Vous êtes sur la page 1sur 7

Materials Science and Engineering A 527 (2010) 14041410

Contents lists available at ScienceDirect

Materials Science and Engineering A


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

3D FEM simulations for the homogeneity of plastic deformation in


AlCu alloys during ECAP
Nahed El Mahallawy a , Farouk A. Shehata b , Mohamed Abd El Hameed b ,
Mohamed Ibrahim Abd El Aal b,c, , Hyoung Seop Kim c
a
b
c

Faculty of Engineering and Materials Science, the German University in Cairo, Cairo, Egypt
Mechanical Design and Production Department, Faculty of Engineering, Zagazig University, Zagazig, Egypt
Department of Materials Science and Engineering, POSTECH (Pohang University of Science and Technology), Pohang 790-784, Republic of Korea

a r t i c l e

i n f o

Article history:
Received 12 June 2009
Received in revised form 6 October 2009
Accepted 10 October 2009

Keywords:
ECAP
AlCu alloys
FEM
Microhardness
Homogeneity of deformation

a b s t r a c t
Equal channel angular pressing (ECAP) is a material processing method that allows very high strains
to be imposed, leading to extreme work hardening and microstructural renement. To investigate the
deformation homogeneity in the transverse direction, rigid-viscoplastic 3D nite element simulations
were conducted for the different numbers of ECAP passes of Al with Cu contents 05%. The simulation
results indicated that the material on the outer side of the die channel undergoes less deformation than
that in the inner side due to the formation of a corner gap. It was also found that the homogeneity increased
with increasing the number of ECAP passes and the copper content due to the decrease in the size of the
corner gap. To verify the 3D nite element simulation results, the microhardness homogeneity across
the transverse direction of the billet was measured. The same trend was observed: the homogeneity
in hardness increased with increasing the number of ECAP passes and Cu contents from 0% to 5%. The
homogeneity of deformation indicated by microhardness and by FEM results was higher for route A
compared with route Bc and increases with the number of ECAP passes. The homogeneity in route A was
higher than that in route Bc by 10% after 2 passes up to 8 passes.
2009 Elsevier B.V. All rights reserved.

1. Introduction
Materials with nanometer or sub micrometer grain size have
received great interest in the last decades because of their unique
mechanical and physical properties and high performance [14].
Recently, several methods of severe plastic deformation (SPD) were
developed to process bulk nanostructured materials with a grain
size from 20 to 200 nm depending on a number of factors [58].
Equal channel angular pressing (ECAP) is considered the most popular among the various SPD processes. The principle of ECAP has
been discussed in many different works [7,8].
However, there are many works on Al 6000 series alloys [9] and
steels [10], but only a few works available on AlCu alloys processed
by ECAP. The microstructure and hardness of Al1.7 wt.% Cu alloy
after ECAP was studied [11], as well as the cyclic deformation and
fatigue properties of Al0.7 wt.% Cu alloy [12], the shear feature
of Al33 wt.% Cu eutectic alloy at 400 C [13], the tensile properties

Corresponding author at: Mechanical Design and Production Department, Faculty of Engineering, Zagazig University, Zagazig, Egypt. Tel.: +20 127195264;
fax: +20 55 2304987.
E-mail addresses: eng mohabdelall@yahoo.com, engmohabdela@postech.ac.kr
(M.I.A.E. Aal).
0921-5093/$ see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2009.10.032

and fracture modes of cast Al0.63 wt.% Cu and Al3.9 wt.% Cu alloys
[14] subjected to ECAP. Recently, Prados et al. studied the tensile
behavior of an ECAP processed Al4% Cu alloy [15]. However, there
is no work concerned with the homogeneity of these AlCu alloys,
as far as the authors know.
The nite element method (FEM) is one of the most important
numerical methods that can be used to explain the deformation
process during the ECAP and related processes [1622]. Most of the
previous studies done on the ECAP by FEM are in two dimensions
(2D) of the plain strain condition. They include the plastic deformation analyses of metallic materials during ECAP [16], the correct
selection of die channel in ECAP [17,18], die design for homogeneous plastic deformation during ECAP [19], FEM analysis of ECAP
in strain rate sensitive metals [20], texture evolution [21,22], modied ECAP processing [23], bending behavior [24], and the die corner
gap formation in ECAP [25]. Also, there is very little work on FEM
that has been done by the three dimensional (3D) method. Recently,
Suo et al. [26] and Basavaraj et al. [27] have done some 3D analyses to trace the homogeneity during the ECAP processes after the
rst pass. Xu et al. [28] and Jiang et al. [29] studied the distribution
of strain in the cross-section of the sample of pure Al and CP-Ti,
respectively, during the 3D FEM simulations for the multiple passes.
However, no work has been found for theoretical investigations and
tracing of the homogeneity of multiple passes during ECAP process.

N.E. Mahallawy et al. / Materials Science and Engineering A 527 (2010) 14041410

1405

numbers of ECAP passes. The strain inhomogeneity index SII along


the transverse direction [26,27] was calculated from the following
equation:
SII =

Fig. 1. (a) ECAP die complete drawing and (b) the shape of the sample in the exit
from the die and the transverse direction (TD) of measuring the microhardness.

In the present work the deformation homogeneity of pure Al,


Al3 wt.% Cu and Al5 wt.% Cu alloys was investigated using 3D FEM
simulations for multiple passes. The effect of the ECAP number of
passes and the Cu content were studied, as well. The simulation
results were veried experimentally using the distribution of the
microhardness along the transverse direction.
2. Experimental
AlCu alloys with 3 and 5 wt.% Cu were cast, and then homogenized at 550 C for 7 days. Commercial purity Al in the form of
as-received ingots was also used. The samples were machined to
15 mm in diameter and 80 mm in length for ECAP. A split die was
constructed with a design that allows the ECAP to occur and to be
repeated in a fast and easy way. The die has a diameter of 15 mm,
channel angle of 110 , and an outer corner angle of 15 , as shown
in Fig. 1. This die gives strain of 0.77 for each pass. The details of the
ECAP process and the mechanical testing were briey mentioned in
a previous work [30]. A lubricant, zinc stearate, was used to reduce
the friction effect between the die and the sample during the ECAP.
The microhardness inhomogeneity index HII was calculated
from the following equation:
HII =

 MaxH MinH 
AvgH

(1)

where MaxH, MinH, and AvgH denote the maximum, minimum,


and average microhardness values along the transverse direction
in the central steady state region in the sample, respectively.

(Maxp Minp)
,
Avgp

(2)

where Maxp, Minp, and Avgp denote the maximum, minimum,


and average effective plastic strains along transverse direction,
respectively.
The workpiece used for analysis was cylindrical in shape with a
diameter of 15 mm and a length of 80 mm. The workpiece geometry
was drawn with the 3D CAD Program Solid Works; it was saved in
the STL le format. The STL format can be loaded into the program
DEFORM-3D used in the current simulations. The die and the punch
were also drawn and loaded into the program in the same way.
The workpiece was considered to be rigid-plastic; the material
characteristics that required dening for the program are shown in
the following ow equation:
T ),
 = (, ,

(3)

where  is the ow stress of the material, is the effective strain,


is the effective strain rate, and T is the temperature of the material. The ow characteristics of the materials used in the simulation
process were obtained from the tensile tests. Fig. 2 shows the true
stresstrue strain relations for the pure Al and the AlCu alloys. The
and T were measured from the tensile tests. The
values of , , ,
stressstrain data were loaded for the different materials in a tabular form for DEFORM-3D and the temperature was considered to
be room temperature, like most experimental works, 20 C.
The die and the punch considered for the experimental and
numerical analyses were made of high strength steel. As the
strength and rigidity of the steel are very high compared to the
workpiece materials, the die and the punch were modeled as a rigid
surface. The inside surface of the die and the outside surface of the
workpiece can be in close contact, depending on the local deforming situations. A nominal value zero for the coefcient of friction
was assumed between these two contacting surfaces. Heat generation due to friction and deformation was ignored. A constant punch
speed of 1 mm/s was imposed on the rigid punch. The workpiece
volume was meshed to 10,000 four-noded elements, which is considered sufcient, as shown in Fig. 3. This number is more than
double the number of elements used in the 3D simulation of the
previous work with the same dimensions of workpiece [29]. The
mesh system was automatically remeshed if the elements became
too distorted during the ECAP simulations.
For the simulation of multiple processing, the last values of
stress and strain in the sample material were adopted as the initial

3. 3D nite element simulation


To approximate the ECAP process in two-dimension, there are
three possible methods, namely: (i) plane strain, (ii) plane stress,
and (iii) axi-symmetric. Two-dimensional plane strain approximation can be used when the thickness of the workpiece is very large;
plane stress approximation can be used when the thickness is very
small. Both the plane strain and the plane stress conditions are not
applicable to the round workpiece used in ECAP because of the difference in workpiece geometries. Axi-symmetric approximation is
not suitable in ECAP because the axes of the channels intersect at
an angle of . In the present work, a 3D model was considered,
in order to study the distribution of the strain accumulated in the
specimen for AlCu alloys and pure Al.
The rigid-viscoplastic 3D nite element simulations were carried out using the commercial software DEFORM-3D V.5.0. The
strain distributions were measured along the transverse direction to compare them with the distributions of the microhardness
and to investigate the homogeneity of the specimens for different

Fig. 2. True stress true strain relation pure aluminum, Al3%Cu alloy and Al5%Cu
alloy.

1406

N.E. Mahallawy et al. / Materials Science and Engineering A 527 (2010) 14041410

values in the simulation of the subsequent processing. The ECAP


processed materials continue to be strain-hardened from the end
states of the previous step, following the ow curves in Fig. 2. This
batch-type method is more realistic for experimental situations
than using the initial continuous dies sets. Most of the simulation
for the multiple steps of ECAP used the continuous-type [31] which
is easy to operate in FEM. The batch-type simulation technique
[21,22] was employed as well to consider real situations well.
The geometry data for the rst step in any new pass was interpolated from the last step in the previous simulation. The boundary
conditions applied to the model were as follows: rst, the die was
modeled as a rigid body. Displacement and rotation in x, y, and z
directions of the die were arrested. Second, the top surface of the
workpiece was in contact with the punch and took load resulting
in the movement of the workpiece. All nodes on the top surface of
workpiece received displacement in the direction of movement of
punch. Route A and route Bc were simulated in the case of Al, and
route A in the case of the AlCu alloys. The modeling of both routes
was made by the positioning option in the program, in which the
workpiece of the last step in each pass was taken as the workpiece
of the rst step in the next pass. The workpiece was positioned in
the die without any rotation about its axis in the case of route A,
and rotated with angle of 90 around its axis after each pass in the
case of route Bc.
4. Results and discussion
4.1. Microhardness distribution and homogeneity
Fig. 3. The meshing used for the workpiec.

The distribution of the microhardness along the traverse direction for the different alloys is used as an index of the inhomogeneity
of the plastic deformation across the samples. Fig. 4 indicates that
the microhardness values increase from the bottom to the top of

Fig. 4. The distribution of microhardness in the transverse direction (a) pure aluminum, (b) Al3%Cu alloy and (c) Al5%Cu alloy.

Fig. 5. The formation of the corner gap in (a) the second pass and (b) the eighth pass in the case of Al3%Cu homogenized alloy.

N.E. Mahallawy et al. / Materials Science and Engineering A 527 (2010) 14041410

1407

Table 1
3-D FEM simulation values of the corner gap angle for aluminum and Al3%Cu alloy.

Fig. 6. Effect of ECAP number of passes and copper content on inhomogeneity index
in microhardness.

the specimens. The lower values of the microhardness at the bottom of the specimens can be attributed to the formation of corner
gaps between the bottom of the specimen and the die during the
ECAP process, as shown in Fig. 5. As the gap made a corner gap in
the bottom of the specimen, the specimen was no longer in contact
with the die in this area. This reduced the deformation in the bottom part of the specimen. However, it should be noticed that the
difference between the values of microhardness from the bottom
to the top of the specimen decreased with the increasing number
of ECAP passes, due to the increase in homogeneity with further
deformation.
There have been analyses on the deformation behavior according to properties of matter in the case of the ECAP process because

Material

Processing route

Number of passes

Corner gap angle

Al
Al
Al
Al
Al
Al3%Cu alloy
Al3%Cu alloy

Route A
Route A
Route A
Route Bc
Route Bc
Route A
Route A

2
8
10
2
8
2
8

35
23
21
33
25
26
15

it has a close relation with the uniform deformation of materials.


In general, materials showing a strain hardening have a large corner gap by the ECAP process and other materials without the strain
hardening have a smaller corner gap, so it can be found out that the
stressstrain hardening of material is one of major parameters of
the occurrence of the corner gap [25]. In addition, the corner gap
serves as a signicant parameter which affects the homogeneous
deformation. For the inhomogeneity index of the microhardness
calculated for the different alloys, see Fig. 6, which indicates that
it decreases with the increasing number of ECAP passes and with
increasing Cu content, i.e. the material becomes more homogeneously deformed. This can be explained by the reduction in the size
of the corner gap with the increasing number of passes, as shown
for Al3%Cu alloy after 8 passes in Fig. 5(b), because strain hardening decreases with strain, shown in ow curves of Fig. 2. Table 1
indicates the values of the corner gap angle after 2 passes and 8
passes for Al3%Cu alloy. It was observed that the value of corner
gap was decreased from 27 after 2 passes to 17 after 8 passes. The
effect of the corner gap on the distribution of the microhardness
was also observed in previous work [32,33]. Fig. 6 indicates that
route A gives more homogeneity in deformation than does route
Bc.

Fig. 7. Effective plastic strain distribution and the formation of corner gap (a) route A during the second pass of pure aluminum and (b) route A at the end of second pass of
pure aluminum.

Fig. 8. Effective plastic strain distribution and the formation of corner gap (a) route A during the tenth pass of pure aluminum and (b) route A at the end of tenth pass of pure
aluminum.

1408

N.E. Mahallawy et al. / Materials Science and Engineering A 527 (2010) 14041410

Fig. 9. Effective plastic strain distribution and the formation of corner gap (a) route Bc during the second pass of pure aluminum and (b) route Bc at the end of second pass
of pure aluminum.

Fig. 10. Effective plastic strain distribution and the formation of corner gap (a) route Bc during the eighth pass of pure aluminum and (b) route Bc at the end of eighth pass
of pure aluminum.

Fig. 11. Effective plastic strain distribution and the formation of corner gap (a) route A during the second pass of Al3%Cu alloy and (b) route A at the end of second pass
Al3%Cu alloy.

4.2. 3D nite element simulations


Figs. 712 show the effective strain distribution for Al processed
by route A, route Bc and Al3%Cu alloy processed by route A for a
different number of ECAP passes. The effective strain distribution
was obtained across the middle plane of the workpiece, where it

is more representative for strain distribution, not on the outer surface of the workpiece as was investigated in previous work [28,29].
It was found that the value of strain at the top of the workpiece
was higher than that in the bottom. The distribution of the strain
along the transverse direction of the sample for Al and AlCu alloys
is shown in Fig. 13. It can be observed that the strain increased

Fig. 12. Effective plastic strain distribution and the formation of corner gap (a) route A during the eighth pass of Al3%Cu alloy and (b) rout A at the end of eighth pass
Al3%Cu alloy.

N.E. Mahallawy et al. / Materials Science and Engineering A 527 (2010) 14041410

1409

Fig. 13. The distribution of effective plastic strain in the transverse direction (a) pure aluminum, (b) Al3%Cu alloy and (c) Al5%Cu alloy.

from the bottom to the top of the sample. This behavior of strain
distribution was observed for Al processed by route A and route
Bc, for different AlCu alloys, and for different numbers of ECAP
passes. The same trend of strain distribution for different processing routes with Al was also observed in previous work with 3D
FEM simulations [2628,33] and 2D FEM simulations [16,18,19,25].
However, Tham et al. [34] indicated different 2D FEM results for Al
6061 alloy, for which the strain was higher at the bottom side of
specimen than at the top of the specimen. The formation of the
corner gap during the ECAP process can be the main reason for
the difference in the effective strain distribution from the top to
the bottom of the workpiece. The formation of the corner gap can
be observed during the ECAP for Al and Al3%Cu alloy, as shown
in Figs. 712 during 3D simulations, which was detected during
the actual ECAP for Al3%Cu alloy, as shown in Fig. 5. The size of
the gap was observed to be reduced with the increasing number
of ECAP passes. The corner gap angle in the case of Al process by
route A was reduced from 35 after 2 passes, down to 23 , and
then to 21 after 8 and 10 passes, respectively. The same behavior
was also observed for Al processed by route BC and for Al3%Cu
alloy of route A, as shown in Figs. 912. Table 1 indicates the values of the corner gap angle in pure Al processed by routes A and
Bc, and Al3%Cu alloy of route A. The formation of the corner gap
was detected in previous nite element works [25,27]. The reduction of the work harden ability that occurs during the ECAP can
explain the reduction of the corner gap with the increasing number of ECAP passes due to the increasing material ability to ll the
die corner.
It is observed that the size of the corner gap was reduced with
increasing Cu content. That is, after 2 passes the corner gap angle
was 35 , 33 and 26 in the cases of Al by route A, Al by route Bc,
and Al3%Cu alloy by route A, respectively. It is also observed that
after 8 passes the corner gap angle becomes 23 , 25 and 15 in the
cases of Al by route A, Al by route Bc, and Al3%Cu alloy by route A,
respectively. In the literature [30], the AlCu alloys were observed
to be more homogenized than the pure Al with the increasing number of ECAP passes. The formation and the reduction in the size of
the corner gap with the increasing number of ECAP passes is in good
agreement with the experimental results that shows that the size

of the corner gap decreases with the number of passes for Al3%Cu
alloy, as shown in Fig. 5. Also, the values of the corner gap angle
from the simulation results in the case of Al3%Cu alloy were very
close to those in the case of the experimental work, as indicated in
Table 1. The inhomogeneity indices of the effective plastic strain for
the different alloys are shown in Fig. 14. It can be observed that the
deformation homogeneity increased with the increasing number
of ECAP passes and the Cu content.
It is very interesting that the homogeneity simulated by FEM in
the case of route A was higher than that in case of route Bc, i.e., route
A homogeneity is higher than that for route Bc by about 10% after
the second pass. This trend stays the same after 6 and 8 passes. This
trend was conrmed from the experimental results, which show
that the homogeneity of route A was higher than that for route Bc
as shown in Fig. 6 except for the second pass. The present experimental and theoretical fact that route A ECAP is better than route
Bc is contradictory to the received knowledge that route Bc is better than route A in terms of grain renement. Therefore, a different

Fig. 14. Effect of ECAP number of passes and copper content on inhomogeneity
index in effective plastic strain.

1410

N.E. Mahallawy et al. / Materials Science and Engineering A 527 (2010) 14041410

ECAP route needs to be selected, depending on the purpose of the


processing, i.e. homogenization or grain renement.
5. Conclusions
In this study, the results of the 3D rigid-viscoplastic nite element analysis of AlCu alloys during ECAP of routes A and Bc
was presented. The simulated results of the plastic effective strain
show the same trend as that of the hardness distribution in the
transverse direction. The FEM simulations prove the experimental
results about the formation of the corner gap and the reduction of
its size with the increasing number of ECAP passes. The effective
plastic strain index of inhomogeneity decreased with the number
of ECAP passes. From the experimental and FEM results, route A was
found to give higher homogeneity than route Bc. The 3D FEM simulation can be successfully used to nd the variations of strain with
the number of passes, processing route, and Cu content in AlCu
alloys.
Acknowledgements
Mohamed Ibrahim Abd El Aal acknowledges the scholarship
of the Egyptian government for his visiting research in POSTECH.
Hyoung Seop Kim acknowledges support by a grant from the Center for Advanced Materials Processing (CAMP) of the 21st Century
Frontier R&D Program funded by the Ministry of Science and Technology, Republic of Korea.
References
[1] R.Z. Valiev, I.V. Alexandrov, Y.T. Zhu, T.C. Lowe, J. Mater. Res. 17 (2002) 58.
[2] H.S. Kim, Scripta Mater. 39 (1998) 10571061.
[3] H.S. Kim, C. Suryanarayana, S.-J. Kim, B.S. Chun, Powder Metall. 41 (1998)
217220.
[4] H.S. Kim, Y. Estrin, Appl. Phys. Lett. 79 (2001) 41154117.

[5] R.Z. Valiev, R.K. Islamgalive, I.V. Alexandrov, Prog. Mater. Sci. 45 (2000)
103189.
[6] Y.H. Jang, S.S. Kim, S.Z. Han, C.Y. Lim, M. Goto, Metal. Mater. Inter. 14 (2008)
171175.
[7] V.M. Segal, Mater. Sci. Eng. A 197 (1995) 157164.
[8] Y. Iwahashi, J. Wang, Z. Horita, M. Nemoto, T.G. Langdon, Scripta Mater. 35
(1996) 143146.
[9] B.S. Moon, H.S. Kim, S.I. Hong, Scripta Mater. 46 (2002) 131136.
[10] H.S. Kim, W.S. Ryu, M. Janecek, S.C. Baik, Y. Estrin, Adv. Eng. Mater. 7 (2005)
4346.
[11] M. Murayama, Z. Horita, K. Hono, Acta Mater. 49 (2001) 2129.
[12] Z.F. Zhang, S.D. Wu, Y.J. Li, S.M. Liu, Z.G. Wang, Mater. Sci. Eng. A 412 (2005)
279286.
[13] J. Wang, S. Kang, H. Kim, Mater. Sci. Eng. A 383 (2004) 356361.
[14] D.R. Fang, Z.F. Zhang, S.D. Wu, C.X. Huanga, H. Zhang, N.Q. Zha, J.J. Li, Mater. Sci.
Eng. A 426 (2006) 305313.
[15] E. Prados, V. Sordi, M. Ferrante, Mater. Sci. Eng. A 503 (2009) 6870.
[16] H.S. Kim, M.H. Seo, S.I. Hong, J. Mater. Proc. Technol. 113 (2001) 622626.
[17] J.-H. Han, H.-J. Chang, K.-K. Jee, K.H. Oh, Metals Mater. Inter. 15 (2009) 439445.
[18] C.J. Lusi Perez, Scripta Mater. 50 (2004) 387393.
[19] S.C. Yoon, P. Quang, S.I. Hong, H.S. Kim, J. Mater. Proc. Technol. 187188 (2007)
4650.
[20] H.S. Kim, M.H. Seo, S.I. Hong, J. Mater. Proc. Technol. 130131 (2004) 497503.
[21] S.C. Baik, Y. Estrin, H.S. Kim, R. Hellmig, H.-T. Jeong, Mater. Sci. Forum 408412
(2002) 697702.
[22] S.C. Baik, Y. Estrin, R.J. Hellmig, H.T. Jeong, H.-G. Brokmeier, H.S. Kim, Z. Metallkd
94 (2003) 11891198.
[23] A.V. Nagasekhar, H.S. Kim, Metals Mater. Inter. 14 (2008) 565568.
[24] S.C. Yoon, A.V. Nagasekhar, H.S. Kim, Metals Mater. Inter. 15 (2009) 215219.
[25] H.S. Kim, M.H. Seo, S.I. Hong, Mater. Sci. Eng. A 291 (2000) 8690.
[26] T. Suo, Y. Li, Y. Guo, Y. Liu, Mater. Sci. Eng. A 432 (2006) 269274.
[27] V.P. Basavaraj, U. Chakkingal, T.S.P. Kumar, J. Mater. Proc. Technol. 202 (2008)
543548.
[28] S. Xu, G. Zhao, Y. Luan, Y. Guan, J. Mater. Proc. Technol. 176 (2006) 251259.
[29] H. Jiang, Z. Fan, C. Xie, Mater. Sci. Eng. A 485 (2008) 409414.
[30] N. El Mahallawy, F. Shehata, M. Abd El Hameed, M. Abd El Al, Mater. Sci. Eng. A
517 (2009) 4650.
[31] H.S. Kim, Mater. Sci. Eng. A503 (2009) 130136.
[32] C. Xu, M. Furukawa, Z. Horita, T.G. Langdon, Mater. Sci. Eng. A 398 (2005)
6676.
[33] P. Leo, E. Cerri, P.P. De Marco, H.J. Roven, J. Mater. Proc. Technol. 182 (2007)
207214.
[34] Y.W. Tham, M.W. Fu, H.H. Hng, M.S. Yong, K.B. Lim, J. Mater. Proc. Technol.
192193 (2007) 121127.

Vous aimerez peut-être aussi