Open Journal of Metal, 2011, 1, 12-15
doi:10.4236/ojmetal.2011.11002 Published Online September 2011 (http://www.SciRP.org/journal/ojmetal)
Copyright © 2011 SciRes. OJMetal
Effect of Hot Accumulative Roll Bonding Process on the
Mechanical Properties of AA5083
Hassan Sheikh, Ebrahim Paimozd
Materials Engineering Department, Malek Ashtar University of Technology, Shahin Shahr, Iran
E-mail: sheikh_scientific@yahoo.com
Received August 10, 2011; revised September 16, 2011; accepted September 25, 2011
Abstract
In this work, accumulative roll bonding (ARB) process as a severe plastic deformation was applied on an
AA5083 sheet up to 6 cycles at the temperature of 300C and at the strain rate of 50 s1. The results of tensile
tests show that the values of the yield stress and the ultimate tensile strength don’t change considerably after
4 cycles. Also, transmission electron microscopy (TEM) micrograph confirmed that the microstructure has
fine (sub) grains with the size of 200 nm - 400 nm.
Keywords: Grain refinement; Aaccumulative Roll Bonding; Mechanical Characterization; Aluminum Alloys
1. Introduction
Reducing the grain size of polycrystalline metallic mate-
rials is an economic way to improve the mechanical pro-
perties such as strength, and toughness [1,2]. The mi-
crostructure with stable fine grains is crucial in the su-
perplastic deformation of materials [3]. Since it is prac-
tically difficult to reduce the grain size of many metallic
materials such as aluminum alloys by a conventional co-
ld working and recrystallization process, several new
methods such as severe plastic deformation (SPD) are
developed to manufacture ultra-fine grain (UFG) materi-
als [2]. SPD processing techniques produce relatively
small quantities of materials, are very difficult to scale
up, and are unlikely to be able to produce at low cost.
The principle of a SPD process includes increasing the
dislocation density and transforming low angle grain
boundaries (LAGBs) into high angle grain boundaries
(HAGBs) [4].The net shape of the sample during proc-
essing is approximately constant, so that there is no
geometric limitation on the applied strain [5].
Among SPD processes, accumulative roll bonding
(ARB) can be performed by a conventional rolling mill
without any special die. Also, the productivity of this
process is relatively high because this process implies the
potential of industrial up-scaling to a continuous produc-
tion of UFGed metallic sheets or plates [6]. An ARB
process is practically limited by technological constraints
such as the occurrence of edge cracking. In this process,
the thickness of the sheet varies between fixed limits,
and very high strains can be accumulated in the material
by repeating the procedure; as a result, a significant
structural refinement can be achieved [6-8]. Saito [6]
used the ARB process at 200C to reach a microstructure
with ultra-fine grains for AA5083. Although this method
arised 551 MPa for the ultimate tensile strength, it con-
siderably reduced the value of ductility to 6%. From his
TEM results, it can be found that there is a high density
of dislocations in grains. Indeed, the true effect of the
grain size on the mechanical properties of this UFG ma-
terial was obscured by the presence of a high density of
dislocations. By applying a heat treatment at the tem-
perature of 180 C for a long time, the density of disloca-
tions decreases considerably [9]. Because of high
strength and low ductility of this alloy at low tempera-
tures, applying the ARB at high strain rates and low
temperatures has many difficulties such as edge cracking
and the necessity of a high power rolling mill. For thin
sheets with the thickness of 1 mm, a high strain can be
applied by using the hot ARB process at a high strain
rate to reach a fine microstructure. This method can be
helpful to scale up the materials with a desirable micro-
structure. Therefore, studying the effect of the hot ARB
on the microstructure and mechanical properties of
AA5083 is the prime objective of the present work.
2. Material and Experimental Procedure
The as-received material used in this study was an
AA5083 sheet of the chemical composition presented in
13
H. SHEIKH ET AL.
Table 1. Figure 1 shows the unique grain color map
recorded on the RD-ND cross-section of the initial sheet
with a mean grain size of 13µm. It can be observed that
the material exhibits an equi-axed polygonal type grain
structure. To achieve a good bonding, the surfaces of the
two sheets with dimensions of 100mm×30mm×1mm
were degreased (in acetone) and wire brushed (by a
stainless steel brush with wires of 0.4mm in diameter).
Plane strain rolling was conducted along the longest di-
mension. After surface treatment, the two sheets were
stacked to make a 2-mm thick specimen and then roll
bonded by 50% reduction in thickness (this corresponds
to a von Misses equivalent strain of 0.8 per pass). After-
wards, the rolled sheet was cut into two pieces along the
transverse direction which produced two sheets of the
length approximately corresponding to the initial length
before rolling. This procedure was repeated up to 6 cy-
cles so that the total equivalent strain of εvM = 4.8 was
achieved. Between the rolling passes, the sheets were
preheated to the temperature of 300C for 5 min and the
rolling was performed without any lubrication. The roll
diameter was 150mm which produced the mean strain
rate of 50 s1 during the rolling. The most severe prob-
lem in the roll bonding step was small edge cracks. In
this study, in order to avoid the propagation of small
edge cracks in the subsequent cycles, the edges of the
roll-bonded sheets were trimmed off.
To appear the interfaces located among the layers, the
RD-ND section of the sample processed by 6 cycles was
heated at the temperature of 125C for 12 h, and then it
was etched in a solution of 10% phosphoric acid in
ethanol. By doing so, Mg2Al3 precipitates locate at high
energy places such as interfaces and the etchant can cor-
rode the precipices; as a result, the interfaces can appear
sharply.
Transmission electron microscopy (TEM) micrograph
was also obtained from the sample processed by 6 cycles
of the ARB process. For doing so, thin foil was prepared
by twin-jet polishing from the mid-thickness plane of the
sample.
The Vickers macrohardness test was taken on the sur-
face using 30kg load. The mechanical properties of the
starting material and the ARBed samples were measured
by tensile tests at room temperature and the strain rate of
0.002 s1. The test specimens were prepared with the
tensile axis parallel to the rolling direction. The dimen-
sions of the specimens were chosen according to the
ASTM-E8 standard.
Table 1. Chemical composition of the material used in this
work (wt%).
Mg Mn Fe Si Cr Cu
4.5 0.71 0.33 0.19 0.058 0.037
3. Results and Discussion
Figure 2 shows the optical micrograph observed in the
RD-ND section of the sample produced by 6 ARB cycles;
the bonded interfaces must be observed across the thick-
ness of the sample. The arrow in Figure 2 shows the
center of thickness. No part of interfaces was not seen
without applying the heat treatment mentioned for etch-
ing. This indicates a good bonding with no delamination
between sheets at each cycle under the present ARB
conditions.
A TEM micrograph related to mid-thickness plane (RD-
TD) of 6 ARB cycles has been represented in Figure 3.
The TEM image shows a feature of (sub) grains with the
Figure 1. Microstructure of the starting material 3.
Figure 2. Optical micrograph of the longitudinal cross-se-
ction of AA5083 sheet after 6 ARB cycles.
Copyright © 2011 SciRes. OJMetal
14 H. SHEIKH ET AL.
size of 200 nm - 400 nm in which the density of disloca-
tions is extremely low. Because of small grain sizes of
severe plastic deformed materials, many previous inves-
tigations have relied on the TEM to study the grain
structures after processing. The selected area diffraction
(SAD) pattern consists of rings of diffraction spots sug-
gesting that the most grain boundaries are separated by
boundaries with high angles of misorientation. It must be
considered that many of the grain sizes quoted from TEM
analysis are underestimates as they include sub-grains [6].
Figure 4 shows the stress-strain curves and the corre-
sponding tensile properties of the ARBed processed
AA5083 sheets. Ultimate tensile strength increases up to
325MPa after two cycles, which is more than 1.25 times
as high in comparison to the starting material. The tensile
and the yield strengths (according to elongation of 0.2%)
increase during the following cycles, but with very
smaller increments for the last cycles. The yield strength
reaches 265MPa after 6 cycles, which is almost 1.7 times
as high as the annealed material. The shapes of the
stress-strain curves are similar in all samples and the
ARBed specimens reach their peak of strength followed
by a small value of necking strain. After 6 ARB cycles,
the elongation decreases to 11.4%, which means that the
material still shows a sufficient ductility. The combina-
tion of elongation, yield and ultimate tensile strength
points out that this material has desirable toughness for
special applications.
The variation of hardness shown in Figure 5 confirms
the results of the tensile tests. Rapid increase in the
hardness and the strength at the strain of up to 1.6 is
mainly due to the work hardening caused by an increase
in dislocation density and formation of sub-grains [10].
Figure 3. TEM micrograph of the sample processed by 6 cy
cles of ARB process.
Figure 4. Nominal stress-strain curves of the AA5083 sheets
processed at different ARB cycles.
Figure 5. Macrohardness variation of AA5083 with increa-
sing the number of ARB process.
During the following cycles, the dislocation density is
almost constant. The occurrence of static and dynamic
recoveries (as two flow softening mechanisms during the
hot rolling and reheating at the temperature of 300C)
causes that the generation and annihilation of disloca-
tions reach an equilibrium state [10]. The increase in
strength after cycle 2 can be attributed to the decrease of
grain size and an increase in the volume fraction of
HAGBs and misorientation. As a result, the strength of
the ARBed samples can be considered as the sum of dis-
location strengthening from LAGBs and grain size
strengthening from HAGBs.
By using the below empirical relationship, the cell size
which is the same value of the (sub) grain size can be
estimated [11]:

0
K
Gb
(1)
Copyright © 2011 SciRes. OJMetal
H. SHEIKH ET AL.
Copyright © 2011 SciRes. OJMetal
15
[1] P. B. Prangnell, J. R. Bowen and A. Gholinia, “The For-
mation of Submicron and Nanocrystalline Grain Structure
by Severe Deformation,” Proceedings of the 22nd Risø´
International Symposium on Materials Science, Roskilde,
2001, pp. 105-126.
Table 2. The values used for the present estimation and the
cell size predicted from Equations 1and 2.
G(GPa) (MPa)σ (MPa)τ (nm)δ K
27 260 104 5 360
[2] Y. S. Kim, S. H. Kang and D. H. Shin, “Effect of Rolling
Direction on the Microstructure and Mechanical Proper-
ties of Accumulative Roll-Bonding (ARB) Processed
Commercially Pure 1050 Aluminum Alloy,” Materials
Science Forum, Vol. 503-504, 2006, pp. 681-686.
doi:10.4028/www.scientific.net/MSF.503-504.681
0.9lnG
K



(2)
Where δ is the cell size, G is the shear modulus, b is the
burgers vector, τ is the resolved shear stress, and τ0 is the
friction stress which is usually neglected since 0
.
The value of τ is calculated by the following equation [11]:
[3] K. F. Zhang and H. H. Yan, “Deformation Behavior of
Fine-Grained 5083 Al Alloy at Elevated Temperature,”
Transactions of Nonferrous Metals Society of China,Vol.
19, 2009, pp. 307-311.
doi:10.1016/S1003-6326(10)60060-X
(3)
M
[4] J. Lee and H. Seok, “Microstructural Evolutions of the Al
strip Prepared by Cold Rolling and Continuous Equal
Channel Angular Pressing,” Acta Materialia, Vol. 50,
2002, pp. 4005-4019.
doi:10.1016/S1359-6454(02)00200-8
where M and σ are the Taylor factor ( the value of 2.5 is
used for the present work) and the flow stress, respec-
tively. The (sub) grain size is predicted from equation. 1
and 2. The values used for the present estimation are
listed in Table 2. According to these values for the sam-
ple processed by 6 cycles of the hot ARB process; the
predicted grain size is 360 nm which has a good agree-
ment with the TEM result. Also, it is possible to ap-
proximate the dislocation density retaining in the grain
interior. The flow stress is related to the dislocation den-
sity (ρ) by the following equation [11]:
[5] R. K. Islamgaliev, N. F. Yunusova, R. Z. Valiev, N. K.
Tsenev, V. N. Perevezentsev and T. G. Langdon, “Char-
acteristics of Superplasticity in an Ultrafine-Grained
Aluminum Alloy Processed by ECA Pressing,” Scripta
Materialia, Vol. 49, 2003, pp. 467-472.
doi:10.1016/S1359-6462(03)00291-4
[6] Y. Saito, H. Utsunomiya, N. Tsuji and T. Sakai, “Novel
Ultra-High Straining Process for Bulk Materials Devel-
opment of the Accumulative Roll-Bonding (ARB) Proc-
ess,” Acta Materialia, Vol. 47, 1999, pp. 579-583.
doi:10.1016/S1359-6454(98)00365-6
12
0Gb

 (4)
where 0
is strain insensitive and can be neglected as
described previously. The dislocation densities for the
present work and Saito’s study were estimated 0.72 ×
1015 and 3.23 × 1015 m
–2, respectively. These results
show that a large part of the obtained yield stress for
Saito’s work can be attributed to the high density of dis-
locations.
[7] N
. Tsuji, Y. Saito, S. H. Lee and Y. Minamino, “ARB
(Accumulative Roll-Bonding) and Other New Techniques
to Produce Bulk Ultrafine Grained Materials,” Advanced
Engineering Materials, Vol. 5, 2003, pp. 338-344.
doi:10.1002/adem.200310077
[8] X. Huang, N. Tsuji, N. Hansen and Y. Minamino, “Mi
Crostructural Evolution During Accumulative Roll-Bond-
ing of Commercial Purity Aluminum,” Materials Science
and Engineering: A, Vol. 340, 2003, pp. 265-271.
doi:10.1016/S0921-5093(02)00182-X
4. Conclusions
[9] K. T. Park, J. H. Park, Y. S. Lee and W. J. Nam, “Com-
parison of Compressive Deformation of Ultrafine-Grain-
ed 5083 Al Alloy at 77 and 298 K,” Metallurgical and
Materials Transactions: A, Vol. 36, 2005, pp. 1365-1368.
doi: 10.1007/s11661-005-0227-8
In this study, a hot ARB process was performed on an
AA5083 sheet up to 6 cycles and equivalent strain of 4.8
was successfully achieved without severe edge cracking.
The results confirmed that the process is effective for the
grain refinement of AA5083. [10] K. T. Park, H. J. Kwon and W. J. Kim, “Microstructural
Characteristics and Thermal Stability of Ultrafine
Grained 6061 Al Alloy Fabricated by Accumulative Roll
Bonding Process,” Materials Science and Engineering: A,
Vol. 316, 2001, pp. 145-152.
doi:10.1016/S0921-5093(01)01261-8
The mechanical properties revealed that the strength of
the sheet considerably increased by the first two ARB
cycles which is attributed to work hardening caused by
an increase in dislocation density and sub-grains. After 6
cycles, (sub) grains with a low internal dislocation den-
sity have a size with the range of 200 - 400 nm.
[11] K. T. Park, J. H. Park, Y. S. Lee and W. J. Nam, “Micro-
structures Developed by Compressive Deformation of
Coarse Grained and Ultrafine Grained 5083 Al Alloys at
77 K and 298 K,” Materials Science and Engineering: A,
Vol. 408, 2005, pp. 102-109.
doi:10.1016/j.msea.2005.07.040
5. References