining

2009

VOL

14

NO

478

Edwards and Ramulu

Characteristics in Ti alloy friction stir welds

3 Macrographs of a 3 mm, b 6 mm, c 9 mm and d 12 mm FSW cross-sections and e photograph of typical weld surface

rotates and translates along the workpiece surface, it


leaves behind tool markings similar to feed marks in
machining. These surface characteristics were characterised as they induce surface stress concentration
factors that may have sever impacts on the mechanical
properties of the joints, particularly high cycle fatigue
life. Figure 8 shows typical cross-section of these tool
markings left behind the FSW tool at the crown of the
weld joint along with typical surface roughness profiles

recorded for the 3 and 9 mm weld samples. For the


3 mm sample, values of Ra and Ry were approximately
22 and 137 mm respectively. For the 12 mm sample, the
Ra and Ry values were approximately 25 and 107 mm
respectively.
Extensive microhardness testing was also conducted
on the welds produced during this study. After the
microstructural examination was complete, the samples
were repolished and two dimensional microhardness

4 Optical micrographs of advancing side base metal to weld nugget transition zones in a 6 mm and c 9 mm welds, and
higher magnication optical micrographs of the same regions in b 6 mm and d 9 mm welds

Science and Technology of Welding and Joining

2009

VOL

14

NO

479

Edwards and Ramulu

Characteristics in Ti alloy friction stir welds

a 12 mm weld; b 3 mm weld
5 Images (SEM) taken from centre of weld nugget

6 Scanning electron images showing grain size distribution through thickness of a 12 mm weld and b 3 mm weld

Discussion

7 Scanning electron image of HAZ in 9 mm weld sample

grids were laid out on the mounts. Figure 9 shows the


results of this microhardness study as contour plots for
each of the weld thicknesses. Figure 10 shows onedimensional plots for mid-thickness microhardness
traverses extending from the base metal on one side of
the weld nugget, all the way through the heat affected
zone, weld nugget and into the base metal on the
opposite side for the 3 and 12 mm weld cases. In Fig. 10,
the approximate locations of the microhardness points
taken in the base material (BM), heat affected zone
(HAZ) and weld nugget (WN) are indicated on the plots.

There were several aspects of the welds produced during


this study which were uncovered via metallurgical
characterisation. First, it was found that these welds
possess an extremely small HAZ, on the order of 200
400 mm, as shown in Figs. 3 and 4. Furthermore, it was
difficult to distinguish between the HAZ and any
TMAZ, even at higher magnification (Fig. 4b and d).
Essentially, the TMAZ in titanium FSW is negligibly
small. This is in strong contrast to aluminium friction
stir welds which possess distinct and rather larger HAZ
and TMAZ. This is likely attributed to the high strength
and low thermal conductivity of titanium compared with
aluminium. In the Ti FSW process, the thermal
conductivity of the material is so low, the heat generated
by the process is not transferred away from the weld
zone fast enough to create the large HAZ that would be
seen in fusion welded joints. Because heat is not
transferred far from the weld zone, the cooler and
stronger material beside the weld zone resists mechanical
deformation that would create a TMAZ, like in
aluminium FSW.
Another characteristic of the titanium FSW nugget is
the grain size distributions (Fig. 6). It was observed that
the grain size is quite uniform across the width of the
FSW nugget at any given depth (Fig. 4). However,
optical and scanning electron microscopy revealed a
clear grain size distribution through the thickness and a
general grain size difference between the welds of

Science and Technology of Welding and Joining

2009

VOL

14

NO

480

Edwards and Ramulu

Characteristics in Ti alloy friction stir welds

8 a photograph of 3 mm weld surface, b surface roughness prole in 3 mm sample, c photograph of 12 mm weld surface and d surface roughness prole in 12 mm sample

different thicknesses. Optical and SEM images have


shown that the grain size in the welds increases with
thickness. Very small grains were observed in the thin
welds and quite large grains were found in the thick
welds. Additionally, there is a clear grain size distribution through the thickness of any given weld section
(Fig. 6). The grains in the welds are relatively large near
the crown, or top, or the weld and become small and
very fine near the root, or bottom, or the weld.
All of these grain size distribution characteristics can
be attributed to the thermomechanical nature of the
process. The grains are large at the top of the weld
because that is where the shoulder of the tool is
generating the most heat. In the root of the weld, the
small tip of the tool generates much less heat, and due to
the low thermal conductivity of the material, the heat
generated at the top by the shoulder is not conducted

into the root. Thus, the material at the root of the joint
is subject to high degrees of thermomechanical work,
with minimal heating, while the material at the top of
the weld is subjected to much greater temperatures,
allowing grain growth post-stirring. The grains in the
region of the weld corresponding to the material directly
under the pin tip are typically extremely fine, even in
nanoscale. The grain size variation in thickness is likely
due to heating and cooling rates. The thinner welds cool
much faster than the thick section welds, restricting
grain growth post-stirring. Additionally, the tools used
in the thinner section welds have a much smaller surface
area, producing less heat than the thick section weld
tools. This also contributes to the smaller grain size.
These microstructural characterisations can also be
used to make a rough estimate of the temperatures in the
weld zone during processing. The SEM image of the

9 Microhardness contour plots for 3, 6, 9 and 12 mm thick FSWs

Science and Technology of Welding and Joining

2009

VOL

14

NO

481

Edwards and Ramulu

10 Single microhardness traverse plots across 3 and


12 mm weld samples at mid-thicknesses of 1?5 and
6 mm respectively

12 mm sample (Fig. 5) shows a transformed beta


microstructure with grain boundary alpha. This suggests
that the weld temperature in this section exceed the beta
transus temperature of 1065uC (1950 F). The SEM
image of the 3 mm sample (Fig. 5) exhibits a much
smaller transformed beta microstructure, with what
could be some untransformed, or primary, alpha. This
suggests that the temperature in the thin section weld is
near the beta transus temperature, but not completely
through it, leaving some untransformed alpha in the
resulting microstructure. The SEM images taken for the
6 and 9 mm samples are similar to that of the 12 mm
sample and all suggest that the beta transus temperature
was exceeded in the weld zone during processing. For all
welds, there was noticeable untransformed alpha in the
HAZ (Fig. 7), indicating that the temperatures in the
HAZ were near, but did not completely exceed the beta
transus temperature, regardless of the weld thickness.
The surface texture left behind the FSW tool is quite
rough (Fig. 6). There are peaks and valleys corresponding to the material flow around the pin tool and under
the shoulder. Most alarming is the sharp shape of the
valleys of these flow lines. These will greatly deteriorate

Characteristics in Ti alloy friction stir welds

the fatigue properties of these welds because these sharp


radii features will cause high stress concentrations and
sites for early crack initiation. Based on the macrographic analysis, a minimum of 500 mm (200 mm for the
feature plus a safety factor of 2?5) is suggested to be
machined off of any Ti FSW surface to prevent fatigue
performance degradations due to the tool markings on
the surface of the welds.
For the microhardness evaluations, the thin welds
show a higher hardness in the weld nugget than the
parent material, while the thicker section welds have a
nearly uniform hardness in the weld nugget compared
with the parent material (Fig. 7). This is also likely due
to the grain size variations in the welds resulting from
the peak temperature and cooling rate differences
between the thick section welds and thin welds. The
thick welds cool at a much slower rate, allowing the
grains to grow to a size more similar to the parent
material resulting in more uniform hardness distributions. Conversely, the grains in the thin welds are
generally smaller than the surrounding parent material
due to the faster cooling rates and lower heat inputs,
resulting in higher microhardness. It is expected that a
proper heat treatment cycle will successfully remove
these microhardness differences in the as welded sections. It was also noted that the thinner welds (Fig. 10a)
show a hardness decrease in the HAZ which is in
agreement with previous microhardness evaluations.
In general, the spindle speed (rev min21) and travel
speed (mm min21) have the most dominate effects on
the quality of the joint. The spindle speed contributes
most to heating and the travel speed has the largest
effect on how well the root of the joint is stirred for
penetration. The most common defects encountered are
a lack of penetration (LoP) defect and small cavities in
the root of the weld. LoP defects are caused by
insufficient heating and deformation at the root of the
joint which leaves behind an unstirred prior joint
interface (Fig. 11). Root cavities (Fig. 11) are caused
by improper combinations of spindle speed and travel
speed. A relatively high spindle speed with a relatively
low travel speed will lead to overstirring of the root and
defect formation. Also, a low spindle speed with a high
travel speed will result in insufficient heating and
working of the root, also leaving behind void defects.
There is a delicate balance among all of the FSW
processing parameters that must be achieved to produce
a defect free joint in Ti6Al4V. Further work needs to

a lack of penetration defect; b root void defect


11 Typical FSW defects

Science and Technology of Welding and Joining

2009

VOL

14

NO

482

Edwards and Ramulu

be conducted to establish relationships between all of the


relevant processing parameters and the resulting quality
characteristics of the joint. Implementing modelling and
experimental measurement techniques to characterise
the thermal gradients and deformation zone formation
during the process will also be valuable for establishing
this relationship, which is essential to identifying the
optimal processing parameter limits for each thickness
in order to ensure that the parameters being used are
sufficiently robust for a production process.

Conclusions
Friction stir welding of Ti6Al4V alloy has been
developed and demonstrated in butt weld thicknesses
ranging from 3 to 12 mm. Extensive metallurgical and
microhardness characterisations have been performed
on welds made under the optimal processing conditions,
which are defined as those that produce defect free
welds. Based on this experimental study, the following
key conclusions were made:
1. Metallurgical characterisations observed grain size
distributions through the thickness of the welds and
between weld thicknesses. However, the grain size was
quite uniform across the width of each weld at any given
depth. In the 6, 9 and 12 mm welds, the microstructure
appears to have crossed the beta transus temperature
due to the presence of alpha prime in the weld nugget
microstructure. The 3 mm weld is likely to have not fully
crossed the beta transus temperature because of remnant
primary alpha in the microstructure.
2. Microhardness results showed a hardness increase
in the thin section welds compared with the surrounding
base material and a more uniform hardness in the thick
section welds. These microhardness characteristics are
likely due to the grain size variations noted previously.
The thin section welds have grains that are smaller than
the parent material while the thick section welds have
much larger grains.

Acknowledgements
The authors of this paper would like to thank the
Boeing Company for the support and Dr D. Sanders

Characteristics in Ti alloy friction stir welds

for his encouragement throughout this research project.


The authors sincere thanks are also extended to
J. Bernath, Edison Welding Institute, for providing
the welded specimens and assistance with process
development.

References
1. G. Welsch, R. Boyer and E. W. Collings (eds.): Materials
properties handbook: titanium alloys; 1994, Materials Park, OH,
ASM International.
2. R. S. Mishra and M. W. Mahoney (eds.): Friction stir welding and
processing; 2007, Materials Park, OH, ASM International.
3. R. S. Mishra and Z. Y. Ma: Mater. Sci. Eng. R, 2005, R50, 1
78.
4. R. Nandan, T. DebRoy and H. K. D. H. Bhadeshia: Prog. Mater.
Sci., 2008, 53, 9801023.
5. T. Trapp, E. Helder and P. R. Subramanian: in Friction
stir welding and processing II, 173178; 2003, San Diego, CA,
TMS.
6. K. V. Jata, P. R. Subramanian, A. P. Reynolds, T. Trapp and
E. Helder: Proc. Int. Offshore and Polar Engineering Conf.,
Toulon, France, May 2004, ISOPE, 2227.
7. W. B. Lee, C. Y. Lee, W. S. Chang, Y. M. Yeon and S. B. Jung:
Mater. Lett., 2005, 59, (26), 33153318.
8. Y. Shang, Y. S. Sato, H. Kokawa, S. H. C. Park and S. Hirano:
Mater. Sci. Eng. A, 2008, A488, 2530.
9. Y. Zhang, Y. S. Sato, H. Kokawa, S. H. C. Park and S. Hirano:
Mater. Sci. Eng. A, 2008, A485, 448455.
10. T. J. Lienert: in Friction stir welding and processing, Chapter 7,
123154; 2007, Materials Park, OH, ASM International.
11. D. Sanders, M. Ramulu, E. Klock-McCook, P. Edwards,
A. Reynolds and T. Trapp: J. Mater. Eng. Perform., 2008, 17,
(2), 187192.
12. D. Sanders, M. Ramulu and P. Edwards: Materialwiss.
Werkstofftech., 2008, 39, (45), 353357.
13. A. J. Ramirez and M. C. Juhas: Mater. Sci. Forum, 2003, 426432,
29993004.
14. E. J. Klock-McCook: Characterization of friction stir welded
and superplasticly formed friction stir welded joints of
titanium, Master thesis, University of Washington, Seattle, WA,
USA, 2005.
15. P. D. Edwards: Experimental and numerical characterization of
friction stir welded and superplastically formed friction stir
welded titanium, Master thesis, University of Washington, Seattle,
WA, USA, 2006.
16. R. John, K. V. Jata and K. Sadananda: Int. J. Fatigue, 2003, 25, (9
11), 939948.
17. A. P. Reynolds, E. Hood and W. Tang: Scr. Mater., 2005, 52, (6),
491494.

Science and Technology of Welding and Joining

2009

VOL

14

NO

483

You might also like