You are on page 1of 11

Materials Science and Engineering A 539 (2012) 194204

Contents lists available at SciVerse ScienceDirect

Materials Science and Engineering A


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

Improving mechanical properties of chip-based aluminum extrudates by


integrated extrusion and equal channel angular pressing (iECAP)
M. Haase a, , N. Ben Khalifa a , A.E. Tekkaya a , W.Z. Misiolek b
a
b

TU Dortmund University, Institute of Forming Technology and Lightweight Construction, Baroper Strae 301, D-44227 Dortmund, Germany
Lehigh University, Institute for Metal Forming, 5 East Packer Avenue, Bethlehem, PA 18015, USA

a r t i c l e

i n f o

Article history:
Received 7 December 2011
Received in revised form 20 January 2012
Accepted 21 January 2012
Available online 28 January 2012
Keywords:
Chip extrusion
Equal channel angular pressing (ECAP)
Aluminum alloy recycling
Die design
Mechanical properties
Microstructure

a b s t r a c t
In order to improve the mechanical properties of proles extruded from aluminum chips, a four turn
equal channel angular pressing tool was integrated into an extrusion die (iECAP die). AA6060 aluminum
alloy turning chips were cold pre-compacted to chip-based billets and hot extruded through the iECAP
die on a conventional forward extrusion press. Mechanical properties and microstructure of the chipbased billets extruded through the iECAP die were investigated and compared to those extruded through
a conventional at-face die and a porthole die. To evaluate the performance of the iECAP processed chipbased proles, conventional cast billets were extruded through the at-face die as a reference material.
To investigate the inuence of temperature on mechanical properties and microstructure of chip-based
proles, the extrusion was performed at 450 C and 550 C.
Tensile tests revealed superior mechanical properties of the chip-based billets extruded through the
iECAP die in comparison to chip-based billets extruded through the at-face and the porthole die as well
as to cast billets extruded through the at-face die.
2012 Elsevier B.V. All rights reserved.

1. Introduction
1.1. Direct extrusion of aluminum chips
The direct recycling of aluminum chips using hot extrusion was
rst proposed and patented by Stern in 1945 [1]. Gronostajski et al.
[2] have investigated the direct conversion of aluminum chips to
nal products using a three step method: granulation of the chips
using a cutting device, cold pre-compaction and nally hot extrusion. The investigations were conducted consecutively with pure
aluminum, AlMg2 and AlCu4 alloys. They produced specimens with
a residual porosity of about 5% after hot extrusion, with an extrusion
ratio of 4:1. Hardness and tensile properties of the chip-based specimen were lower compared to extruded cast billets. Gronostajski
et al. [3] proposed the following factors to contribute signicantly
to the bonding quality of aluminum and aluminum alloy chips
with an introduced consolidating phase: (i) the amount, form
and size of the consolidating phase; (ii) the degree of neness of
the aluminum and aluminum alloy chips; (iii) the cold pressing

Corresponding author. Tel.: +49 231 755 2654; fax: +49 231 755 2489.
E-mail addresses: Matthias.Haase@iul.tu-dortmund.de (M. Haase),
Nooman.Ben Khalifa@iul.tu-dortmund.de (N. Ben Khalifa),
Erman.Tekkaya@iul.tu-dortmund.de (A.E. Tekkaya), wzm2@lehigh.edu
(W.Z. Misiolek).
0921-5093/$ see front matter 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2012.01.081

parameters; (iv) the shape of the extrusion dies; (v) the degree of
reduction; (vi) the lubrication method and the lubricants used; and
(vii) the temperature and rate of extrusion. The results of direct
recycling of machining chips consisting of different alloys without previous granulation were published by Fogagnolo et al. [4].
Hot and cold pre-compaction processes before hot extrusion were
compared. They found that only the combination of hot compaction
and hot extrusion led to a sufcient chip bonding for low extrusion
ratio (6.25:1), whereas for higher extrusion ratio (25:1) both precompaction methods led to a sufcient bonding of the extruded
chips. For higher extrusion ratio, the difference in ultimate tensile strength (UTS) between hot and cold pre-compacted billets
after extrusion was negligible. Therefore, Fogagnolo et al. [4] proposed the combination of cold pre-compaction and hot extrusion
to be the most promising process in terms of cost to benet ratio.
Tekkaya et al. [5] have studied the recycling of AA6060 aluminum
alloy chips using cold pre-compaction and hot extrusion. The inuence of different chip geometries, produced by milling and turning
operations, on tensile properties of the extruded chip-based billets
was investigated. During extrusion, a two-feeder porthole die with
an extrusion ratio of 34:1 was used in order to break oxide layers covering the chips and achieve good bonding of pure metal. The
mechanical properties showed comparable results to extruded cast
billets, independent of chip geometry and chip production method.
Generally, the introduced plastic strain during extrusion is
dened by the extrusion ratio. Increasing the extrusion ratio will

M. Haase et al. / Materials Science and Engineering A 539 (2012) 194204

result in an increase of the introduced plastic strain but also in an


increase of the extrusion force [6]. However, the selection of the
die design for a given extrusion ratio can provide a different level
of strain and therefore different conditions for chip bonding [5].
1.2. Equal channel angular pressing as a consolidation tool
Severe plastic deformation (SPD) processes introduce ultralarge plastic strain into bulk metals which results in the formation
of an ultra-ne grained microstructure with grain sizes below 1 m
and grain boundaries of high misorientation angles [6]. A well
known SPD process is equal channel angular pressing (ECAP), also
known as equal channel angular extrusion (ECAE), which was introduced and patented by Segal in 1977 [7]. Through simple shear,
ECAP is able to introduce plastic strain into the processed material without changing the cross-section of the workpiece [6]. An
overview of the principles of ECAP is given by Valiev and Langdon
[8].
Xiang et al. [9] have compared the microstructure and mechanical properties of aluminum powders consolidated in the ECAP
process at elevated temperatures and in the process of hot extrusion. The hardness value of the specimens consolidated with ECAP
was higher compared to the extruded samples. Xiang et al. [9]
related this to the ne microstructure and a high density of dislocations introduced into the workpiece. Xia et al. [10] compared
the effect of ECAP on microstructure and mechanical properties of
both pure aluminum powder previously consolidated with back
pressure ECAP (BP-ECAP) and cast aluminum billets. The consolidated workpiece showed a ner grain structure and higher tensile
strength. Balog et al. [11] found that the application of back pressure was inevitable when using ECAP as a consolidation tool for
Al particles in order to avoid surface cracks of the processed specimen. Luo et al. [12] recycled titanium machining chips directly
by using BP-ECAP at elevated temperature. After two passes, full
density and good bonding were achieved. According to Luo et al.
[12], the oxide layers gradually cracked with repeated deformation,
potentially leading to an additional dispersion strengthening.
ECAP has its own limitations, rst, the length to diameter ratio
of the workpiece is limited to a critical value to prevent bending.
Second, the ram of the press has a limited travel distance, which
limits the length of the workpiece. Third, due to inhomogeneous
microstructure and the appearance of cracks near both ends of the
workpiece, material has to be cut off in these regions. Fourth, the
process is labor intensive because the workpiece must be reinserted
into the die after every pass, making industrial application difcult
[13].
Paydar et al. [14] have proposed the combined process of
forward extrusion and ECAP (FE-ECAP) for the consolidation of
commercial pure aluminum particles. The method was carried out
at 200 C with a ram speed of 0.2 mm/s and an extrusion ratio of
7.1:1. Superior hardness, tensile strength and comparable ductility were achieved with the new method compared to extruded
ingot samples and samples consolidated with forward extrusion.
Paydar et al. [14] related this to a ner average grain size. Paydar et al. [15] also proposed the combined process of ECAP before
forward extrusion (ECAP-FE) for consolidation of commercial pure
aluminum particles. As an advantage to the previously presented
concept, the extrusion after ECAP automatically creates back pressure. However, this concept led to a lower ductility at similar
strength compared to FE-ECAP. Ying et al. [16] investigated the solid
state recycling of AZ91 Mg alloy machining chips through the process of cold compaction, hot compaction, extrusion and single pass
ECAP. The average grain size was proven to be much smaller compared to cast AZ91 alloy processed by extrusion and ECAP. Results
were attributed by Ying et al. [16] to the dispersion of oxide contaminants. Without single pass ECAP, the extruded cast workpiece

195

showed superior strength and ductility compared to the recycled


ones. After applying ECAP, the recycled workpiece showed higher
strength but lower ductility compared to the cast billet processed
under the same conditions.
Recently, Orlov et al. [17] have presented a two turn ECAP tool
integrated into a conventional extrusion tool, based on the principles of a patent by Estrin et al. [18]. They produced bars of ZK60 Mg
alloy, which showed improved strength and ductility compared to
the initial state at an extrusion ratio of 19:1.
2. Tool design
In this paper, the concept of integrated extrusion and ECAP
(iECAP) was modied in order to improve the mechanical properties of proles extruded from aluminum chips. For this purpose,
the number of ECAP turns was increased to four and the angle
between the channels was reduced. Mechanical properties and
microstructure of chip-based proles produced with the iECAP die
were investigated and compared to chip-based proles produced
with conventional hot extrusion using both the at-face die and the
porthole die. Cast billets were also extruded through the at-face
die to provide a reference material.
The die was designed for the use on a conventional hydraulic
extrusion press. A schematic illustration of the die is presented in
Fig. 1.
The tool consists of ve parts with part one including the conventional extrusion die and parts two to ve including the four
ECAP steps. The extrusion part was fabricated as a prechamber die
for a prole with a rectangular cross section of 20 mm 20 mm,
having a corner radius of rprole = 3.1 mm. The resulting extrusion
ratio (ER) for this prole is 8.7:1, the resulting true strain after
the extrusion step can be calculated to Extrusion = ln (ER) 2.16
[19]. Four ECAP turns were implemented into the die, equivalent
to routes CAC in conventional ECAP, where C indicates a sample rotation of 180 and A no rotation between consecutive passes
around extrusion direction axis [20]. Back pressure is automatically
applied to the processed material after every ECAP turn when the
material ows to the die wall in the proposed tool with the exception of the last ECAP turn. The channel angle was fabricated with
= 90 according to Nakashima et al. [21], who showed that imposing a very intense plastic strain is important in order to achieve
a ne grained microstructure. The die was designed without an
outer arc of curvature, which leads to the highest strain in every
ECAP turn for dies with = 90 [22]. The channel displacement was
designed as K = wc = 20 mm, where wc is the channel width, referring to Raab [23], who proposed K = dc for round specimens, where
dc is the channel diameter, in order to achieve strain homogeneity
in the cross section of specimens processed by ECAP with parallel
channels. A graphite based guiding tool was used behind the last
ECAP step and the die exit to maintain straight material ow.
3. Experimental materials and methods
3.1. Experimental material and processing
The material used in the following experiments was aluminum
alloy AA6060, provided by apt Hiller GmbH. The chemical composition of the material (see Table 1) was analyzed by apt Hiller
GmbH by optical emission spectroscopy using a Thermo Scientic
ARL 3460 Metals Analyzer.
The aluminum was processed in three consecutive steps. First,
aluminum chips were produced by turning of the cast billet. Second,
the chips were cold compacted on a hydraulic press into chip-based
billets. Third, the billets were forward extruded on a hydraulic
extrusion press at elevated temperature using the billet-to-billet

196

M. Haase et al. / Materials Science and Engineering A 539 (2012) 194204

Fig. 1. Schematic illustration of the four turn iECAP die and the corresponding prole cross section. is the channel angle and K is the channel displacement. All given
dimensions are in mm.
Table 1
Chemical composition of the AA6060 aluminum alloy used in the study.
Si

Fe

Cu

Mn

Mg

Cr

Zn

Ti

Others

Al

0.45

0.208

0.016

<0.05

0.372

0.001

0.005

0.014

0.005

Balance

extrusion method without removing the discard. The three processing steps are illustrated in Fig. 2(a)(c).
The chips were produced from as-received cast bars using a
turning device without lubrication. The turning parameters are
given in Table 2.
After turning, the chips had an average hardness value of
82 HV. The chips were cold compacted in a steel tube on a
hydraulic press with an inner diameter of dsteel tube = 60 mm. The
procedure of lling the steel tube and compacting the chips was
repeated several times to achieve a billet length of at least 60 mm.
An average billet density of 80% was achieved with a compaction
force of 500 kN. The extrusion was conducted on the hydraulic

Table 2
Turning parameters for the fabrication of the chips and geometric chip
characterization.
Turning parameters
Cutting speed
Feed
Cutting depth

vc = 400 m/min
f = 0.4 mm
ap = 2.25 mm

Chip geometry
Length
Width
Thickness

lchips 11 1.7 mm
wchips 7.6 1.2 mm
tchips 1.05 0.35 mm

extrusion press Collin LPA250t, which has a maximum extrusion


force of 2.5 MN. This particular press has a container diameter of
d0 = 66 mm. The length of the remaining discard for every extrusion
cycle is constant and 30 mm.

Fig. 2. Processing of the AA6060 aluminum alloy (a) turning tool used for chip fabrication and produced chips, (b) hydraulic press for cold compaction of the chips and
cold-compacted billet and (c) extrusion press and extruded prole.

M. Haase et al. / Materials Science and Engineering A 539 (2012) 194204

197

Fig. 3. Tools used for the investigation of chip extrusion and the corresponding material extracted from the die after extrusion (a) conventional at-face die, (b) four-feeder
porthole die and (c) iECAP die.

Table 3
Fabricated specimens and corresponding process parameters.
Specimen
Reference (Ref.)
Flat-face die, low temperature (FL)
Flat-face die, high temperature (FH)
Porthole die, low temperature (PL)
Porthole die, high temperature (PH)
iECAP die, low temperature (EL)
iECAP die, high temperature (EH)

Die
Flat-face
Flat-face
Flat-face
Porthole
Porthole
iECAP
iECAP

Billet temperature
450 C
450 C
550 C
450 C
550 C
450 C
550 C

Material
Cast
Chips
Chips
Chips
Chips
Chips
Chips

To evaluate the performance of the chip-based billets extruded


through the designed iECAP die, chip-based billets were also
extruded through a conventional at-face die and a four-feeder
porthole die. Both dies were used to manufacture solid proles
with a rectangular cross section of 20 mm 20 mm, resulting in the
extrusion ratio of 8.6:1. The tools and the corresponding material
removed from the dies are shown in Fig. 3.
As reference specimen, conventional cast billets of the same
alloy as the chips, with a diameter of dbillet = 62 mm and a length
of lbillet = 85 mm, were extruded through the at-face die. All billets were homogenized in an electrical furnace for 6 h at 550 C. In
order to investigate the inuence of billet temperature on mechanical properties and microstructure of the chip-based extrudates, the
billets were extruded at 550 C and 450 C. The chip-based billets
extruded at 450 C and the cast billets were cooled in the furnace
after homogenization and before extrusion. Ram speed was set to
1 mm/s, die and container temperature to 450 C. The specimen
notation for the combinations of dies used, billet temperature and
material are shown in Table 3.

of 2.5 103 s1 at room temperature on a Zwick/Roell Z250 tensile test machine. Tensile test specimens were fabricated parallel
to extrusion direction (ED) by milling in accordance with EN ISO
6892-1:2009 standard and pulled to failure. All reported results
are an average of at least three tensile tests. The dimensions of the
tensile specimens are given in Fig. 4. The rst tensile specimen for
each condition was machined out of the extruded prole after at
least 1000 mm from the proles front end.
Vickers hardness was measured with an applied loading force
of 1.961 N (i.e. HV0.2) and a holding time of 10 s at room temperature. The measurement was conducted on a Struers Duramin-1
with a Vickers diamond indenter in accordance with DIN EN ISO
6507-1:2005 standard. Specimens were taken from the EDTD
plane, mechanically ground using SiC paper (grit 500, 1000, 2400
and 4000 for 180 s each) and polished for 600 s with colloidal silica. The mean values of HV are an average of at least six hardness
measurements.

3.2. Mechanical characterization

The microstructure was characterized for all extruded proles


using light optical microscopy (LOM) under polarized light on Zeiss
Axio Imager.M1m with Zeiss AxioCam MRc. Specimens were taken
from the EDTD plane, mechanically ground and nally polished
with colloidal silica as it was done for the hardness measurements.

In order to compare the mechanical properties of the extruded


proles, tensile tests and hardness measurements were conducted. Tensile tests were performed with an initial strain rate

3.3. Microstructure

Fig. 4. Tensile test specimen (light gray) machined out of the extruded prole (dark gray). L0 is the gage length for the strain sensor, all given dimensions are in mm.

198

M. Haase et al. / Materials Science and Engineering A 539 (2012) 194204

A discard of 30 mm is left in the press container after the extrusion of each billet. For easier comparison of the initial steps of
extrusion for different dies, the curves in Fig. 5 are aligned to the
end-point of the extrusion stroke and the extrusion force is shown
as a function of the remaining ram displacement. The maximum
extrusion force of the cast billet extruded through the at-face die
was 0.73 MN at a remaining ram displacement of 106 mm. The
chip-based billets extruded through the at-face and porthole die
showed a maximum extrusion force of 0.73 MN and 1.7 MN at a
remaining ram displacement of 85 mm and 76 mm, respectively.
The chip-based billet extruded through the iECAP die had a maximum extrusion force of 1.98 MN at a remaining ram displacement
of 78 mm.
Fig. 5. Extrusion force vs. remaining ram displacement. Billets were extruded at a
billet temperature of 550 C and a ram speed of 1 mm/s through different dies.

The polished specimens were electrolytically etched using Barkers


reagent [24] with an applied voltage of U = 25 V dc for 120 s on a
Struers LectroPol-5. The average grain size was measured using the
linear intercept method in accordance with ASTM E112-96 (2004)
standard.

4. Results

4.2. Material ow in iECAP and imposed strain


In order to reveal the material ow within the iECAP die, the
material was taken out of the die and the container of the extrusion
press using caustic soda, cut in the EDND plane, ground, polished,
etched with Barkers reagent and investigated under polarized
light. Fig. 6 shows a collage of images of the aluminum processed
with the iECAP die revealing dead metal zone (DMZ) formation.
The presented gure (Fig. 6) is a collage of images, so the overall
dimensions are slightly incorrect due to cutting losses of 1 mm
for each cut. Regions of different grain sizes can be observed.

4.1. Extrusion force


The extrusion force is one of the characteristic factors in the
extrusion process. Therefore, it is very advantageous to investigate
the extrusion force for new die designs and to compare them to
known die sets. Chip-based billets of the same length (117 mm)
were prepared and extruded at a billet temperature of 550 C and
a ram speed of 1 mm/s through the iECAP die, the porthole die and
the at-face die. For comparison, a cast billet of the same length
was extruded through the at-face die. Cast billets were extruded
through all die sets to ll the dies before performing all the experiments. The extrusion forces related to the different die sets are
shown in Fig. 5.

4.3. Bonding quality of chips recycled with the iECAP die


During the extrusion process, the single chips were welded
together. No voids between the chips can be observed for the chipbased specimen extruded through the iECAP die in Fig. 7(a). The
appearance is comparable to the cast billet extruded through the
iECAP die shown in Fig. 7(c). The shape of the single chips can be
made visible within the extrudate using Barkers reagent, as shown
in Fig. 7(b). When treating the extruded cast billet with the same
reagent, no signicant difference can be seen in the macroscopic
appearance of the extruded cast billet (Fig. 7(d)).

Fig. 6. Collage of the microstructure images in the EDND plane of the iECAP die and magnications.

M. Haase et al. / Materials Science and Engineering A 539 (2012) 194204

199

maximum true strain value of 0.24, the EL sample has a maximum


true stress value of 174 MPa.
The FH and PH specimens have maximum true stress values of
169 MPa and 178 MPa at corresponding true strain values of 0.15
and 0.16, respectively. The EH specimen has a maximum true stress
value of 195 MPa at a corresponding true strain value of 0.24.
5. Discussion
5.1. Extrusion force

Fig. 7. Chip-based and cast specimens processed with the iECAP die and cut in the
EDTD plane. (a) and (c): after grinding and polishing. (b) and (d): after grinding,
polishing and etching. The presented images are collages of smaller localized images.
ED is upwards.

4.4. Microstructure analysis


The microstructures of the processed specimens, labeled
according to Table 3, are shown in Fig. 8 for the specimen extruded
at 450 C and Fig. 9 for the specimen extruded at 550 C.
The microstructure analysis of the reference specimen reveals
an elongated grain structure and a thick zone of peripheral coarse
grains (PCG). The FL specimen shows equiaxed and elongated grains
with the PCG zone smaller compared to the reference specimen. The
FH specimen shows a similar microstructure to the FL sample with
a bigger PCG zone. The PL specimen has a similar microstructure to
the PH specimen with equiaxed and elongated grains at the center
of the prole, corresponding to the seam weld line, and equiaxed
grains elsewhere. The PCG zone is smaller for the PL specimen. The
EL specimen has equiaxed grains without a PCG-zone while the EH
sample shows equiaxed grains and a PCG zone similar in size to the
PH sample.
The dimension of the PCG zone increases for every die set with
increasing temperature, however, no increase in average grain size
could be measured for all chip-based specimens when PCG was
not taken into account. The iECAP processed specimens had an
average grain size of 31.8 m, similar to the chip-based specimens
processed with the at-face or porthole die.
4.5. Hardness measurement
The results of the Vickers hardness measurement across the
specimen width within the EDTD plane for the specimens
extruded at 450 C and 550 C are shown in Fig. 10(a) and (b),
respectively.
The reference specimen has the lowest and the FL specimen has
the highest average hardness value of 41 HV and 53 HV, respectively. The PL specimen has an average hardness value of 44 HV,
while the hardness value of the EL specimen is 46 HV. The FH
specimen has 49 HV, the highest average hardness value among
the specimens extruded at 550 C. The PH and EH specimen have
an average hardness value of 48 HV and 43 HV, respectively.
4.6. Tensile test
The true stress vs. true strain curves for the specimen extruded
at 450 C and 550 C are presented in Fig. 11(a) and (b), respectively.
The reference specimen has a maximum true stress value of
167 MPa at a true strain value of 0.15. The FL and PL specimens
have maximum true stress values of 162 MPa and 168 MPa at
corresponding true strain values of 0.14 and 0.17, respectively. At a

As it can be seen in Fig. 5, the extrusion of the cast billet begins


earlier compared to the chip-based billets, although all billets have
the same initial length. The density of chip-based billets achieved
by cold compaction was 80%. The chip-based billets are further
compacted in the press container before the extrusion takes place.
A signicant increase in extrusion force during upsetting can be
observed for the chip-based billet extruded through the at-face
die. This can be related to the effect of hot compaction before upsetting of the chip-based billet in the press container. While extruding
aluminum through the porthole or the iECAP die, a small aluminum
shell is formed in the press container. Because of the formation of
the shell during lling the dies, the initial extrusion force does not
start at zero at the next extrusion cycle due to additional friction
within the press container.
The earlier initialization of upsetting for the chip-based billet
extruded through the at-face die compared to the other die sets
is due to slightly higher initial billet density of the billet extruded
through the at-face die.
The maximum extrusion forces of cast and chip-based billets
extruded through the at-face die are similar. The theoretical
length of the chip-based billet at full compaction is 20% shorter
compared to the length of the cast billet due to its 80% density. This
indicates that a higher extrusion force is necessary for the extrusion
of chip-based material under the same conditions.
The porthole die and the iECAP die show higher maximum
extrusion forces than the at-face die. For the porthole die, this
can be related to two different mechanisms. The rst is the force
needed to divide the aluminum into four single strings. The second
is the increase in friction force due to the increase in friction surfaces compared to the at-face die. For the iECAP die, the increase
in extrusion force can be related to the four shear zones present
within the ECAP steps and to the increase in friction surfaces. For
this die, all friction surfaces after the prechamber can be seen as
one single bearing surface of complex geometry leading to the nal
prole geometry.
5.2. Material ow in iECAP and imposed strain
It can be observed in Fig. 6 that four dened dead metal zones
appear in the ECAP steps of the iECAP die. Eivani and Karimi Taheri
[25] investigated the effect of DMZ formation on imposed strain
and pressing force during conventional ECAP using upper bound
analysis. It was proposed that the formation of the deformation
zone and therefore of the DMZ, depends on the friction condition
and the channel angle. They assumed the DMZ to act as a die wall
with an outer arc of curvature, leading to an angle of deformation
zone ( ). According to Eivani and Karimi Taheri [25], this angle can
be calculated as:

 = 2 cot

1m

1+m

(1)

where m is the friction factor and is the channel angle. The presented iECAP tool has a channel angle of = 90 in every ECAP step.
Assuming the friction factor to be m = 1, the deformation zone angle

200

M. Haase et al. / Materials Science and Engineering A 539 (2012) 194204

Fig. 8. Collages of the microstructures of the specimens extruded at 450 C with corresponding magnications. Specimens were taken from the EDTD plane, ED is upwards.

can be calculated to be  = 90 in every ECAP step by using Eq. (1).


An increasing angle of deformation zone is resulting in a decrease
of the total imposed strain [22]. The introduced equivalent plastic
strain for a single pass was proposed by Eivani and Karimi Taheri
[25] as:
tot

1
=
3

 
2

1m
+
1+m

2 cot1



1m

1+m

1+m
2

(2)

For m = 1, the introduced total plastic strain is tot 0.9 in one


ECAP step which leads to a total imposed plastic strain in the ECAP

part of the iECAP die of ECAP 3.6. As observed from the revealed
microstructure in Fig. 6(a), the actual deformation angle seems less
than  = 90 . However, a zone of ultra ne grains between coarse
grains occurs along the theoretical curvature related to an angle of
 90 , as it can be seen in Fig. 6(b), so shear along this curvature
is assumed.
5.3. Bonding quality of chips recycled with the iECAP die
Referring to Gronostajski et al. [3], extrusion ratio, extrusion
temperature and die shape are signicant process parameters

M. Haase et al. / Materials Science and Engineering A 539 (2012) 194204

201

Fig. 9. Collages of the microstructures of the specimens extruded at 550 C and of the reference specimen with corresponding magnications. Specimens were taken from
the EDTD plane, ED is upwards.

controlling the quality of the chip bonding. Taking into account


the investigation of Fogagnolo et al. [4], the processing route of
cold compaction and hot extrusion through a at-face die was not
leading to sufcient chip bonding for an extrusion ratio of 6.3:1.
The iECAP die used in the experiments had an extrusion ratio of
8.7:1, the investigated at-face die and porthole die of 8.6:1.
The usage of the at-face die led in some cases to a peeled off surface of the extruded prole as it is shown in Fig. 12, which conrms
the results of the referred investigation.

All chip-based billets extruded through the porthole die and


the iECAP die of comparable extrusion ratio showed sufcient chip
bonding without surface defects.
For all die sets used in the performed experiments, the bonding quality increased with increasing temperature. These results
conrm the results of Ceretti et al. [26], who presented a new procedure for the identication of the extrusion welding criterion.
They simulated aluminum seam welding in porthole die extrusion by performing a at rolling experiment of a sandwich of two

202

M. Haase et al. / Materials Science and Engineering A 539 (2012) 194204

Fig. 10. Results of hardness measurement over the specimens width in the EDTD plane (a) specimens extruded at 450 C, (b) specimens processed at 550 C and reference
specimen.

Fig. 11. Tensile test results of the processed specimens. (a) Specimens extruded at 450 C and (b) Specimens extruded at 550 C and reference specimen.

rectangular AA6061 aluminum alloy specimens in order to bond


them. They varied the testing temperature and the rolling ratio in
order to investigate the effect of temperature and pressure on the
welding quality. In the experiment of Ceretti et al. [26], higher temperatures led to a better solid state bonding, measured by optical
investigation and micro-hardness tests in the normal direction to
the bonding line.
Taking into account the geometry of the die, the welding quality
increased when changing the die set from at-face die to porthole
die and then to the iECAP die. This can be related to the longer

deformation zone. Referring again to Ceretti et al. [26], it was shown


that a better solid state bonding was achieved for a higher rolling
ratio, i.e. higher pressure and strain. By using the porthole die,
a higher amount of pressure affects the chip bonding due to the
pressure occurring in the welding chamber and a higher amount of
strain is introduced into the specimen due to the additional shear
when the material ow is divided into the four single strings. Therefore, the bonding quality increases using a porthole die instead of a
at-face die. The iECAP die needed the highest extrusion force of the
investigated die sets which leads to a high amount of pressure on
the chips during extrusion. In addition, every ECAP turn within the
die leads to a back pressure for the previous turn and to additional
strain through shear. Referring to Xia et al. [27], when consolidating particles using ECAP, the application of back pressure leads to
a synthesis of fully dense bulk material. The combination of high
pressure, additional strain and back pressure explains the superior
bonding quality of the iECAP processed specimen.

5.4. Microstructure analysis

Fig. 12. Peeled off prole surface during extrusion through the conventional atface die for chip extrusion at an extrusion ratio of 8.6:1.

The microstructure of extruded chip-based billets differs from


the microstructure of extruded cast billets (reference specimen)
in terms of shape and size of the grains (Figs. 8 and 9). Chipbased billets extruded through the at-face die have a smaller grain
size compared to cast billets extruded through the same die. The
microstructure analysis revealed a mixture of equiaxed and elongated grains for chip-based proles produced with the at-face die.
When using the porthole or the iECAP die for chip extrusion, the
grains become equiaxed and the PCG zones become smaller compared to chip-based or cast billets extruded through the at-face
die.

M. Haase et al. / Materials Science and Engineering A 539 (2012) 194204

Fig. 13. Prevented grain growth due to chip boundary in a chip-based specimen, ED
is upwards.

A positive effect concerning the extrusion of chips instead of cast


material is the prevention of grain growth, as chip boundaries act
as natural barriers between grains, which can be seen in Fig. 13.
Although the single chips are not visible after extrusion, the
chip structure still exists and could be made visible through
metallographic techniques. It is assumed that only a partial breaking of the oxide layer between the single chips occurs, which
prevents the chips from perfect welding and leads to the shown
effect. Comparing the microstructure of the chip-based proles
with the reference specimen in the EDTD plane, the average grain
size of chip-based proles is 68% smaller compared to cast billets extruded through the at-face die. This can be related to the
dispersion of oxide contaminants, as it was assumed by Ying et al.
[16], and to the prevention of grain growth due to chip boundaries
as proposed by the current authors.
5.5. Hardness measurement
The overall hardness distribution over the EDTD plane is rather
homogeneous for all specimens, the average standard deviation of
each specimen is lower than 2.5. When increasing the extrusion
temperature from 450 C to 550 C, the hardness value increased
for the chip-based specimen extruded through the porthole die by
9.1%, while it decreased by 7.5% and 6.5% for the chip-based
specimen extruded through the at-face and the iECAP die, respectively. All chip-based specimens have a higher hardness compared
to the cast billet extruded through the at-face die. This can be
related to the smaller average grain size of the chip-based specimen
compared to the extruded cast billet. Additionally, Luo et al. [12]
suggested an additional dispersion strengthening due to gradual
fragmentation of the oxide layers between chips.
The hardness of an individual chip after turning and before
cold-compaction and hot extrusion was 82 HV. As a result of
hot extrusion, all extruded chip-based proles exhibited a lower
average hardness compared to the chips after turning. The chipbased billets extruded through the at-face die were exposed to the
lowest strain and therefore to the lowest temperature in the deformation zone during extrusion for both extrusion temperatures of
450 C and 550 C compared to the chip-based billets extruded
through the porthole die and iECAP die. This resulted in the highest
hardness for the FL specimen and the FH specimen for the corresponding extrusion temperatures.
5.6. Tensile test
The maximum true stress values and the corresponding true
strain values for the FL specimen and PL specimen are comparable
to the reference specimen. Having a comparable maximum true
stress value, the EL specimen shows a 60% higher corresponding
true strain value than the reference specimen. The high ductility of
the EL specimen benets in case of further mechanical processing
of the extruded proles in terms of their forming limit.

203

The FH specimen shows a comparable maximum true stress


value to the reference specimen at the same true strain value.
Slightly higher true stress (6.6%) and true strain (6.7%) values can
be observed for the PH specimen. The EH sample shows superior
strength and ductility, having a 16.7% higher maximum true stress
value at a true strain value 60% higher compared to the reference
specimen.
The tensile test results strongly depend on the extrusion temperature. The maximum true stress value of the specimen extruded
at 550 C are always higher than of those extruded at 450 C with
the same die set. This is most signicant for the iECAP die processed
specimen with an increase in maximum true stress of 12.1% compared to an increase of 6% and 4.3% for the porthole die and
at-face die processed specimen, respectively, when increasing the
extrusion temperature from 450 C to 550 C. The maximum true
strain level increased by 0.1 for the at-face die and decreased by
0.1 for the porthole die processed specimen when the temperature
was changed from 450 C to 550 C. No inuence on the true strain
value could be measured for the iECAP die processed specimens.
For the chip-based specimen extruded at 550 C, an increase in the
maximum true stress value of 5.3% and 10% can be observed
when changing the die set from the at-face to the porthole and
from the porthole to the iECAP die, respectively. For the investigated parameters, a change of the die set leads to a more signicant
change in strength and ductility than a change of temperature.
No correlation between the values of the hardness measurement
(Fig. 10) and the tensile test results (Fig. 11) can be observed for the
chip-based specimens. The hardness values represent the material
response of an individual chip or eventually two chips and the chip
boundary, as the dimension of the indentation is smaller compared
to the dimension of a chip. The tensile properties are representing the material response as an average of the chips in the whole
deformation zone of the tensile test specimen. This includes the
mechanical properties and microstructure of the individual chips,
which correlate to the hardness values, and in addition the bonding
between the single chips and the contribution of the PCG zone.
The microstructure of the chip-based specimen extruded
through the at-face die at 450 C (Fig. 8, FL) appears to result in
superior tensile results compared to the cast reference specimen
(Fig. 8, Ref.). However the tensile properties are slightly inferior for
the chip-based specimen. This can be related to insufcient chip
bonding when extruding chip-based billets through the at-face die
at the low extrusion ratio of 8.6:1. Tensile properties of the chipbased specimens increased when the die set was changed from
at-face die to porthole die and then to iECAP die (Fig. 11). These
more complex extrusion dies provided deformation routes resulting in better chip bonding, more equiaxed grains and a decrease in
the size of the PCG zone. The inferior tensile results of the porthole
die compared to the iECAP die are due to the presence of the four
welding lines where the material strings of the four portholes were
welded together in the welding chamber.
6. Conclusion
A four turn equal channel angular pressing tool integrated
into an extrusion die (iECAP die) was designed to improve the
mechanical properties of proles extruded from aluminum chips.
Microstructure and mechanical properties of chip-based billets
extruded through the iECAP die were compared to chip-based billets extruded through the at-face die and the porthole die as well
as to cast billets extruded through the at-face die. The following
conclusions can be drawn:
- At the low extrusion ratio of 8.7:1, the deformation path of the
at-face die did not guarantee sound chip bonding. However, the
porthole and the iECAP die provided deformation conditions for

204

M. Haase et al. / Materials Science and Engineering A 539 (2012) 194204

the same extrusion ratio resulting in successful solid state recycling of aluminum chips. The superior bonding quality of the chips
extruded through the iECAP die can be related to a high amount of
pressure affecting the chips, additional strain and back pressure
due to the ECAP turns.
The analysis of the material ow in the iECAP die revealed
dead metal zone formation in the corners of the ECAP steps,
related to friction between aluminum and the die walls. The dead
metal zone is considered to act as an extension of the die wall,
leading to a deformation zone angle of  90 . Additional strain
of ECAP 3.6 was introduced into the material in the four ECAP
parts of the iECAP die.
For chip-based billets extruded through the iECAP die, no chip
boundaries could be observed after extrusion. After etching, the
chip boundaries could be made visible again. It is assumed that
chip welding is partially prevented due to remaining oxide layers
on the chips. The remaining oxide layers are assumed to be obstacles for the grain growth over the chip boundaries and potentially
lead to dispersion strengthening.
The chip-based billets extruded through the iECAP die and the
porthole die, respectively, showed equiaxed small grains, while
the chip-based billets extruded through the at-face die showed a
combination of equiaxed and elongated grains with a bigger PCG
zone compared to the other die sets.
Using the iECAP die instead of the at-face or the porthole die
as a tool for solid state recycling of aluminum machining chips,
under the presented conditions, leads to improved chip bonding
and superior strength and ductility.
Compared to cast billets extruded through the at-face die, hardness measurement and tensile tests revealed superior mechanical
properties of chip-based billets extruded through the iECAP die.

Acknowledgments
The support for W.Z. Misiolek as Mercator Visiting Professor
at TU Dortmund University has been provided by the German
Research Foundation (DFG) while he has been also supported by
the Loewy Family Foundation at Lehigh University in Bethlehem,
PA, USA through Loewy Professorship. The corresponding author

acknowledges the nancial support granted by the Graduate School


of Energy Efcient Production and Logistics in Dortmund.
References
[1] M. Stern, U.S. Patent 2,391,752 (1945).
[2] J.Z. Gronostajski, J.W. Kaczmar, H. Marciniak, A. Matuszak, J. Mater. Process.
Technol. 64 (1997) 149156.
[3] J. Gronostajski, H. Marciniak, A. Matuszak, J. Mater. Process. Technol. 106 (2000)
3439.
[4] J.B. Fogagnolo, E.M. Ruiz-Navas, M.A. Simn, M.A. Martinez, J. Mater. Process.
Technol. 143144 (2003) 792795.
[5] A.E. Tekkaya, M. Schikorra, D. Becker, D. Biermann, N. Hammer, K. Pantke, J.
Mater. Process. Technol. 209 (2009) 33433350.
[6] A. Azushima, R. Kopp, A. Korhonen, D.Y. Yang, F. Micari, G.D. Lahoti, P. Groche,
J. Yanagimoto, N. Tsuji, A. Rosochowski, A. Yanagida, CIRP Ann. Manuf. Technol.
57 (2008) 716735.
[7] V.M. Segal, Patent of the USSR, No. 575892 (1977).
[8] R.Z. Valiev, T.G. Langdon, Prog. Mater. Sci. 51 (2006) 881981.
[9] S. Xiang, K. Matsuki, N. Takatsuji, M. Tokizawa, T. Yokote, J. Kusui, K. Yokoe, J.
Mater. Sci. Lett. 16 (1997) 17251727.
[10] K. Xia, X. Wu, T. Honma, S.P. Ringer, J. Mater. Sci. 42 (2007) 15511560.
[11] M. Balog, F. Simancik, O. Bajana, R. Guillermo, Mater. Sci. Eng. A 504 (2009) 17.
[12] P. Luo, H. Xie, M. Paladugu, S. Palanisamy, M.S. Dargusch, K. Xia, J. Mater. Sci.
45 (2010) 46064612.
[13] C. Xu, S. Schroeder, P.B. Berbon, T.G. Langdon, Acta Mater. 58 (2010) 13791386.
[14] M.H. Paydar, M. Reihanian, E. Bagherpour, M. Sharifzadeh, M. Zarinejad, T.A.
Dean, Mater. Lett. 62 (2008) 32663268.
[15] M.H. Paydar, M. Reihanian, E. Bagherpour, M. Sharifzadeh, M. Zarinejad, T.A.
Dean, Mater. Des. 30 (2009) 429432.
[16] T. Ying, M. Zheng, X. Hu, K. Wu, Trans. Nonferrous Met. Soc. China 20 (2010)
604607.
[17] D. Orlov, G. Raab, T.T. Lamark, M. Popov, Y. Estrin, Acta Mater. 59 (2011)
375385.
[18] Y. Estrin, H. Ferkel, R.J. Hellmig, T. Lamark, M.V. Popov, German Patent
DE102005049369 (2008).
[19] M. Bauser, G. Sauer, K. Siegert, Strangpressen, Aluminium-Verlag, Dsseldorf,
2001.
[20] A. Rebhi, T. Makhlouf, N. Njah, Y. Champion, J.-P. Couzini, Mater. Charact. 60
(2009) 14891495.
[21] K. Nakashima, Z. Horita, M. Nemoto, T.G. Langdon, Acta Mater. 46 (1998)
15891599.
[22] Y. Iwahashi, Z. Horita, M. Nemoto, T.G. Langdon, Acta Mater. 45 (1997)
47334741.
[23] G.I. Raab, Mater. Sci. Eng. A 410411 (2005) 230233.
[24] R.K. Roy, S. Das, J. Mater. Sci. 41 (2006) 289292.
[25] A.R. Eivani, A. Karimi Taheri, Comp. Mater. Sci. 42 (2008) 1420.
[26] E. Ceretti, L. Fratini, F. Gagliardi, C. Giardini, CIRP Ann. Manuf. Technol. 58 (2009)
259262.
[27] K. Xia, X. Wu, Scripta Mater. 53 (2005) 12251229.

You might also like