You are on page 1of 9

Materials Science and Engineering A 397 (2005) 376384

Microstructure and mechanical properties of friction stir welded SAF


2507 super duplex stainless steel
Y.S. Sato a, , T.W. Nelson b , C.J. Sterling b , R.J. Steel c , C.-O. Pettersson d
a

Department of Materials Processing, Graduate School of Engineering, Tohoku University, 6-6-02 Aramaki-aza-Aoba,
Sendai 980-8579, Japan
b Mechanical Engineering Department, Brigham Young University, 435 CTB, Provo, UT 84602, USA
c MegaStir Technologies, 275 West 2230 North, Provo, UT 84604, USA
d Sankvik Materials Technology, Sandviken S-811 81, Sweden
Received 18 January 2005; received in revised form 18 February 2005; accepted 25 February 2005

Abstract
The microstructure and mechanical properties of friction stir (FS) welded SAF 2507 super duplex stainless steel were examined. High-quality,
full-penetration welds were successfully produced in the super duplex stainless steel by friction stir welding (FSW) using polycrystalline
cubic boron nitride (PCBN) tool. The base material had a microstructure consisting of the ferrite matrix with austenite islands, but FSW
refined grains of the ferrite and austenite phases in the stir zone through dynamic recrystallisation. Ferrite content was held between 50 and
60% throughout the weld. The smaller grain sizes of the ferrite and austenite phases caused increase in hardness and strength within the stir
zone. Welded transverse tensile specimen failed near the border between the stir zone and TMAZ at the retreating side as the weld had roughly
the same strengths as the base material.
2005 Elsevier B.V. All rights reserved.
Keywords: Duplex stainless steel; Friction stir welding; Mechanical properties; Microstructure

1. Introduction
Duplex stainless steels have a mixed microstructure consisting of ferrite (bcc) and austenite (fcc) phases [1,2]. When
duplex stainless steels have the optimum phase balance,
which is usually approximately equal proportions of ferrite
and austenite phases, they exhibit higher resistance to stress
corrosion cracking and higher strength than austenitic stainless steels [1]. Taking advantages of these positive factors,
duplex stainless steels are widely used in the oil and gas,
petrochemical, pulp and paper, and pollution control industries [1].
It is well known that the duplex stainless steels exhibit
good weldability [1], but the melting and solidification associated with fusion welding processes destroy the favorable
duplex microstructure of these stainless steels. Microstruc

Corresponding author. Tel.: +81 22 795 7353; fax: +81 22 795 7352.
E-mail address: ytksato@material.tohoku.ac.jp (Y.S. Sato).

0921-5093/$ see front matter 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2005.02.054

ture of the wrought duplex stainless steels has a pronounced


orientation of austenite islands in the ferrite matrix, parallel and transverse to the rolling direction, but fusion welding
produces a microstructure consisting of coarse ferrite grains,
and both intergranular and intragranular austenite phases in
the weld metal and heat affected zone (HAZ) [15]. In general, the volume fraction of ferrite is much higher than that of
austenite in the weld metal and HAZ. These changes in microstructure cause the loss of low-temperature notch toughness and corrosion resistance in the weld [1,2]. To alleviate
these problems, careful control of the weld metal composition
and temperature are often required during welding.
Friction stir welding (FSW) is a solid-state joining process
developed and patented by The Welding Institute (TWI) in
UK in 1991 [6]. Since inception, FSW had been restricted to
the lower melting temperature materials, such as aluminum
(Al) and magnesium (Mg) alloys [716]. However, over the
past 5 years, much progress has been made in FSW of high
temperature materials by numerous investigators [1725].

Y.S. Sato et al. / Materials Science and Engineering A 397 (2005) 376384

377

These studies have reported that FSW achieves similar grain


refinement in the stir zone of the steels as those observed in
aluminum. Additionally, FSW does not accompany melting
and solidification, alleviating the formation of porosity and
adverse phase transformations during the welding process.
This situation suggests that the favorable duplex microstructure of the duplex stainless steel should be maintained, but
there have been few papers dealing with FSW of duplex stainless steel.
The present study used SAF 2507 super duplex stainless
steel as the base material. This particular super duplex stainless steel exhibits significantly higher corrosion resistance
than early grades of duplex stainless steels [1]. The nominal
chemical composition of the alloy used in this investigation
is Fe25Cr7Ni3.5Mo0.25NWCu (wt.%). The present
study applied FSW using polycrystalline cubic boron nitride
(PCBN) tool to SAF 2507 super duplex stainless steel. The
postweld microstructure and mechanical properties of the FS
weld were examined.

2. Experimental procedures
The base material used in the present study is a commercial
SAF 2507 (UNS 2750) super duplex stainless steel, 4 mm
in thickness. Test plates were prepared with dimensions of
300 mm in length and 100 mm in width. Light sanding of
the top surface and joint butt surface to remove oxide and
contaminants prior to welding was performed using 80 grit
emery cloth. Before FSW, each coupon was degreased with
a methanol solvent.
FSW tool used in this study had a shoulder diameter of
25 mm with the pin being 3.8 mm in length. The shoulder and
pin section of the tool were manufactured from solid PCBN.
A locking collar was used to hold the PCBN and transfer
torque from a tungsten carbide shank as shown in Fig. 1.
FS welds were completed on a vertical milling machine fitted with servomotors and control system. Since the machine
and tool were exposed to high temperatures during FSW, a
liquid cooled tool holder equipped with telemetry system to
broadcast tool temperature was used for the weld trails. An
argon atmosphere was introduced through a gas cup around
the tool at a flow rate of 2.8 105 mm3 /s (1 m3 /h) to avoid
the surface oxidation. A 3.5 tilt was applied to the tool during FSW. The welding direction (WD) was identical to the
rolling direction (RD) of the plate. The welding parameters
were: rotational speed of 450 rpm and weld travel speed of
1 mm/s.
Microstructure in the weld was examined by optical microscopy (OM) and orientation imaging microscopy
(OIMTM ) [26,27]. Sample for OM examination were electrolytically etched in a 10 wt.% oxalic acid solution at 30 V
for 20 s. Details of the microstructure were examined by
OIM. Cross section for OIM was cut perpendicular to the
welding direction and then electrolytically polished in 20 ml
HClO4 + 180 ml C2 H5 OH solution at 223 K (50 C). Crys-

Fig. 1. PCBN friction stir welding tool assembly.

tallographic data were obtained from several regions, which


are illustrated as solid squares in Fig. 2. The notation CEN
means the weld centre, which was defined as the centre of region swept by the shoulder. The notations ASn and RSn
indicate that the location of analysis is n mm away from the
weld centre at the advancing and retreating sides, respectively. These notations are used throughout the present paper.
Crystallographic data collection by OIM was performed in a
PHILIPS XL30-SFEG scanning electron microscope (SEM),
operating at 30 kV under step size of 0.6 m. Each observation area was 150 m 150 m. Crystallographic data were
expressed by phase map with grain boundaries and {1 1 1}
pole figure. In the phase map, austenite and ferrite phases
were colored with gray and white, respectively, and the thick
and thin black lines show grain boundaries with misorientation exceeding 15 and misorientation between 3 and 15 ,
respectively. Coordinate axes of {1 1 1} pole figures are WD
(or RD for the base material), transverse direction (TD) and
normal direction (ND) of the plate, as shown in Fig. 2.
Vickers hardness test was conducted on the cross section
perpendicular to the welding direction, using a Vickers indenter with a 9.8 N load for 10 s, to examine the distribution

378

Y.S. Sato et al. / Materials Science and Engineering A 397 (2005) 376384

Fig. 2. Schematic illustration of locations analyzed by OIM on cross section perpendicular to the welding direction.

of hardness originating with the duplex microstructure consisting of ferrite and austenite. Transverse tensile specimens
were removed perpendicular to the welding direction and
prepared in accordance with ASTM E8. Tensile tests were
carried out at room temperature on a 445 kN MTS tensile
testing machine at a crosshead speed of 0.05 mm/s. A 51 mm
extensometer was used to determine the 0.2% offset yield
strength.

3. Results and discussion


3.1. Microstructure distribution of the weld
Low-magnification overview of friction stir welded SAF
2507 duplex stainless steel is presented in Fig. 3. In the cross
section, the left- and right-hand sides of the weld center are
consistent with retreating and advancing sides of the rotating

Fig. 3. Cross section perpendicular to the welding direction of friction stir welded super duplex stainless steel 2507. An arrow shows the fracture location for
the transverse tensile test.

Fig. 4. Optical micrograph and phase map obtained by OIM of the base material.

Y.S. Sato et al. / Materials Science and Engineering A 397 (2005) 376384

tool, respectively [2833]. The stir zone is seen around the


weld center. The border between the stir zone and the thermomechanically affected zone (TMAZ) is very distinct on the
advancing side, while it is more diffuse on the retreating side.
It is apparent that the weld interior exhibits a high degree of
continuity and no defects.
An optical micrograph and a phase map of the as-received
base material are shown in Fig. 4. The base material has a
typical microstructure of wrought duplex stainless steels consisting of ferrite matrix with austenite islands. OIM analysis
revealed that the austenite islands contained a higher number
of grain boundaries (mostly twin type boundaries) than the
ferrite. OIM analysis also revealed that the ferrite content was
about 51%. Average grain sizes of austenite and ferrite phases
in the base material were about 4.3 and 5.1 m, respectively.
Optical microstructures of regions A, B, C and D
shown in Fig. 3 are indicated in Fig. 5. Region B lies on the
weld centre, and region D is located on the border of the stir

379

zone and TMAZ. Regions A and C are located around 2 mm


away from the weld centre at the retreating and advancing
sides, respectively. Region B has the microstructure consisting of ferrite matrix with the more elongated austenite islands.
Austenite islands of region B look finer than those of the base
material. Region A has the similar microstructure to region
B, while region C seems to contain finer austenite islands
than region B. Distribution of the austenite islands is finest in
the stir zone at the advancing side, as shown in micrograph
of region D. In this region, D, the austenite in the stir zone
exhibits an average grain size of 2.2 m lying immediately
adjacent to elongated austenite islands in the TMAZ.
Phase maps of regions RS4.5, CEN and AS4.5 in the weld
are shown in Fig. 6. All regions consist of a ferrite matrix with
the austenite islands similar to that of the base material. Distribution of the austenite islands in regions RS4.5 and CEN
is similar to that in the base material, but the austenite islands
contain more grain boundaries than the base material. Region

Fig. 5. Optical microstructures of regions A, B, C and D shown in Fig. 2.

380

Y.S. Sato et al. / Materials Science and Engineering A 397 (2005) 376384

Fig. 6. Phase maps of regions RS4.5, CEN and AS4.5 in the weld.

AS4.5 has more finely distributed austenite islands than the


base material, and the austenite islands of this region lie along
the border of the stir zone/TMAZ. Again, the average grain
size of the austenite phase was finest in this region, averaging
2.2 m. In all regions, the grain size of the austenite phase
was smaller than that of the ferrite phase.
Phase map of region AS5.5, located just outside the stir
zone, is shown in Fig. 7. This has the roughly same morphology of austenite islands as the base material. However,
the ferrite matrix in this region contains a relatively higher
density of sub-boundaries than that in the base material. Av-

erage grain size of the austenite phase in this region was about
4.0 m, which is the roughly same as that of the base material, while average grain size of the ferrite phase was about
4.0 m, smaller than that of the base material.
Grain size profiles of the austenite and ferrite phases are
indicated in Fig. 8. The austenite phase exhibits a smaller
grain size than the ferrite phase throughout the weld. In the
stir zone, grain sizes of austenite and ferrite phases decrease
from the retreating side towards the advancing side. The finest
austenite and ferrite grains are observed in the stir zone near
the TMAZ. Ferrite content profile across the stir zone in the
weld is shown in Fig. 9. Ferrite content varies between 50 and
60% across the weld and TMAZs, which is slightly higher
than that of the base material. The higher ferrite content in
the stir zone is attributed to the exposure to high temperatures during FSW and the fact ferrite is more stable than the
austenite at the higher temperatures in the duplex stainless
steel [1].
{1 1 1} pole figures of the ferrite and austenite phases in
several regions in the weld are presented in Fig. 10. The base

Fig. 7. Phase map of region AS5.5 in the weld.

Fig. 8. Grain size profiles of the austenite and ferrite phases in the weld.

Y.S. Sato et al. / Materials Science and Engineering A 397 (2005) 376384

Fig. 9. Profile of ferrite content in the weld.

material has the weak {1 1 0}1 1 2 orientation in the austenite phase and the relatively strong {0 0 1}1 1 0 orientation
in the ferrite phase, where {h k l} is the crystallographic plane
perpendicular to the ND and u v w is the crystallographic
direction parallel to RD. Regions RS4.5, CEN and AS4.5
have different texture components, both in the austenite and
ferrite phases, from those observed in base material. These
changes in texture are indicative of the intense deformation
of material during FSW [29,30,32]. In region AS5.5, on the
other hand, the texture component in the austenite phase is
roughly the same as that of the base material, while the ferrite phase exhibits a different texture component unlike that
of the base material.
3.2. Microstructural evolution during FSW
During FSW, the duplex stainless steel is heated up to
high temperatures created by frictional heating arising from

381

rotation of the welding tool. It is well known that the microstructure of duplex stainless steel changes to the fully
ferritic structure when the steel is exposed to temperatures
higher than about 1573 K (1300 C) [1]. Numerical calculation using the present chemical composition and the thermochemical database [34] estimated that formation of the
fully ferritic microstructure required temperature exceeding
about 1600 K (1327 C) in the present steel. The formation of
the fully ferritic microstructure is followed by nucleation of
austenite phase at grain boundaries of the ferrite grains during the weld cooling cycle. Therefore, the stir zone should
have the coarse ferrite matrix with intergranular austenite
phases, if it is heated up to temperatures higher than 1600 K
(1327 C) during FSW. In the present study, however, several
microstructural analyses show that all regions in the weld
have the microstructures consisting of the ferrite matrix with
austenite islands, although the distribution and morphology
of the austenite islands are inhomogeneous in the weld. These
results indicate that the maximum temperature during FSW
did not exceed 1600 K (1327 C).
During FSW, the material in the stir zone simultaneously
experiences large shear stresses along the pin tool surface, as
well as exposure to high temperatures [29,30,32]. A previous study [33] investigating material flow of FSW using an
in situ marker technique illustrated that material located at
the advancing side of the stir zone is displaced around the
entire circumference of the pin tool, ending up at roughly
the same location it began, displaced only slightly forward
of its original position. As a result, this report suggests that
material in the stir zone at the advancing side receives the
most severe deformation, which would tend to break up the
austenite islands producing a smaller grain size. This corresponds well with the microstructures reported herein, in that
the finest austenite grains are observed along the advancing
side of the stir zone.

Fig. 10. {1 1 1} pole figures of the austenite and ferrite phases in the base material (BM), regions RS4.5, CEN, AS4.5 and AS5.5 in the weld.

382

Y.S. Sato et al. / Materials Science and Engineering A 397 (2005) 376384

The grain size profile (Fig. 8) showed that the austenite and
ferrite grains in the stir zone were smaller than those in the
base material. Additionally, the phase maps (Fig. 6) showed
that both the austenite and ferrite phases in the stir zone did
not exhibit a heavily deformed microstructure, e.g. many low
angle grain boundaries. Both the grain size profile and phase
maps suggest that dynamic recrystallisation occurred both
in the austenite and ferrite phases during FSW. It is generally known that dynamic recrystallisation easily occurs in the
austenitic stainless steels, while ferritic steels hardly experience dynamic recrystallisation because the ferrite phase has
a high stacking fault energy [35].
In the case of the duplex stainless steels, however, deformation is localized in the ferrite matrix at high temperatures, because the ferrite phase is relatively weaker than
the austenite phase [36,37]. Consequently, the recrystallised
grains are often formed in ferrite phase more easily than
in austenite phase [3639]. Some research [36,37] suggests
that the recrystallised grains in the ferrite phase are formed
by continuous dynamic recrystallisation, which is characterized by strain-induced progressive rotation of subgrains
with little boundary migration. In the present study, the duplex stainless steel experienced plastic deformation by the
rotating tool at relatively high temperatures during FSW. As
such, it is likely that the ferrite matrix in the stir zone undergoes continuous dynamic recrystallisation through the same
scenario.
On the other hand, the morphology of austenite islands
in the stir zone was much different from that of the base
material, as shown in Figs. 5 and 6. This suggests that the
austenite islands also experienced intense plastic strain during FSW, leading to dynamic recrystallisation in the austenite
phase.
After dynamic recrystallisation, the recrystallised grains
grow during the on-cooling thermal cycle [40,41]. As mentioned above, since the ferrite phase is more likely to undergo
dynamic recrystallisation than the austenite phase, as a result
of the high temperature deformation of FSW, it is likely that
the recrystallised ferrite grains would grow early than those
of the austenite phase. This is likely the reason why the ferrite
phase exhibits a larger grain size than the austenite phase in
the stir zone.
Region AS5.5, located just outside the stir zone, had the
similar morphology of austenite islands to the as-received
base material. Grain size and texture of the austenite phase
in region AS5.5 were roughly the same as those in the base
material, but the ferrite phase had the different texture components than that of the base material. This result suggests
that only the ferrite matrix underwent deformation and recrystallisation during FSW because the ferrite phase has a
lower flow stress at elevated temperature than the austenite phase in the duplex stainless steel, as mentioned above.
It is generally known in Al alloys that dislocations introduced into the TMAZ rearranged or migrate at temperature producing a recovered microstructure [4245]. However, the ferrite phase in region AS5.5 did not exhibit a de-

Fig. 11. Hardness profile across the stir zone in the weld.

formed microstructure. This result suggests that the ferrite


phase in this region may undergo sufficient plastic strains
to induce continuous dynamic recrystallisation as a result
of the significant localisation of deformation in the ferrite
phase.
3.3. Mechanical properties and effect of microstructure
on mechanical properties
A typical transverse hardness profile of a FSW in 2507
super duplex stainless steel is indicated in Fig. 11. Given the
fact that the ferrite content in the stir zone is roughly uniform
(Fig. 9), the increase of hardness in the stir zone suggests that
the hardness profile is related to the grain sizes of ferrite and
austenite phases in the weld (Fig. 8).
Transverse tensile properties of the weld are shown in
Fig. 12(a). The as-FSW 2507 super duplex exhibits roughly
the same 0.2% offset yield and ultimate tensile strengths as
the base material, with the exception of the elongation. Total elongation to failure based on the standard 51 mm gauge
length was roughly 50% of the base material. However, given
the amount of reduction in area as observed in Fig. 12(b),
the actual percentage of ductility of the FSW specimens is
likely much higher than reported. All tensile failures occurred roughly 7 mm from the weld centre at the retreating side, i.e. near the border of the stir zone and TMAZ,
as shown in Fig. 12(b). This is consistent with the data presented in Fig. 11, where the failures tend to move toward the
TMAZ as a result of the higher hardness, which is proportional to strength in the metallic materials [9,46], of the stir
zone.
It should be pointed out that the tensile specimens used in
this study did not have uniform thickness (see Fig. 3) across
the weld. Typically, the thinnest section of a FSW is located
at the centerline of the weld, as a result of the tilt angle used
during welding. This is a reason why the transverse tensile
samples consistently fractured near the border of the stir zone
and TMAZ, rather than in the base material, which exhibited
the lowest hardness.

Y.S. Sato et al. / Materials Science and Engineering A 397 (2005) 376384

383

Fig. 12. (a) Transverse tensile properties of the weld and (b) cross section of the fractured tensile sample.

4. Conclusions
The present study examined the microstructure and mechanical properties of FSW in SAF 2507 super duplex stainless steel. FSW using PCBN tool produced high-quality
welds in the super duplex stainless steel. FSW significantly
refined the ferrite and austenite phases through dynamic recrystallisation. The smaller ferrite and austenite grains created increased hardness and strength in the stir zone. As a
result, weld transverse tensile failures consistently occurred
near the border of the stir zone and TMAZ, exhibiting roughly
the same yield and ultimate tensile strengths as the base material.

Acknowledgements
The authors are grateful to Mr. J.W. Pew and Mr. J.N.
Ostler for technical assistance and acknowledge Prof. K.
Ishida and Dr. I. Ohnuma for the thermochemical calculation. They also wish to thank Prof. H. Kokawa, Dr. S.H.C.
Park and Dr. J.-Q. Su for their helpful discussion.

References
[1] ASM Specialty Handbook Stainless Steels, ASM International, Materials Park, 1994, pp. 383388.
[2] J. Hilkes, K. Bekkers, Weld. J. 74 (1995) 5154.
[3] H. Matsunaga, Y.S. Sato, H. Kokawa, T. Kuwana, Sci. Technol.Weld.
Joining 3 (1998) 225232.

[4] Y.S. Sato, H. Kokawa, T. Kuwana, Sci. Technol. Weld. Joining 4


(1999) 4149.
[5] Y.S. Sato, H. Kokawa, Scr. Mater. 40 (1999) 659663.
[6] C.J. Dawes, W.M. Thomas, Weld. J. 75-3 (1996) 4145.
[7] M.W. Mahoney, C.G. Rhodes, J.G. Flintoff, R.A. Spurling, W.H.
Bampton, Metall. Mater. Trans. A 29A (1998) 19551964.
[8] Y.S. Sato, H. Kokawa, M. Enomoto, S. Jogan, T. Hashimoto, Metall.
Mater. Trans. A 30A (1999) 31253130.
[9] Y.S. Sato, H. Kokawa, Metall. Mater. Trans. A 32A (2001)
30233031.
[10] B. Heinz, B. Skrotzki, Metall. Mater. Trans. B 33B (2002) 489498.
[11] K.A.A. Hassan, A.F. Norman, D.A. Price, P.B. Prangnell, Acta
Mater. 51 (2003) 19231936.
[12] J.-Q. Su, T.W. Nelson, R. Mishra, M. Mahoney, Acta Mater. 51
(2003) 713729.
[13] S.H.C. Park, Y.S. Sato, H. Kokawa, Scr. Mater. 49 (2003) 161166.
[14] H.J. Liu, H. Fujii, M. Maeda, K. Nogi, J. Mater. Process. Technol.
142 (2003) 692696.
[15] Y.S. Sato, Y. Sugiura, Y. Shoji, S.H.C. Park, H. Kokawa, K. Ikeda,
Mater. Sci. Eng. A A369 (2004) 138143.
[16] Y.S. Sato, M. Urata, H. Kokawa, K. Ikeda, Mater. Sci. Eng. A A354
(2003) 298305.
[17] W.M. Thomas, P.L. Threadgill, E.D. Nicholas, Sci. Technol. Weld.
Joining 4 (1999) 365.
[18] T.J. Lienert, W.L. Stellwag, B.B. Grimmett, R.W. Warke, Weld. J.
82 (2003) 1s.
[19] A.P. Reynolds, W. Tang, T. Gnaupel-Harold, H. Prask, Scr. Mater.
48 (2003) 12891294.
[20] K. Okamoto, S. Hirano, M. Inagaki, S.H.C. Park, Y.S. Sato, H.
Kokawa, T.W. Nelson, C.D. Sorensen, Proceedings of the 4th International Symposium on Friction Stir Welding, Park City, USA, May
2003, TWI, CD-ROM.
[21] S.H.C. Park, Y.S. Sato, H. Kokawa, K. Okamoto, S. Hirano, M.
Inagaki, Scr. Mater. 49 (2003) 11751180.
[22] S.H.C. Park, Y.S. Sato, H. Kokawa, K. Okamoto, S. Hirano, M.
Inagaki, Scr. Mater. 51 (2004) 101105.

384

Y.S. Sato et al. / Materials Science and Engineering A 397 (2005) 376384

[23] C.D. S Sorensen, Proceedings of Friction Stir Welding Symposium


at the 14th International Society of Offshore and Polar Engineers
(ISOPE 2004), vol. IV, Toulon, France, May 2328, 2004, pp. 8
14.
[24] A. Ozekcin, H.W. Jin, J.Y. Koo, N.V. Bangaru, R. Ayer, G. Vaughn,
R. Steel, S. Packer, Int. J. Offshore Polar Eng. 14 (2004) 284288.
[25] Y.S. Sato, T.W. Nelson, C.J. Sterling, Acta Mater. 53 (2004)
637645.
[26] B.L. Adams, S.I. Wright, K. Kunze, Metall. Trans. A 24A (1993)
819831.
[27] D.J. Dingley, D.P. Field, Mater. Sci. Technol. 13 (1997) 6978.
[28] K. Colligan, Weld. J. 78 (1999) 229s237s.
[29] Y.S. Sato, H. Kokawa, K. Ikeda, M. Enomoto, S. Jogan, T.
Hashimoto, Metall. Mater. Trans. A 32A (2001) 941948.
[30] D.P. Field, T.W. Nelson, Y. Hovanski, K.V. Jata, Metall. Mater. Trans.
A 32A (2001) 28692877.
[31] T.U. Seidel, A.P. Reynolds, Metall. Mater. Trans. A 32A (2001)
28792884.
[32] S.H.C. Park, Y.S. Sato, H. Kokawa, Metall. Mater. Trans. A 34A
(2003) 987994.
[33] B. London, M. Mahoney, W. Bingel, M. Calabrese, R.H. Bossi,
D. Waldron, in: K.V. Jata, et al. (Eds.), Friction Stir Welding and
Processing II, TMS, Warrendale, 2003, pp. 312.

[34] Scientific Group Thermodata Europe: SGTE Solution Database version 2.1, 1999.
[35] T. Maki, T. Furuhara, K. Tsuzaki, ISIJ Intern. 41 (2001) 571579.
[36] K. Tsuzaki, H. Matsuyama, M. Nagano, T. Maki, J. Jpn. Inst. Met.
54 (1990) 878887.
[37] W. Reick, M. Pohl, A.F. Padilha, ISIJ Intern. 38 (1998) 567571.
[38] Y. Maehara, Trans. Iron Steel Inst. Jpn. 27 (1987) 705.
[39] J. Keichel, G. Gottstein, J. Foct, Mater. Sci. Forum 318320 (1999)
785792.
[40] Y.S. Sato, M. Urata, H. Kokawa, Metall. Mater. Trans. A 33A (2002)
625635.
[41] Y. Li, L.E. Murr, J.C. McClure, Mater. Sci. Eng. A A271 (1999)
213223.
[42] Y.S. Sato, H. Kokawa, M. Enomoto, S. Jogan, Metall. Mater. Trans.
A 30A (1999) 24292437.
[43] Y.S. Sato, M. Urata, H. Kokawa, K. Ikeda, M. Emonoto, Scr. Mater.
45 (2001) 109114.
[44] Y.S. Sato, S.H.C. Park, H. Kokawa, Metall. Mater. Trans. A 32A
(2001) 30333042.
[45] Y.S. Sato, M. Urata, H. Kokawa, K. Ikeda, Scr. Mater. 47 (2002)
869873.
[46] M.F. Ashby, D.R.H. Jones, Engineering Materials 1, Pergamon Press,
Oxford, 1980, p. 105.

You might also like