You are on page 1of 7

Journal of Materials Processing Technology 210 (2010) 2231–2237

Contents lists available at ScienceDirect

Journal of Materials Processing Technology


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

Production of wire via friction extrusion of aluminum alloy machining chips


W. Tang ∗ , A.P. Reynolds 1
Department of Mechanical Engineering, University of South Carolina, 300 Main Street, Columbia, SC 29208, USA

a r t i c l e i n f o a b s t r a c t

Article history: This paper describes the production of fully consolidated wires from aluminum alloy AA2050 and alu-
Received 1 February 2010 minum alloy AA2195 machining chips via the friction extrusion process. In this work, the extrusion rate
Received in revised form 16 June 2010 was linearly related to the power input/die rotational speed. Hot crack and cold crack defects appeared
Accepted 16 August 2010
on wires produced using either too high or too low a power input/die rotational speed. The extruded
wire microstructure consisted of fully equiaxed, recrystallized grains. Average grain size in the wires
generally increases with the increase of die rotational speed/extrusion power. Micro-hardness was homo-
JEL classification:
geneous across the wire transverse cross-sections. Positive response to post-extrusion heat treatment
81.20.Hy
was observed with increasing extrusion power indicating increasing temperature and in process solu-
Keywords: tion heat treatment of the wire with sufficient extrusion power. The ductility of defect free wire was
Friction extrusion demonstrated by the absence of cracking in 5T bend tests.
Recycling © 2010 Elsevier B.V. All rights reserved.
Wire

1. Introduction cussed. The effect of cold press pressure prior to melting process in
aluminum chips recycling was studied by Mashhadi et al. (2009).
There are two primary methods of recycling aluminum chips Aluminum chips were granulated, cold compacted and extruded at
to produce a useful product: the so-called ‘conventional’ method the temperature range of 500–550 ◦ C by Gronostajski and Matuszak
and the direct conversion method (Tekkaya et al., 2009). The con- (1999), and additional sintering process was added before the final
ventional method requires melting of the material to be recycled, extrusion of aluminum chips recycling by Samuel (2003). 6061 alu-
casting of a billet, and then hot extrusion of the billet to form a minum blocks were directly hot extruded from compacted chips
consolidated product in wire or rod form. The direct conversion by Tekkaya et al. (2009). Al powder was added as a soft matrix in
method utilizes conventional hot extrusion which may or may the 7075 aluminum chips direct conversion process, and the influ-
not be preceded by a static pressing stage. Compared with con- ence of Al powder content and extrusion temperature in products
ventional recycling, the direct conversion of aluminum scrap into mechanical properties were discussed by Sherafat et al. (2009).
compact metal may result in 40% material, 26–31% energy and Friction stir processing (FSP) and friction stir welding (FSW) are
16–60% labor savings (Gronostajski and Matuszak, 1999). In the solid state processes for microstructure modification and joining,
past decade, research works of aluminum chips recycling were car- respectively. Both rely on plastic dissipation for heat genera-
ried out with both ‘conventional’ and direct conversion methods. tion and may be considered severe plastic deformation processes.
Various aluminum turning scraps were melted at 800 ◦ C to recover FSW was patented by The Welding Institute in 1991 (Thomas et
aluminum metal with the protective salt flux under nitrogen atmo- al., 1991). In both technologies, welding and processing, a rotat-
sphere by Xiao and Reuter (2002). In that paper, it also showed that ing, non-consumable tool, is used to produce a plasticized layer
the difficulty of recycling the selected aluminum scrap depends on which undergoes severe plastic deformation and heating. Gener-
scrap type, scrap size distribution, contaminant, and the ratio of ally speaking, application of FSP to metals results fine grains in
surface area to body volume. The recyclability of aluminum chips the processed zone due to substantial deformation at hot work-
was experimentally studied by Puga et al. (2009) by using differ- ing temperatures. In addition, FSP has been used to close porosity
ent melting techniques and the influence of chips preparation in in castings as shown by Fuller et al. (2007) in an application to
the aluminum alloy recovery rate and dross production was dis- Ni–Al bronze. In 1993 another friction based process, friction extru-
sion, was patented by The Welding Institute (Thomas et al., 1993).
This patent was allowed to lapse in the early part of this decade.
However, there are some attractive aspects of the friction extrusion
∗ Corresponding author. Tel.: +1 803 777 1279; fax: +1 803 777 0106.
process and it may yet prove to be an industrially useful technique
E-mail addresses: tang@cec.sc.edu (W. Tang), reynolds@cec.sc.edu
(A.P. Reynolds). for the achievement of several purposes, e.g. recycling of machin-
1 1
Tel.: +1 803 777 9548; fax: +1 803 777 0106. ing waste, consolidation of powder product and, potentially, as a

0924-0136/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.jmatprotec.2010.08.010
2232 W. Tang, A.P. Reynolds / Journal of Materials Processing Technology 210 (2010) 2231–2237

Fig. 2. Friction stir wire extrusion setup.

FSW system control computer with 10 Hz data collection rate. The


process torque and power were recorded by a torque transducer
mounted on the machine spindle (which drives the rotating die):
these data were recorded with a 60 Hz sampling rate. All extrusion
runs were performed using a constant extrusion force of 17.8 kN
Fig. 1. Friction extrusion apparatus: (a) rotating extrusion die and (b) extrusion
billet chamber.
(4000 lbs). The die rotation rates used for extrusion of the 2195 and
2050 wires are shown in Table 1. For each extrusion run the bil-
let chamber was charged with 10 g of clean, dry machining chips.
method of mechanical alloying. The friction extrusion process may During each extrusion run less than 1/4 of the charge was trans-
be considered a variant of the direct conversion method of recycling formed into wire (about 150 mm or 6 ). The wire length was limited
since a consolidated product is produced without melting and cast- by the headspace above the die. The remainder of the chips was
ing. In this paper, we describe the direct production of AA2050 and consolidated into a disk in the billet chamber.
AA2195 aluminum wires by friction extrusion of machining chips.
2.2. Materials
2. Experimental procedures
The materials extruded were AA2050 and AA2195 aluminum
2.1. Friction extrusion alloy chips. Both materials are Al–Cu–Li–Mg–Ag alloys with sim-
ilar compositions: production of wire from these materials is of
The key components of the friction extrusion device, shown in interest for use in free-form fabrication processes and for use
Fig. 1, are a 25.4 mm (1 ) inner diameter stationary ‘billet’ chamber as a welding consumable. The nominal composition of 2195 is
and a 25 mm (0.98 ) diameter rotating die. Both parts are fabricated Al–4.0Cu–1.0Li–0.5Mg–0.4Ag–0.12Zr. The nominal composition of
from H13 tool steel. The rotating die has a scrolled face and a central alloy 2050 is Al–3.55Cu–0.4Mg–1.0Li–0.45Ag–0.1Zr–0.35Mn. Rel-
through hole. The scroll is similar to patterns machined into friction atively uniform and clean chips were milled from 2050-T8 and
stir welding tool shoulders designed for operation at a 0◦ angle of 2195-T8 aluminum plates without any lubricant or cutting fluid.
attack (Thomas et al., 1991). The central hole has a diameter of Representative chips used in the wire extrusion process are shown
2.5 mm (0.098 ) which defines the resulting wire diameter. The die, in Fig. 3.
as pictured, is rotated in a counter-clockwise direction so that the A post-extrusion heat treatment (PEHT) of 15 h at 160 ◦ C was
scrolls move material toward the center hole (die exit). applied to the extruded wires. The PEHT is similar to the artificial
The die is actuated by a friction stir welding (FSW) machine aging step used to produce T8 tempers in the alloy plates. However,
operated in a Z-axis load control mode (MTS-Process Development
System). This implies that the entire extrusion process occurs under Table 1
a constant pressure and the extrusion rate and die advance rate may Friction extrusion parameters.
vary during the process. This is in contrast to a “normal” extru- Extrusion run # Material Die rotation Extrusion
sion process wherein the extrusion rate is held constant and the rate (rpm) force (kN)
pressure may vary to maintain the specified rate. The Z-axis of the
1 2050 100 17.8
machine is coincident with the extrusion axis hence the Z-force is 2 2050 150 17.8
equivalent to the extrusion force. The billet chamber is clamped 3 2050 200 17.8
onto a stainless steel back plate. The wire extrusion process is best 4 2050 250 17.8
described as a back extrusion wherein the rotating die is pressed 5 2050 300 17.8
6 2050 400 17.8
into the billet chamber under load control and the wire is extruded 7 2195 100 17.8
up through the central hole in the die. A schematic of the friction 8 2195 150 17.8
stir wire extrusion setup is shown in Fig. 2. 9 2195 200 17.8
During the friction extrusion process, the rotational speed, Z- 10 2195 250 17.8
11 2195 300 17.8
force (extrusion force), and die movement were recorded by the
W. Tang, A.P. Reynolds / Journal of Materials Processing Technology 210 (2010) 2231–2237 2233

Fig. 4. Macro-views of wires produced by friction extrusion. (a) Alloy/rpm from top
to bottom: 2050/400 rpm, 2195/200 rpm, 2195/100 rpm. (b) 2050 wire (250 rpm)
after 5T bend.
Fig. 3. Milled aluminum chips for direct conversion to wire via friction extrusion.
Each unit on the scale represents 1 mm.
Fig. 5 shows higher magnification views of representative
defective wires. The wires extruded at 300 and 400 rpm exhib-
in normal industrial practice the aging step is typically preceded ited closely spaced, circumferential, surface, cracks as shown in
by a solution heat treatment, water quenching, and stretching to Fig. 5(a). The cracks were not observed over the entire wire
several percent plastic deformation. surface. In the 2050 wires extruded at 300 and 400 rpm, the
cracks appeared after 125 mm and 75 mm of extrusion, respec-
2.3. Metallography and mechanical testing tively. The 2195 wire produced at 300 rpm exhibited cracking after
90 mm of wire was extruded. Fig. 5(b) shows the surface of a
Wires were initially inspected visually. Subsequently, trans- 2195 wire extruded at 100 rpm. In this wire, the surface defect
verse and longitudinal cross-sections of the extruded wires were appears to be a result of cold tearing and individual tears do not
ground, polished and etched with Keller’s Reagent (190 ml water, extend around the entire circumference; however, the tearing is
2 ml HF, 3 ml HCl, and 5 ml HNO3 ). Microstructures were examined present along the entire length of the wire. Also evident on the
by optical microscopy. The grain size of each wire was measured 100 rpm wire is the low pitch twist mentioned in the previous
using the mean linear intercept method (ASTM, 2004). paragraph.
Vickers micro-hardness tests were performed on the wire
transverse cross-section. On each wire a hardness traverse was 3.2. Extrusion rate and power
made along an entire diameter with an indentation spacing of
0.254 mm (10 or 11 indents), a load of 200 g, and a loading time Typical curves of die movement, Z-axis load, and die rotational
of 10 s (HV200). Hardness was measured in as extruded and post- speed during extrusion are shown in Fig. 6. The die rotation rate is
extrusion heat treated conditions. Bend tests were carried out on quite steady and in accord with the commanded value during the
selected, as extruded, wire (i.e. prior to PEHT). Wires were bent extrusion process. The Z-axis force was fairly stable although some
around a mandrel with a radius of 12.5 mm (approximately five spikes can be observed. Also, there is a substantial spike in load dur-
times the wire thickness). ing the highly transient, initial portion of the process (near 10 s).

3. Results

3.1. Wire quality vs. extrusion parameters

Defect free wires were produced from both 2050 and 2195
chips at intermediate die rotation rates (150, 200, and 250 rpm).
At all other die rotation rates, surface flaws were observed visu-
ally. Fig. 4(a) shows representative good quality and poor quality
wires. The top wire in the figure is a 2050 wire produced with a die
rotation rate of 400 rpm. Cracks may be observed over about 50%
of the wire surface. The middle wire is 2195 produced at 200 rpm:
the entire surface is quite smooth but a slight twist can be seen
especially on the left. The wire at the bottom of the figure is 2195
produced at 100 rpm. The low rpm wire exhibits cold tearing along
the entire length of the wire and a more pronounced twist. In gen-
eral, the twist is not obvious on any of the wires made at die rotation
rates greater than or equal to 250 rpm but becomes more apparent
with decreasing rpm below 250. All the defect free wires exhibit
Fig. 5. Extruded wire surface defects caused by excessively high or low die rotation
good ductility in the as-extruded condition as evidenced by their rates (wire diameter of 2.5 mm sets scale). (a) 2050 wire extruded with 400 rpm die
performance in bend tests (Fig. 4(b) for example). rotation rate and (b) 2195 wire extruded with 100 rpm die rotation rate.
2234 W. Tang, A.P. Reynolds / Journal of Materials Processing Technology 210 (2010) 2231–2237

Fig. 6. Processing parameters feedback in friction wire extrusion.

Fig. 8. Extrusion rate and power vs. die rotation rate.

The graph shown is divided into two regimes at approximately 20 s


into the process. Before 20 s, it is assumed that the process of chip study, the calculated rate of increase of extrusion rate with power
consolidation is taking place. After 20 s, wire extrusion is occurring. is approximately 0.004 (mm s−1 )/W.
This division is rationalized as follows. The mass of the initial billet
chamber charge and the fully consolidated alloy density are known. 3.3. Wire microstructure
Also, the initial position of the die relative to the bottom of the bil-
let chamber is known. At approximately 20 s, given the known die Transverse cross-sections of 2050 and 2195 wires are shown
diameter and position, it is determined that the chamber volume is in Fig. 9. Fig. 9(a) is a cross-section of the 2195 wire extruded at
very close to that of 10 g of the alloy in the fully consolidated state. 100 rpm. The wire is consolidated everywhere except at the outer
The volume swept out by the die over the remaining portion of the edge where evidence of cold tearing can be observed (see the top
process (die area × die movement) is equal to the volume of wire right of the figure). Fig. 9(b) shows a completely consolidated 2195
produced (wire cross-sectional area × wire length). Therefore, the wire produced with a die rotation rate of 250 rpm. Substantial
extrusion rate will be equal to the wire length divided by the time cracking can be seen in the cross-section of the 2050 wire extruded
required for the second part of the process. In the case of Fig. 6, the at 400 rpm (Fig. 9(c)). The cross-sections are consistent with the
extrusion time is approximately 60 s. surface appearance of the wires produced at the various die rotation
Fig. 7 shows representative wire length vs. time for the 2195 rates.
extrusions produced using die rotation rates from 100 to 300 rpm. Higher magnification micrographs of transverse and longitudi-
The extrusion rates in mm/s are equal to the slopes of the curves. nal cross-sections of the 2050 wire produced with a die rotation
The curves in Fig. 7 do not have strictly constant slopes (the slopes rate of 250 rpm are shown in Fig. 10(a) and (b), respectively. Simi-
generally increase slightly with increased extrusion length) so, the lar sections of all the other wires were also examined. The grain
extrusion rates plotted vs. die rotation rates in Fig. 8 are average structures appear equiaxed on both the longitudinal and trans-
values. Also plotted in Fig. 8 is the extrusion power vs. die rota- verse sections exhibiting no elongation in the extrusion direction
tion rate. Both the extrusion rate and power increase linearly with (the wire longitudinal axis). The mean linear intercept grain sizes
increasing die rotation rate. Also, the values for 2195 and 2050 are (obtained from the transverse sections) are plotted vs. die rotation
nearly identical and fall on the same trend lines: not surprising rates for all of the 2050 and 2195 wires in Fig. 11. Generally, the
considering the similarity in their compositions. Since both power grain size increases with increase of die rotation rate. However, for
and extrusion rate have direct, linear, relationships to the rota- the 2050 wires, a plateau near a grain size of 11 ␮m is observed for
tion rate, they also have a direct, linear, relationship to each other. rotation rates greater than or equal to 300 rpm.
For the die geometry, extrusion pressure, and alloys used in this
3.4. Wire hardness and response to post-extrusion heat treatment

Fig. 12 shows the average hardness for all wires before and after
post-extrusion heat treatment. The results are based on 10 or 11
evenly spaced Vickers micro-hardness indentations made across
the transverse cross-section of each wire both before and after post-
extrusion heat treatment (e.g. Fig. 9(a)). In the figure, the error bars
are representative of one standard deviation. With one exception
the standard deviations are very small: less than 3% of the mean.
In the case with the greatest variability (2050 wire produced at
150 rpm and post-extrusion heat treated) the coefficient of varia-
tion is still less than 10%. Though average hardness values changed
with die rotational speed and heat treatment, hardness values in
a single wire cross-section remained relatively constant from the
center to the edge. Several observations can be made based on the
data in the figure. The as-extruded hardness values for the 2050
wires and the 2195 wires are very similar for a given die rotation
rate: there is more separation for the post-extrusion heat treated
Fig. 7. 2195 wire length vs. extrusion time for different wire rotation rates. results. The as-extruded hardness values of the wires produced at
W. Tang, A.P. Reynolds / Journal of Materials Processing Technology 210 (2010) 2231–2237 2235

Fig. 9. Transverse cross-sections of wire extruded with different die rotation rates. (a) 2195 wire extruded with 100 rpm rotational speed, (b) 2195 wire extruded with
250 rpm rotational speed, (c) 2050 wire extruded with 400 rpm rotational speed.

100 rpm are the lowest: for higher die rotation rates there is very lit- The observed die rotation rate limits are almost certainly functions
tle difference in hardness. At all rotation rates greater than 100 rpm, of the parameters which have not been varied during this study,
post-extrusion heat treatment results in an increase of the average e.g. die geometry, extrusion force, and alloy composition. How-
hardness while at 100 rpm, post-extrusion heat treatment causes a ever, it is likely that the underlying, critical factor is the extrusion
slight decrease in hardness. The post-extrusion heat treated hard- temperature. As for FSW, it is very difficult to access the actual
ness values increase rapidly with increasing die rotation rate up temperature in the process zone. For friction extrusion, the heat
to 300 rpm for the 2050 and 250 rpm for the 2195. Above these source is contained in the deforming volume and it is not possi-
rotation rates, the hardness values reach a plateau. ble to introduce a thermocouple into this region. However, relative
The hardness and grain size data indicate a high level of homo- values of extrusion temperature can be deduced from the extruded
geneity across the wire cross-sections. This is likely due to the wire grain size when the material is the same. As in FSW, the higher
fact that after sufficient deformation to ensure consolidation is the processing temperature the bigger the grain size. (A measure
achieved, the microstructure and properties are then determined of the average welding temperature in friction stir welding is rel-
primarily by the temperature at which the final extrusion takes atively easy to obtain from thermocouples embedded in the tool
place. As in FSW, this final deformation temperature dictates the pin.) Therefore, by comparing grain size and grain size change with
grain size and (for precipitation hardening alloys) the level of solute rotational rate trends of the extruded wires we can obtain a basic
super saturation. idea of how the temperature varies with friction wire extrusion
processing parameters.
4. Discussion The relationship between wire grain size and die rotation rate
(Fig. 11) is very similar to that typically observed between friction
It is apparent that the process window for production of good stir weld (FSW) nugget grain size and friction stir weld tool rotation
quality wires by the friction extrusion process (as practiced in this rate. In Fig. 11, both wire grain sizes increased with the increase
study) is bounded by upper and lower values of die rotation rate. of die rotational speeds, and the increase of 2050 wire grain size
2236 W. Tang, A.P. Reynolds / Journal of Materials Processing Technology 210 (2010) 2231–2237

Fig. 12. Average Vickers hardness of 2050 and 2195 wires as functions of die rotation
rate. Error bars indicate one standard deviation.

increased from 800 to 1800 rpm. The computational fluid dynam-


ics (CFD) modeling results by Long and Reynolds (2006) showed
flow field maximum temperature increased with the increase of
rotational velocity. The particular mechanism of grain formation is
not settled as various researchers propose that continuous dynamic
recrystallization (Jata and Semiatin, 2000), geometric dynamic
recrystallization (Fonda et al., 2004), or static recrystallization (Sato
et al., 2002) followed by grain growth are responsible for the final
microstructure. Of course, all other things being equal, each of these
mechanisms predict increasing grain size with increasing temper-
ature. It is also believed that during FSW, with increasing tool
rotation rate, a limiting temperature is attained. That is, the tem-
perature of the deforming material does not increase further with
increasing rotation rate, which has been observed in both exper-
imental temperature measurement (Gerlich et al., 2008) and CFD
Fig. 10. Transverse (a) and longitudinal (b) cross-sections of 2050 wire. (a) 2050
wire transverse cross-section microstructure (250 rpm) and (b) 2050 wire longitu-
computer modeling (Long and Reynolds, 2006). This limiting tem-
dinal section microstructure (250 rpm). perature may be related to rapidly declining flow stress (limiting
plastic dissipation) or changes in the friction conditions (limiting
reached to a plateau when the die rotation reached to 250 rpm. strain) near the bulk or incipient melting temperatures. Achieve-
Studies in FSW of tool rotational speed affect nugget grain size of ment of such a temperature limit may account for the grain size
7010 aluminum (Hassan et al., 2003) and 2219 aluminum (Long et plateau observed for the alloy 2050 wires which were produced at
al., 2007) showed the same trends as in Fig. 11. In FSW, it is generally 250, 300, and 400 rpm. Such a plateau may exist in 2195 for rota-
accepted that, all other things being equal, higher tool rotation rate tion rates ≥300 rpm; however, rates greater than 300 rpm were not
leads to higher temperature which is, in turn, the cause of increased attempted for the 2195.
grain size. Sato et al. (2002) measured nugget temperature with Information regarding the temperature attained during the fric-
thermocouples in 6063 aluminum FSW, and the peak tempera- tion extrusion process may be gleaned from the hardness data. It
ture increased from 400 ◦ C to 500 ◦ C when tool rotational speed can be seen in Fig. 12 that the hardness of post-extrusion aged 2195
is approximately constant for die rotation rates ≥250 rpm. A sim-
ilar trend is observed for alloy 2050 at rotation rates ≥300 rpm.
These trends are likely indicative of achievement of full solution
heat treatment during the extrusion process for die rotation rates
in the plateau region. Again, this is a phenomenon which is mirrored
in the friction stir welding process wherein FSW weld nuggets in
precipitation hardening aluminum alloys may achieve full T6 hard-
ness after post weld aging if the nugget temperature is near the
solution heat treatment temperature (Reynolds et al., 2005).

5. Conclusions and summary

(1) Defect free 2050 and 2195 aluminum wire can be directly pro-
duced from machining chips by the friction extrusion process.
Limits on the process appear to be related to the extrusion tem-
perature which, if too low, results in cold tearing and, if too high,
causes what appears to be a form of hot cracking.
(2) The wire extrusion rate is directly related to the extrusion
Fig. 11. Wire grain size as a function of die rotation rate. power for the range of parameters examined in this paper.
W. Tang, A.P. Reynolds / Journal of Materials Processing Technology 210 (2010) 2231–2237 2237

(3) The microstructure of the friction extruded wire is character- Hassan, K.A.A., Prangnell, P.B., Norman, A.F., Price, D.A., Williams, S.W., 2003. Effect
ized by equiaxed, recrystallized, grains: the average grain size of welding parameters on nugget zone microstructure and properties in high
strength aluminum alloy friction stir welds. Science and Technology of Welding
increases with increasing power up to some limiting value. and Joining 8, 257–268.
(4) The hardness of the wires is relatively homogeneous: a suffi- Jata, K.V., Semiatin, S.L., 2000. Continuous dynamic recrystallization during friction
ciently high extrusion temperature is required to promote good stir welding of high strength aluminum alloys. Scripta Materialia 43, 743–749.
Long, T., Reynolds, A.P., 2006. Parametric studies of friction stir welding by commer-
response to post-extrusion heat treatment by artificial aging. cial fluid dynamics simulation. Science and Technology of Welding and Joining
(5) Good bend ductility was exhibited by the as-extruded wires 11, 200–208.
which were produced using an intermediate die rotation Long, T., Tang, W., Reynolds, A.P., 2007. Process response parameter relationships
in aluminum alloy friction stir welds. Science and Technology of Welding and
rate. Joining 12, 311–317.
Mashhadi, H., Amini, Moloodi, A., Golestanipour, M., Karimi, E.Z.V., 2009. Recycling
While good quality wire has been produced using the friction of aluminium alloy turning scrap via cold pressing and melting with salt flux.
Journal of Materials Processing Technology, 3138–3142.
extrusion process, in order to improve the utility of the process, it
Puga, H., Barbosa, J., Soares, D., Silva, F., Ribeiro, S., 2009. Recycling of aluminium
will be necessary to increase the attainable extrusion rates with- swarf by direct incorporation in aluminium melts. Journal of Materials Process-
out entering the die rotation rate regime which causes overheating. ing Technology 209, 5195–5203.
This will necessitate research into effects of die geometry and extru- Reynolds, A.P., Tang, W., Khandkar, Z., Khan, J.A., Lindner, K., 2005. Relationships
among weld parameters, hardness distributions, and temperature histories in
sion force on extrusion rate and quality. alloy 7050 friction stir welds. Science and Technology of Welding and Joining
10, 190–199.
Samuel, M., 2003. A new technique for recycling aluminium scrap. Journal of Mate-
References rials Processing Technology 135, 117–124.
Sato, Y.S., Urata, M., Kokawa, H., 2002. Parameters controlling microstructure and
ASTM Standard #E112, 2004. Standard test methods for determining average grain hardness during friction-stir welding of precipitation-hardenable aluminum
size. alloy 6063. Metallurgical and Materials Transactions A 33, 625–635.
Fonda, R.W., Bingert, J.F., Colligan, K.J., 2004. Development of grain structure during Sherafat, Z., Paydar, M.H., Ebrahimi, R., 2009. Fabrication of Al7075/Al, two phase
friction stir welding. Scripta Materialia 51, 243–248. material, by recycling Al7075 alloy chips using powder metallurgy route. Journal
Fuller, M.D., Swaminathan, S., Zhilyaev, A.P., McNelley, T.R., 2007. Microstructural of Alloys and Compounds 487, 395–399.
transformations and mechanical properties of cast NiAl bronze: effects of fusion Tekkaya, A.E, Schikorra, M., Becker, D., Biermann, D., Hammer, N., Pantke, K., 2009.
welding and friction stir processing. Materials Science and Engineering A 463, Hot profile extrusion of AA-6060 aluminum chips. Journal of Materials Process-
128–137. ing Technology 209, 3343–3350.
Gerlich, A., Yamamoto, M., North, T.H., 2008. Local melting and tool slippage dur- Thomas, W.M., Nicholas, E.D., Needham, J.C., Church, M.G. Templesmith, P., Dawes,
ing friction stir spot welding of Al-alloys. Journal of Materials Science 43, C.J., 1991. International Patent Application No. PCT/GB92/02203 and GB Patent
2–11. Application No. 9125978.9.
Gronostajski, J., Matuszak, A., 1999. The recycling of metals by plastic deformation: Thomas, W.M., Nicholas, E.D., Jones, S.B., 1993. US Patent # 5,262,123.
an example of recycling of aluminum and its alloy chips. Journal of Materials Xiao, Y., Reuter, M.A., 2002. Recycling of distributed aluminum turning scrap. Min-
Processing Technology 92–93, 35–41. erals Engineering 15, 963–970.

You might also like