Sei sulla pagina 1di 9

Materials and Design 32 (2011) 1948–1956

Contents lists available at ScienceDirect

Materials and Design


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

Severe plastic deformation of copper and Al–Cu alloy using multiple channel-die
compression
A.K. Parimi, P.S. Robi, S.K. Dwivedy ⇑
Mechanical Engineering Department, Indian Institute of Technology Guwahati, Guwahati 781 039, India

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

Article history: Severe plastic deformation studies of copper and Al–Cu alloy by multiple channel-die compression tests
Received 2 April 2010 were investigated. The materials were tested under plane strain condition by maintaining a constant
Accepted 30 November 2010 strain rate of 0.001/s. Extensive grain refinement was observed resulting in nano-sized grains after severe
Available online 4 December 2010
plastic deformation with concomitant increase in flow stress and hardness. The microstructural investi-
gation of the severely deformed materials was investigated using optical microscope and scanning elec-
Keywords: tron microscope. Shear band formation was identified as the failure mechanism in the two phase Al–Cu
A. Non-ferrous metals and alloys
alloy. The results indicate difficulty in obtaining severe plastic deformation for alloys having two phase
A. Nano material
F. Plastic behavior
micro-structure.
Ó 2010 Elsevier Ltd. All rights reserved.

1. Introduction deformation (T = 400 °C) up to a true strain of 6.0 and reported


the following unusual behavior. (a) The yield stress dependence
In recent years, bulk nano-structured materials (NSM) pro- was non-monotonic. (b) The yield stress of each subsequent defor-
cessed by the methods of severe plastic deformation (SPD) have at- mation step was always less than the flow stress at the end of the
tracted interest among material scientists and engineers. Using previous one. (c) The strain hardening as a function of strain in
SPD approach, materials can be deformed to very large strains each deformation step was similar.
thereby enhancing the physical and mechanical properties. Valiev Li and Blum [6] used multiple uniaxial compression technique
et al. [1] carried an extensive research on different methods of along three orthogonal axis to impart SPD in copper up to a true
SPD for obtaining NSM. The advantages of these methods being strain of 7.8. The increased strength and strain rate sensitivity by
that ultra-fine grained materials with enhanced strength, higher this technique was similar to that obtained through ECAP process.
thermal stability, increased fatigue life, high damage tolerances, Li et al. [7] studied the strain rate sensitivity of copper at 373 K as a
absence of porosity high strain rate super plasticity can be ob- function of prestrain by multiple uniaxial compression technique.
tained. Applications of these materials include light weight struc- The deformation behavior of ultrafine-grained (UFG) and coarse-
tures for automotive and aerospace industries, sports goods like grained (CG) copper produced by severe plastic deformation was
tennis rackets having high strength to weight ratio, medical im- reported by Li et al. [8]. Kundu et al. [9] used multiple compression
plants, high strength micro springs and gears, etc. technique in a channel-die to impart severe plastic deformation to
Saito et al. [2] studied the accumulative roll-bonding process for copper at ambient and elevated temperatures up to an equivalent
producing ultra-fine grained bulk aluminum. Vorhauer and Pippen strain of eight. Zhu et al. [10] reported the properties like strength,
[3] studied the influence of type and path of deformation on the ductility, toughness, fatigue strength, wear resistance, thermal sta-
microstructural evolution during severe plastic deformation. They bility, corrosion resistance, magnetic and optical properties of dif-
used plane strain compression technique in cyclic/multiple chan- ferent nanostructured materials obtained by SPD. Sauvage et al.
nel compression die similar to plane strain rolling condition. They [11] investigated the role of interphase boundaries on the grain
imparted SPD through repeated pressing of pure Al and Cu through size reduction mechanisms in a Cu–Cr composite. Experiments
a channel-die. Becker and Lalli [4] studied the texture evolution on 70–30 brass in a channel-die by Nourbakhsh and Vujic [12]
and the effect on grains in channel-die compression process. indicated work hardening rate to be high for low and high strains
Salishchev et al. [5] investigated the micro-structure evolution whereas for medium strains the rate of work hardening was low.
and mechanical behavior of the commercial pure titanium. The influence of the micro-structure and the misorientation rela-
Multiple channel-die compression tests were carried out at warm tionship between grains on the mechanical properties in the Cu
specimens processed by equal-channel angular extrusion up to
⇑ Corresponding author. Tel.: +91 361 258 2670; fax: +91 361 2690762. 16 passes was investigated by Torre et al. [13]. Ghosh [14]
E-mail address: dwivedy@iitg.ernet.in (S.K. Dwivedy). deformed precipitation hardened aluminium alloys along three

0261-3069/$ - see front matter Ó 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.matdes.2010.11.074
A.K. Parimi et al. / Materials and Design 32 (2011) 1948–1956 1949

principal axes in successive operations up to a cumulative true plate was used. Since the previous strain history of this material
strain of eight. The tests performed at different strain rates and was not known, it was annealed at 450 °C for 2 h. For two phase
temperatures resulted in elongated grains. Vanderesse et al. [15] material, Al–Cu alloy cast in the form of slab was used as the start-
introduced new high temperature channel-die compression equip- ing material. In order to reduce the chemical non-homogeneity like
ment designed to operate under inert atmosphere up to 800 °C for coring and segregation during solidification, the cast slab was an-
Zircalloy-4. Kapoor and Chakravartty [16] studied the deformation nealed at 425 °C for 4 h in a resistance heated muffle furnace and
behavior of ultra-fine grained Al–Mg alloy produced by equal- then water cooled.
channel angular pressing after deforming the material to equiva- From the annealed plates of Cu and Al–Cu alloy, rectangular
lent prestrains of 8 and 13. Wang and Ma [17] studied the strain samples with dimensions of 12.1 mm  6 mm  6 mm were pre-
hardening and strain rate hardening behavior of nano-structured pared by wire cut electrical discharge machining process. The
and ultrafine-grain metals based on some experimental findings. chemical composition analysis of both the materials were deter-
Wadsack et al. [18] studied the development of micro-structure mined using Energy Dispersive X-ray analysis (EDS). EDS spec-
and thermal stability of nano-structured Chromium processed by trums obtained at a very low magnification is shown in Fig. 4.
SPD methods. Kucukomeroglu [19] investigated the effect of Quantitative analysis using high purity cobalt as the standard re-
equal-channel angular extrusion on mechanical and wear proper- vealed that Cu samples were 99.9% pure and the composition of
ties of eutectic Al–12Si alloy. aluminum alloy was Al–5.5 wt.%Cu.
In spite of the high interest in SPD methods, very less research Multiple channel-die compression experiments were carried
has been done on the cyclic/multiple channel-die compression out on an Instron make 8801 model servo-hydraulic controlled
method of SPD. With few exceptions, majority of the SPD experi- Universal testing machine (UTM). The samples were deformed at
ments were carried out on commercially pure aluminium or cop-
per, where the material exists as a single phase. The micro-
structure of majority of the metallic alloys consist of two phases
with varying properties. The effects of the micro-structure on the
severe plastic deformation have not been investigated in detail.
The present work was taken up to investigate the characteristics
of single phase and two phase alloys deformed by multiple chan-
nel-die compression method. Two different materials, pure copper
and Al–5.5%Cu alloy, were tested using this technique. The flow
characteristics of these materials were investigated during each
deformation pass.

2. Experimental procedures

Fig. 1 shows schematically the various steps in a multiple chan-


nel-die compression process. After each compression pass, the
sample is rotated by 90°. After three pass, the sample attains the
original orientation. Fig. 2 shows the dimensions of the die block
used for the experiments which consist of two split die blocks,
top and bottom punches, four bolts with nuts and washers. The
photographs of the sample placed in one half of the split die as well
as the die punch assembly are shown in Fig. 3.
Using this set up, the bulging of the specimen during deforma-
tion could be avoided and the material deformed under plane
strain condition. Also, the grinding of the sample after each com-
pression cycle reported in the work of Li and Blum [6] was not
necessary.
For studying the deformation behavior of single phase material,
specimen from commercially available electrical grade copper Fig. 2. Dimensions of the split die block and punches.

Fig. 1. Schematic representation of the sequence of multiple channel-die compression process.


1950 A.K. Parimi et al. / Materials and Design 32 (2011) 1948–1956

Fig. 3. Photographs of (a) the sample placed in one half of the split die and (b) die punch assembly.

Fig. 4. EDS analysis image showing the composition of (a) Cu and (b) Cu–Al alloy.

a constant true strain rate of 0.001 s1 for each pass. For this the tor position at each time interval during each pass was recorded.
velocity of the UTM actuator (v) was controlled based on the The effective plastic strain and true stress during the compression
instantaneous actuator position (h) as per the following test was determined by the following relationships.
relationships. Effective plastic strain:
Velocity of the ram:  
2  1=2
v ¼ e_  L0  expðe_  tÞ ð1Þ
e ¼ e21 þ e22 þ e23 ð4Þ
3

dh ¼ v  dt ð2Þ True stress:

r ¼ instantaneous load=instantaneous cross-sectional area: ð5Þ


h ¼ hprev ious  ðdhÞcurrent ð3Þ
It was not possible to determine the exact instantaneous cross-
where e_ is the true strain rate, and t is the instantaneous time, L0 is sectional area of the specimen during the deformation process.
the initial height of the specimen. The external load and the actua- Hence, the calculation of the true stress was based on the assump-
A.K. Parimi et al. / Materials and Design 32 (2011) 1948–1956 1951

Fig. 5. (a) A typical load vs. deformation curve, and (b) a typical true stress vs. true strain. Curve during compression test for annealed copper sample.

tion that the material deforms homogenously under plane strain


condition. The cross-sectional area was taken as w  t, where, the
width (w) was calculated using constant volume relationship,
and t is the thickness.
Tests were carried out with initial height of 12.1 mm and it was
compressed to a height of 6 mm corresponding to a strain e = 0.80,
following the sequence shown in Fig. 1. Experiments were re-
peated till the material failed. Total 12 samples were deformed
for Cu and 8 samples for the Al–Cu alloy.
Samples corresponding to each pass was sectioned using a low
speed precision saw and metallographically polished for observa-
tion under optical microscope (OM) and scanning electron micro-
scope (SEM). The copper samples were etched using a solution of
Fig. 6. A typical work hardening rate (dr/de) vs. strain (e) plot for copper.

Fig. 7. True stress vs. cumulative equivalent true strain graphs of copper after 4, 8 and 10 passes are shown in (a)–(c), respectively.
1952 A.K. Parimi et al. / Materials and Design 32 (2011) 1948–1956

25 ml NH4OH + 25 ml water + 25–50 ml H2O2 (3%) where as the 3. Results and discussion
Al–Cu alloy samples were etched using Keller’s reagent before
microstructural investigation. The grain size of the materials were 3.1. Copper
determined by line intercept method. The composition of the sec-
ond phases in the Al–Cu alloy was determined by SEM–EDS. The A typical load vs. displacement plot obtained during the chan-
hardness of the sample after each pass was determined using a nel-die compression testing is shown in Fig. 5a. The r vs. e plot ob-
Vicker’s micro-hardness tester (Make: BUEHLER, Model: Micromet tained from the above data is shown in Fig. 5b. This is similar to
2101). Microhardness was measured by applying an indentation that reported in literature [9]. The plot shows a linear increase in
load of 10 g for 10 s. The average diagonal length of the square true stress up to a true strain of 0.05 followed by plastic deforma-
impression formed was used for the determination of the hardness tion. After attaining a maximum stress at a strain of 0.3, the stress
value. Each value of the hardness reported is for an average of 6 decreases marginally or fluctuates till a strain of 0.5 is attained.
readings. Further deformation results in increase in the flow stress.
This behavior can be explained as follows. Since a very small
clearance is present between the specimen and the die, when the
yield stress is reached the material deforms sideways till it makes
contact with the die walls. During this stage, the frictional stresses
are absent between the side walls of the die and the billet. Further
advancement of the punch results in plane strain deformation of
the material. During this stage, two phenomena can occur: (i) the
material work hardens resulting in the increase in the stress and
(ii) the frictional forces come into picture as a result of the con-
straint imposed by the die walls during the deformation. Both
these result in increase in the stress with increase in strain. How-
ever, it is evident that in this region the stress increases with in-
crease in strain at a decreasing rate. This is due to the decrease
in contact area of the sample with the die walls up to a certain
amount of deformation (where the contact area becomes a square)
beyond which the contact area again starts increasing. A simple
calculation reveals that the minimum contact area is attained at
Fig. 8. Variation of hardness with cumulative equivalent strain for copper samples. a strain of 0.42.

Fig. 9. SEM micrographs of copper samples (a) before deformation (b) after 5 pass and (c) after 9 pass.
A.K. Parimi et al. / Materials and Design 32 (2011) 1948–1956 1953

Fig. 6 shows the typical work hardening rate (dr/de) vs. strain increase in the stress for strain beyond 0.42 can be attributed
(e) plotted up to a strain of 0.42. This figure was plotted by drawing to the increase in the stress due to friction as well as the fact that
a trend line from the experimental values of stress and strain. At the deforming material comes in contact with the die walls along
the onset of deformation, the work hardening rate is very high. the width direction thereby partly resisting the metal flow along
The rate of work hardening decreases at a fast rate with increase this direction.
in strain. After a certain amount of strain, the work hardening rate After each deformation pass, the equivalent true strain changes.
reaches almost zero. The increase in stress with further strain is It may be noted that after the first three passes, the material attains
mainly due to the frictional stress between the billet and die. The the initial geometry where the effective strain becomes zero.
Hence the true stress (r) vs. cumulative equivalent strain is plot-
ted. Fig. 7 shows the plot obtained for a copper sample with
10 passes. It is observed that the maximum flow stress during
(111)
the deformation gradually increases with each pass. After every
three passes, the maximum flow stress decreases for most of the
samples. The subsequent deformations after every three passes
indicate increase in maximum flow stress. For copper samples,
while a drop in flow stress beyond a cumulative equivalent strain
(200) of five was observed by Kundu et al. [9], in the present work it was
(220) observed at an equivalent value of seven.
Fig. 8 shows the hardness values plotted for different cumula-
tive equivalent strain corresponding to the different passes. The
hardness of the material increased from 96 VHN to 140 VHN after
Fig. 10. X-ray diffraction pattern of the copper sample after 10 passes. Scanning the first pass, i.e., 46% increase. Through the subsequent passes re-
was done in 2h values between 35° and 75°. sulted in an increasing trend, the hardness values were found to be

Fig. 11. (a) Typical load vs. deformation curve, (b) typical true stress–strain curve and (c) work hardening rate (dr/de) vs. strain plot for Al–Cu alloy.
1954 A.K. Parimi et al. / Materials and Design 32 (2011) 1948–1956

Fig. 12. True stress vs. cumulative equivalent true strain plots after (a) second pass and (b) after third pass during multiple compression tests for Al–Cu alloy.

Fig. 13. Al–Cu alloy samples shown during different stages of multiple compression (a) undeformed (b) after 1 pass (c) after 2 passes and (d) after 3 passes.

fluctuating. This is mainly due to the fact that the hardness values
for different passes were obtained from different samples since it
was not possible to carry out the experiment on the same sample.
The maximum hardness obtained was 177 VHN after eight passes,
i.e., 84% increase in hardness compared to the undeformed sample.
After eight passes, the samples became extremely brittle result-
ing in the formation of fine macroscopic cracks on the surface of
the copper specimen. After an equivalent strain of 8.5, i.e., after
10 passes, the material physically cracked. The SEM micrographs
of the copper in the undeformed condition, after the 5th pass
and after the 9th pass are shown in Fig. 9a–c, respectively. The
average grain size of copper in the undeformed condition was
25 lm. After five passes the average grain size reduced to almost Fig. 14. Variation of micro-hardness with cumulative equivalent true strain for
2 lm. After nine passes, extensive grain refinement resulting in different number of passes of Al–Cu alloy.

nano-structured material was observed. Though the bulk copper


scanning 2h between 35° and 75°. The first three peaks (43.2°, 50.2°
samples appeared to be nano-structured, the size could not be
and 73.9°corresponding to reflections from the (1 1 1), (2 0 0) and
determined from the SEM images.
(2 2 0) planes) of Cu crystals were analyzed. For this, the XRD data
The crystallite size of the Cu samples after 10 passes was deter-
was collected at a step size of 0.02° s1 with 10 s sampling time at
mined by X-ray diffraction technique using Cu Ka radiation
each step for the respective peaks.
(k = 1.54 Å) at a tube rating of 40 kV and 40 mA using a commercial
From the XRD analysis, using Eq. (6), the crystallite size of the
X-ray diffractometer (Bruker D8 Advance). The size of the copper
copper samples after 10 pass was found to be 21.84 nm. This indi-
crystallites was estimated from the full width at half maximum
cates that nano-structured bulk copper material could be obtained
(FWHM) of the respective XRD peaks using the Scherrer equation
by the multiple channel-die compression technique.
[21]:

0:9k 3.2. Al–Cu alloy


t¼ ð6Þ
B Coshhkl
Fig. 11a–c show the typical plots of load vs. deformation, true
where t is the crystallite size, k is the wavelength of the X-rays, hhkl stress vs. true strain and work hardening rate (dr/de) vs. strain
the Bragg angle, and B is the FWHM. Fig. 10 shows the X-ray diffrac- for the Al–Cu alloy obtained from the compression test up to a
tion pattern obtained for the copper samples after 10 passes while strain of 0.65. The trends of the variation of these plots are similar
A.K. Parimi et al. / Materials and Design 32 (2011) 1948–1956 1955

Fig. 13 shows the photographs of Al–Cu alloy for various passes.


After the second pass, cracks were observed on the surface of the
samples. From the stress–strain curve, it can be observed that this
cracking of the sample occurs when the cumulative equivalent true
strain value is around 1.5. After the third pass the material was
found to be extremely brittle and was completely cracked.
The variation of micro-hardness with cumulative equivalent
true strain for different number of passes of Al–Cu alloy is shown
in Fig. 14. A drastic increase in the hardness value for the Al–Cu al-
loy was observed after the first pass. The hardness of the material
increased by 59% after the first pass, whereas there was only a mar-
ginal increase in hardness during the subsequent deformations, i.e.
7% and 2% increase during the second and third passes respec-
tively. The brittleness of the material increased resulting in the for-
mation of the macroscopic defects after the 2nd and 3rd pass.
Fig. 15. SEM micrograph of Al–Cu alloy showing two different phases.
Analysis of the SEM image shown in Fig. 15 revealed an average
grain size of 50 lm. The micrograph shows a continuous brittle eu-
to that of copper specimen. Unlike the copper samples, the Al–Cu tectic mixture at grain boundary regions around the ductile pri-
alloy specimen could be strained only up to three passes, i.e. to a mary aluminum (a-phase). The composition of the eutectic
cumulative strain of 2.5. mixture analyzed by EDS was found to be CuAl2. The optical
Earlier plain strain compression study on two phase structured micrographs shown in Fig. 16 indicates fracture of the eutectic
Brass [12] revealed that the material could be strained only to a constituents after the 1st pass. The primary phase is also found
cumulative strain of 2.3. The plots of true stress vs. cumulative to have elongated mainly at the mid-height regions. Observation
equivalent true strain for different passes are shown in Fig. 12. of the etched sample after second pass resulted in fine grain sized
The stress–strain plots during compression testing of the Al–Cu al- primary phase (photograph not included). The un-etched optical
loy during the first pass exhibited trend similar to that of the cop- micrograph Fig. 16c shows the eutectic constituents have been
per alloy. fragmented and distributed uniformly in the primary aluminum
From the true stress vs. cumulative equivalent true strain plots phase. Grain refinement of the primary aluminum phase was
for Al–Cu alloy, the maximum flow stress increased from 320 MPa observed during the deformation. After 1st pass, grain size was
after the first pass to 375 MPa after the second pass. found to be 40 lm and after 3rd pass grain size reduced to around

Fig. 16. Optical microscope images of Al–Cu alloy at different stages (a) undeformed sample, (b) after 1 pass, (c) after 2 passes and (d) after 3 passes showing localized shear
band formation.
1956 A.K. Parimi et al. / Materials and Design 32 (2011) 1948–1956

20–25 lm. Localized shear bands were observed after the 3rd 7. The strengthening in the two phase material during the defor-
(Fig. 16d) pass. mation process is by work hardening.
The micro-structure of the Al–Cu alloy (Fig. 15) showing the 8. It is difficult for Al–Cu two phase materials to undergo large
two constituent phases are in agreement with the binary equilib- plastic deformation. After a true strain of 1.5, the material
rium phase diagram of Al–5.5%Cu alloy [20]. As per the phase dia- developed cracks.
gram, the primary phase (a-phase) is a solid solution of copper in
aluminum and the secondary phase, is an eutectic with a composi-
tion of 67%Al–33%Cu. The eutectic is a mixture of ductile a-phase Acknowledgement
and an intermetallic, h, in the relative ratio of 0.73:1. Intermetallic
materials being brittle fail at very low strain values. During the The authors thank Mr. Kaustubh Acharyya, Centre for Nano-
plastic deformation, the primary phase deforms extensively com- technology Indian Institute of Technology Guwahati for his help
pared to the secondary phase. This large variation in the deforma- in one of the experiment.
tion characteristics results in breakage of the brittle constituent
during the plastic strains and at the same time nucleates defects. References
The formation of these defects is mainly due to the differential flow
characteristics of the two phases in material. The shear band for- [1] Valiev RZ, Islamgaliev RK, Alexandrov IV. Bulk nano-structured materials from
severe plastic deformation. Prog Mater Sci 2000;45:103–89.
mation at localized regions in the material results in the failure [2] Saito Y, Tsuji N, Utsunomiya H, Sakai T, Hong RG. Ultra-fine grained bulk
of the material during subsequent deformation. Shear band forma- aluminium produced by accumulative roll-bonding process. Scripta Mater
tion was also observed in two phase brass [12] for medium strains. 1998;39:1221.
[3] Vorhauer A, Pippan R. The influence of type and path of deformation on the
Though grain refinement of the a-phase was observed in the Al–Cu microstructural evolution during severe plastic deformation. In: Zehetbauer M,
alloy, they could not be deformed for very large strains due to Valiev RZ, editors. Nanomaterials by severe plastic deformation. Weinheim:
physical failure of the samples. Wiley VCH; 2004. p. 684–90.
[4] Becker R, Lalli LA. Texture evolution in channel-die compression part II: effects
The results indicate that in single phase copper material the ini- of grains which shear. Textures Microstruct 1991;14–18:145–50.
tial strengthening is a result of work hardening whereas the harden- [5] Salishchev GA, Zherebtsov SV, Mironov SY, Myshlayev MM, Pippan R.
ing during the later part is due to reduction in grain size and that Formation of a submicrocrystalline structure I titanium during successive
uniaxial compression in three orthogonal directions. In: Zehetbauer M, Valiev
single phase materials can be successfully subjected to large defor-
RZ, editors. Nanomaterials by severe plastic deformation. Weinheim,
mation by SPD techniques with corresponding increase in strength. Germany: Wiley VCH; 2004. p. 691–7.
For two phase materials, the strengthening is mainly due to the work [6] Li YJ, Blum W. Strain rate sensitivity of Cu after severe plastic deformation by
multiple compression. Phys Status Solidi A 2005;202:119–21.
hardening which occurs during the first stage of the deformation
[7] Li YJ, Zeng XH, Blum W. On the elevated temperature deformation behavior of
process. Since large deformation is not possible for two phase Al– polycrystalline Cu subjected to predeformation by multiple compression.
Cu alloy, due to the nucleation of defects, strengthening due to grain Mater Sci Eng A 2008;483–484:547–50.
refinement cannot be achieved in these material. It can therefore be [8] Li YJ, Zeng XH, Blum W. Transition from strengthening to softening by grain
boundaries in ultrafine-grained Cu. Acta Mater 2004;52:5009–18.
concluded that two phase materials cannot undergo severe plastic [9] Kundu A, Kapoor R, Tewari R, Chakravarty JK. Severe plastic deformation of
deformation. copper using multiple compression in a channel die. Scripta Mater
2008;58:235–8.
[10] Zhu YT, Lowe TC, Langdon TG. Performance and applications of nanostructured
4. Conclusions materials produced by severe plastic deformation. Scripta Mater
2004;51:825–30.
[11] Sauvage X, Jessner P, Vurpillot F, Pippan R. Nanostructure and properties of a
Multiple channel-die compression test were carried out on sin-
Cu–Cr composite processed by severe plastic deformation. Scripta Mater
gle phase pure copper as well as two phase Al–Cu alloy. The flow 2008;58:1125–8.
stress, hardness and micro-structure of these materials were inves- [12] Nourbakhsh S, Vujic D. High strain plane-strain deformation of 70–30 brass in
tigated under severe plastic deformation. The main findings and a channel die. Acta Metall 1986;34:1083–90.
[13] Torre FD, Lapovok R, Sandlin J, Thomson PF, Davies CHJ, Pereloma EV.
conclusions are as follows: Microstructures and properties of copper processed by equal channel angular
extrusion for 1–16 passes. Acta Mater 2004;52:4819–32.
1. Single phase copper material could be deformed to a cumula- [14] Ghosh AK. Method for producing a fine-grain Aluminium alloy using three axis
deformation. US Patent 4721537; 1988.
tive equivalent true strain of 7 without any defect formation. [15] Vanderesse N, Desrayaud C, Girard-Insardi S, Darrieulat M. Channel die
2. The maximum true stress value for the copper increased from compression at high temperature. Mater Sci Eng A 2008;476:322–32.
470 MPa during the initial pass to 640 MPa during the 6th pass. [16] Kapoor R, Chakravartty JK. Deformation behavior of an ultrafine-grained Al–
Mg alloy produced by equal-channel angular pressing. Acta Mater
3. Nano-structured bulk copper material could be achieved by the 2007;55:5408–18.
multiple channel-die compression technique. [17] Wang YM, Ma E. Strain hardening, strain rate sensitivity, and ductility of
4. The initial strengthening in the single phase copper is a result of nanostructured metals. Mater Sci Eng A 2004;375–377:46–52.
[18] Wadsack R, Pippan R, Schedler B. Development of microstructure and thermal
work hardening whereas the hardening during the later part is stability of nanostructured chromium processed by SPD. In: Zehetbauer M,
due to grain refinement. Valiev RZ, editors. Nanomaterials by severe plastic deformation. Weinheim,
5. The two phase Al–Cu alloy could not be subjected to severe Germany: Wiley VCH; 2004. p. 655–9.
[19] Kucukomeroglu T. Effect of equal-channel angular extrusion on
plastic deformation due to the fragmentation of the eutectic
mechanical and wear properties of eutectic Al–12Si alloy. Mater Des
mixture. 2010;31:782–9.
6. Localized shear band formation in the soft matrix was identified [20] ASM Handbook, Alloy Phase Diagrams, vol. 03. ASM International; 1992.
as the final cause of failure of the two phase material. [21] Cullity BD. Elements of X-ray diffraction. 2nd ed. Addison-Wesley; 1978.

Potrebbero piacerti anche