You are on page 1of 11

Experimental Investigation on

Xun Liu
S.M. Wu Manufacturing Research Center,
Joining Dissimilar Aluminum
Department of Mechanical Engineering,
University of Michigan, Alloy 6061 to TRIP 780/800 Steel
Ann Arbor, MI 48109

Shuhuai Lan1
Through Friction Stir Welding
S.M. Wu Manufacturing Research Center,
Department of Mechanical Engineering,
Friction stir welding (FSW) technique has been successfully applied to butt joining of
University of Michigan,
aluminum alloy 6061-T6 to one type of advanced high strength steel (AHSS), transforma-
Ann Arbor, MI 48109
tion induced plasticity (TRIP) 780/800 with the highest weld strength reaching 85% of
e-mail: lans@umich.edu
the base aluminum alloy. Mechanical welding forces and temperature were measured
under various sets of process parameters and their relationships were investigated, which
also helped explain the observed macrostructure of the weld cross section. Compared
Jun Ni with FSW of similar aluminum alloys, only one peak of axial force occurred during the
S.M. Wu Manufacturing Research Center,
plunge stage. Three failure modes were identified during tensile tests of weld specimens,
Department of Mechanical Engineering,
which were further analyzed based on the microstructure of joint cross sections. Interme-
University of Michigan,
tallic compound (IMC) layer with appropriate thickness and morphology was shown to
Ann Arbor, MI 48109
be beneficial for enhancing the strength of Al–Fe interface. [DOI: 10.1115/1.4030480]

Keywords: friction stir welding, dissimilar Al 6061/TRIP steel, welding force, joint
strength, Al–Fe interface

1 Introduction quality was evaluated through both tensile tests and microstruc-
ture analysis. These experimentally measured force and tempera-
Increasing demand for lightweight vehicles promotes applica-
ture data can serve as verifications of analytical and numerical
tions of multimaterial structures [1], which necessitates develop-
models for FSW of dissimilar materials, which would subse-
ment of reliable and cost-effective joining methods for dissimilar
quently be beneficial in selection and optimization of process pa-
lightweight materials. One of these typical pairs of materials is
rameters. Besides, the range of force and temperature values
aluminum alloy and steel, which are highly difficult to be welded
would be helpful for related fixture designs on FSW equipment,
together due to their great differences in physical and mechanical
which is critical for achieving successful joints especially for
properties. Moreover, the large amount of brittle IMCs formed
FSW process.
during traditional fusion welding process will severely deteriorate
the joint and initiate a fast rupture under applied stress [2–6].
The solid state nature of FSW process provides it with certain 2 Experiment
advantages over traditional fusion welding methods. Not only can
The entire experimental setup is shown schematically in Fig. 1.
it avoid solidification related problems but also the low heat input
TRIP 780/800 steel sheets with the thickness of 1.4 mm were pro-
associated with this process can effectively inhibit formation of
vided from U. S. Steel Corporation. Thickness of the aluminum
brittle IMC, which makes it a promising solution for dissimilar
alloy Al6061-T6511 is 1.5 mm and its chemical compositions and
material joining. Several studies have been carried out on using
mechanical properties are listed in Table 2. Below the workpiece
FSW for butt joining sheets of aluminum alloy and steel and they
is a replaceable backing plate where a hole with diameter of 1 mm
were summarized in Table 1 shown below.
was drilled for mounting thermocouple. A type K thermocouple
From the table it can be seen that materials in the steel side of
was positioned at the back surface of the aluminum workpiece
all the aforementioned works are either mild steel or 304 stainless
and was 1 mm away from the abutting edge. Workpiece and re-
steel, whose yield strength are no more than 400 MPa. So far, few
placeable backing plate were assembled into a specially designed
open literatures have reported FSW of aluminum alloy to AHSSs,
fixture for securely clamping the workpiece both vertically and
which is more desirable in lightweight vehicle structures. TRIP
laterally. All of these were further mounted onto a dynamometer
steel is one type of AHSS. It has both a high mechanical strength
(Kistler 9255B). The dynamometer can measure the mechanical
and work hardening rate based on a certain volume fraction of
welding force in both axial direction Fz and the direction along
retained austenite in its microstructure, which can be transformed
joint line Fx. All FSW experiments were displacement controlled
into martensite during plastic deformation [16]. In this study, we
and performed on the high stiffness M.S. Machining Center.
investigated the feasibility on FSW of Al 6061 to TRIP 780/800
Considering the unsymmetrical feature of FSW process, steel
steel, whose yield strength is 780 MPa and is more than three
was put in the advancing side where tool feed rate and peripheral
times of that of Al 6061. Mechanical welding force exerted on the
tool velocity are positively combined as admissible joint configu-
FSW tool was measured during plunge stage. Thermal history of
rations according to Watanabe et al. [15].
the midpoint of the weld line in aluminum side was also measured
The FSW tool used in this study comprised a conical tapered
and related to the macrostructure of the weld cross section. Joint
nonthreaded pin. Detailed tool dimensions are shown in Fig. 2.
Based on Watanabe et al. [15], optimum joint quality can be
1
Corresponding author. achieved by placing the tool closer to aluminum side instead of a
Contributed by the Materials Division of ASME for publication in the JOURNAL OF
ENGINEERING MATERIALS AND TECHNOLOGY. Manuscript received May 5, 2014; final
symmetric configuration. To quantify this important factor, the pa-
manuscript received April 17, 2015; published online May 11, 2015. Assoc. Editor: rameter of tool offset is introduced and defined as the distance
Said Ahzi. between the FSW tool axis and the faying surface of the two

Journal of Engineering Materials and Technology OCTOBER 2015, Vol. 137 / 041001-1
C 2015 by ASME
Copyright V

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jemta8/934059/ on 04/09/2017 Terms of Use: http://www.asme.o


Table 1 Literatures on FSW of dissimilar aluminum alloys to steel with butt joint configuration

Literature Al alloy Steel Thickness (mm) Rotational speed; welding speed Joint efficiency IMC

Uzun et al. [7] Al 6013-T4 304 stainless steel 4 800 rpm; 80 mm/min 70% N/A
Ghosh et al. [8] Pure Al 304 stainless steel 2.5 1000 rpm; 50 mm/min 82% Fe3Al
Tanaka et al. [9] Al7075-T6 Mild steel 3 400–1200 rpm; 100 mm/min 60% FeAl3
Lee et al. [10] Al 6056-T4 304 stainless steel 4 800 rpm; 80 mm/min N/A FeAl4
Dehghani et al. [11] Al 5186 St 52 mild steel 3 355 rpm; 14–80 mm/min 90% FeAl6, Fe2Al5
Chen and Kovacevic [12] Al6061 AISI 1018 steel 6 914 rpm; 140 mm/min 38.6% Fe4Al13, Fe2Al5
Kovacevic and Jiang [13]
Chen [14] Al6061-T651 SS400 mild steel 6 550, 800 rpm; 54, 72, 90 mm/min 76% N/A
Watanabe et al. [15] Al5083 SS400 mild steel 2 100–125 rpm; 25 mm/min 86% FeAl, FeAl3

Fig. 1 Schematic illustration of the FSW experimental


configuration
Fig. 2 Detailed dimensions of the FSW tool (unit: mm)
materials. Larger tool offset means the tool is shifted more into
aluminum. Since certain fraction of the FSW tool needs to remain were performed on the MTS Insight ten tensile machine at a strain
immersed in the steel side, which will be subjected to severe fric- rate of 103. Microstructural analysis using both optical micros-
tional conditions, the FSW tool is required to be made of refrac- copy and scanning electron microscopy (SEM) was performed on
tory materials with superior thermal and wear resistance at high the joint cross sections perpendicular to the weld line. The metal-
temperatures, such as tungsten carbide [17], tungsten-rhenium lurgical samples were ground and polished according to prepara-
[18], Si3N4 [19], and polycrystalline cubic boron nitride (PCBN) tion guidelines for aluminum alloys considering their lower
[20]. In our study, tungsten carbide with 10% cobalt content was hardness. Energy dispersive spectrometry (EDS) technique has
selected for its combined properties of high hardness, relatively further been utilized for determination of the Al–Fe interfacial
good fracture toughness and much lower cost compared with layer composition.
PCBN.
During initial investigations, three levels of rotational speeds
were attempted and acceptable joint could hardly be obtained
3 Results and Discussion
under the lowest level of 600 rpm regardless of choices of other
parameters. A visible groove adjacent to the joint line can be 3.1 Mechanical Welding Force. Typical curves of the axial
observed on aluminum top surface, which suggests leakage of ma- and traverse force experienced by FSW tool during the whole pro-
terial from the inside weld nugget. This could be accounted for by cess are shown in Fig. 3. During the plunge stage where the rotat-
the insufficient heat generation rate associated with the low rota- ing tool was gradually inserted into workpiece, axial pressing
tional speed. Therefore only the two higher levels of rotational force Fz increased rapidly in the beginning until it arrived at an in-
speed, 1200 rpm and 1800 rpm were further investigated under termediate plateau, which continued for a short period, and then
three levels of welding speed and two levels of tool offset. The increased again until it reached to the peak value when the tool
studied welding conditions are listed in Table 3, where N, v, and shoulder started abrading against the upper surface of steel. Dur-
offset are abbreviations for rotational speed, welding speed and ing dwell stage when the tool was held rotating in position, axial
tool offset, respectively. force Fz decreased a little and lateral welding force Fx remained
Tensile specimens perpendicular to the weld line were prepared negligible. During the final welding stage where the tool trans-
according to the ASTM E8 standard. The welded area was located lated along the weld line to finish the whole joint, the axial force
almost right in the center of the specimen. All of the tensile tests Fz decreased and gradually arrived at a stable value. On the other

Table 2 Chemical composition and mechanical properties of the Al6061-T6511 alloy

Chemical composition Mechanical properties

% Si % Mg % Fe % Cr % Cu % Zn % Ti YS (MPa) UTS (MPa)

0.6 1 0.35 0.42 0.27 0.12 0.08 245 283

041001-2 / Vol. 137, OCTOBER 2015 Transactions of the ASME

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jemta8/934059/ on 04/09/2017 Terms of Use: http://www.asme.o


Table 3 Process conditions for FSW of Al 6061 to TRIP steel force Fx, which is the forced required for translating FSW tool
along weld line, was negligible when welding aluminum. In con-
FSW Rotational Welding Tool offset trast, Fx was much larger when welding aluminum to TRIP steel.
condition speed (N): rpm speed (v): mm/min (offset): mm Interpretation for both the second and third differences could
again be based on the fact that TRIP steel possesses much stronger
1 1200 30 1.03
2 1200 60 1.03
mechanical properties, which make it more difficult to be confined
3 1200 90 1.03 within the weld and stirred along the joint line.
4 1200 30 1.63 Displacement controlled FSW process will change the mechan-
5 1200 60 1.63 ical welding force exerted onto the FSW tool based on different
6 1200 90 1.63 process conditions. In order to evaluate their effects, four charac-
7 1800 60 1.03 teristic values were selected for quantitative description of the
8 1800 90 1.03 force curves. The first two are the axial force Fz during plunge
9 1800 120 1.03 stage and one corresponds to the transient plateau state appears
10 1800 60 1.63 and the other is the peak axial force. The third and last ones are
11 1800 90 1.63
12 1800 120 1.63
the stable axial force Fz and lateral moving force Fx reached dur-
ing welding stage.
Figure 4 compares the effects of tool rotational speed and tool
offset on the axial force during plunge stage. It can be clearly seen
hand, lateral moving force Fx increased rapidly but after certain that higher rotational speed and tool offset can reduce both the
period of fluctuations also reached to a stable state. Compared plateau and peak axial force to a large extent. Raising the rota-
with the works of Trimble et al. [21], Soundararajan et al. [22], tional speed can increase the shear strain rate of the material sur-
and Park [23] where similar aluminum alloys were friction stir rounding the pin and also the rubbing frequency between the pin
welded, three differences in the force curves can be observed. surface and the workpiece, which therefore promote both plastic
First, only one peak axial force existed during the plunge stage for and frictional heat generation and make it easier to initiate the
welding dissimilar materials, whereas two peaks occurred for welding. Larger tool offset implies smaller volume of the pin in
welding same aluminum alloy. This could be explained from the the stronger steel side, which accordingly requires smaller force to
aspect that as FSW tool pin gradually inserted into same alumi- plunge the tool against the workpiece.
num material, heat generated by plastic deformation and friction Figure 5 showed the stable axial force Fz reached during weld-
would increase to a certain point where material deformation re- ing stage, which indicates the required pressure for confining the
sistance was reduced to a much larger degree. Consequently, the material within the weld when translating the tool, were compared
required force for pushing the pin further down can be greatly under different values of welding speed, rotational speed, and tool
reduced. A second peak was then generated afterward when the offset. It can be observed that overall Fz increased slightly with
whole shoulder of FSW tool abraded the workpiece. On the other the traverse welding speed, whereas the relationship was weak in
hand, in this study the volume of tool pin immersed in TRIP steel the range of the studied welding speeds. On the other hand, larger
side continuously grew larger during its downward motion. Since rotational speed and tool offset can effectively reduce axial force,
deformation resistance of this TRIP steel is much higher than alu- which suggested that the stirred materials in the weld nugget were
minum, which can be indicated by their difference in yield more compliant and easily controlled under these conditions.
strength, heat generated by aluminum side can only compensate Lateral moving force Fx during steady-state welding stage
for the additionally required force to stir steel. Accordingly, a under different sets of process parameters was compared in Fig. 6.
short equilibrium state was obtained, which was quickly broken Effect of welding speed on Fx is even smaller. Higher rotational
when the tool shoulder started contacting the surface of alumi- speed can reduce Fx, but the effect is not as significant as that for
num. Peak axial force was generated as the tool shoulder abraded Fz. Tool offset is still an efficient parameter for reducing Fx.
against steel surface. Second, when welding similar aluminum A schematic diagram for illustration of the force distribution on
alloys, axial force Fz can be greatly reduced by 35–60% [21,23] FSW tool during the process was shown in Fig. 7. Axial force
during welding stage whereas no more than 20% reduction can be comes from two regions: the surface of the pin and the shoulder.
observed when welding Al to TRIP steel. Finally, lateral moving Each of them consists of both aluminum and steel sections. If the

Fig. 3 Typical curves of axial and traverse forces experienced by the FSW tool during the
whole process

Journal of Engineering Materials and Technology OCTOBER 2015, Vol. 137 / 041001-3

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jemta8/934059/ on 04/09/2017 Terms of Use: http://www.asme.o


Fig. 4 Effects of tool rotational speed and tool offset on the value of axial force Fz during
plunge stage: (a) plateau axial force and (b) peak axial force

total area of shoulder surface is denoted as As, the fraction of alu- values of these three variables, which means aluminum deforma-
minum region on shoulder as fs, end surface area of pin as Apz, pe- tion stress occupies a greater portion of the total Fx and Fz and
riphery surface area of the pin as Apx, and the corresponding leads to their lower values. On the other hand, in the following
fractions of aluminum part are fpz and fpx, the following equations session it can be shown that welding speed v had an insignificant
can be established: effect on the maximum temperature during the process. Influence
of welding speed on strain rate is also negligible compared with
ðð ðð that of rotational speed, which can be shown by calculating the ra-
Fz ¼ FðzÞ ðzÞ
s þ Fp ¼ rAl dA þ rFe dA tio of the average material moving speed induced by rotation to
ðð
fs As
ðð
ð1fs ÞAs that induced by translation
þ rAl dA þ rFe dA (1) ð rs
fpz Apz ð1fpz ÞApz r  2pN  2prdr=prs2
vrotation 0 2 2pN
ðð ¼ ¼ prs (3)
ðxÞ ðzÞ ðzÞ
vtranslation v 3 v
Fx ¼ Ffs þ FðxÞ
pp ¼ lAl FAl þ lFe FFe þ rAl dA
fpx Apx
ðð ðð ðð where rs is the shoulder radius of the FSW tool, N is the rotational
þ rFe dA ¼ lAl rAl dA þ lFe rFe dA speed, and v is the welding speed. After substituting the lowest
ð1fpx ÞApx fs As ð1fs ÞAs rotational speed of 1200 r/min, the highest welding speed of
ðð ðð 120 mm/min and the shoulder radius of 6.35 mm into this equa-
þ rAl dA þ rFe dA (2) tion, the ratio can be calculated to be 900, which means strain rate
fpx Apx ð1fpx ÞApx of the material is primarily determined by rotational speed. As a
result, welding speed changes stress distribution to a small extent
ðxÞ ðxÞ
where Ffs represents the frictional force on the shoulder and Fpp and the axial and lateral moving force can only be slightly
is the deformation resistance of plasticized material exerted on the affected.
pin. l is the frictional coefficient and r is the flow stress, which is
a function of temperature, strain, and strain rate during FSW pro- 3.2 Thermal History. Figure 8 showed the thermal history at
cess. Since higher rotational speed can elevate temperature distri- 1 mm point in aluminum side and it can be seen that welding
bution in the nugget, which can reduce stress terms in Eq. (1) speed had an insignificant effect on the maximum temperature
therefore the axial force. Values of fs , fpz , and fpx are determined experienced by this point. However, larger welding speed can
by the parameter of tool offset. Larger tool offset leads to higher directly reduce the length of thermal history. Another special

Fig. 5 Effects of various process parameters on the axial force Fig. 6 Effects of various process parameters on lateral moving
Fz during stable welding stage force Fx during stable welding stage

041001-4 / Vol. 137, OCTOBER 2015 Transactions of the ASME

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jemta8/934059/ on 04/09/2017 Terms of Use: http://www.asme.o


of the FSW tool pin. Under the axial pressure of the FSW tool,
this portion of steel was squeezed and extruded into the aluminum
matrix, which finally exhibited as a steel strip embedded in the
retreating side. However, this steel strip was located within the
contour of the pin, which indicated that it was formed behind the
tool pin as the tool moved forward rather than formed simultane-
ously along the side of the pin. This back deposition helped to
explain the observed two peaks in the thermal history of the 1 mm
point at the back surface of aluminum. In the nugget, particles
with different sizes and morphologies were scattered around,
which can be shown through SEM under higher magnification to
be either sheared-off steel fragments encompassed by intermetal-
lics or simply intermetallics.
The larger inclination angle of Al–Fe interface in the advancing
Fig. 7 Schematic diagram for illustration of the force distribu- side than the original cone angle of the tool pin could be explained
tion on FSW tool (side view parallel to the weld line)
from the aspect of temperature distribution. Near the upper sur-
face of workpiece, great amount of heat is generated due to fric-
feature of this thermal history is that unlike FSW of similar mate- tion on the tool shoulder. This amount of heat can be conducted
rials, roughly two temperature peaks appeared during this dissimi- downward and added to the heat generated mainly through plastic
lar material joining process. Material flow and heat convection deformation around the periphery of the FSW tool pin. Upper
needed to be considered to explain this phenomenon. As can be region of the weld can therefore receive both greater amount of
shown in the later session of the macro-image of weld cross sec- conduction heat and plastic deformation heat based on the larger
tion, a steel strip has been extruded from the bottom of advancing pin radius. The elevated temperature can therefore reduce the steel
side and deposit in the wake of the tool. Steel has a much smaller deformation resistance and promote the penetration of stirred over
thermal diffusivity than aluminum, which means it has a much aluminum into steel side. As can be shown in Fig. 10, under the
higher ability to store thermal energy than to do heat conduction. lowest welding speed of 30 mm/min in (a), the Al–Fe interface
When the steel was deposited backward, its associated heat can still exhibited a linear morphology. As the welding speed was
again make a contribution to the temperature raise even when the increased to 60 mm/min, the interface started to show certain
tool already passed. In other words, the first peak was reached as degree of nonlinearity with a smaller slope in the upper region.
the FSW tool arrived at this position, followed by a small temper- The highest welding speed of 90 mm/min in (c) will lead to an
ature drop due to heat dissipation but then a second peak was obvious distinction of slopes in different areas and the inclination
obtained through deposition of steel as the tool moved continu- angle is even more flat in the top. This effect of welding speed
ously forward. On the other hand, during FSW of similar alumi- could be explained from the aspect of thermal history. As shown
num alloys, the convective heat associated with the deposited in the previous session, larger welding speed can directly reduce
aluminum will be quickly dissipated in the whole matrix due to the high temperature duration period in the workpiece, which
heat conduction. Therefore, two peaks can be observed during this means less amount of frictional heat can be conducted downward
dissimilar aluminum to steel welding whereas only one peak tem- within the limited time scope. Leit~ao et al. [24] also reported that
perature exists in the thermal history for welding similar alumi- diminishing the welding speed can obtain a more uniform temper-
num alloys. ature distribution.
In order to evaluate the effects of other process parameters,
including rotational speed and tool offset, this Al–Fe interface in
3.3 Macrostructure of the Cross Section. A typical optical the advancing side was linearly fitted and the average slope was
macro-image of the weld cross section perpendicular to the abut- calculated for each welding condition. The results are shown in
ting edge is shown in Fig. 9, which reveals a good quality joint Fig. 11. It can be observed that larger welding speed can increase
containing neither visible pores nor cracks. Three features can be the inclination angle regardless of different values of rotational
extracted for characterizing microstructure evolution during the speed and tool offset. Higher rotational speed resulted in a steeper
process. In the advancing side, an evident interface between alu- interface, which suggests a more uniform temperature distribu-
minum and steel can be observed and the inclination angle was tion. Equivalently, higher rotational speed can compensate the
larger than the original cone angle of the tool. The inclination uneven temperature effect caused by larger welding speed and at
angle is defined as the angle between the vertical direction and the the same time elevate the whole temperature level. On the other
direction of this interface. In the retreating side, certain amount of hand, the influence of tool offset is not regular. Overall, smaller
steel has been peeled off from the advancing side due to the action tool offset can increase the inclination angle of the interface,
which could be attributed to by the involvement of larger fraction
of steel in the weld zone. Since steel has a relatively low value of
thermal conductivity, the temperature tends to have a less uniform
vertical distribution.
Regarding weld discontinuities, unlike the “wormhole” defect
where void exists at the root of advancing side [25], which is
more commonly observed in FSW of similar aluminum alloys,
during FSW of dissimilar steel to aluminum investigated in this
study, the advancing side always exhibits a continuous interface
without appearance of evident porosity or cracks. On the contrary,
visible discontinuities can exist on the retreating side as shown in
Fig. 12, which is always associated with the stirred over steel
strip. This could be considered from two aspects. One is that this
portion of steel is extruded over and flows upward along with the
aluminum in the wake of the tool. This upward flow, however,
Fig. 8 Thermal history at 1 mm position in Al side under differ- scarcely involves any interaction between aluminum matrix and
ent welding speeds with the rotational speed of 1800 rpm and the steel strip, which therefore hardly makes contribution to
tool offset of 1.63 mm the bonding of the two materials and is likely to generate defects.

Journal of Engineering Materials and Technology OCTOBER 2015, Vol. 137 / 041001-5

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jemta8/934059/ on 04/09/2017 Terms of Use: http://www.asme.o


Fig. 9 Typical view of the cross section normal to the weld line (process condition: rotational
speed of 1800 rpm, welding speed of 60 mm/min, and tool offset of 1.63 mm)

Fig. 10 Effect of welding speed on the inclination angle of Al–Fe interface in the advancing
side under the same rotational speed of 1200 rpm and tool offset of 1.63 mm: (a) 30 mm/min,
(b) 60 mm/min, and (c) 90 mm/min

The other aspect related to defects formation is the difference of promotes formation of porosity and microcracks between steel
the thermal expansion coefficient between aluminum and steel. and aluminum.
Since the coefficient of thermal expansion for aluminum is much
higher than steel, even if they fill up the weld at high temperature, 3.4 Joint Strength and Failure Mode. The highest ultimate
aluminum will tend to shrink more during cooling, which tensile strength (UTS) of FSW joints obtained in this study was
240 MPa, which is about 85% of the base Al alloy. One of the ten-
sile specimens obtained under rotational speed of 1800 rpm, weld-
ing speed of 90 mm/min and tool offset of 1.63 mm is shown in

Fig. 11 Effects of different welding speed, rotational speed, Fig. 12 Defects generated during the FSW process (condition:
and tool offset on the inclination angle of Al–Fe interface in the rotational speed of 1800 rpm, welding speed of 120 mm/min,
advancing side and tool offset of 1.63 mm)

041001-6 / Vol. 137, OCTOBER 2015 Transactions of the ASME

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jemta8/934059/ on 04/09/2017 Terms of Use: http://www.asme.o


Fig. 13. Top and side views of the specimen before tensile test
were shown in the upper two images. The lower three images was
the same specimen after fracture and the third image showed a
bottom view of the joint where the initial butting edge of the two
workpiece can be recognized. Necking phenomenon can be
clearly observed in the heat affected zone of aluminum side. Frac-
ture path is away from the weld nugget and the original faying
surfaces of these two materials have been effectively welded to-
gether. Loss of strength in the aluminum heat affected zone is
likely to be caused by the dissolution and over-aging of fine
precipitates in the Al 6061 after T6 aging.
Relationships between the weld tensile strength and process pa-
rameters are shown in Figs. 14 and 15. In Fig. 14 where the rota-
tional speed was 1200 rpm, it can be observed that a higher
welding speed was likely to reduce the joint strength if the FSW Fig. 14 Tensile strength of the weld specimen under different
tool was shifted less into aluminum side. In contrast, under a tool offsets and welding speeds (rotational speed: 1200 rpm)
larger tool offset, a higher welding speed turned out to be benefi-
cial for the joint quality. Similar trend can also be observed in
1200 rpm and tool offset was 1.63 mm. As the welding speed
Fig. 15 where a larger rotational speed of 1800 rpm was applied.
gradually increased from 30 mm/min to 90 m/min as shown from
Besides, larger tool offset was shown to help increase the weld
Figs. 17(a)–17(c), a growing distance between the tail of steel
reliability with smaller heights of error bars for tensile strength.
strip and upper surface of aluminum can be observed. According
Furthermore, in the range of welding and rotational speeds inves-
to tensile test results, the joint strength also increased from
tigated in this study, a larger tool offset can overall increase the
Figs. 17(a)–17(c), which implied that smaller stirred over steel
tensile strength except the case where rotational speed is
strip can yield stronger joint. Figure 18 showed two samples
1800 rpm and welding speed is 60 mm/min. It can be anticipated
welded under the same process conditions, where the rotational
that if the tool is further moved into aluminum, the tensile strength
speed is 1800 rpm, welding speed is 120 mm/min and the tool off-
would start to drop at a certain stage. In other words, there exists
set is 1.03 mm. The UTS of the specimen in Fig. 18(a) was
an optimum tool offset to achieve the highest strength. Under a
approximately 36.7 MPa smaller than that in Fig. 18(b). Length of
lower welding speed and higher rotational speed, this optimized
the stirred over steel strip was approximately the same in these
tool offset value can be smaller and more fraction of the pin in
two cases, which indicated that steel strip that was vertically
steel side can yield better strength.
upward or tilted toward retreating side is more harmful to the joint
Three types of failure modes can be identified. Except necking
strength than the steel strip with an inclination angle toward the
and fracture in aluminum side as shown in Fig. 13, the other two
nugget.
are both related to the interface between this hooking steel strip
and the aluminum matrix. Most of the fractures occurred along
the outside boundary of the steel strip, which was marked out as 3.5 Microstructure Analysis. Initially, the hooking steel in
position 1 in Fig. 9 and the rest occurred along the inside bound- aluminum side was anticipated to be beneficial to the joint
ary show as position 2. Joint strength of weld specimen was strength in the sense that it can be considered as an enhancing rib
closely related to the morphology and size of this hooking steel structure in a relatively soft matrix. However, tensile tests showed
strip. If the vertical part of this steel strip was so long that it pene- an opposite results. In order to understand the underlying mecha-
trated the upper surface of the aluminum alloy in the retreating nisms for the current failure modes, SEM was used to analyze the
side, the fracture would take place at the inside boundary of steel Al–Fe interface at different locations of the weld. Figure 19
strip, as shown in Fig. 16, which greatly deteriorated the joint and showed the SEM results of the cross section in Fig. 9. A thin layer
resulted in the lowest strength. of IMC with a different color from either aluminum or steel can
In most cases, the steel strip was covered beneath the upper sur- be observed along the Al–Fe interface in the advancing side along
face of aluminum and the specimens fractured along the outside base steel as well as the inside boundary of steel strip. Moreover,
boundary, as shown in Figs. 17 and 18. Effects of different weld- the thickness of this intermetallic (IMC) layer was less than 1 lm.
ing speeds on the morphology and size of the hooking steel strip In contrast, along the outside boundary of the steel strip no inter-
observed on the joint cross section and corresponding joint metallic layer can be identified. This implied that either no inter-
strength were compared in Fig. 17, where the rotational speed was metallic was formed or the thickness of this layer was in the order
of nanometers. Together with tensile tests results, it can be

Fig. 13 Tensile specimen before and after breakage (process


conditions: rotational speed of 1800 rpm, welding speed of Fig. 15 Tensile strength of the weld specimen under different
90 mm/min, and tool offset of 1.63 mm) tool offsets and welding speeds (rotational speed: 1800 rpm)

Journal of Engineering Materials and Technology OCTOBER 2015, Vol. 137 / 041001-7

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jemta8/934059/ on 04/09/2017 Terms of Use: http://www.asme.o


Fig. 16 Specimen broken along the inside boundary of the steel strip (another sample from
the process conditions with rotational speed of 1800 rpm, welding speed of 60 mm/min, and
tool offset of 1.63 mm)

Fig. 17 Effects of welding speeds on the size of stirred over steel strip and joint strength
under rotational speed of 1200 rpm and tool offset of 1.63 mm: (a) 30 mm/min, (b) 60 mm/min,
and (c) 90 mm/min

concluded that a layer of intermetallic with appropriate thickness against the stress intensity, which as a result would enhance joint
can contribute to the joint strength despite of its general brittle quality.
mechanical properties. The underlying mechanism can be consid- In order to further validate this statement, the fractured necking
ered from mechanics analysis. Stress singularity exists at dissimi- specimen shown in Fig. 13 was also metallurgically prepared and
lar material interface, which is liable to initiate cracks under put under SEM and the results were shown in Fig. 20. In this case,
loading and cause failures. On the other hand, the presence of sheared off steel fragments are more finely distributed and diffi-
high strength intermetallic layer can alleviate this effect and resist cult to recognize through lower magnification optical microscope.

Fig. 18 Effects of inclination of the stirred over steel strip on joint strength obtained under
the same process conditions with the rotational speed of 1800 rpm, welding speed of 120 mm/
min and tool offset of 1.03 mm: (a) UTS of 180.8 MPa and (b) UTS of 144.1 MPa

041001-8 / Vol. 137, OCTOBER 2015 Transactions of the ASME

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jemta8/934059/ on 04/09/2017 Terms of Use: http://www.asme.o


Fig. 19 SEM analysis results for the cross section shown in Fig. 9

Moreover, the height of the steel strip is much smaller than that in a great extent. Elemental mapping through EDS has also been
Fig. 19. Most importantly, presence of intermetallic layer can be conducted for the Al–Fe interface at both the tip of steel strip and
observed to exist along the Al–Fe interface not only in the advanc- the advancing side. The results were shown in Figs. 21 and 22,
ing side but also at the stirred over steel strip, which accordingly respectively, and these concentration profiles again indicated the
verified that a suitable intermetallic layer can contribute the joint interdiffusion phenomenon of aluminum and iron atoms across
strength. A vortexlike structure was revealed at the tip of this the interface.
stirred over steel strip, which can be regarded as micro-interlocks Compositions of the Al–Fe interface were also measured at sev-
between steel and aluminum and enhance the interface strength to eral positions and the results were listed in Table 4. The atom ratio

Fig. 20 SEM analysis results for the cross section of the fractured necking specimen in
Fig. 13

Journal of Engineering Materials and Technology OCTOBER 2015, Vol. 137 / 041001-9

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jemta8/934059/ on 04/09/2017 Terms of Use: http://www.asme.o


Fig. 21 Elemental mapping of the Al–Fe interface in the advancing side, which corresponds
to position 1 in Fig. 20

Fig. 22 Elemental mapping of the Al–Fe interface at the tip of the stirred over steel strip,
which corresponds to position 4 in Fig. 20

Table 4 Elemental composition of the Al–Fe interface at differ- • A stirred over steel strip embedded in the aluminum can be
ent locations (atm %) observed on the weld cross section. Three failures modes
were identified. Except necking and fracture in aluminum
Tip of steel strip Advancing side side, the other two are related to the outside and inside boun-
(position 4 in Fig. 20) (position 1 in Fig. 20) daries of this stirred over steel strip.
• Intermetallic layer with an appropriate thickness and mor-
Position Al (%) Fe (%) Al (%) Fe (%)
phology can contribute to the joint strength.
1 60.07 35.41 24.41 75.59
2 76.83 20.08 24.70 73.25 Acknowledgment
3 67.66 30.31 25.63 72.12
This work was supported by the CERC-CVC U.S.–China Pro-
gram of Clean Vehicle under Award No. DE-PI0000012 and
National Science Foundation (Grant No. 1266088, Investigation
between aluminum and iron is much higher at the tip of the steel of Electro-Plastic Effect on Advanced High Strength Steels and
strip than that in the advancing side. Considered together with the Its Application in Friction Stir Joining of Dissimilar Material).
Al–Fe diagram, Al2Fe or Al13Fe4 was indicated to be the interme-
tallic phases that existed at the tip of steel strip. On the other References
hand, for the major Al–Fe interface at the advancing side the com- [1] Sakiyama, T., Murayama, G., Naito, Y., Saita, K., Oikawa, Y. M. H., and Nose,
position was more likely to be Fe3Al. T., 2013, “Dissimilar Metal Joining Technologies for Steel Sheet and Alumi-
Another feature in Fig. 20 was the presence of a small crack at num Alloy Sheet in Auto Body,” Nippon Steel & Sumitomo Metal Corp.,
the inside boundary of steel strip, which could probably be intro- Tokyo, Nippon Steel Technical Report No. 103, pp. 91–98.
duced during the tensile loading test or have already formed asso- [2] Ogura, T., Saito, Y., Nishida, T., Nishida, H., Yoshida, T., Omichi, N., Fuji-
moto, M., and Hirose, A., 2012, “Partitioning Evaluation of Mechanical Proper-
ciated with the welding process. In either situation, it was ties and the Interfacial Microstructure in a Friction Stir Welded Aluminum
suggested that this small crack can sustain the tensile test before Alloy/Stainless Steel Lap Joint,” Scr. Mater., 66(8), pp. 531–534.
the actual fracture took place at the heat affected zone of [3] Das, H., Basak, S., Das, G., and Pal, T. K., 2013, “Influence of Energy Induced
From Processing Parameters on the Mechanical Properties of Friction Stir
aluminum. Welded Lap Joint of Aluminum to Coated Steel Sheet,” Int. J. Adv. Manuf.
Technol., 64(9–12), pp. 1653–1661.
4 Conclusions [4] Movahedi, M., Kokabi, A. H., Seyed Reihani, S. M., Cheng, W. J., and Wang,
C. J., 2013, “Effect of Annealing Treatment on Joint Strength of Aluminum/
FSW technique was successfully applied to joining of Al 6061 Steel Friction Stir Lap Weld,” Mater. Des., 44, pp. 487–492.
alloy to TRIP 780/800 steel and the maximum tensile strength [5] Movahedi, M., Kokabi, A., Reihani, S., and Najafi, H., 2012, “Effect of Tool
Travel and Rotation Speeds on Weld Zone Defects and Joint Strength of Alu-
obtained is about 85% of the base Al alloy. Effects of different minium Steel Lap Joints Made by Friction Stir Welding,” Sci. Technol. Weld.
sets of process parameters were studied in aspects of mechanical Join., 17(2), pp. 162–167.
welding force, temperature distribution, and joint strength. Micro- [6] Uematsu, Y., Kakiuchi, T., Tozaki, Y., and Kojin, H., 2012, “Comparative
structure analysis was also carried out on weld cross sections to Study of Fatigue Behaviour in Dissimilar Al Alloy/Steel and Mg Alloy/Steel
Friction Stir Spot Welds Fabricated by Scroll Grooved Tool Without Probe,”
understand joint failure mechanism. Main findings of this study Sci. Technol. Welding Join., 17(5), pp. 348–356.
are summarized as follows: [7] Uzun, H., Dalle Donne, C., Argagnotto, A., Ghidini, T., and Gambaro, C.,
2005, “Friction Stir Welding of Dissimilar Al 6013-T4 To X5CrNi18-10 Stain-
• Only one peak axial force instead of two was observed during less Steel,” Mater. Des., 26(1), pp. 41–46.
the plunge stage. [8] Ghosh, M., Kar, A., Kumar, K., and Kailas, S., 2012, “Structural Characterisa-
• Higher rotational speed and larger tool offset can effectively tion of Reaction Zone for Friction Stir Welded Aluminium-Stainless Steel
Joint,” Mater. Technol.: Adv. Perform. Mater., 27(2), pp. 169–172.
reduce mechanical welding force. Welding speed has an in- [9] Tanaka, T., Morishige, T., and Hirata, T., 2009, “Comprehensive Analysis of
significant effects on either the welding force of peak temper- Joint Strength for Dissimilar Friction Stir Welds of Mild Steel to Aluminum
ature but is directly related to the length of thermal history. Alloys,” Scr. Mater., 61(7), pp. 756–759.

041001-10 / Vol. 137, OCTOBER 2015 Transactions of the ASME

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jemta8/934059/ on 04/09/2017 Terms of Use: http://www.asme.o


[10] Lee, W.-B., Schmuecker, M., Mercardo, U. A., Biallas, G., and Jung, S.-B., [18] Barnes, S., Bhatti, A., Steuwer, A., Johnson, R., Altenkirch, J., and Withers, P.,
2006, “Interfacial Reaction in Steel–Aluminum Joints Made by Friction Stir 2012, “Friction Stir Welding in HSLA-65 Steel: Part I. Influence of Weld Speed
Welding,” Scr. Mater., 55(4), pp. 355–358. and Tool Material on Microstructural Development,” Metall. Mater. Trans. A,
[11] Dehghani, M., Amadeh, A., and Akbari Mousavi, S., 2013, “Investigations on the 43(7), pp. 2342–2355.
Effects of Friction Stir Welding Parameters on Intermetallic and Defect Formation [19] Ohashi, R., Fujimoto, M., Mironov, S., Sato, Y., and Kokawa, H., 2009, “Effect
in Joining Aluminum Alloy to Mild Steel,” Mater. Des., 49, pp. 433–441. of Contamination on Microstructure in Friction Stir Spot Welded DP590 Steel,”
[12] Chen, C. M., and Kovacevic, R., 2004, “Joining of Al 6061 Alloy to AISI 1018 Sci. Technol. Weld. Join., 14(3), pp. 221–227.
Steel by Combined Effects of Fusion and Solid State Welding,” Int. J. Mach. [20] Park, S. H. C., Sato, Y. S., Kokawa, H., Okamoto, K., Hirano, S., and
Tools Manuf., 44(11), pp. 1205–1214. Inagaki, M., 2009, “Boride Formation Induced by PCBN Tool Wear in
[13] Kovacevic, R., and Jiang, W. H., 2004, “Feasibility Study of Friction Stir Weld- Friction-Stir-Welded Stainless Steels,” Metall. Mater. Trans. A, 40(3), pp.
ing of 6061-T6 Aluminium Alloy With AISI 1018 Steel,” Proc. Inst. Mech. 625–636.
Eng., Part B, 218(10), pp. 1323–1331. [21] Trimble, D., Monaghan, J., and O’Donnell, G., 2012, “Force Generation During
[14] Chen, T., 2009, “Process Parameters Study on FSW Joint of Dissimilar Metals Friction Stir Welding of AA2024-T3,” CIRP Ann., 61(1), pp. 9–12.
for Aluminum–Steel,” J. Mater. Sci., 44(10), pp. 2573–2580. [22] Soundararajan, V., Zekovic, S., and Kovacevic, R., 2005, “Thermo-Mechanical
[15] Watanabe, T., Takayama, H., and Yanagisawa, A., 2006, “Joining of Aluminum Model With Adaptive Boundary Conditions for Friction Stir Welding of Al
Alloy to Steel by Friction Stir Welding,” J. Mater. Process. Technol., 178(1–3), 6061,” Int. J. Mach. Tools Manuf., 45(14), pp. 1577–1587.
pp. 342–349. [23] Park, K., 2009, “Development and Analysis of Ultrasonic Assisted Friction Stir
[16] Kuziak, R., Kawalla, R., and Waengler, S., 2008, “Advanced High Strength Welding Process,” Ph.D. thesis, University of Michigan, Ann Arbor, MI.
Steels for Automotive Industry,” Arch. Civil Mech. Eng., 8(2), pp. 103–117. [24] Leit~ao, C., Louro, R., and Rodrigues, D., 2012, “Using Torque Sensitivity
[17] Jafarzadegan, M., Feng, A., Abdollah-Zadeh, A., Saeid, T., Shen, J., and Analysis in Accessing Friction Stir Welding/Processing Conditions,” J. Mater.
Assadi, H., 2012, “Microstructural Characterization in Dissimilar Friction Stir Process. Technol., 212(10), pp. 2051–2057.
Welding Between 304 Stainless Steel and st37 Steel,” Mater. Charact., 74(1), [25] Reynolds, A. P., 2008, “Flow Visualization and Simulation in FSW,” Scr.
pp. 28–41. Mater., 58(5), pp. 338–342.

Journal of Engineering Materials and Technology OCTOBER 2015, Vol. 137 / 041001-11

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jemta8/934059/ on 04/09/2017 Terms of Use: http://www.asme.o

You might also like