Sie sind auf Seite 1von 9

Materials and Design 87 (2015) 405413

Contents lists available at ScienceDirect

Materials and Design


journal homepage: www.elsevier.com/locate/jmad

Microstructure and residual stress distributions in friction stir welding of


dissimilar aluminium alloys
Hamed Jamshidi Aval
Department of Mechanical Engineering, Babol University of Technology, Shariati Avenue, PO Box 484, Babol, Iran

a r t i c l e

i n f o

Article history:
Received 19 January 2015
Received in revised form 4 August 2015
Accepted 11 August 2015
Available online 14 August 2015
Keywords:
Friction stir welding
Dissimilar aluminium alloys
Residual stress
Microstructures

a b s t r a c t
The aim of this investigation was to study the effect of welding heat input and postweld natural aging on residual
stress, microstructure, and precipitation distribution in different zones of dissimilar friction stir welding of 8 mm
thick plates of AA6082-T6 and AA7075-T6. It was found that atomic diffusion occurs at the interface of the materials in the stir zone of the joints. Transmission electron microscopic investigations showed that reprecipitation of
ne GuinierPreston zone, , and precipitates resulted in increased micro-hardness in the SZ after natural
aging. An increase in welding heat input resulted in decreased maximum tensile residual stress and increased
size of the tensile residual stress region. Natural aging within the SZ and thermo-mechanical affected zone
resulted in 1520 MPa reduction of the residual stress in these zones.
2015 Elsevier Ltd. All rights reserved.

1. Introduction
Friction stir welding (FSW) is a solid-state welding technique, developed by The Welding Institute (TWI), UK, which has demonstrated very
high suitability for the joining of dissimilar aluminium alloys. As there is
no melting during this process, all the defects related to the presence of
brittle interdendritic and eutectic phases are eliminated; thus, friction
stir welds may be produced with superior properties when compared
with fusion welds [1]. The strength and microstructure of friction stir
welded joints are dependent on the employed welding parameters
such as rotational and welding speed, geometric parameters of the
tool and the work piece, and initial material temperature. In this regard,
some studies on dissimilar FSW of aluminium alloys including AA6xxx
and AA7xxx series were reported in the literatures [25]. For instance,
Steuwer et al. [6] have examined the effect of process parameters on
residual stress in dissimilar FSW of AA5083AA6082. They showed
that the rotational speed of the welding tool affects the residual stress
in the weld more signicantly when compared with the transverse
speed. Moreover, they reported that the residual stress on the AA5083
side is larger when compared with the AA6082 side. Jonckheere et al.
[7] studied the torque, temperature, and hardening precipitation evolution of 4.7 mm thick similar and dissimilar friction stir welds between
6061-T6 and 2014-T6 aluminium alloys. They found that the power required to perform dissimilar welds lies in between the torques required
to perform similar welds of both alloys, and so do the temperatures.
Furthermore, they reported that an increase in temperature causes an
increase in precipitate's radius and a decrease in precipitate's volume

E-mail address: h.jamshidi@nit.ac.ir.

http://dx.doi.org/10.1016/j.matdes.2015.08.050
0264-1275/ 2015 Elsevier Ltd. All rights reserved.

fraction leading to lower yield strength. Ouyang and Kovacevic [8]


studied the material ow and microstructural evolution in the dissimilar friction stir welded AA6061AA2024 joint. They classied the nugget
zone into three different regions: the mechanically mixed region
(MMR) characterized by relatively dispersed particles of different
alloy constituents, the stirring-induced plastic ow region (SPFR)
consisting of alternative vortex-like lamellae of the two Al-alloys, and
the unmixed region (UMR) consisting of ne equiaxed grains of the
6061 aluminium alloy. They reported that the degree of material
mixing, the thickness of the deformed aluminium alloy lamellae, and
the material ow patterns depend on the related positions in the nugget
zone and the processing parameters.
Prime et al. [9] studied the residual stress prole of 25.4 mm thick
dissimilar friction stir welded AA7050AA2024. They reported that
the stress in the test specimen peaked only at about 32 MPa and had
the conventional M prole with tensile stress peaking in the heataffected zone (HAZ) outside the weld. Da Silva et al. [10] have evaluated
the effect of joining parameters on the mechanical properties,
microstructural features, and material ow of 3 mm thick dissimilar
friction stir welded AA2024AA7075. They found that the HAZ at the
retreating side showed the minimum hardness value. Moreover, they
have reported that high rotation speed yielded intense onion ring-like
mixing pattern in the stir zone (SZ), but a reduced surface integrity
and limited material mixing at low rotation speeds has been observed.
Cavaliere et al. [11] studied the mechanical and microstructural properties of dissimilar 2024 and 7075 aluminium sheets joined by FSW. They
reported that the presence of the FSW line reduces the fatigue behavior,
but the comparison to the parent material is acceptable.
Cole et al. [12] have considered the inuences of tool offset and
toolwork piece interface temperature during dissimilar FSW of

406

H. Jamshidi Aval / Materials and Design 87 (2015) 405413

Table 1
The alloy chemical composition (wt.%).

Table 3
Welding parameters used in the experiments.

Alloy

Zn

Mg

Cu

Mn

Si

Fe

Ti

Cr

Al

Designation

Rotational speed (rpm)

Welding speed (cm/min)

AA6082-T6
AA7075-T6

0.20
5.21

1.12
2.30

0.14
1.32

0.91
0.10

1.20
0.21

0.34
0.37

0.08
0.12

0.04
0.18

Balancing
Balancing

Sample A
Sample B
Sample C
Sample D

1000
1000
1200
1200

9
12
9
12

Table 2
Mechanical properties of the base material from tensile tests.
Alloy

Yield stress (MPa)

Tensile strength (MPa)

AA6082-T6
AA7075-T6

288
498

337
560

AA6061AA7075. They reported that the weld tool offsets into the
AA7075 side, and increasing amount of AA7075 stirred into the nugget
increases the measured tensile strength. Guo et al. [13] studied the effects of heat input on the material ow, microstructure, and mechanical
properties of the dissimilar friction stir welded AA6061AA7075. They
found that in conditions that AA6061 is located in the advancing side,
better mixing of the material in the SZ was achieved. Furthermore,
they reported that the tensile strength of the joints increases with the
decrease of heat input and corresponds very well to the minima in the
micro-hardness prole. Bala Srinivasan et al. [14] investigated the
microstructural aspects and stress corrosion cracking (SCC) behavior
of dissimilar friction stir welded AA7075AA6056 joint. They found
that the mechanical mixing of the materials at elevated temperatures
in the SZ has led to the evolution of a ne-grained recrystallized structure, but the dwell time at these high temperatures was insufcient
for diffusion of elements during friction stir processing. Moreover,
they reported that though the weld nugget is resistant to SCC, the
thermo-mechanical affected zone (TMAZ)/HAZ region of AA7075 is
prone to SCC in 3.5% sodium chloride (NaCl) solution at lower strain
rate than 107/s. pekolu and am [15] have considered the effects of
initial temper conditions and postweld heat treatment (PWHT) on the
microstructure and mechanical properties of the dissimilar friction stir
welded AA7075AA6061 joint. They showed that PWHT could be used
in order to improve the joint properties for both O and T6 joints.
AA6082 and AA7075 Al alloys are increasingly used in automotive,
rail transportation and aerospace industries. Joining of these two alloys
is required in some of these applications, while AA7075 and AA6082

alloys cannot be welded by conventional welding processes, as fusion.


FSW as a solid state welding process is capable of producing highquality welds when optimized parameters are used. In this work, the
effects of rotational and linear speeds of tool on residual stress prole
in dissimilar FSW of AA6082-T6 and AA7075-T6 are examined, and
the developed microstructures and precipitation distribution in the
weld zone as well as the mechanical properties of the joints are evaluated. Accordingly, X-ray diffraction (XRD) residual stress analysis and tensile testing together with optical metallography and transmission
electron microscopy (TEM) are performed to assess the effects of
process parameters on welded joints.

2. Experimental procedures
In this study, aluminium alloy plates AA7075-T6 and AA6082-T6 of
8 mm thickness were investigated. The composition in weight percent
and mechanical properties of both alloys are presented in Tables 1 and
2. Welds were produced using a FSW machine in load control with a
tilt angle and axial load of 2 and 12 kN, respectively. The weld conguration had the AA6082 alloy placed on the advancing side and the
AA7075 alloy on the retreating side. Single-pass friction stir butt welds
were made using a tool made of H13 steel with a shoulder 23 mm in diameter and triangular frustum pin. The dimensions of the tool used in
this study are shown in Fig. 1. The tool rotational speeds of 800, 1000,
1200, and 1400 rpm and linear speeds of 9, 12, and 15 cm/min were applied in the experiments, but only the welds made by using the rotating
speeds of 1000 and 1200 rpm and the linear speeds of 9 and 12 cm/min
were satisfactory. Therefore, the last rotational and linear speeds were
investigated in this study. The welded samples were named according
to Table 3. The thermal history of various regions in the advancing
and retreating sides of the joints was measured using six K-type
thermocouples with a 0.25 mm diameter wire as shown in Fig. 2.

Fig. 1. The tool geometry used in this study.

H. Jamshidi Aval / Materials and Design 87 (2015) 405413

407

Table 4
Measured peak temperature and grain sizes of stir zone in the advancing and retreating
sides of various samples.
Sample

A
B
C
D

Fig. 2. Schematic view of the samples employed in welding experiments and the thermocouple positions.

To determine the microstructure of the joints, the welded samples


were transversely sectioned and polished. The optical metallography
was performed on polished samples using a reagent composed of 3 ml
nitric acid (HNO3), 6 ml hydrouoric acid (HF), 6 ml hydrochloric acid
(HCl), and 150 ml H2O. In addition, thin sheets for TEM were prepared
by means of mechanical polishing (thinning the sample piece to about
0.2 mm) and double jet electropolishing. The mechanical polished samples were electropolished using a solution of 25% HNO3 in methanol at a
voltage of 12 V and temperature of 20 2 C. Microstructural characterization was performed using an optical microscope Olympus PME3
and Tecnai G2 transmission electron microscope operating at 200 kV.
Besides, distribution of elements in the SZ was studied by Electron

Grain size of SZ (m)

Peak temperature (C)

Advancing side

Retreating side

Advancing side

Retreating side

7
4.9
8.4
6.1

5.9
3.4
7.2
4.7

471
446
493
463

460
431
481
449

Probe Micro-Analyzer (EPMA). The linear intercept method based on


ASTM: E112-13 was used to nd the average grain size.
The Vickers micro-hardness of the joints was measured along the
mid-thickness of the plates with a 100 gf loading for 15 s. Tensile testing
was conducted on the specimens with a 50 mm gauge length and
12.5 mm width which were prepared according to ASTM: E8M-13
standard at a constant cross-head speed of 1 mm/min. The deformation
behavior of tensile samples was studied by a digital image correlation
(DIC) method. The longitudinal and transversal residual stress distributions at 30 m depth from the top surface of the samples and perpendicular to welding direction were measured by XRD technique using CrK
radiation with the X-ray tube operating at 20 kV and 4 mA target
currents.
3. Results and discussion
The macrostructure of cross-section of the abovementioned specimens is shown in Fig. 3. It can be seen that no onion ring feature was
found in the SZ. From Fig. 3(c) and (d) it is apparent that higher rotation
speeds and lower traverse speeds result in the material mixing of complex patterns. However, this phenomenon is not more sensible at the
lower rotational speed of 1000 rpm. As shown in Table 4, the better

Fig. 3. Macrostructure of welded samples; a) sample A (1000 rpm, 9 cm/min), b) sample B (1000 rpm, 12 cm/min), c) sample C (1200 rpm, 9 cm/min), and d) sample D
(1200 rpm, 15 cm/min).

408

H. Jamshidi Aval / Materials and Design 87 (2015) 405413

Fig. 4. Zinc content in samples A and C, measured in the points indicated in Fig. 3.

mixing of the materials in the weld nugget may be attributed to the


higher heat generation and peak temperature of samples A and C than
the others. Although, mixing of the materials qualitatively can be
detected due to the different etching behaviors of the base materials,
to evaluate the distribution of elements in the cross-section of welds,
EPMA analysis was performed. The quantitative analysis of zinc in samples A and C on the points specied in Fig. 3(c) and (d) is shown in Fig. 4.
Comparing the results, it is possible to conrm that zones 1 and 9 correspond to the base materials AA6082 and AA7075, respectively. The zinc
content of zones 2 and 3 suggests that these zones are composed basically of AA6082, while zones 7 and 8 composed basically of AA7075.
The zinc content of regions 4, 5, and 6 suggests that these zones are
formed by a mixture of both materials. The zinc content between two
base materials can be related to atomic diffusion at the interface of
these joints. The materials in the weld nugget experience frictional
and plastic deformation heating and intense plastic deformation during
FSW. Therefore, the diffusion rate of zinc atoms due to severe plastic deformation of the material and high peak temperature in the weld nugget

Fig. 5. Microstructure of different zones of sample B; a) BM of AA6082, b) BM of AA7075, c) HAZ of AA6082, d) HAZ of AA7075, e) TMAZ of AA6082, f) TMAZ of AA7075, g) SZ of AA6082,
and h) SZ of AA7075.

H. Jamshidi Aval / Materials and Design 87 (2015) 405413

409

Fig. 6. a) Micro-hardness vs. aging time of sample B, b) comparison between the micro-hardness of samples B and C after 365 days of natural aging.

can result in the increase of the diffusion rates higher than that in the
static condition. Similar results for the atomic diffusion in the interface
of samples during FSW of Al and Mg alloys have been previously reported [16].
Fig. 5 illustrates the detailed microstructure in various zones of
sample B. Fig. 5(a) and (b) presents the base metal microstructure of
AA6082 and AA7075, respectively. The average grain sizes of AA6082
and AA7075 are approximately 85 m and 65 m, respectively. The microstructures of the HAZ in both sides are similar to the base materials
as shown in Fig. 5(c) and (d). Fig. 5(e) and (f) represents the TMAZ on
the advancing and retreating sides, respectively. The SZ and HAZ were
bounded by the TMAZ, where bent and elongated grains were observed.
As can be seen in the images, the bent and elongated TMAZ grains in the
advancing side due to large relative deformation in this zone [17] are
more visible than those in the retreating side. The microstructure of
the SZ experiences high temperature and severe plastic deformation,
thus the microstructure of this zone is characterized by recrystallized
equiaxed grains, while the grain size in the AA7075 side is ner than
that in the AA6082 side (Fig. 5(g), (h) and Table 4). The SZ grain size
difference was also reported by Guo et al. [13] and Bala Srinivasan
et al. [14] in dissimilar FSW of 6000 and 7000 series aluminium alloys.
They attributed this difference to the initial grain size of the base

materials [14] and the higher content of second phase particles in


AA7075 alloy [13]. Besides, the SZ grain size in the FSW also was affected
by plastic deformation rate and by welding temperature [18,19].
Therefore, the coarser SZ grain size of AA6082 may be due to ner
base material grain size and higher alloying elements (and higher second phase particles) of AA7075 than AA6082 and higher temperature
in the advancing side (Table 4). As seen in Table 4, increasing ratio of
rotational to linear speed enhances the SZ grain size and peak temperature in both advancing and retreating sides; for example, the grain sizes
of the SZ in the AA6082 side of sample B with welding parameters
1000 rpm and 12 cm/min and sample C with welding parameters
1200 rpm and 9 cm/min are 4.9 m and 8.4 m, respectively. The plastic
deformation rate and welding temperature are two parameters that act
inversely on dynamic recrystallized grain size. In other words, the ne
SZ grain size is produced by higher plastic deformation rate and lower
welding temperature. Although increasing rotational to linear speed
induced more plastic deformation rate, the peak temperature effect on
the SZ grain size is dominant on deformation rate at the range of
rotational and linear speeds investigated in this paper.
Fig. 6(a) shows the micro-hardness prole of sample B measured
along mid-thickness of the joints at six different natural aging periods
after FSW. The base materials AA6082-T6 and AA7075-T6 show the

410

H. Jamshidi Aval / Materials and Design 87 (2015) 405413

Fig. 7. TEM bright eld images of different zones in sample B; a) base material of AA7075-T6, b) base material of AA6082-T6, c) TMAZ of AA7075 side, d) TMAZ of AA6082 side, e) SZ of
AA7075 side, and f) SZ of AA6082 side.

average Vickers micro-hardness values of about 95HV0.1 and 165HV0.1,


respectively. The hardness prole changes in the SZ and TMAZ after natural aging were more signicant than the other regions. The increase in
micro-hardness value after 365 days of natural aging within the SZ and
TMAZ of the AA6082 side and the SZ of the AA7075 side compared to
micro-hardness value of these zones in as welded condition is noticeable. However, no obvious changes were observed in the TMAZ of the
AA7075 side. The change in hardness in the SZ and TMAZ of welded
samples is due to reprecipitation after natural aging. The microhardness of samples B and C after 365 days of natural aging is compared
in Fig. 6(b), and related TEM images of various regions from the base
metal to the dynamically recrystallized zone of sample B are summarized in Fig. 7. As can be seen in sample C, with the highest heat input,
softening takes places at higher distances and maximum microhardness in both sides of sample C, which is lower than those in sample
B. It is worth noting that the SZ hardness decrease in the AA7075 side at
2 mm from the weld center line of sample C is related to AA6082 ow to
the AA7075 side due to stirring effect of pin as shown in Fig. 3(c). As it

was found previously described in the literature, in the AlMgSi


6XXX series aluminium alloys, the decomposition of supersaturated
solid solution (SS) proceeds in the following steps: the formation of
needle-like zones along the 100Al crystal direction; the ordering of
their structure and the formation needles; the transition from
needles to rods; and the formation of equilibrium phase [20]. In
the AlZnMg(Cu) system (7XXX alloys) with Mg:Zn lower than 0.5,
the precipitation sequence is generally reported as SS Guinier
Preston (GP) zones (MgZn) (MgZn2) [21]. Most of its strength
is associated with precipitation of GP zones and and semi-coherent
phases. In the base material of AA7075-T6 (Fig. 7a), the TEM micrographs revealed a ne distribution of precipitates 1090 nm in size. In
previous studies by Rhodes et al. [22] and Su et al. [23], the precipitates
with a size less than 20 nm are GP zones and the coarser precipitate is .
Besides, the AA6082-T6 base material presents needle-type precipitates
(Fig. 7b), and the size of the precipitates was found to be in a range of
1050 nm length. The morphology of these precipitates indicates that
they are type [24,25]. In the TMAZ of the AA7075 side, the small

H. Jamshidi Aval / Materials and Design 87 (2015) 405413

Fig. 8. Schematic representation of the precipitation sequence within the SZ and TMAZ.

Fig. 9. Residual stress prole of samples B and C; a) longitudinal residual stress, b) transverse residual stress, and c) longitudinal residual stress vs. aging time of sample B.

411

412

H. Jamshidi Aval / Materials and Design 87 (2015) 405413

Fig. 10. a) The strain maps, and b) the local mechanical properties of different weld sub-zones for sample B.

precipitates had dissolved and only round particles with an average diameter in the order of 70250 nm remain (Fig. 7(c)). Fig. 7(d) illustrates
that needle-type precipitates from base material AA6082 became
coarser and wider spaced with a round precipitate in the order of
60 nm diameter, and rod morphology precipitates 80250 nm in length
and 915 nm in width in the TMAZ. With respect to size, morphology,
and studies by Sato et al. [26] and Ohmori et al. [27], the rod precipitates
are dened as precipitates and round precipitates belong to the Si and
phases. The type and morphology changes of the TMAZ precipitations
in response to the heat input and the deformation rate from FSW are
reasons why this zone shows an abrupt hardness decrease as compared
to the base materials. As shown in Fig. 7(e), the SZ of the AA7075 side

Table 5
Measured temperature of different zones of samples B and C.
Sample

B
C

Temperature (C)
Advancing side (AA6082 side)

Retreating side (AA7075 side)

SZ

TMAZ

HAZ

SZ

TMAZ

HAZ

446
493

411
463

307
369

431
481

389
448

295
356

shows three different precipitates which were identied as GP zones,


round type , and rod shape precipitates. Besides, the SZ in the
AA6082 side presents ne round type and Si and rod precipitates
(Fig. 7(f)). The increase in micro-hardness in the SZ after natural aging
can be related to reprecipitation of ne precipitates and and GP
zones. The schematic representation of the precipitation sequence within the SZ and TMAZ is shown in Fig. 8. Curves 1 and 2 indicate the
welding thermal cycles in the TMAZ and SZ, respectively. In the TMAZ,
growth of th base metal precipitates (GP and in AA7075 and in
AA6082) occurred, whereas reprecipitation of , GP and occurred
in the SZ during cooling.
Fig. 9 shows the residual stress prole of samples welded with the
lowest and highest heat inputs, that are, samples B and C, respectively.
As shown in Fig. 9, for both samples longitudinal residual stress in the
center of the weld in both advancing and retreating sides has tensile
nature and away from the weld center rst increases up to maximum
tensile stress near the edge of the tool then in farther distance becomes
compressed. The overall behavior of transversal residual stress is similar
to longitudinal residual stress, but in the weld center line, transversal
stress has compression nature and maximum residual stress is
approximately two times less than the longitudinal residual stress.
The residual stress proles of the longitudinal and transverse directions

H. Jamshidi Aval / Materials and Design 87 (2015) 405413

are asymmetric, and the maximum tensile residual stress occurs in the
AA7075 side. Another important result (Fig. 9) is that with increasing
welding heat input, the maximum tensile residual stress decreases,
and the size of the tensile residual stress region on either side of the
weld line increases; for example, the maximum tensile residual stresses
of the AA6061 side of samples B and C are 104 MPa and 97 MPa, respectively. Besides, the sizes of the tensile residual stress region in samples B
and C are 48 mm and 67 mm, respectively. The results of the longitudinal residual stress measured in sample B after two different natural
aging periods are shown in Fig. 9(c). It can be seen that there is a noticeable decrease in the residual stress from 7 to 365 days of natural aging
within the SZ and TMAZ. During natural aging, the microstructural modication due to the vacancies and solute atom ow may cause the creep
behavior and stress relaxation [28]. Therefore, the lower residual stress
of the SZ and TMAZ after 365 days of natural aging may be attributed to
stress relaxation in these zones due to the natural aging phenomena.
The strain maps correspond to strain distribution at maximum load,
and the related local mechanical properties of different welding
subzones for sample B achieved by DIC method are shown in Fig. 10.
According to the strain maps, the related macrostructure of tensile
sample is shown in Fig. 10(a). The tensile sample fracture in the TMAZ
of the AA6082 side where minimum hardness and tensile residual stress
in the weld zone are observed is illustrated in Figs. 6 and 9. It can be seen
that the residual stress distribution at the region between the HAZ of the
AA6082 and AA7075 sides is a function of the local mechanical properties and the uctuating inhomogeneous temperature distributions
during and after welding. In other words, the SZ, TMAZ, and HAZ of
the AA6082 side with lower residual stress prole show higher peak
temperature (Table 5) and lower micro-hardness and yield stress than
the related regions in the AA7075 side (Figs. 6 and 10(b)). Besides,
AA6082 at moderate temperatures has a rather lower yield stress than
AA7075; therefore, the residual stress build up in the AA6082 side will
be lower than the AA7075 side [6,29].
4. Conclusion
The inuence of rotational and linear speeds of triangular frustum
pin on residual stress and microstructure of dissimilar friction stir
welded AA6082-T6 and AA7075-T6 is examined by optical microscopy,
TEM, micro-hardness, and XRD residual stress tests. The results can be
summarized as follows:
1- Zinc distribution across the weld shows that with increasing welding
heat input per unit length, the mixing of the material in the weld nugget of the joints is performed more efciently, and due to severe plastic
deformation of the material and high-peak temperature in the weld
nugget, the atomic diffusion occurs at the interface of the materials.
2- In all welding heat input, the grain size in the AA7075 side is ner than
that in the AA6082 side. The peak temperature effect on the SZ grain
size is dominant on deformation rate.
3- Coarsening and dissolution of the strengthening precipitates present
in the AA6082 and AA7075 base metals result in softening in the
weld zone. Besides, reprecipitation of ne GP, , and precipitates results in an increase of micro-hardness in the SZ after natural aging.
4- The residual stress distribution at different zones is a function of the
local mechanical properties and the temperature distributions of different welding subzones. An increase in welding heat input resulted
in decreased maximum tensile residual stress and increased size of
the tensile residual stress region on both sides of the weld line.
5- The longitudinal residual stress in the SZ has tensile nature, and away
from the weld center rst increases up to maximum tensile stress near
the edge of the tool, then in farther distance becomes compressed,
while transverse residual stress in the weld center line has compression nature and maximum residual stress is approximately two
times less than the longitudinal residual stress.

413

References
[1] W.M. Thomas, E.D. Nicholas, J.C. Needham, M.G. Murch, P. Templesmith, C.J. Dawes,
Friction Stir Welding, 1991.
[2] P. Cavaliere, A. De Santis, F. Panella, A. Squillace, Effect of welding parameters on
mechanical and microstructural properties of dissimilar AA6082AA2024 joints
produced by friction stir welding, Mater. Des. 30 (2009) 609616.
[3] R. Palanivel, P. Koshy Mathews, N. Murugan, I. Dinaharan, Effect of tool rotational
speed and pin prole on microstructure and tensile strength of dissimilar friction
stir welded AA5083-H111 and AA6351-T6 aluminum alloys, Mater. Des. 40
(2012) 716.
[4] Youbao Song, Xinqi Yang, Lei Cui, Xiaopeng Hou, Zhikang Shen, Yan Xu, Defect features and mechanical properties of friction stir lap welded dissimilar AA2024
AA7075 aluminum alloy sheets, Mater. Des. 55 (2014) 918.
[5] P. Cavaliere, F. Panell, Effect of tool position on the fatigue properties of dissimilar
20247075 sheets joined by friction stir welding, J. Mater. Process. Technol. 206
(2008) 249255.
[6] A. Steuwer, M.J. Peel, P.J. Withers, Dissimilar friction stir welds in AA5083AA6082:
the effect of process parameters on residual stress, Mater. Sci. Eng. A 441 (2006)
187196.
[7] Caroline Jonckheere, Bruno de Meester, Anne Denquin, Aude Simar, Torque, temperature and hardening precipitation evolution in dissimilar friction stir welds between
6061-T6 and 2014-T6 aluminum alloys, J. Mater. Process. Technol. 213 (2013)
826837.
[8] J.H. Ouyang, R. Kovacevic, Material ow and microstructure in the friction stir butt
welds of the same and dissimilar aluminum alloys, J. Mater. Eng. Perform. 11
(2002) 5163.
[9] M.B. Prime, T.G. Herold, J.A. Baumann, R.J. Lederich, D.M. Bowden, R.J. Sebring, Residual stress measurements in a thick, dissimilar aluminum alloy friction stir weld, Acta
Mater. 54 (2006) 40134021.
[10] A.A.M. da Silva, E. Arruti, G. Janeiro, E. Aldanondo, P. Alvarez, A. Echeverria, Material
ow and mechanical behaviour of dissimilar AA2024-T3 and AA7075-T6 aluminium
alloys friction stir welds, Mater. Des. 32 (2011) 20212027.
[11] P. Cavaliere, R. Nobile, F.W. Panella, A. Squillace, Mechanical and microstructural behaviour of 20247075 aluminium alloy sheets joined by friction stir welding, Int. J.
Mach. Tools Manuf. 46 (2006) 588594.
[12] E.G. Cole, A. Fehrenbacher, N.A. Dufe, M.R. Zinn, F.E. Pfefferkorn, N.J. Ferrier, Weld
temperature effects during friction stir welding of dissimilar aluminum alloys
6061-T6 and 7075-T6, Int. J. Adv. Manuf. Technol. 71 (2014) 643652.
[13] J.F. Guo, H.C. Chen, C.N. Sun, G. Bi, Z. Sun, J. Wei, Friction stir welding of dissimilar
materials between AA6061 and AA7075 Al alloys effects of process parameters,
Mater. Des. 56 (2014) 185192.
[14] P. Bala Srinivasan, W. Dietzel, R. Zettler, J.F. dos Santos, V. Sivan, Stress corrosion
cracking susceptibility of friction stir welded AA7075AA6056 dissimilar joint,
Mater. Sci. Eng. A 392 (2005) 292300.
[15] Gven pekolu, Grel am, Effects of initial temper condition and postweld heat
treatment on the properties of dissimilar friction-stir-welded joints between
AA7075 and AA6061 aluminum alloys, Metall. Mater. Trans. A 45 (2014) 30743087.
[16] Y.S. Sato, S.H.C. Park, M. Michiuchi, H. Kokawa, Constitutional liquation during dissimilar friction stir welding of Al and Mg alloys, Scr. Mater. 50 (2004) 12331236.
[17] H.J. Liu, J.J. Shen, Y.X. Huang, L.Y. Kuang, C. Liu, C. Li, Effect of tool rotation rate on
microstructure and mechanical properties of friction stir welded copper, Sci.
Technol. Weld. Join. 14 (2009) 577583.
[18] L. Fratini, G. Buffa, CDRX modelling in friction stir welding of aluminium alloys, Int. J.
Mach. Tools Manuf. 45 (2005) 11881194.
[19] G. Buffa, L. Fratini, R. Shivpuri, CDRX modelling in friction stir welding of AA7075-T6
aluminum alloy: analytical approaches, J. Mater. Process. Technol. 191 (2007)
356359.
[20] G.E. Totten, D.S. MacKenzie, Handbook of Aluminum: Physical Metallurgy and
Processes, CRC Press, 2003.
[21] G.W. Lorimer, R. Nicholson, Further results on the nucleation of precipitates in the
AlZnMg system, Acta Metall. 14 (1966) 10091013.
[22] C.G. Rhodes, M.W. Mahoney, W.H. Bingel, R.A. Spurling, C.C. Bampton, Effects of friction stir welding on microstructure of 7075 aluminum, Scr. Mater. 36 (1997) 6975.
[23] J.Q. Su, T.W. Nelson, R. Mishra, M. Mahoney, Microstructural investigation of friction
stir welded 7050-T651 aluminium, Acta Mater. 51 (2003) 713729.
[24] G.A. Edwards, K. Stiller, G.L. Dunlop, M.J. Couper, The precipitation sequence in Al
MgSi alloys, Acta Mater. 46 (1998) 38933904.
[25] C.A.W. Olea, L. Roldo, J.F. dos Santos, T.R. Strohaecker, A sub-structural analysis of
friction stir welded joints in an AA6056 Al-alloy in T4 and T6 temper conditions,
Mater. Sci. Eng. A 454455 (2007) 5262.
[26] Y.S. Sato, H. Kokawa, M. Enomoto, S. Jogan, Microstructural evolution of 6063 aluminum during friction-stir welding, Metall. Mater. Trans. A 30 (1999) 24292437.
[27] Y. Ohmori, L.C. Doan, K. Nakai, Ageing processes in AlMgSi alloys during continuous heating, Mater. Trans. 43 (2002) 246255.
[28] W. Woo, Z. Feng, X.-L. Wang, C.R. Hubbard, Neutron diffraction measurements of
time-dependent residual stresses generated by severe thermomechanical deformation, Scr. Mater. 61 (2009) 624627.
[29] G.Z. Quan, G.S. Li, Y. Wang, W.Q. Lv, C.T. Yu, J. Zhou, A characterization for the ow
behavior of as-extruded 7075 aluminum alloy by the improved Arrhenius model
with variable parameters, Mater. Res. 16 (2013) 1927.

Das könnte Ihnen auch gefallen