Вы находитесь на странице: 1из 9

Materials Science & Engineering A 556 (2012) 175183

Contents lists available at SciVerse ScienceDirect

Materials Science & Engineering A


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

Friction-stir dissimilar welding of aluminium alloy to high strength steels:


Mechanical properties and their relation to microstructure
R.S. Coelho a,b,n, A. Kostka c, J.F. dos Santos d, A. Kaysser-Pyzalla a,b
a

Helmholtz-Zentrum Berlin f
ur Materialien und Energie GmbH, 14109 Berlin, Germany
Former at Max-Planck-Institut f
ur Eisenforschung GmbH, 40237 D
usseldorf, Germany
c
Max-Planck-Institut f
ur Eisenforschung GmbH, 40237 D
usseldorf, Germany
d
Helmholtz-Zentrum Geesthacht GmbH, Zentrum f
ur Materialforschung und K
ustenforschung, Institute of Materials Research, Materials Mechanics, Solid-State Joining Processes,
21502 Geesthacht, Germany
b

a r t i c l e i n f o

abstract

Article history:
Received 2 March 2012
Received in revised form
19 June 2012
Accepted 20 June 2012
Available online 6 July 2012

The use of light-weight materials for industrial applications is a driving force for the development of
joining techniques. Friction stir welding (FSW) inspired joints of dissimilar materials because it does
not involve bulk melting of the basic components. Here, two different grades of high strength steel
(HSS), with different microstructures and strengths, were joined to AA6181-T4 Al alloy by FSW. The
purpose of this study is to clarify the inuence of the distinct HSS base material on the joint efciency.
The joints were produced using the same welding parameter/setup and characterised regarding
microstructure and mechanical properties. Both joints could be produced without any defects.
Microstructure investigations reveal similar microstructure developments in both joints, although
there are differences e.g. in the size and amount of detached steel particles in the aluminium alloy (heat
and thermomechanical affected zone). The weld strengths are similar, showing that the joint efciency
depends foremost on the mechanical properties of the heat and the thermomechanical affected zone of
the aluminium alloy.
& 2012 Elsevier B.V. All rights reserved.

Keywords:
Friction stir welding
Dissimilar joint
Aluminium alloy
High strength steel
Microstructure
EBSD
Mechanical properties

1. Introduction
Energy and environmental issues in transportation systems
have a strong impact on material selection and on the
development of joining techniques [17]. The incorporation of
light-weight materials in many structures (e.g. automotive, shipbuilding and aerospace) allow a reduction of weight and consequently fuel consumption. In this regard, dissimilar joints
between light-weight materials such as aluminium alloys (Al
alloy) and steels are receiving increased interest both in science
and industrial application [820]. However, the substitution of
one material for another is not straightforward and highly
efcient products require appropriate joining processes.
Dissimilar fusion welding between Al alloy and steels is a
challenge in welding control because of the large differences in
melting temperature and physical and mechanical properties of
the alloys involved. The process often results in complex weld
pool shapes, inhomogeneous solidication microstructures and
segregations in addition, the extremely low solubility of Fe in Al

n
Materialien und
Corresponding author at: Helmholtz-Zentrum Berlin fur
Energie GmbH, 14109 Berlin, Germany. Fax: 49 30 8062 15752.
E-mail address: rodrigo.coelho@helmholtz-berlin.de (R.S. Coelho).

0921-5093/$ - see front matter & 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.msea.2012.06.076

leads to the formation of brittle and excessive Al-rich FexAly


intermetallic phases [21,22] which are detrimental for the
mechanical properties of the joint [812,20,22].
Friction stir welding (FSW) is based on extreme plastic
deformation in the solid-state where no associated bulk melting
is involved [2326]. At early stages of the process development,
FSW appears especially attractive for joining Al alloys and other
light-weight materials like Mg alloys [2431]. This is connected
with two main reasons:
(1) the process prevents melting and solidication, minimising
residual stresses, cracking, porosity and loss of volatile
solutes;
(2) the plastic deformation (stirring) of such light-weight materials (e.g. Al and Mg alloys) can be realized using relatively
simple welding tools (e.g. made of tool steel).
For an application of the FSW process on steels and titanium,
an optimisation of tool material and geometry is needed; continuing attempts can be found in the literature [3236]. However,
FSW application for these materials is still limited by the cost of
effective available tools [37].
In the case of dissimilar joints, FSW appears as a suitable and
promising process to minimise problems related to materials

176

R.S. Coelho et al. / Materials Science & Engineering A 556 (2012) 175183

incompatible with respect to melting temperatures and the


formation of brittle intermetallic phases [27,38]. Recent studies
have shown that welds between dissimilar materials such as Mg
alloys to Al alloys [3944], Mg alloys to steels [4547], Al alloys to
Titanium [48] and Al alloys to steels [1320] can be produced by
FSW. The high quality of dissimilar welds produced between Al or
Mg alloys to steels or titanium in a butt-joint conguration is
associated to the smart idea of positioning the tool pin centre
shifted towards the Al or Mg alloys (xed on the retreating side)
[1420,48] or xing Al or Mg alloys on the top side in the case of
lap joints [13,4547]. Thus, the tool barely makes contact with
the steel and minimum tool wear occurs, improving the cost
efciency of the process.
In the present study, two grades of high strength steel (HSS)
with signicantly different microstructure, and strengths were
selected to be joined to AA6181-T4 Al alloy by FSW. In order to
access the inuence of the distinct HSS base material on joint
efciency, the joints were produced by applying the same welding parameters and by shifting the tool pin centre towards the Al
alloy. Early investigations were conducted in one of these joints
and presented elsewhere [14]. Those studies [14] were focussed
on the analysis of the complex material ow based on microstructural observations applying SEM-EBSD technique. Here, we
discuss the inuence of distinct HSS base material on the joint
efciency and microstructure formation. We show that independent of the HSS chosen the joint efciency is determined by the
heat-affected zone of the Al alloy, which controls the mechanical
properties of the joints.

2. Experiments
2.1. Materials
Commercially available materials that are suitable for automotive structures and reinforcement parts were selected for this
study. DP600 and HC260LA HSS plates were chosen to be joined
to AA6181-T4 Al alloy by FSW. The chemical composition and the
mechanical properties of the steels and the Al alloy are presented
in Tables 1 and 2, respectively. It can be seen that the Al alloy
tensile strength is substantially less than those of the HSSs.
The microstructure of the base materials (BMs) is presented in
Fig. 1. The individual microstructural characteristics can be
summarised as following:
(a) the AA6181-T4 Al alloy shows the typical large grains slightly
elongated with a length of about 70 mm;
(b) the HC260LA HSS shows a-Fe (ferrite) grains with a size of
about 40 mm and presence of pearlite on the grain
boundaries;
(c) the dual phase steel DP600 HSS shows a-Fe grains with a size
of about 15 mm (appear dark in Fig. 1c) and the presence of
martensite as a second phase (appears bright in Fig. 1c).

2.2. Welding procedure


The joints were produced in a butt-joint conguration at the
Helmholtz-Zentrum Geesthacht, Germany, using a gantry system
equipped with a mechanical clamping table [50]. The welding
setup is schematically illustrated in Fig. 2(a) and the process
parameters follow in Table 3.
A FSW tool, the TungstenRhenium WRe25, consisting of a
13 mm diameter concave shoulder and a 5 mm cylindrical nonthreaded pin, was selected for this welding conguration. The pin
edge was offset about 1 mm from the weld centre line towards
the AA6181-T4 Al alloy (Fig. 2b). Minimum wear is expected in
this conguration once the pin is plunged into the softer AA6181T4 Al alloy and does not come in contact with the HSS.

2.3. Microstructure assessment


The microstructure was investigated using optical microscopy
(OM), scanning electron microscopy (SEM) electron backscatter
diffraction (EBSD) technique and transmission electron microscopy (TEM). SEM-EBSD characterisations were conducted using a
Zeiss Neon 40 eld emission gun SEM equipped with the Hikari
EDAX/TSL EBSD system and a Jeol JSM-6490 Tungsten lament
SEM equipped with the Pegasus EDAX/TSL EBSD system. The
analyses were carried out on the sample cross section and on the
top weld surface. The samples were extracted by spark erosion a
few centimetres behind the start point of the weld, avoiding any
problems related with the process that might not be completely
stable at its start. The specimen preparation for OM and SEM
consisted of standard metallographic procedure followed by a
short etching in a 5% Nital solution. Consequently, the grain
boundaries in the HSS microstructure were highlighted. The
assessment of the Al alloy microstructure was conducted mainly
by EBSD. The specimens for EBSD were prepared in the standard
metallographic way with careful nal polishing and without
chemical etching. Such sample preparation allows the assessment
of both the HSS and the Al alloy microstructure [7,14].
TEM work was performed using a Jeol JEM-2200FS operated at
200 kV. The specimens were prepared from regions of interest
along the Al alloyHSS interface using a Jeol JEM-9320 focussed
ion beam system (FIB) operated at 30 kV.

Table 2
Obtained yield, strength and strain values for each material selected for this study.
The values are an average of at least three tensile tests conducted for each
material.
Materials

Yield strength (MPa) Tensile strength (MPa) Elongation (%)


259 75
397 71
625 72

AA6181-T4 1277 5
HC260LA
3077 1
DP600
3227 5

267 1
317 1
247 1

Table 1
Chemical composition of the materials selected for this study (wt.%).
Materials
*
AA6181-T4

Si
0.85

Fe
0.25

Cu
0.06

Mn
0.09

Mg
0.74

Cr
0.01

Ni
0.002

Zn
0.012

Materials
**
HC260LA
***
DP600

Cmax
0.10
0.10

Mnmax
0.60
1.40

Simax
0.50
0.15

Pmax
0.025
0.07

Smax
0.025
0.008

Almin
0.015
0.02

Timax
0.15

Nmax

0.009

[49]
norm DIN EN 10268, Edition 10.06.
nnn
norm SEW 097 Part 1.
nn

Ti
0.023

Al
Bal.

R.S. Coelho et al. / Materials Science & Engineering A 556 (2012) 175183

177

Fig. 1. Micrographs of the alloy microstructures for each base materials involved in the presented work: (a) AA6181-T4 Al alloy, (b) HC260LA HSS and (c) DP600 HSS.

2.4. Mechanical tests

Fig. 2. Schematic illustration of joint setup studied here (a) Al alloy and steel
joined by FSW in a butt-joint conguration and (b) the relation of real microstructure and tool offset position applied to produce the welds analysed in this
study [14].

Table 3
Welding parameters.

Universal microhardness tests and tensile tests were conducted to assess the mechanical properties of the joints. Regarding the tensile tests, standard at specimens with gauge sections
of 1.5 mm  10 mm  45 mm were extracted by spark erosion
cutting from the base materials and the weld in the transversal
direction of the welding direction. At least three samples from
each base material and from the joints were tested. All welded
specimens were machined on their surface in order to remove the
marks and the crown features left by the tool shoulder and to
avoid any inuences of them on the results. Afterwards, failure
mechanisms of the at tensile test specimens were studied in the
SEM by a fracture surface analysis.
Microhardness was measured on the specimen cross section
according to DIN 50359-1:1997, Universal hardness (HU) standard, applying a 0.02 N load. The measurements were performed
in three different cross section depths in order to check for
through-thickness variance. The measurements crossed all
regions of interest, from the Al alloy side through to the HSS
side. One specimen for each joint setup was tested.

Main welding parameters


Travel speed (mm/s)
Rotation speed (rpm)
Down-force (kN)

8.0
1600
5.0

Main tool properties


Pin diameter (mm)
Shoulder diameter (mm)
Pin length (mm)
Pin offset Dy (mm)
Tool material
Pin feature

5.0
13
1.35
1.0
WRe25
No-threaded

3. Results
3.1. Microstructure of the joints
On the macroscopic scale, the as-welded joints revealed a good
weld surface quality containing neither macro voids/cracks nor
imperfections regarding the weld alignment (Fig. 2b).
Through the weld cross section, both analysed joints revealed
the same microstructure features showing no evidence of mixing
between the Al alloy and the HSS. In both cases, a small amount of

178

R.S. Coelho et al. / Materials Science & Engineering A 556 (2012) 175183

detached particles of HSS transported into the Al alloy was observed


(Fig. 3). However, comparing both joints, it is evident that the softer
HC260LA HSS shows a more strongly deformed interface (dash line
in Fig. 3b and d) and slightly larger HSS detached particles than the
harder DP600 HSS (arrows in Fig. 3a and c).
A systematic investigation of the welding setup and material
ow of the joint analysed in this paper was previously conducted on
DP600 HSS [14]. Some insight into the material ow and microstructure formation can be visualised and summarised by the energy
dispersive X-ray (EDX) analysis shown in Fig. 4. These results were
obtained for the joint using the HC260LA HSS and conrm that there
is no evidence of mixing between both materials and also highlight
the presence of the vortex-like structures on the advancing side of
the tool (Zn mapEDX analysis, Fig. 4d). The vortex features
becomes smooth approaching the bottom part of the cross section
and implies the movement of the HSS detached particles.

A close view of both joint interfaces reveals also very similar


features (Fig. 5). The high shear strain and friction heating during
the process supports the intermetallic phase formation and the
interlocking between both materials. In both cases, crack-free
bonding between both materials occurs. The microstructure at
the interface is complex and non-smooth, characterised by ne
equiaxed a-Fe (ferrite) grains and a very thin layer of intermetallic FexAly compounds formed by chemical reactions and diffusion between the Al and Fe.
Complementary TEM investigations (Fig. 6) revealed in details
the intermetallic compound present in both analysed joints. The
intermetallic compounds form very thin stripes (around 50 nm
thick) embedded approximately 500 nm deep into the steel
matrix. Fig. 6(c), which reveals the Fe2Al5-phase, summarises
several attempts of unambiguous phase identication via electron
diffraction from selected areas of the interfaces of both joints. The

Fig. 3. Cross section overview of both investigated joints: (a) Al alloy to HC260LA HSS and (b) the thermo-mechanically deformed interface; (c) Al alloy to DP600 HSS and
(d) the thermo-mechanically deformed interface.

Fig. 4. EDX analysis of element distributions: (a) analysed region, (b) Al distribution in blue, (c) Fe distribution in pink and (d) Zn distribution highlighting the vortex-like
structure on the advancing side in red. (For interpretation of the references to colour in this gure legend, the reader is referred to the web version of this article.)

R.S. Coelho et al. / Materials Science & Engineering A 556 (2012) 175183

179

Fig. 5. Secondary electron SEM micrographs of joint interface highlighting similar features: intermetallic formation and non-smooth interface resulting in mechanical
interlocking between both materials; (a) Al alloy to HC260LA HSS and (b) Al alloy to DP600 HSS.

Fig. 6. TEM investigation of the Al alloyHSS interface: (a) Al alloy to DP600 HSS, (b) Al alloy to HC260LA HSS and (c) selected area electron diffraction pattern from the
Fe2Al5 reaction product on the interface between Al alloy and DP600 HSS.

low volume fraction of the intermetallic phase, the presence of


chemical gradients (up to 20 at% on 300 nm) and the crystallographic lattice strain made the phase identication very
challenging.

change crossing the thermomechanically affected regions. The joint


produced with the softer HSS (Fig. 7a) shows a thermomechanical
area with length of about 4 mm, while in the other joint it is less
than 1 mm (Fig. 7b). The highest values in both cases were observed
at the weld interface.

3.2. Mechanical properties


3.2.1. Microhardness assessments
Microhardness proles were measured on the weld cross section
at three different depths: top, middle and bottom (Fig. 7). For both
investigated joints, the results do not show any variance crossing
the Al alloy welding regions: base material (BM), heat-affected
zone (HAZ), thermo-mechanically-affected zone (TMAZ) and stir
zone (SZ). On the other hand, the HSS side shows a hardness

3.2.2. Tensile properties (stressstrain curves)


Fig. 8 shows the typical tensile stressstrain curves of all
investigated specimens. The curves can be considered as an
average property since at least three specimens were prepared
for each BM and joint, and all of them revealed the same trend.
The black and blue curves plotted in Fig. 8 are for Al alloy and HSS,
respectively. Comparing both HSS curves, differences in yield

180

R.S. Coelho et al. / Materials Science & Engineering A 556 (2012) 175183

Fig. 7. Universal microhardness measurements horizontally along the Al alloy and the HSS side: (a) Al alloy to HC260LA HSS and (b) Al alloy to DP600 HSS.

Table 4
Obtained yield, strength and strain values for each joint investigated in this study.
Joint

Yield strength
(MPa)

Tensile strength
(MPa)

Elongation
(%)

112 72
119 72

2007 8
211 7 2

87 1
77 1

Al AlloyHC260LA
Al alloyDP600
n

One specimen slide off during the test, therefore only 2 were considered.

in Fig. 8(a) and (b) which have been chosen for improved clarity
in the presentation of results.

Fig. 8. Typical stressstrain curves obtained by tensile test for each BMs (black
and blue curves) and joints (red curves) analysed in this study: (a) Al alloy to
HC260LA HSS and (b) Al alloy to DP600 HSS. (For interpretation of the references
to colour in this gure legend, the reader is referred to the web version of this
article.)

strength and ultimate tensile strength are evident, even though


both of them show high elongation to fracture (see Table 2).
Typical tensile stressstrain curves of both Al alloyHSS joints
are ploted in red on the diagrams. The results show that the joint
behaviour follows the tendency of the Al alloy stressstrain curve.
The main ndings regarding the tensile strengths of the joints are
summarised in Table 4. It can be concluded that both joints reveal
almost the same yield strengths as well as tensile strengths.
Additionally, regarding the elongation to fracture, both joints also
reveal similarities showing almost 30% of the Al alloy BM
property (see Table 2). Please note the different scales for stress

3.2.3. Fractography analyses


After the tensile test the fractured samples were submitted to
fractography examinations in order to assess the origin of the
specimen failure and also to determine which fracture mechanisms occurred. Both analysed samples show the same fracture
features, hence, only the results obtained for the harder DP600
HSS joint are presented.
Fig. 9(a) shows a top view micrograph of the fractured sample
highlighting the expected schematic welding tool position (tool
offset position). The analysis reveals that the failure occurs at the
retreating side of the Al alloy crossing the BM-HAZ-TMAZ zones
far away from the joint interface and from the weld centre point.
Analysis of the fractured surface through the cross section reveals
a ductile fracture mode characterised by the presence of equiaxed
dimples (Fig. 9b and c).
Additional investigations of the top view reveal that few cracks
were found close to the joint interface (Fig. 9d). The analysis
suggests that those cracks which were found propagated following the HSS particles [13]. Chemical composition mapping conducted through EDX-analysis conrms that the cracks are
connected to the HSS particles (Fig. 9eg).
Additional information regarding the joint failure region (Al alloy
BM-HAZ-TMAZ) were obtained by EBSD analysis. Those most
relevant regions were analysed in the as-welded samples cross
section and the results reveal that a large difference in grain
morphology and distribution can be seen crossing these regions
(arrow in Fig. 10a). While the Al alloy BM shows elongated grains
with a size of 70 mm (Fig. 1a), the Al alloy SZ shows characteristics of
ne equiaxed grains 5 mm diameter in size. Further, top surface
analysis conducted on the interface between welded zone and the Al
alloy BM reveals the strong grain deformation (arrows in Fig. 10b)
caused by the movement of the FSW tool shoulder [14]. The
analysis, conducted on the top weld surface after standard metallographic sample preparation (approx.  200 mm layer removal),
reveals that the surface shear layer caused by the shoulder

R.S. Coelho et al. / Materials Science & Engineering A 556 (2012) 175183

181

Fig. 9. SEM micrographs showing the fractured tensile test specimen: (a) the schematic illustration of the relationship sample and tool position; (bc) the microscopic
analysis of the fractured surface and (dg) the EDX element distribution analysis highlighting the cracking presence close to the joint interface (dg).

Fig. 10. Inverse pole gures maps from the critical interface Al alloy BM-HAZTMAZ-SZ: (a) cross section and (b) top analysis [14]. The arrows highlight the
gradient in grains characteristics.

movement extends several hundred micrometres (300 mm) from


the top weld surface.

4. Discussion
In this study, dissimilar joints between Al alloy and two grades of
HSS were investigated in terms of microstructure and mechanical

properties. The purpose of the study is to clarify the inuence of


distinct HSS BMs on the joint efciency. The results have shown that
both investigated samples show the same microstructure and
properties since a very similar microstructure evolution occurred
during the welding process. The welding parameter chosen (tool
offset position) meant that severe plastic deformation occurred
mainly in the Al alloy and the pin barely moved into the HSS, which
is consistent with the microstructure characteristics presented here.
The microstructure formation in FSW is directly connected to the
material ow, and in the presented study, both joints revealed the
same material ow feature (Fig. 4) [14].
In terms of microstructure, it is evident that joints produced
with the softer steel (HC260LA) reveal a slightly larger deformed
interface region in the Al-alloy and a larger amount/size of the
HSS detached particles (Fig. 3). Such differences were also
observed in the hardness proles presented in Fig. 7(a) where
the HSS side shows a strong hardness gradient which is connected
with different levels of thermo-deformation. The highest hardness
values in both cases were observed in the weld interface where
the smallest HSS grains were formed together with the thin strips
of intermetallic compounds.
Regardless of the HSS used, the microstructure features of the
joints were characterised by ne equiaxed a-Fe (ferrite) grains
with a small amount of thin intermetallic Fe2Al5 compounds and
rough interface Al alloyHSS. The very small quantity, complex
morphology (thin stripes of intermetallic reaction products were
incorporated with steel BM) and strong chemical gradients made
unambiguous phase identication of the intermetallic compounds
very difcult. Only in the joint with the DP600 HSS it was possible
to report presence of the Fe2Al5-phase. In terms of the binary Al
Fe equilibrium diagram, a couple of intermetallic phase compounds, FeAl3 (y-phase), Fe2Al5 (Z-phase) and FeAl2 (x-phase),
might be expected to be formed [21]. Springer et al. [20]
investigated the formation of intermetallic reaction layers which
form at the interface of FSW joints between low C (0.12 wt% C)
steel and both pure Al (99.5%) and Al-5 wt.% Si. They claimed that
in the as-welded state no intermetallic reaction product could be
observed at the interface, and only after annealing the intermetallic compounds make detailed investigations possible.
Analysis of the stressstrain tensile curves revealed no inuence of the joint interface on the joint efciency. The results

182

R.S. Coelho et al. / Materials Science & Engineering A 556 (2012) 175183

revealed that the joint curves (red in Fig. 8) follow the tendency of
the Al alloy curve showing almost the same yield and strength. By
evaluating the results presented in Tables 2 and 4, it becomes
clear that the measured ultimate tensile strength for the joints
does not reach the yield strength of the HSS. This fact suggests
that all deformation during the tensile test took place solely in the
Al alloy. Additionally, no evidence of plastic deformation was
observed in micrographs of the tensile specimens, which suggests
once again that the Al alloy is responsible for the overall efciency
of the joint.
Fractography analysis showed that failure always occurs far
away from the joint interface on the retreating side crossing the
interface of the Al alloy BM-HAZ-TMAZ-SZ (Fig. 9). Those regions
(arrow in Fig. 10a) reveal the presence of highly deformed grains
and evidence of sub-grain development. No signicant differences
in hardness were observed crossing those regions (Fig. 7). The
strong differences in a grain size distribution and grain shape
shown in Fig. 10 originate from each region of the joint (BM-HAZTMAZ-SZ) being exposed to different thermo-mechanical cycles.
The gradual accumulation of strain accompanied by the frictional
heating leads to the nucleation and growth of new grains during the
ongoing deformation process (non-homogeneous dynamic recrystallisation). While accumulated plastic deformation increases the
mechanical strength of the TMAZ and SZ, the grains located in the
HAZ reveal relatively small defect densities (concentration of
dislocations and sub-grains) and control the mechanical behaviour
of the joint as the weakest elements of the structure.
Additionally, the fracture analysis suggests that the crack
propagation follows the HSS inclusions. These ne-grained HSS
inclusions are surrounded by the intermetallic compound and act
as stress concentrators upon load. The control of their amount is
crucial for joint efciency. Hence, the welding setup and parameter chosen in this study were crucial for the good acquired
mechanical properties by controlling the tool pin position and
consequently the amount of HSS detached particles and the
interface formation (Figs. 3 and 5) [13].

(e) Failure analysis showed that cracks propagated following the


detached steel particles. Hence, is crucial for the joint efciency
to control the amount of observed inclusions in the AlSZ.

Acknowledgements
The authors gratefully acknowledge the Helmholtz Association
of German Research Centres for nancial support via the virtual
institute Improving Performance and Productivity of Integral
Structures through Fundamental Understanding of Metallurgical
Reactions in Metallic Joints (VI-IPSUS). The authors also would
like to thank Mr. Martin Preilowski for help in sample preparation
during his undergraduate studies at MPIE.

References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]

5. Conclusions
In the present study, friction-stir dissimilar joints between two
grades of HSS (DP600 and HC260LA) and AA6181-T4 Al alloy were
produced applying the same welding parameters and setup (tool
offset position). The studies focussed on the inuence of distinct
HSS BMs on the joint efciency. The conducted analysis can be
summarised as follows:

[17]
[18]
[19]
[20]
[21]
[22]
[23]

(a) Due to the tool offset position, the complex material ow in


the stir zone mainly involves the Al alloy. Crossing all weld
regions in the Al alloy side (BM-HAZ-TMAZ-SZ), strong differences in grain size distribution and shape were observed.
(b) A complete and crack-free bonding between both materials
was observed. The complex and non-smooth interface consists of very ne a-Fe (ferrite) grains and thin strips of Fe2Al5
intermetallic compound. The grain renement and defect
formation in the Fe-TMAZ result in signicant strain
hardening.
(c) The joint efciency showed an ultimate tensile strength equal
to 80% of the Al alloy BM. No evidence of plastic deformation
was observed on the HSS side in either joint. All deformation
during the tensile test took place solely in the Al alloy.
(d) The interface Al alloy BM-HAZ-TMAZ was revealed to be the
weakest element of the joint and is where the fracture always
occurred. The region is characterised by a strong microstructure
gradient and by properties associated with incomplete recovery/
recrystallisation processes.

[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]

T.A. Barnes, I.R. Pashby, J. Mater. Process. Technol. 99 (2000) 6271.


T.A. Barnes, I.R. Pashby, J. Mater. Process. Technol. 99 (2000) 7279.
A. Kochan, Assembly Automation 22 (1) (2002) 2935.
G. Michalos, S. Makris, N. Papakostas, D. Mourtzis, G. Chryssolouris, CIRP J.
Manuf. Sci. Technol. 2 (2010) 8191.
M. Marya, Mater. Sci. Forum 580-582 (2008) 155158.
N.S. Ermolaeva, M.B.G. Castro, P.V. Kandachar, Mater. Des. 25 (2004)
689698.
R.S. Coelho, Joining of Light-Weight Materials by Friction Stir Welding and
Bochum, Germany, 2008.
Laser Beam Welding, Ph.D. Thesis, Ruhr-Universitat
R. Borrisutthekul, T. Yachi, Y. Miyashita, Y. Mutoh, Mater. Sci. Eng. A 467
(2007) 108113.
A. Mathieu, R. Shabadi, A. Deschamps, M. Suery, S. Mattei, D. Grevey, E. Cicala,
Opt. Laser Technol. 39 (2007) 652661.
P. Peyre, G. Sierra, F. Deschaux-Beaume, D. Stuart, G. Fras, Mater. Sci. Eng. A
444 (2007) 327338.
L. Agudo, D. Eyidi, C.H. Schmaranzer, E. Arenholz, N. Jank, J. Bruckner,
A.R. Pyzalla, J. Mater. Sci. 42 (2007) 42054214.
L. Agudo Jacome, S. Weber, A. Leitner, E. Arenholz, J. Bruckner, H. Hackl,
A.R. Pyzalla, Adv. Eng. Mater. 11 (5) (2009) 350358.
R.S. Coelho, A. Kostka, S. Sheikhi, J.F. dos Santos, A.R. Pyzalla, Adv. Eng. Mater
10 (10) (2008) 961972.
R.S. Coelho, A. Kostka, J.F. dos Santos, A.R. Pyzalla, Adv. Eng. Mater 10 (12)
(2008) 11271133.
T. Tanaka, T. Morishige, T. Hirata, Scr. Mater. 61 (2009) 756759.
T. Watanabe, H. Takayama, A. Yanagisawa, J. Mater. Process. Technol. 178
(2006) 342349.
C.M. Chen, R. Kovacevic, Int. J. Mach. Tools Manuf. 44 (2004) 12051214.
H. Uzun, C.D. Donne, A. Argagnotto, T. Ghidini, C. Gambaro, Mater. Des. 26
(2005) 4146.
W.-B. Lee, M. Schmuecker, U.A. Mercardo, G. Biallas, S.-B. Jung, Scr. Mater. 55
(2006) 355358.
H. Springer, A. Kostka, J.F. dos Santos, D. Raabe, Mater. Sci. Eng. A 528 (2011)
46304642.
O. Kubaschewski, Iron-Binary Phase Diagrams, Springer-Verlag, Berlin, 1982.
H. Springer, A. Kostka, E.J. Payton, D. Raabe, A. Kaysser-Pyzalla, G. Eggeler,
Acta Mater. 59 (2011) 15861600.
W.M. Thomas, E.D. Nicholas, J.C. Needham, M.G. Murch, P. Temple-Smith, C.J.
Dawes, Friction Stir Butt Welding, International Patent Application no. PCT/
GB92/02203, December 1991.
R.S. Mishra, Z.Y. Ma, Mater. Sci. Eng. R 50 (2005) 178.
R. Nandan, T. DebRoy, H.K.D.H. Bhadeshia, Prog. Mater. Sci. 53 (2008)
9801023.
D. Lohwasser, Z. Chen (Eds.), Friction Stir Welding: From Basics to Applications, Published by Woodhead Publishing Limited and CRC Press LLC, 2010.
L.E. Murr, J. Mater. Eng. Perform. 19 (8) (2010) 10711089.
C. Leita~ o, B. Emlio, B.M. Chaparro, D.M. Rodrigues, Mater. Des. 30 (2009)
32353242.
N.T. Kumbhar, S.K. Sahoo, I. Samajdar, G.K. Dey, K. Bhanumurthy, Mater. Des.
32 (2011) 16571666.
U.F.H.R. Suhuddin, S. Mironov, Y.S. Sato, H. Kokawa, C.-W. Lee, Acta Mater. 57
(2009) 54065418.
W. Xunhong, W. Kuaishe, Mater. Sci. Eng. A 431 (2006) 114117.
M.P. Miles, T.W. Nelson, R. Steel, E. Olsen, M. Gallagher, Sci. Technol. Weld.
Join. 14 (3) (2009) 228232.
T. Saeid, A. Abdollah-zadeh, T. Shibayanagi, K. Ikeuchi, H. Assadi, Mater. Sci.
Eng. A 527 (2010) 64846488.
M. Ghosh, K. Kumar, R.S. Mishra, Mater. Sci. Eng. A 528 (2011) 81118119.
H.K.D.H. Bhadeshia, T. DebRoy, Sci. Technol. Weld. Join. 14 (3) (2009)
193196.
S. Mironov, Y.S. Sato, H. Kokawa, Acta Mater. 57 (2009) 45194528.
R. Rai, A. De, H.K.D.H. Bhadeshia, T. DebRoy, Sci. Technol. Weld. Join. 16 (4)
(2011) 325342.

R.S. Coelho et al. / Materials Science & Engineering A 556 (2012) 175183

[38] T. DebRoy, H.K.D.H. Bhadeshia, Sci. Technol. Weld. Join. 15 (4) (2010)
266270.
[39] R.S. Coelho, A. Kostka, H. Pinto, A. Rothkirch, J. Dos Santos, A.R. Pyzalla.
Proceedings of the 2008 Denver X-ray Conference, Colorado, USA, 48 August
2008 (manuscript for electronic publication).
[40] A. Kostka, R.S. Coelho, J. dos Santos, A.R. Pyzalla, Scr. Mater. 60 (2009) 953956.
[41] Y. Yong, Z. Da-tong, Q. Cheng, Z. Wen, Trans. Nonferrous Met. Soc. China 20
(2010) s619s623.
[42] P. Venkateswaran, Z.-H. Xu, X. Li, A.P. Reynolds, J. Mater. Sci. 44 (2009)
41404147.
[43] M.A. Mod, A. Abdollah-zadeh, F. Malek Ghaini, Mater. Des. 36 (2012)
161167.

[44]
[45]
[46]
[47]

183

A.C. Somasekharan, L.E. Murr, Mater. Charact. 52 (2004) 4964.


Y. Weia, J. Lib, J. Xiong, F. Huang, F. Zhang, Mater. Des. 33 (2012) 111114.
Y.C. Chen, K. Nakata, Mater. Des. 30 (2009) 39133919.
S. Jana, Y. Hovanski, G.J. Grant, Metall. Mater. Trans. A 41A (2010)
31733182.
[48] U. Dressler, G. Biallas, U.A. Mercado, Mater. Sci. Eng. A 526 (2009) 113117.
[49] J.F. dos Santos, in: G. Amancio , J.F. dos Santos (Eds.), Proceedings of an
International Seminar on Friction Stir Welding of Steels, 2223 November
2007, Geesthacht, Germany.
[50] J. Bacalhau, S. Sheikhi, J.F. dos Santos, in: G. Amancio, J.F. dos Santos (Eds.),
Proceedings of an International Seminar on Friction Stir Welding of Steels,
2223 November 2007, Geesthacht, Germany.

Вам также может понравиться