Anda di halaman 1dari 11

Xin Wu1

Mem. ASME
Mechanical Engineering,
Wayne State University,
Detroit, MI 48202
e-mail: xwu@eng.wayne.edu

Jianhui Shang2
Mem. ASME
Hirotec America,
Auburn Hills, MI 48326
e-mail: jshang@ewi.org

An Investigation of Magnetic
Pulse Welding of Al/Cu
and Interface Characterization
This paper investigated the effect of magnetic pulse welding (MPW) condition (welding
power, surface scratches, and contamination) on the establishment of welding between
aluminum and copper tubes, and the associated welding mechanisms. The results showed
that higher applied power and surface scratches in tangential direction were in favor for
good weld, and oil on the surface prevented welding. Direct evidences were obtained on
local interface melting under a high welding power. CuAl intermetallics with different
atomic ratios were identified by energy dispersion spectrum (EDS) chemical analysis and
by microscratching test. The mechanisms of MPW and the process improvement were
discussed. [DOI: 10.1115/1.4027917]
Keywords: magnetic pulse welding, electromagnetic welding, aluminumcopper welding,
surface contamination, microstructure characterization

Introduction
A high power magnetic pulse (MP) generated from electromagnetic coils can be used for material processing, such as MP
cutting, forming, and welding. The magnetic pulse welding
(MPW), also called electromagnetic welding, applies a high pressure pulse force on the mating interface and achieve a strong
bonding without physical contact between tool (here electromagnetic coil) and workpiece, thus bring various advantages. It is generally recognized that the MPW allows welding dissimilar
materials, including thermodynamically incompatible materials
which otherwise is required for fusion welding. It prevents the use
of a consumable electrode, and it eliminates heat-affected zone
and improve welding quality and processing repeatability. In addition, it is also considered as an environmental-friendly material
process since no postcleaning or finishing is needed as often seen
in many other welding processes. In recent years, for energy saving and environment protection in automotive manufacturing,
there is a high demand to weld dissimilar materials for lightweight
structures, and to weld electrical power components made of copper and aluminum (e.g., power cables, connectors and terminals)
in electrical vehicles that is difficult to be welded by fusion welding, due to very high electrical and thermal conductivities. For
these reasons this welding technique has been recognized as a viable approach to join same or dissimilar materials and attract new
theoretical and practical interests [1].
The initial work on electromagnetic force generation techniques
and its application in metal forming and welding were developed
in the former Soviet Union, see a comprehensive handbook by
Belyy et al. [2]. There are some earlier work published on welding
system development, including the MPW study in vacuum with
preheating by Strizhakov [3], the MPW with arc heating by
Yablochnikov [4], the design of apparatus and inductors by
Yablochnikov [5,6] for welding large diameter thin-walled pipes.
In recent years the study on MP forming and welding of lightweight materials has received increased interest, and a series of
papers were published by Daehns group at Ohio State University,
among which Tamhane et al. [7] studied on sample size effect on
the ductility of a ring expansion process, Daehn et al. [8,9]
1

Corresponding author.
Present address: Edison Welding Institute, Columbus, OH 43221.
Contributed by the Manufacturing Engineering Division of ASME for publication
in the JOURNAL OF MANUFACTURING SCIENCE AND ENGINEERING. Manuscript received
February 8, 2013; final manuscript received June 24, 2014; published online August
6, 2014. Assoc. Editor: Wei Li.
2

reported the improved formability in electromagnetic sheet forming, Thomas et al. [10] reported the forming limit testing and analysis on aluminum tube expansion, Fenton and Daehn [11]
performed a modeling work on electromagnetic sheet forming,
Golowin et al. [12] developed a new type of electromagnetic actuator with uniform pressure, and Zhang et al. [13,14] studied on the
microstructures at the vicinity of the welded AlCu interface
using electron backscattering diffraction method (EBSD) and
showed that severely deformed and recrystallized fine-grained
materials produced near the welded interface.
In terms of the workpiece materials to be joined and the welding process conditions, MPW of Al to Al was studied by Shribman et al. [15], who reported that the strength of the welds
reached that of the base metal (A7075-T6). The effect of process
parameters on welding aluminum sheets was reported by Kore
et al. [16], who showed that for a given discharge energy the shear
strength of the welds reached the maximum value at an optimum
coil standoff distance, and the geometry of the coil also has important effect on the product strength. On welding Al to Cu,
Marya et al. [17] and Marya and Marya [18] studied the microstructures and temperatures at the aluminumcopper interface,
and the results showed that a hard copper rich intermetallic phase
with the same composition as the equilibrium c-Cu2Al was
formed. These, along with the observed interfacial voids, were
used as the evidences of Al melting, and a simple analytical model
was used for estimating the interfacial temperatures that provided
further support for interface melting. Aluminum to steel welding
was reported by Aizawa et al. [19], who studied the welding
parameters for several aluminum alloys (A1050, A2017, A3004,
A5182, A5052, A6016, and A7075) to weld with a steel. For
welding Al to Mg, Ben-Artzy et al. [20] reported the formation of
intermetallic phases with different compositions, and suggested
that a rapid solidification occurred within a thin melted layer at
the interface, and an energy balance analysis was provided to indicate that there was enough energy to melt a thin interfacial layer.
For MPW of Ti to Al, Marya and Gerard [21] provided a brief discussed on Ti welding in this overview paper. Many efforts have
been made on modeling MPW process, see a review article by
El-Azab et al. [22]. Although MP force is more readily to be
applied on conductive metals that allow induced current to generate repulsive magnetic field and force to the applied one from the
coil, the MP force can also be used for consolidating powders and
joining nonmetals, nonconductive materials indirectly through
conductive capitulation, such as that in powder consolidation
[23,24]. Another material joining technique with great similarity

Journal of Manufacturing Science and Engineering


C 2014 by ASME
Copyright V

OCTOBER 2014, Vol. 136 / 051002-1

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

to the MPW is explosive welding, and the research results in this


area are also helpful in understanding the MPW process, but will
not be reviewed here. In the above mentioned studies, although
the processing conditions of MPW have been studied extensively,
the surface topological feature and contamination on the welding
process has not been addressed, which should also affect the welding strength.
Significant efforts have been made to understand the mechanisms responsible for material joining at a high speed impact, but
due to the highly nonequilibrium condition of the process the preexisting knowledge in material process and behaviors, thermodynamics and kinetics under an equilibrium or quasi-equilibrium
conditions may not be able to fully explain the current complex
phenomena, and controversy speculations on observation interpretations often exist. While MPW has been commonly viewed by
majority of the papers as a solid-state welding process, which is
mainly based on observations that no heat affected zone exists,
several relatively recent papers used more advanced techniques
on analyzing interface microstructure and chemistry, and on thermal analysis that suggested the possibility of interface melting,
see previously mentioned works [17,18,20]. In addition, Stern and
Aizenshtein [25] reported the possibility of melting and solidification near joint interface based on the observation that along the
bonding interface either a discontinuous pocket string or a continuous transition layer were formed, which was explained either due
to grain refinement from local melting followed by rapid solidification, or due to the formation of intermetallic phases (for dissimilar materials). On the other hand, the work by Zhang et al. [13] on
EBSD study of aluminum MPW showed different findings of a
recrystallized/refined microstructure from heavily deformed
grains with high grain boundary angles, implying that the microstructure evolution was still within the solid state phase and the
melting point was not reached. It is notable that both melting/
rapid solidification and solid-state deformation/recrystallization
can produce a fine-grained microstructure with low dislocation
density and with high-angle grain boundaries; despite the two are
totally different processes at different temperature ranges. However, if a heavily deformed grain shape and high dislocation density were observed, the process must be in solid state, and neither
recrystallization nor melting would occur, otherwise the deformation microstructure would disappear. It is commonly accepted
[26] that recrystallization occurs after certain critical degree of
cold work followed by heating to 1/32/3 absolute melting temperature Tm, and for a pure metal the recrystallization temperature
would be lower, normally 0.4 Tm. It is not so clear weather it is
possible that under a very high energy input rate and strain rate,
within very short time period, a deformed system can remain its
heavily deformed state without recrystallization, or a much higher
superheating temperature is need for melting to occur. Thus, it is
interesting to further investigate the welding conditions and the
involved three debatable welding processes: (1) melting followed
by solidification/crystallization, (2) deformation followed by
recrystallization (without melting), (3) there is no sufficient heat
and time of heating for recrystallization or even recovery, so that
the stored energy from electromagnetic deformation will not be
released, especially when the time is very short. Another interesting issue in MPW is the formation of wavy interface, which has
been widely observed, and commonly considered as a result of
shear instability similar to that observed in fluid dynamics, see
Nassiri et al. [27] and Ben-Artzy et al. [28], but the effect of original surface morphology on the interface waviness and bonding
strength is not reported.
In this paper, MPW of Cu/Al tube is further studied with the
focus on the effect of initial surface conditions (surface topology
and contamination) on the joint establishment, which is an issue
not being previously reported. The welded interfaces were characterized, and the new phase formation was identified and analyzed,
and more importantly, some direct evidences on interface melting
were obtained. The mechanisms of PMW and the means of process improvement are discussed.
051002-2 / Vol. 136, OCTOBER 2014

Experimental Condition
Electromagnetic Welding Setup, Specimens, and Welding
Conditions. A MP generator used for this study is made by
Hirotec America (Model Pulsar MP-30I9 Research Edition). It
consists of a capacitor bank and a high voltage cabinet for charging the capacitors, capable of generating 30 kJ at a charging voltage of 9 kV. The MP coil system is shown in Figs. 1(a)1(d),
which consists of a 5-turn coil in connection with the MP generator and a magnetic field concentrator to intensify the magnetic
force applied onto the aluminum tube wall.
The as-received tubes of aluminum alloy AA6063-O and pure
copper C110 (abbreviated as Al and Cu thereafter) were machined
by lathe turning on both inner and outer surfaces over the joint
lengths. While the Al tube has uniform cross section over length
of 1.5 mm wall thickness (see Fig. 1(e)), the Cu tube has different
wall thicknesses in its end region, following an earlier design by
Shang [29] for supporting the squeezing force and for axial alignment, resulting in three AlCu weld-zones, see Fig. 1(f). Before
and during welding both Al and Cu tubes were firmly clamped
from outside of the welding zone, but inside the welding zone the
two tube ends were free from constraints in radial and axial
directions.
Three initial copper tube surface conditions were prepared for
investigating their effects on welding

Surface condition-A: The as-machined tube surface, produced by lathe turning with the finishing surface containing
tangential scratches over its length. See Figs. 2(a) and 2(b).
Surface condition-B: After the lathe turning (condition-A)
additional manual sanding was performed on the Cu outer
surfaces with 200-grit coarse sand papers along the axial
direction, which replaced the original tangential lathe
scratches (Fig. 2(b)) with a new set of axial scratches on the
Cu outer surface (Fig. 2(c)), in order to investigate the surface
topology effect.
Surface condition-C: After the lather turning (condition-A) a
silicon-based high-viscosity lubricant oil was applied over
the welding zone surfaces, to produce an artificially contaminated interface (exaggerated) before welding, in order to
investigate if the high impact force can break down the contamination layer to establish weld.

Three applied voltages of 4.3 kV, 5.2 kV, and 6.0 kV were used
for charging the battery bank, which is proportional to the energy
and power discharged to the coil and is indirectly related to the
welding MP force when other settings are identical. The coil current and output energy were measured during welding. The actual
energy delivered to the workpieces can be much reduced due to
the loss of magnetic field in the coil-workpiece coupling, and the
Joule heat in the coil loop. The energy received by the workpiece
is dissipated mainly in the forms of plastic deformation, the
induced current within the workpiece, and the frictional work at
the Cu/Al interface, all of which contribute the heating of
workpiece.
Welding Quality Evaluation. The welding quality was evaluated by peeling test and by interface microstructure examination.
For the peeling test two parallel cuts in 5-mm spacing were made
on the Al tube wall along the axial direction. A plier was used to
clamp the Al cut strip from its chopped-off end, and manually
twisted/rolled along the welded strip top. The welding quality was
qualitatively categorized as welded if the peeled surfaces
showed that the fracture mainly occurred within one of the base
metals, or as not-welded if the observed separation mainly
occurred at the original CuAl contact surfaces, associated with
limited resistance to peel the strip off.
Microstructure examination on the sectioned/polished centralplane surfaces was performed to see if the Al and Cu are tightly
connected without gap, or if a gap exists between Al and Cu. By
image processing the percentage of the bonded interface length
Transactions of the ASME

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 1 The electromagnetic system (a) and its assembly schematics (b), consisting of an electrical coil (c), a field concentrator made of copper (d), the tubular workpieces Al (e), and Cu
with special end geometry (f)

directions, as well as von Mises effective strain over the tube axial
position x were calculated based on the following formulations:


dx  tx
(1)
eh x ln
d0  t0
 
tx
(2)
et x ln
t0




A0
d0  t0
t0
ln

eh et (3)
ex x ln
Ax
dx  tx tx

0:5
(4)
ee 2=3e2x e2h e2t

Fig. 2 As-machined Al and Cu tube samples before welding (a)


and the produced surface scratches on the Cu tubes in hoop
direction (b). The further sanded surface is shown in (c).

over the total interface length was obtained as another way to


evaluate the welding quality.

Measurement of Strain Distribution and Plastic Work. The


cross-sectional images were also used for dimension and strain
measurement. The binary images were input into a user-developed
MATLAB program, and the axial and radial coordinates (x, y) of Al
and Cu boundaries were identified and obtained. From the outer
diameter d and thickness t of Al and Cu at each x-coordinate (pixel)
the average strain components in hoop, thickness and axial
Journal of Manufacturing Science and Engineering

where the subscripts x, h, t stand for axial, hoop, and thickness


directions, respectively, and d(x), t(x), and A(x) are the outer
diameter, thickness, and the cross-sectional area at section x; the
subscript 0 stands for their initial values. The work done per
volume, w(x), is the product of the effective strain and stress; here
the mean of yield strength and ultimate strength Y from references
was used as the ideal work without considering strain hardening
and strain rate effect on the flow stress. The total plastic work W
of entire deformation zone was obtained by further integrating the
unit energy per volume w over weld zone volume V over the x
range [a, b]
ee
re dee  Yee x
(5)
wx
W

b
a

wx  Axdx

pdx  tx  tx  Y  ee xdx (6)

OCTOBER 2014, Vol. 136 / 051002-3

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

This work is one part of the total energy dissipation in the MPW,
to be used and discussed for later estimation of the heat
generation.
Microstructure Characterization. The welding conditions
were repeated for producing samples not damaged by peeling test,
for microstructure examination. The welded samples were slowly
cut along the central axis with a bench saw under water cooling,
to avoid possible microstructural change. Samples were ground to
the finishing step with 4000-mesh metallurgical sand paper under
water cooling, followed by cloth polishing to 0.05 lm Al2O3 powder abrasives water suspension in the finishing pass. For each sample the entire weld zone was photographed with a stereo
microscope (for low magnification image) and with a metallurgical microscope (by Olympus), both allows mounting a digital
camera with image software (OptixCam OCView).
Additional examinations on microstructures and fractured
surfaces were performed under a scanning electron microscopy
(SEM, model Hitachi-2400). The chemical composition distribution across the interfaces was analyzed with an EDS probe, and
the obtained chemical compositions were compared with the equilibrium CuAl binary phase diagram. In addition, in order to characterize the local mechanical properties near the interface region,
microscratch test was performed under a constant load and with a
microindenter that slided across the interface on a sectioned and
polished surface, and the scratched line width was measured to
reflect the strength variation near the interface region relative to
the two base metals.

Results
General Features of the MPW and the Effect on Applied
Welding Energy. A typical recorded coil current vs. time curve,
performed at 5.2 kV charging voltage, is shown in Fig. 3(a). The
current curve shows a sinusoidal shape with rapid decay in the

amplitude, and the maximum current of 178 kA was reached at


2  105 s at the first peak, and the measured total welding output
energy was 11.2 kJ. Varying the charging voltage has resulted in
different peak current and output energy, but the curve shape
remained the same. The peeled and also central-sectioned sample
at 5.2 kV, see in Fig. 3(b), shows significant diameter reduction
within the welding zone, for both Al and Cu, and the magnitude
of the diameter reduction increases toward Al tube end (from left
to right). The Al tube wall is almost in full contact with the Cu
tube; except for the Cu corner transition zone-II. Figure 3(c)
shows that the peeled/fractured Al surface is very rough, and certain amount of aluminum materials left over on the Cu surface.
Due to the strong bonding significant force and torque were
needed to peel the Al strip apart from the welded interface. This
welded result applies to all surface conditions A and B. However, the oil-contaminated sample (surface condition C) was not
welded in the sense that during sample sectioning one half of the
specimen was broken from the Al and Cu weld surface. Another
nonbroken half had less than 10% of bonded length, and was easily separated. The percentage of the welded length over total interface length on the sectioned surface is given in Table 1.
Weld Zone Deformation and Bonded Length. For surface
condition A Fig. 4 shows the selected axial cross-sectional images
under different welding voltage that were used for the measurement of strains, and Figs. 5(a)5(c) show the measured distributions of true strain components of Cu and Al at 5.2 kV as an
example. It can be seen that toward the Al end the strain increases
with x, but in the transition zone II the high strain is due to the
over calculation of its wall thickness under the current thickness
measurement/calculation method. Figure 5(d) shows the effective
strain distribution for all surface conditions and the applied charging voltages. By integration of the plastic work for all cases
Fig. 5(e) shows that the total plastic works of Al, Cu and their sum
increase with increasing the charging voltage. Figure 5(f) shows the
percentage welded length increases with voltage, for surface condition A. For surface condition B the welded length is slightly lower
than A, and for the contaminated surface the welded length is much
lower than condition A, based on limited samples available.
In above true strain calculation using Eqs. (5) and (6), the yield
stresses were from available data [30,31], and an average value of
the yield stress and tensile strength was taken, for copper (C110)
at 144.5 MPa and for aluminum (AA6063-O) at 82.7 MPa. The
error in the strain measurement, based on the transition zone-II
strain measurement using current coordinate for tilted tube dimensions for actual wall thickness strain, is in the range of 615%, and
the average treatment for the actual nonuniform deformation (that
neglects the redundant work and underestimates the actual plastic
work). This first-order approximation provides a means to estimate the total energy received by the workpiece. It is well know
that during plastic deformation more than 90% of the received
external work is dissipated in the form of heat, so that this measurement provide a very conservative estimation of the heating
energy in the MP welding using this simple and straightforward
method of geometrical measurement, which can be useful for
Table 1 Peeling test results and bonded percentage lengths
for different welding conditions. The percentage is the bonded
length over the total interface length estimated from micrographs of sectioned surfaces.
Cu surface condition
A

Fig. 3 The measured current wave for welding at 5.2 kV (a),


and the resulting specimen cross section (b) showing diameter
reduction of Al and Cu, and the peeled fracture surface (c)
showing very rough aluminum inner surface, indicating strong
bonding and ductile fracture within Al

051002-4 / Vol. 136, OCTOBER 2014

B
C

Scratches in
hoop direction
Scratches in
axial direction
Oil-contaminated
over A

4.6 kV

5.2 kV

6.0 kV

Welded (55%) Welded (63%) Welded (73%)


N/A

Welded (59%)

N/A

N/A

N/A

Not welded
(10%)

Transactions of the ASME

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Comparing samples with and without oil contamination under


the same applied voltage (6.0 kV), the contaminated surface did
not establish welded interface in majority contact area, under similar condition or actually has slightly higher effective strain, likely
due to a reduced interface friction from oil lubrication, so that the
Al tube can move in axial direction more freely.
This indicates the oil contaminated surface cannot be easily
broken by MP pressure, even with significant plastic deformation.
This is different from the oxidation layer that can be broken by
MP force as reported by previous study [32]. Thus, for MPW a
clean surface is still required to establish a strong bonding, but a
quantitative description of the cleanness requirement needs to be
further studied.
The Effect of Surface Topology and Scratch Orientation
Scratches in Hoop Direction. For the sample with lathe turning
scratches in hoop direction and welded at 6.0 kV the optical
micrographs are shown in Fig. 7, with the weld zone shown in
three consecutive segments, labeled as L1, L2, and L3, respectively. It shows that wavy interfaces were formed, and both the
amplitude and wavelength of the waves increase toward the freeend of Al tube (to right). The measured average wavelength was
about 190 lm, which is close to lathe machining scratch pitch distance shown on the outer surface machine scratches; thus, it is
necessary to further check if this correlation is just a coincidence,
and if so the sample with initial scratches in axial direction should
have the similar wavelength.

Fig. 4 Selected images of sectioned samples under three


applied voltages, and the binary images of Cu and Al used for
strain calculation

further interface temperature analysis (not provided here). Other


heat sources not considered include Joule heat from induction current and interfacial frictional heat at the interface. A direct multiphysics energy calculation is very complex involving energy
conversion and transformation from the capacitor to the primary
coil, the secondary coil, the field concentrator of unknown efficiency, and finally to the induced field at the workpiece that cannot be conveniently calculated and directly measured.
The Effect of Surface Contamination. For the oilcontaminated sample (surface condition C) very weak bonding
strength was obtained even under the highest applied voltage
(6.0 kV). For majority areas the interface bonding was not established. The observation of the separated Al and Cu surfaces shows
that

On the Al outer surface, Figs. 6(a) and 6(b), the lath turning
scratches are clearly seen (since no solid contact on it), but
the spacing of the scratches is increasing toward Al tube end
(right side), consistent with the axial strain distribution, see
Fig. 5, indicating the contacting sequence was from Cu zoneI (with a large OD and small clearance), then passed the transition zone-II and developed increased shear displacements
in zone-III.
On the Al inner surface, Figs. 6(c) and 6(d), the majority of
the welding zone turned to black, apparently the originally
thick and transparent viscous oil turned to dry powders (carbon black) after MPW, suggesting that the temperature was
high enough to decompose the oil. In addition, the areas with
most black powders were within the middle 1/3 of the welding zone (zone-III), suggesting that the temperature and pressure of the middle zone was the highest over the entire
processing zone. The lathe turning scratches were clearly
seen on both Al and Cu mating surfaces, but with many damages and deformations.

Journal of Manufacturing Science and Engineering

Scratches in Axial Direction. The longitudinally scratched


specimen (surface condition B), after welded and sectioned along
the axial direction, is shown in Fig. 8. In this case only the Cu
inner surface were sanded in longitudinal direction, while the Al
inner surface was still with lathe scratches in tangential direction,
but Cu has much higher strength than Al. The results show that

With the pre-existing surface markers of Cu and Al in the


perpendicular cross-over arrangement, the produced interface
waves were relatively random, and no obvious periodic feature was recognized.
For the fractured interface region, the shape profiles of the
fractured two edges in Cu and Al sides do not match with
each other, though with some similarity. This suggests that
the plastic deformation occurred after separation, so that the
edge profile and location alignment have been changed and
shifted.

The original handmade sanding scratches in axial direction on


Cu surface and in tangential direction on Al did not show a clear
periodic feature. From the view of interface bonding, if the two
mating surfaces have the same surface scratch orientations a larger
interface area can be generated that is in favor of all three possible
bonding mechanisms, i.e., the mechanical mixing/interlocking,
physical bonding, and chemical reaction. In contrast, for the
scratches in perpendicular directions between the two welding
surfaces, the effective surface contact area and associated welding
effect was reduced.
The wavy surfaces have been widely reported in other MPW
studies. Based on Ben-Artzy et al. [28], the wavy interface formation in MPW was explained by KelvinHelmholtz instability
mechanism: under the MP force the reflected shock waves interact
with the welding collision point at the interface, where interferences are the source for the wave initiation. Based on this mechanism, the initial surface feature (here the lathe turning scratches
on both mating surfaces) served as the initiator of formed waves
and affected the welded wavy wavelength within certain range,
similar to that in unstable vibration under a periodic external
force; but in the case of the two mating surfaces scratched in perpendicular directions, and especially when the stronger scratch
sets were not coincide with the MP shear direction the more random wavelengths were produced.
OCTOBER 2014, Vol. 136 / 051002-5

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 5 The distributions of the true strain components along the axial distance from Cu head
end (at x 5 0) to Al tube free end (at x 5 20 mm), for Al (a) and Cu (b) at 5.2 kV. Also shown are
the effective strains for Al and Cu for this specimen (c), and for the specimens welded at different voltages and surface preparation conditions (d). The plastic works (e) and the bonded
length fraction (f) are plotted for different welding voltages.

Interface Reaction and Evidence of Melting. The new phase


formation requires a rapid mass transportation by diffusion over
certain time period, but the first MP peak is only in 105 s, so the
peak temperature and interface deformation become critically
important to provide a high diffusion rate for the interface reaction, and the combined thermomechanical condition determines
the type and amount of the reaction product and its morphology/
microstructure.
Mechanical mixing is one way to reduce diffusion path for new
phase formation, as seen in Fig. 9, where the wavy interface contained a 1040 lm thickness of the new phase with distinct color/
contrast from the parent Al and Cu, and some Cu base metals
have been completely wrapped within the new phase pocket.
Mechanical mixing is a necessary condition in forming such a discontinuous new phase across the interface (because atomic diffusion path must produce a continuous profile). In addition, in Fig.
9(b) some microcracks can be seen within the new phase region,
which are stopped at the boundaries of the more ductile base metals (this is different from the polishing scratches as the latter are
more straight and continuous crossing different phases), indicating
the new phase is more brittle than the base metals.
051002-6 / Vol. 136, OCTOBER 2014

Interface Melting. The direct evidence of interface melting was


obtained for a sample welded at 6.0 kV, see Fig. 10, where a round
pore was found inside the new phase region, and by refocusing to
the bottom of the pore it shows the smooth inner surface of the
pore with cracks. This pore cannot be formed from pull-out or
intrude-in of an inclusion during welding or sample sectioning/
polishing (otherwise a sharp edge and surface would be produced). In addition, from its round shape, smooth edge, and glassy
pore surface, it cannot be formed by high-pressure impact in solid
state (otherwise the pore will be crashed to a collapsed shape
rather than a round shape). Thus, it is more reasonable to believe
it was formed from liquid phase. The process involved a thermally
assisted diffusion process at a high temperature for minimizing
the surface energy. Two more evidences of melted sites were
found on the same sample from SEM observation; see Fig. 11, in
which some pores from trapped gas and a glassy pore surface
appearance can be clearly seen. This morphology can only be produced from melting. In addition, several cracks exist on the pore
surface, indicating a high thermal stress during rapid solidification. All the three pores shown have a size in the range of
2030 lm in diameter. Within the current welding operation
Transactions of the ASME

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 8 Interfaces of the sample with the original copper surface


sanded in axial direction. No correlation of the failed interface
edge and the original lathe-turned pitch can be identified.

Fig. 6 The sample with oil contamination (not welded): (a) and
(b) Al outer surface; (c) and (d) Al inner surface; and (e) and (f)
Cu outer surface. The right side shows the local enlargement of
the boxed area.

The equilibrium AlCu phase diagram (shown in the Appendix)


may be a helpful reference. When heated to above 800 K and
cooled down to room temperature, intermetallic compounds Alm:Cun can form with various Stoichiometric atomic ratios, including
AlCu, Al9Cu11, and Al2Cu3, as labeled at the bottom of the phase
diagram. Under current nonequilibrium condition involving a high
pressure and very short time the actual phase formation may not
exactly follow this diagram but with some offset of phase region.
From the SEMEDS results, the measured Al:Cu ratio of the new
phase in the interface region was within the range of 1:1 to 2:3
range. From the above observation and analysis we can conclude
that there existed a chemical reaction and AlCu intermetallic
formation.
Microscratching test was performed for a sample welded at
5.2 kV. We did not use microharness test with the consideration
that it often has data scattering and more repeats are required for
data averaging, so it is not suitable for small area measurement
containing a high hardness gradient. In contrast, a microscratching
test can be much simpler and informative for the current purpose
to compare the relative properties of the new phase and base metals at the small area across the interface. Vickers micro-indenter
was used with one of the diagonal axis aligned along the scratching direction running across the interface, under a 120 gf load and
a constant speed of 2 lm/s on the polished surface, see Fig. 14.
The produced scratch was much narrower (4 lm) in width within
the new phase region than that in Cu base metal (1214 lm) and
in Al base metal (14 lm), as marked in the figure, indicating that
the intermetallic phase is much harder/stronger than the two base
metals. Also shown are the multiple cracks in the new phase
region, again indicating this new compound is brittle.

Discussions
Fig. 7 A sectioned sample welded at 6.0 kV (top), and the
microstructures over the three consecutive segments L1, L2,
and L3 from left to right

window only three melting sites were found, all under the highest
welding voltage (6.0 kV).
Characterization of New Phase Formed on the Welding
Interface. SEM micrographs of two welded interface regions are
shown in Fig. 12, where a transition zone between two base metals can be clearly seen. SEMEDM analysis was performed across
the interface for the two samples under a spot operation mode in
order to collect more signals over longer time period for better
accuracy, and the results are shown in Fig. 13, given the new
phase a chemical composition in the range of 6080 at. % Cu and
balanced Al.
Journal of Manufacturing Science and Engineering

On the Interface Reaction Process. For the current material


system and MPW conditions, the observed phenomena may lead
to the following understanding of welding process
Upon application of MP force, large amount of energy, at
a very high energy density and energy input rate, was
added into the aluminum tube and resulted in Al wall
impact with the outer surface of Cu tube. The kinetic
energy converted to surface deformation energy and adiabatic heating at a high rate, with no sufficient time for heat
transfer, causing rapid temperature rising in the welding
zonewithin a thin layer of aluminum and copper contacting surfaces.
Mechanical mixing of two metals at the interface occurred
as the result of severe plastic deformation. Surface topology
has important effect on the mechanical mixing and wavy
interface formation.
OCTOBER 2014, Vol. 136 / 051002-7

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 9 New compound phase formed at the welding interface region that mixes with base
materials (a) and (b), and contains microcracks (b), observed under optical microscope;
incomplete intermetallic phase formation at a wavy pocket is seen in (c), with the Al wave front
converted to intermetallic but the tail still remained as base Al

front of the welding zone, which may cause microstructural


damage.
Continued heat transfer and cooling to RT may result in
thermal stress development, and the final welded interface
often contains certain separation zone, with some local
cracks within the brittle intermetallic zone.

Fig. 10 The optical micrograph of a sample welded at 6.0 kV,


and a round pore with a smooth surface was observed (a). By
refocusing at the pore bottom (b) some surface cracks were
revealed.

Interface temperature increases significantly, and depending


on the magnitude of applied energy interface melting may
occur.
New intermetallic compounds formed, which involved mass
transportation and interdiffusion of Cu and Al atoms in solid
phase or liquid phase. This is a thermally activated process
requiring certain time to start and finish. The thickness of
the intermetallic phase was in 05 lm, depending on the
applied energy and the surface conditions.
Due to the sequential process along the welding length, the
earlier formed brittle intermetallic phase may subject to further shear deformation from later impact deformation in
051002-8 / Vol. 136, OCTOBER 2014

On Welding Mechanisms. The experimental observations suggest that multiple joining mechanisms co-exist and operate simultaneously. These mechanisms include mechanical interlocking (at
macron and micron scales), physical bonding (atomic adhesion from
fresh new surface contacts), and chemical bonding/reaction (new
compound formation). The contribution of each mechanism depends
on the material pairs to be welded, the applied energy level and the
surface topological condition. Chemical bonding from intermetallic
formation has two opposite effects on bonding strength: on the one
side, it greatly increases the bonding strength; on the other hand, it
tends to form microcracks under the condition of a high temperature
gradient, a high thermal stress, and a large plastic deformation, especially when the intermetallic layer is too thick.
The observations indicate that (1) the formation of intermetallic
phase at the interface can occur by solid-state reaction without
melting, which plays important role in bonding strength; (2) interface local melting is possible, but is not widely observed and is
not a necessary condition for intermetallic formation. Thus, the
dominating MPW mechanism is still a solid-state process.
The surface roughness/scratch orientation not only affects mechanical interlocking and physical bonding but also it affects the
local contact stress as well, which changes the chemical reaction
Transactions of the ASME

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 11 The sample welded under 6.0 kW power (a), with three local melting zones shown in
(b) and (c) as evident by the smooth glassy pore surfaces with cracks; (d) the local enlargement of one pore shown in (b).

Fig. 12 SEM micrographs of two interface regions and their local enlargements. The rough
Cu surface was from chemical etching.

driving force and kinetics. Thus, further study can be helpful on


optimization of workpiece surface topology as one of the process
design element. As for the surface contamination, it plays as a barrier for all three joining mechanisms, and give a sensitive but negative impact to the weld establishment, so that a clean surface or
Journal of Manufacturing Science and Engineering

coated surface with favorable chemical composition can be one


way to enhance the weldability; on the other hand, this bonding
prohibiting layer can provide a useful means for special welding
design of selective patterns containing combined welding and
nonwelding features.
OCTOBER 2014, Vol. 136 / 051002-9

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 15 The AlCu binary phase diagram, computed with commercial software and with available thermodynamics database

Fig. 13 SEMEDM chemical analysis on two samples, along


the lines crossing the interface (top), and the copper atomic
percentage distribution along the scanned line spots (bottom)

(5) The strain and deformation energy by MP force were measured from geometry of welded specimens as a means of
low bond estimation of total heat generation, but local heat
at the interface needs to be further analyzed.
(6) Welding mechanisms responsible for MPW have been further discussed based on the experimental observations. The
direction of process improvement is provided in terms of
power application and interface preparation.

Acknowledgment
The authors like to express their sincere thanks to Professor
Glenn Daehn, who offered a lightening discussion and lab tour at
the early stage of this study.
Funding support of this study was from Hirotec America and
Wayne State University internal fund.
Fig. 14 Scratch test with the use of Vickers indenter and operated at 120 gf and 2 lm/s, with one of the diagonal axis aligned
along the moving direction. The circled area contains a few
microcracks.

Conclusion
(1) AlCu tubes have been successfully welded by applying
electromagnetic pulse force over Al tube outer surfaces,
under different electromagnetic power input levels and surface conditions. Welding was not established on oil contaminated surfaces.
(2) The processing parameters that affect welding strength
include not only the chemical property and compatibility of
two joining materials, the power used in the electromagnetic welding, but also the surface topological conditions
(the surface roughness, scratch orientations, and contamination). A correlation between the wavelength of lathe turning
scratches and that of welded interface waves were
observed.
(3) Intermetallic compounds formed at the welded interface,
and the microstructure, chemical composition, and mechanical properties were examined, identified, and tested.
(4) Evidences on liquid phase formation have been obtained
under a high power welding condition, which produced
smooth and glassy pore surface on the trapped air bubbles,
and with surface cracks.
051002-10 / Vol. 136, OCTOBER 2014

Nomenclature
d(x), t(x), A(x) deformed tube outer diameter, thickness, and
cross-sectional area at x
d0, t0, A0 initial tube outer diameter, thickness, and crosssectional area
w(x) plastic work per volume at x
W total plastic work over welding length
ee(x) true effective (von Mises) strain at x
et(x) true strain component in thickness direction at x
ex(x) true strain component in axial direction at x
eh(x) true strain component in hoop direction at x

Appendix: The Equilibrium AlCu Phase Diagram


The equilibrium AlCu phase diagram is shown in Fig. 15, generated using commercial software COMPUTHERM [33] and published binary phase diagram database.

References
[1] Kochan, A., 2000, Magnetic Pulse Welding Shows Potential for Automotive
Applications, Assem. Autom., 20(2), pp. 129131.
[2] Belyy, I. V., Fertik, S. M., and Khimenko, L. T., 1977, Electromagnetic Metal
Forming Handbook (A Translation of the Russian Book: Sprvochnik Po
Magnitno-Impul Snoy Obrabotke Metallov), Vischa Shakola, Kharikov State
University, Kharkov, Ukraine (Translated by M. M. Altynova and G. S. Daehn,
Columbus, OH).
[3] Strizhakov, E. L., 1981, Magnetic Pulse Welding in Vacuum With Preheating, Weld. Prod., 28(2), pp. 1718.

Transactions of the ASME

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

[4] Yablochnikov, B. A., 1983, Magnetic Pulse Welding With Arc Heating,
Autom. Weld. USSR, 36(2), pp. 5457.
[5] Yablochnikov, B. A., 1983, Apparatus for Magnetic Pulse Welding Large
Diameter Thin-Walled Pipes, Autom. Weld. USSR, 36(4), pp. 4144.
[6] Yablochnikov, B. A., 1983, An Inductor for Magnetic Pulse Welding Large
Diameter Thin-Walled Pipes, Autom. Weld. USSR, 36(1), pp. 5053.
[7] Tamhane, A. A., Altynova, M. M., and Daehn, G. S., 1996, Effect of Sample Size
on Ductility in Electromagnetic Ring Expansion, Scr. Mater., 34(8), pp. 13451350.
[8] Daehn, G. S., Vohnout, V. J., and DuBois, L., 1999, Improved Formability
With Electromagnetic Forming: Fundamentals and a Practical Example, Sheet
Metal Forming Technology, M. Y. Demeri, ed., The Minerals, Metals & Materials
Society, Warrendale, PA, pp. 105116.
[9] Seth, M., Vohnout, V. J., and Daehn, G. S., 2005, Formability of Steel Sheet
in High Velocity Impact, J. Mater. Process. Technol., 168(3), pp. 390400.
[10] Thomas, J. D., Seth, M., Daehn, G. S., Bradley, J. R., and Triantafyllldis, N.,
2007, Forming Limits for Electromagnetically Expanded Aluminum Alloy
Tubes: Theory and Experiment, Acta Mater., 55(8), pp. 28632873.
[11] Fenton, G. K., and Daehn, G. S., 1998, Modeling of Electromagnetically
Formed Sheet Metal, J. Mater. Process. Technol., 75(13), pp. 616.
[12] Golowin, S., Kamal, M., Shang, J. H., Portier, J., Din, A., Daehn, G. S., Bradley,
J. R., Newman, K. E., and Hatkevich, S., 2007, Application of a Uniform Pressure Actuator for Electromagnetic Processing of Sheet Metal, J. Mater. Eng.
Perform., 16(4), pp. 455460.
[13] Zhang, Y., Babu, S. S., Zhang, P., Kenik, E. A., and Daehn, G. S., 2008,
Microstructure Characterisation of Magnetic Pulse Welded Aa6061-T6 by Electron Backscattered Diffraction, Sci. Technol. Weld. Joining, 13(5), pp. 467471.
[14] Zhang, Y. A., Babu, S. S., and Daehn, G. S., 2010, Interfacial UltrafineGrained Structures on Aluminum Alloy 6061 Joint and Copper Alloy 110 Joint
Fabricated by Magnetic Pulse Welding, J. Mater. Sci., 45(17), pp. 46454651.
[15] Shribman, V., Stern, A., Livshitz, Y., and Gafri, O., 2002, Magnetic Pulse
Welding Produces High-Strength Aluminum Welds, Weld. J., 81(4), pp.
3337. Available at: http://www.aws.org/wj/apr02/feature1.html
[16] Kore, S. D., Date, P. P., and Kulkarni, S. V., 2007, Effect of Process Parameters
on Electromagnetic Impact Welding of Aluminum Sheets, Int. J. Impact Eng.,
34(8), pp. 13271341.
[17] Marya, M., Priem, D., and Marya, S., 2003, Microstructures at Aluminum
Copper Magnetic Pulse Weld Interfaces, Materials Science Forum, T. Chandra,
J. M. Torralba, and T. Sakai, eds., Trans Tech Publications, D
urnten, Swizerland,
pp. 40014006.
[18] Marya, M., and Marya, S., 2004, Interfacial Microstructures and Temperatures
in AluminiumCopper Electromagnetic Pulse Welds, Sci. Technol. Weld.
Joining, 9(6), pp. 541547.

Journal of Manufacturing Science and Engineering

[19] Aizawa, T., Kashani, M., and Okagawa, K., 2007, Application of Magnetic
Pulse Welding for Aluminum Alloys and SPCC Steel Sheet Joints, Weld. J.,
86(5), pp. 119s124s. Available at: http://www.aws.org/www/wj/2007/05/
WJ_2007_05.pdf
[20] Ben-Artzy, A., Stern, A., Frage, N., and Shribman, V., 2008, Interface
Phenomena in AluminiumMagnesium Magnetic Pulse Welding, Sci. Technol. Weld. Joining, 13(4), pp. 402408.
[21] Marya, S., and Gerard, P., 2006, Welding and Joining of Titanium and Its
Alloys. Trends and Developments in France, Rare Metal Mater. Eng., 35, pp.
1520.
[22] El-Azab, A., Garnich, M., and Kapoor, A., 2003, Modeling of the Electromagnetic Forming of Sheet Metals: State-of-the-Art and Future Needs, J. Mater.
Process. Technol., 142(3), pp. 744754.
[23] Ivanov, V., Kotov, Y. A., Samatov, O. H., B
ohme, R., Karow, H. U., and Schumacher, G., 1995, Synthesis and Dynamic Compaction of Ceramic Nano Powders by Techniques Based on Electric Pulsed Power, Nanostructured
Materials, 6(14), pp. 287290.
[24] Ahn, J. H., Kim, Y. J., and Kim, B. K., 2006, NiZrTiSiSn/Cu Metallic
Glass Composites Prepared by Magnetic Compaction, Materials Letters,
60(2930), pp. 37473751.
[25] Stern, A., and Aizenshtein, M., 2002, Bonding Zone Formation in Magnetic
Pulse Welds, Sci. Technol. Weld. Joining, 7(5), pp. 339342.
[26] Callister, W. D., Jr., and Rethwisch, D. G., 2012, Fundamentals of Materials
Science and Engineering, 4th ed., Wiley, New York.
[27] Nassiri, A., Chini, G., and Kinsey, B., 2014, Spatial Stability Analysis of
Emergent Wavy Interfacial Patterns in Magnetic Pulsed Welding, CIRP Ann.
Manuf. Technol., 63(1), pp. 245248.
[28] Ben-Artzy, A., Stern, A., Frage, N., Shribman, V., and Sadot, O., 2010, Wave
Formation Mechanism in Magnetic Pulse Welding, Int. J. Impact Eng., 37(4),
pp. 397404.
[29] Shang, J., 2006, Electronmagnetically Assisted Sheet Metal Stamping, Ph.D.
dissertation, Ohio State University, Columbus, OH.
[30] C110, http://www.matweb.com/search/DataSheet.aspx?MatGUID07ed2c8eea
29427abd450157263ef28c
[31] AA6063-O, http://www.matweb.com/search/DataSheet.aspx?MatGUID6c41
b0bdea564d20b9fd287c03c97923
[32] Kore, S., Imbert, J., Zhou, Y., and Worswick, M., 2010, 19Electromagnetic
Pulse Welding of Magnesium to Aluminium Sheets, Welding and Joining of
Magnesium Alloys, L. Liu, ed., Woodhead Publishing, Cambridge, UK, pp.
367379.
[33] CompuTherm, http://www.computherm.com/index.php?routeproduct/product
&path71&product_id25

OCTOBER 2014, Vol. 136 / 051002-11

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Anda mungkin juga menyukai