You are on page 1of 9

Materials Science and Engineering A 504 (2009) 40–48

Contents lists available at ScienceDirect

Materials Science and Engineering A


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

Investigation of the effect of strain rate and temperature on the deformability


and microstructure evolution of AZ31 magnesium alloy
Ismael Abdel Maksoud a,∗ , H. Ahmed b , Johannes Rödel a
a
German University in Cairo, New Cairo City Main Entrance Al Tagamoa Al Khames, Cairo, Egypt
b
Arcelor Mittal - OCAS NV | Pres. J.F. Kennedylaan 3 9060 Zelzate, Belgium

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

Article history: Magnesium, being the lightest available structural material, is a primary candidate for automotive and
Received 11 June 2008 aviation industries; however, the difficulty of deforming magnesium is a limiting factor. This research work
Received in revised form 12 October 2008 aims to investigate the deformability of AZ31 magnesium alloy employing various deformation conditions.
Accepted 16 October 2008
The results are used to correlate the effect of changing the deformation conditions on the resulting material
behavior and microstructure. Tensile testing experiments were conducted at a wide variety of deformation
Keywords:
conditions to determine the constitutive behavior of AZ31 magnesium alloy. The effect of changing the
AZ31 magnesium alloy
deformation conditions on the resulting microstructure is also examined and quantified by measuring the
Strain rate
Temperature
volume fraction and size of the dynamically recrystallized grains. Hardness measurements were employed
Dynamic recrystallization to determine the softening behavior of the material during deformation due to dynamic recrystallization.
Constitutive behavior The findings indicate a decrease in the peak stress and hardness and an increase in the fraction and size
Microstructure of dynamically recrystallized grains with decreasing strain rate and increasing deforming temperature.
Temperature plays a major role in determining the fraction and size of dynamically recrystallized grain
whereby, at a certain strain rate, an increase in temperature by 50 ◦ C results in a change in the fraction of
dynamically recrystallized grains by ∼30%. Further correlations were developed based on these findings
to optimize the deformation behavior.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction complicated geometries of magnesium alloys but casting provides


low-mechanical strength due to the presence of gas porosities [5,6].
The commercial use of magnesium alloys is a growing trend Consequently, to avoid the disadvantages of casting, it is neces-
in automotive and aerospace industries [1–3]. This is attributed sary to improve the deformability of magnesium alloys by using
towards the low density of magnesium which makes it one of the secondary manufacturing processes such as rolling and extrusion
lightest available structure materials. The low density of magne- [7].
sium alloys is an important factor to decrease the fuel consumption In order to be able to carry out the secondary manufacturing
which in turn saves the natural energy resources and reduce gas processes in magnesium, it is necessary to understand the mate-
exhaust emissions [2–4]. In addition, the low density and con- rial behavior at different deformation conditions. This entails the
siderably high-yield strength of magnesium alloys are favorable need to develop the constitutive behavior that describes the mate-
properties in designing various utilities like casings for electronic rial behavior at various deformation conditions. There are two
devices [5,6]. approaches to determine the constitutive behavior relation of AZ31
In spite of the above-mentioned advantages and applications, magnesium alloy. The first is describing the behavior in terms of
the use of pure magnesium in industry is limited because of its low microstructure, deformation conditions and various materials con-
deformability by cold working, due to its hexagonal closed-packed stants related to the properties of the material [8–10]. The second
(HCP) crystal structure. Manufacturing of magnesium in open air approach describes the behavior in terms of solely the deformation
at high temperatures can be one of the limiting factors due to its conditions and various material constants [2,11–13]. Eq. (1), com-
high reactivity with oxygen making it highly flammable. Casting monly known as Arrhenius type relation, Eqs. (2) and (3), known
is the most dominant manufacturing process used in producing as hyperbolic sine relation, can be used to describe the constitutive
behavior of metals under high temperatures [2]:
 −Q 
∗ Corresponding author. 
ε̇ = A  n exp (1)
E-mail address: som3a.baggio@gmail.com (I.A. Maksoud). RT

0921-5093/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2008.10.033
I.A. Maksoud et al. / Materials Science and Engineering A 504 (2009) 40–48 41

 −Q 
2.1. Tensile testing
ε̇ = A exp(ˇ) exp (2)
RT
 −Q  The test conditions that were employed in this work are at tem-
ε̇ = A(sinh ˛)n exp (3) peratures ranging from 25 and 404 ◦ C and strain rates between
RT 1 × 10−4 and 1 × 10−2 s−1 . To ensure repeatability and accuracy,
each test condition was repeated at least three times. The tensile
where ε̇ is the strain rate, A , A , A, n , n, ˇ and ˛ are material
specimens were manufactured from extruded rods that according
constants,  is the applied stress, Q is the activation energy for
to the DIN 50125 standard. It is worth mentioning that the extruded
deformation, R is the molar gas constant and T is the temperature.
rods were provided by Magnesium Elektron Co., UK as in kind sup-
In AZ31 magnesium alloy, Yi et al. [14] reported a decrease in
port.
the ultimate tensile stress at a strain rate of 5.5 × 10−4 s−1 from
Once fracture occurred, the specimens were quenched immedi-
307 to 70 MPa as the temperature increased from 25 to 250 ◦ C,
ately, in less than 5 s, in a water bath to prevent any microstructure
respectively, while Jäger et al. [15] reported a stress reduction at
change (static recrystallization or grain growth) [7]. For precise
1.3 × 10−4 s−1 from 343 to 14 to 7 MPa as the temperature increased
temperature measurement, one sample was manufactured with a
from 25 to 350 to 400 ◦ C, respectively. The mechanical testing
hole in the middle and the specimen temperature was measured
experiments carried out in AZ31 magnesium alloys [2,5,7,8,11,16]
using a type K thermocouple placed in this hole. Relations were
show a decrease in the material strength and an increase in the
developed, subsequently, to correlate the measured furnace tem-
ductility at elevated temperatures and low-strain rates. Guo et al.
perature to the sample actual temperature and were subsequently
[2] had also studied the effect of deformation conditions on the
used in this research.
peak stress of AZ31 magnesium alloy using compression tests and
reported that dynamic recrystallization (DRX) is the responsible
2.2. Metallography examination
mechanism for decreasing the peak stress at high temperatures and
low-strain rates. Other research work had also emphasized the role
After the specimens were quenched in a water bath, a longitu-
of DRX, the decrease of critical resolved shear stress (CRSS) for non
dinal sample, away by 2 cm from the neck, was cut. The samples
basal slip systems, which in turn activate these systems, and dislo-
were mounted, ground, polished and etched (3 g malic acid, 95 ml
cations annihilation on the reduction of the mechanical strength at
distilled water and 2 ml nitric acid). Optical microscopy examina-
high temperatures and low-strain rates [5,14,17].
tion was carried out using Zeiss Axio Imager Polarizer Microscopy
Barnett [11] established an Arrhenius function to account for
where at least eight micrographs were captured for each sample.
the constitutive behavior of AZ31 magnesium alloy with strain rate
sensitivity and activation energy of 0.15 and 145 kJ mol−1 , respec-
2.3. Hardness measurements
tively. Recently, Beer et al. [17] examined the microstructure of
compressed AZ31 magnesium alloy, at a temperature of 350 ◦ C and
Hardness measurements were conducting using Buehler
a strain rate of 0.01 s−1 , by studying the percentage and size of DRX
Micromet 2104 machine, hardness tests were conducted to the
grains at a strain up to 100% for wrought and cast AZ31 magnesium
mounted samples applying a load of 100 g. Hardness tests were
alloy and reported that the wrought material achieve 100% fraction
employed at least five times for each sample and the average was
of DRX grains and approximately 11 ␮m DRX grain size as the strain
then calculated.
reaches 100%.
It is worth mentioning that the constants used to determine the
constitutive behavior relation, and the volume fraction and size of 3. Results
DRX grains based on tensile tests are rarely found in literature. The
quantification of the effect of deformation conditions, employing 3.1. Tensile testing
tensile testing conditions, on the resulting change in peak stress
and DRX grains has not been reported previously in literature. Fig. 1 shows the engineering stress strain curve of two extreme
This paper summarizes the findings of the research conducted to conditions (highest strain rate and lowest temperature versus low-
understand the workability and deformability of AZ31 magnesium
alloy. Extensive tensile experiments were carried out to under-
stand the constitutive behavior at various ranges of deformation
conditions, namely: temperature and stain rate and determine the
effect of changing these conditions on the resulting microstructure
and the material hardness. Based on these results, various relations
were developed to understand the effect of changing the deforma-
tion conditions on the deformability of AZ31 magnesium alloy.

2. Experimental work

Tensile testing was carried out using a Zwick/Roell Z100


machine to understand the constitutive behavior of AZ31 magne-
sium alloy with the following chemical composition (wt%): 3.0 Al,
1.0 Zn, 0.3 Mn, bal. Mg. The resulting microstructure was investi-
gated and examined on the quenched samples after fracture. Before
conducting tensile testing, the specimens were annealed in the
furnace for 8 h at 400 ◦ C followed by furnace cooling to ensure
the same initial microstructure. This step was carried out to help
in decreasing the residual stresses resulted during manufactur-
ing. Fig. 1. Engineering stress stain curves of two extreme deformation conditions.
42 I.A. Maksoud et al. / Materials Science and Engineering A 504 (2009) 40–48

Fig. 2. Effect of temperature on the peak stress of AZ31 magnesium alloy. Fig. 3. Effect of strain rate on the peak stress of AZ31 magnesium alloy.

of strain rate on the room temperature peak stress values can be


est strain rate and highest temperature) where the engineering
neglected since the values are almost identical at different strain
peak (maximum) stress is illustrated by two dashed circles. The
rates.
use of true peak stress, rather than yield stress, is more relevant
Table 1 shows the percentage reduction in peak stress as the
since it accounts for the true maximum stress that can be applied
temperature increases, where room temperature data are consid-
to the material during deformation. True peak stress is also the
ered as the reference points. For example, as shown in Table 1,
minimum stress required to continue deformations at high elonga-
as the temperature increases by 40 ◦ C (321 to 361 ◦ C), the peak
tions.
stress percentage decreases by 8% (82% to 74%) from the ref-
After converting the engineering peak stress to true peak stress
erence point, 295 MPa. This can be useful in developing useful
[18], the average true peak stress was calculated from at least three
correlations to predict the effect of changing the deformation
samples and the relationship between true peak stress and vari-
conditions on the change in peak stress as will be explained
ation in deformation conditions is established and can be seen in
later.
Figs. 2 and 3.
As can be seen in Fig. 2, the peak stress decreases with increasing
the temperature. This is due to the fact that increasing the temper- 3.2. Microstructure
ature leads to decreasing the CRSS for pyramidal and prismatic slip
systems and thereby an increase in the number of activated slip sys- The initial microstructure, after the annealing process, is shown
tems [2]. As the CRSS decreases and more slip systems are activated, in Fig. 4 where very few dynamically recrystallized grains could
there is an increase in the dislocations density moving on favorable be observed in addition to the deformation by twinning which can
slip systems resulting in higher probability of recovery. The phe- result during the manufacturing process of the specimens. The posi-
nomenon of reduction of peak stress with increasing temperature tion of DRX grains surrounding the grain boundaries as reported
and decreasing strain rate can also be attributed to the fact that by Mwembela [19] is due to the high-energy levels at the grain
the newly formed DRX grains result in removal of substructures boundaries forming a necklace structure.
[7]. From Figs. 5 and 6, it can be concluded that twinning plays a role
As can be seen from Fig. 3, the peak stresses decreases with in the deformation of AZ31 magnesium alloy at room temperature
decreasing strain rates. This was explained by Guo et al. [2] where and the number of observed twins increases with decreasing the
it was found that increasing strain rates, result in an increase in strain rate. However and as can be seen in Fig. 7, at high temper-
the dislocations density at grain boundaries resulting in high-stress atures, parallel twin lines could not be observed. This conclusion
concentration regions. This phenomenon could also be related to agrees also with previous research work indicating that twinning
dynamic recrystallization whereby there is more time for nucle- may play an important role at relatively low temperatures only
ation and growth events at low-strain rates. However, the effect [5,9,10,13,20].

Table 1
The relationship between temperature and percentage reduction in peak stress.

T (◦ C) ε̇ = 1 × 10−2 s−1 ε̇ = 1 × 10−3 s−1 ε̇ = 1 × 10−4 s−1

 peak Pct. reduction (%)  peak Pct. reduction (%)  peak Pct. reduction (%)

25 (Ref.) 295.6 Ref. 292.1 Ref. 287.8 Ref.


213 143.3 51.5 117.8 59.7 84.9 70.5
263 107.3 63.7 82.0 71.9 60.1 79.1
269 97.6 67.0 79.1 72.9 56.5 80.4
321 77.7 73.7 59.8 79.5 40.5 85.9
361 54.5 81.6 40.2 86.2 24.1 91.6
404 39.3 86.7 24.4 91.7 13.9 95.2
I.A. Maksoud et al. / Materials Science and Engineering A 504 (2009) 40–48 43

Fig. 4. Optical microstructure of AZ31 magnesium alloy before deformation. Fig. 6. Optical microstructure of AZ31 magnesium alloy at T = 25 ◦ C and ε̇ = 1 ×
10−3 s−1 .

A comparison between Figs. 7 and 8 shows the strain rate also 4. Discussions and analysis
affect the fraction of DRX grains. As the strain rate increases, the
fraction of DRX grains decreases. In Fig. 9, the DRX grains could 4.1. Constitutive behavior of AZ31 magnesium alloy
be clearly observed as fine and equiaxed areas. By comparing
Figs. 7 and 9, it can be seen that as the temperature increases, The need for constitutive equations is to be able to quantify
the fraction of DRX grains also increases. It should be noted that the AZ31 magnesium alloy behavior at different deformation con-
at 213 ◦ C, the effect of dynamic recrystallization could not be ditions, temperatures and strain rates, in order to optimize its
observed. deformability. Eqs. (1) and (3) are the main relations used to

Fig. 5. Optical microstructure of AZ31 magnesium alloy at T = 25 ◦ C and ε̇ = 1 × Fig. 7. Optical microstructure of AZ31 magnesium alloy at T = 263 ◦ C and ε̇ = 1 ×
10−2 s−1 . 10−2 s−1 .
44 I.A. Maksoud et al. / Materials Science and Engineering A 504 (2009) 40–48

Fig. 10. Relation between logarithms of peak stress and strain rate (Eq. (4)).

tal measurements. This was proved by Barnett [11] where Eq. (1)
can only be used for peak stress values between 20 and 90 MPa.
To emphasize Barnett [11] findings, both equations are discussed
showing their average percentage error.
Fig. 8. Optical microstructure of AZ31 magnesium alloy at T = 263 ◦ C and ε̇ = 1 ×
10−4 s−1 .
4.1.1. Arrhenius equation
determining the constitutive behavior of AZ31 magnesium alloy 4.1.1.1. Stress exponent (n ). After some mathematical manipula-
[2,11–13,19,21,22]. Both equations do not depend on microstructure tions of Eq. (1), Eq. (4) was derived:
parameters such as initial grain size making them more applicable
 
1 ∂ log 
in industrial fields since the maximum stress, required for defor- = (4)
n ∂ log ε̇
mation, is a function of deformation conditions (T, ε̇), material T

properties (Q, n) and one constant (A) only. By calculating the reciprocal of the slope of each linearly fitted
Although Eq. (1) can be easily applied in comparison to Eq. (3), line in Fig 10, the n value was calculated at each deformation condi-
the reported stress is not as accurate in comparison to experimen- tion. As can be seen from Fig. 10, the stress exponent is independent
at room temperature values so these data were excluded from the
sequence of obtaining n . The average n value was subsequently
calculated. As can be noted in Fig. 10, the slope of the linearly fit-
ted line changes with increasing temperature which could be an
indication of the change in the deformation mechanism [23].
The value of the stress exponent presented in this work is 7
which agrees with the findings of Barnett [11], which is approxi-
mately 7 but is higher than the findings of Huang et al. [12] which
is 5.3. However, it was reported that high-stress exponent values
could be due to the presence of fine oxides in the specimens [2].
Based on literature findings, it was found that a stress exponent
with the value of 7 is a climb controlled deformation mode [9,23].
This agrees with the findings of Ishikawa et al. [9] and also it is
common to have climb controlled deformation mode at elevated
temperatures [11].

4.1.1.2. Activation energy. After further mathematical manipula-


tions of Eq. (1), Eq. (5) was derived:
 
 ∂ ln 
Q =nR (5)
∂1/T
ε̇

By calculating the slope of each linearly fitted line in Fig. 11,


obtaining average values and then multiplying this average by n ,
obtained previously, and the molar gas constant, the Q value was
calculated.
As can be noted in Fig. 11, the activation energy increases with
Fig. 9. Optical microstructure of AZ31 magnesium alloy at T = 361 ◦ C and ε̇ = 1 ×
decreasing the strain rate. The value of the activation energy pre-
10−4 s−1 . sented in this work is 153 kJ mol−1 which nearly agrees with the
I.A. Maksoud et al. / Materials Science and Engineering A 504 (2009) 40–48 45

4.1.2. Hyperbolic sine equation


By replacing each , from the derived equation from Arrhenius
relation, by sinh(˛), where ˛ is assumed to be 0.052 MPa−1 based
on Mwembela et al. [19,21] and for the activation energy com-
parison with other literature [19]. The following equations can be
derived:
 
1 ∂ log(sin ˛)
= (8)
n ∂ log ε̇
T
 
∂ ln(sin ˛)
Q = nR (9)
∂1/T
ε̇

ln Z = n ln(sin ˛) + ln A (10)

The value of the stress exponent presented in this work is 2.3,


which lies slightly higher than the findings of Mwembela [19] which
is approximately 2 and slightly lower than the findings of Liu et al.
[13] which is 3. This slight difference could be attributed due to the
effect of the initial grain size [13].
The value of the activation energy presented in this work is
Fig. 11. Relation between natural logarithms of peak stress and temperature recip- 153 kJ mol−1 which nearly agrees with the findings of Barnett
rocal (Eq. (5)).
[11], Huang et al. [12] and Liu et al. [13] which is 147, 145 and
145 kJ mol−1 , respectively. It also agrees with the activation energy
findings of Barnett [11], Huang et al. [12] and Liu et al. [13] which obtained from the Arrhenius relation previously.
was 147, 145 and 145 kJ mol−1 , respectively.
4.1.3. Validation
4.1.1.3. Pre-exponential constant (A ). After mathematical manipu- The percentage error between measured and predicted peak
lations of Eq. (1), Eq. (6) was derived: stress is calculated using Eq. (11) and the total average error is
ln Z = n ln  + ln A (6) calculated in Eq. (12):

where Z is the Zener–Hollomon parameter combining the effect of Measured stress − Predicted stress
Error (%) = × 100 (11)
strain rate and temperature shown in Eq. (7): Measured stress
Q 
1 
18
Z = ε̇ exp (7) Average error (%) = |Error (%)| (12)
RT 18
i=1
By calculating the y-intercept of each linearly fitted line in
Fig. 12, obtaining average values and calculating the exponential Barnett [11] reported that the Arrhenius equation is suitable only
of the average, the A values were calculated. in the stress range between 20–90 MPa but in this work, it is suitable
The A values proved to vary dramatically with changing the for stresses between 24 and 107 MPa and within 263 and 361 ◦ C.
temperature. However, the Arrhenius equation exhibits further The average percentage error is equal to 12%. The hyperbolic sine
sensitivity to the exponent parameters (Q, T) rather than the pre- function shows an average error less than 7%.
exponential constant (A ).
4.2. Microstructure evolution

4.2.1. Volume fraction of DRX grains


To determine the volume fraction (Xv ) of DRX grains, the ASTM
standard number E562-02 was employed. From Figs. 13 and 14, it
can be seen that increasing the temperature and decreasing the
strain rate leads to an increase in Xv of dynamically recrystallized
grains. The increase in the fraction recrystallized is justified by the
softening of the material. At low-strain rates, there is more time
for nucleation and growth events, in comparison to higher strain
rates. By comparing the differences in gaps between the curves in
Figs. 13 and 14, it can be seen that the temperature plays an impor-
tant role than strain rate in affecting the volume fraction of DRX
grains.
Analogous to the quantifications done on percentage reduction
in the peak stress, Table 2 also summaries the percentage gain in
the volume fraction of DRX grains as the stain rate decreases.
A comparison of the fraction of dynamically recrystallized grains
data to reported values, shows that the values obtained in this work
are about one-third of the existing values [7,17]. This could be due to
the fast quenching technique used, which hinders any static recrys-
Fig. 12. Relation between natural logaithms of Zener–Hollomon parameter and tallization. From Table 2, it could be concluded that the percentage
peak stress to the power of stress exponent (Eq. (6)). gain in the volume fraction (Xv ) of DRX grains is mainly dependant
46 I.A. Maksoud et al. / Materials Science and Engineering A 504 (2009) 40–48

Fig. 13. Effect of temperature on the volume fraction of dynamically recrystallized Fig. 15. Effect of temperature on dynamically recrystallized grain diameter of AZ31
grains of AZ31 magnesium alloy. magnesium alloy.

Fig. 14. Effect of strain rate on the volume fraction of dynamically recrystallized Fig. 16. Effect of strain rate on dynamically recrystallized grain diameter of AZ31
grains of AZ31 magnesium alloy. magnesium alloy.

on temperature, while the strain rate does not affect the fraction of for grain growth at higher temperatures. At lower strain rates, there
DRX grains significantly. is more time for the nucleation and growth events. Several research
work found also the same phenomena [9,10,17]. By comparing the
4.2.2. DRX grain size differences between the curves in Figs. 15 and 16, it can be seen
To determine the DRX grain diameter, the ASTM standard E112- that the temperature plays an important role than the strain rate in
96 was used. From Figs. 15 and 16, it can be seen that increasing affecting the DRX grain size.
the temperature and decreasing the strain rate increases the size of Analogous to the quantifications done on the percentage reduc-
DRX grains. This was expected since there is a higher driving force tion in the peak stress and percentage gain in the DRX volume

Table 2
The relationship between temperature and percentage gain in volume fraction of dynamically recrystallized grains.

T (◦ C) ε̇ = 1 × 10−2 s−1 ε̇ = 1 × 10−3 s−1 ε̇ = 1 × 10−4 s−1

Fraction rex. Pct. gain (%) Fraction rex. Pct. gain (%) Fraction rex. Pct. gain (%)

213 (Ref.) 2.7 Ref. 5.3 Ref. 7.4 Ref.


269 7.3 167.5 9.4 77.9 12.2 63.8
321 13.2 385.8 16.2 205.4 18.8 153.7
361 22.2 717.9 24.2 357.5 26.8 260.1
404 31.3 1052.9 33.2 527.3 35.3 374.5
I.A. Maksoud et al. / Materials Science and Engineering A 504 (2009) 40–48 47

Table 3
The relationship between temperature and percentage gain in dynamically recrystallized grain size.

T (◦ C) ε̇ = 1 × 10−2 s−1 ε̇ = 1 × 10−3 s−1 ε̇ = 1 × 10−4 s−1

Drex (␮m) Pct. gain (%) Drex (␮m) Pct. gain (%) Drex (␮m) Pct. gain (%)

213 (Ref.) 1.15 Ref. 1.93 Ref. 2.64 Ref.


263 2.71 117.1 3.46 78.9 4.30 57.1
269 2.86 128.4 3.62 87.6 4.58 67.2
321 4.93 294.0 5.84 202.6 7.00 147.8
361 7.65 537.7 8.81 356.3 9.99 250.1
404 10.65 802.3 12.02 522.2 13.16 363.7

fraction, Table 3 also summaries the percentage gain in the DRX


grain size as the stain rate decreases, where the missing point
(T = 213 ◦ C and ε̇ = 0.01 s−1 ) is extrapolated, according to Fig. 16,
to be 1.15 ␮m.
Fig. 17 shows the effect of deformation conditions and the
corresponding Zener–Hollomon parameter (Z), on the DRX grain
diameter. Using an inverse power law to fit the data, the parame-
ters in Eq. (13) are calculated where A is 139.56 and n is 0.13. This is
close to the findings of Beer et al. [17] (A = 165.5 and n = 0.126) and
Guo et al. [2] (A = 109.4 and n = 0.105). As can be seen from Fig. 17, as
Z increases (strain rate increases and temperature decreases), the
DRX grain size decreases. From Fig. 17, it can be noted that the grain
size in this work is the smallest; this could be due to the very fast
quenching applied during the course of this work (less than 5 s),
which hinders grain growth and static recrystallization. This can
also be attributed to the difference in the initial grain size, where
Watanabe et al. [10] found that the smaller initial grain size leads
to finer dynamically recrystallized grain size:

DRXdiameter = AZ −n (13)
Fig. 18. Average hardness at various temperatures.
4.3. Hardness

It was found that the hardness measurements are not sensi- 4.4. Operation and correlation curves
tive to the changes in the strain rate. It was also noted that as the
temperature increases, the hardness of the material decreases as Several correlations between the percentage reduction of peak
shown in Fig. 18. The decrease in hardness decreases with increas- stress and percentage gain of fraction and size of recrystallized
ing temperature could be attributed to the high fraction of DRX grains can be seen in Figs. 19 and 20. It can be noted in Fig. 19
grains. For example and as can be seen from Fig. 18, as the temper- that at a strain rate of 1 × 10−3 s−1 , the percentage gain of the vol-
ature increased from 213 to 321 ◦ C, the hardness values decreased ume fraction and diameter of DRX grains is almost equivalent as
from to 85 to 81 Hv . the temperature changes. Fig. 20 is a summary of various tables
presented in this paper to quantify the effect of changing the defor-
mation parameters on the percentage change in the peak stress

Fig. 17. Effect of temperature on dynamically recrystallized grain diameter of AZ31 Fig. 19. Correlations between size and the fraction of dynamically recrystallized
magnesium alloy. grains at 1 × 10−3 s−1 .
48 I.A. Maksoud et al. / Materials Science and Engineering A 504 (2009) 40–48

In AZ31 magnesium alloy, it has been found that at high tem-


perature and low-strain rates the peak stress is reduced while
the volume fraction and size of dynamically recrystallized grains
increases. This phenomenon was attributed to the dynamic recrys-
tallization process which is agreed on by other research work.
However as shown during the course of this work, it has been found
that the temperature plays a major role, rather than strain rates, in
activating the dynamic recrystallization process during deforma-
tion.
Two constitutive equations were developed to quantify the
effect of changing the deformation conditions on the flow stress
in AZ31 magnesium alloy which is summarized in Eqs. (16) and
(17), with an average error of 12% and 7%, respectively:
  153090 0.142
= 37.361 × ε̇ exp (16)
RT
  152080 0.441
 = 19.231 sinh−1 5.961 × 10−9 ε̇ exp (17)
RT
Fig. 20. Correlations between the peak stress with size and the fraction of dynami-
cally recrystallized grains at 1 × 10−3 s−1 . where  is the peak stress (MPa), ε̇ is the strain rate (s−1 ), R is the
molar gas constant (J mol−1 K−1 ) and T is the deformation temper-
ature (K).
and corresponding change in the fraction of dynamically recrystal-
lized grains. For example, if the temperature increases from 213 to Acknowledgments
300 ◦ C, there is a reduction in the peak stress by about 42% and a
gain the in the Xv and size of DRX grains by about 150% where the The authors gratefully acknowledge Magnesium Elektron Co.,
data recorded at a deformation temperature of 213 ◦ C were used as UK for providing the material as an in-kind support to this research
the reference point to determine the percentage change. work. The authors wish to acknowledge the German university in
It is of crucial importance to develop mathematical relations, at Cairo for providing access to their workshop and experimental facil-
certain strain rate, to describe the effect of changing deformation ities. The discussions with Martyn Alderman and Tims Wilks at
temperature on the resulting peak stress and volume fraction and Magnesium Elektron are gratefully acknowledged.
size of DRX grains. These relations can be used in an industrial set-
ting to correlate the effect of the resulting change in the process References
parameters on the resulting change in peak stress and fraction of
[1] H.Z. Ye, X.Y. Liu, J. Alloys Compd. 402 (2005) 162–169.
dynamically recrystallized grains. Hence, Eqs. (14) and (15) were
[2] Q. Guo, H.G. Yan, H. Zhang, Z.H. Chen, Z.F. Wang, J. Mater. Sci. Technol. 21 (2005)
developed based on fitting the data shown in Fig. 20. The two rela- 1349–1354.
tions are good approximations of the percentage reduction in peak [3] F. Kaiser, J. Bohlen, D. Letzig, K.-U. Kainer, A. Styczynski, C. Hartig, Adv. Eng.
stress and percentage gain in fraction and size of DRX grains at Mater. 5 (2003) 891–896.
[4] A.A. Luo, JOM 54 (2002) 42–48.
1 × 10−3 s−1 , assuming the peak stress and fraction of dynamically [5] D.L. Yin, K.F. Zhang, G.F. Wang, W.B. Han, Mater. Sci. Eng. A 392 (2005) 320–325.
recrystallized grains at 213 ◦ C as the reference points and T is the [6] X. Cao, M. Jahazi, J.P. Immarigeon, W. Wallace, J. Mater. Process. Technol. 171
temperature in ◦ C: (2006) 188–204.
[7] J.C. Tan, M.J. Tan, Mater. Sci. Eng. A 339 (2003) 124–132.
% reduction in peak stress [8] H. Watanabe, H. Tsutsui, T. Mukai, M. Kohzu, S. Tanabe, K. Higashi, Int. J. Plast.
17 (2001) 387–397.
= −396 + 3.4T − 9.14 × 10−3 T 2 + 9 × 10−6 T 3 (14) [9] K. Ishikawa, H. Watanabe, T. Mukai, Mater. Sci. 40 (2005) 1577–1582.
[10] H. Watanabe, H. Tsutsui, T. Mukai, K. Ishikawa, Y. Okanda, M. Kohzu, K. Higashi,
Mater. Trans. JIM 42 (2001) 1200–1205.
[11] M.R. Barnett, J. Light Met. 1 (2001) 167–177.
% gain in Xv or size of DRX grains [12] G. Huang, L. Wang, G. Huang, F. Pan, Mater. Sci. Forum. 488–489 (2005) 215–218.
[13] Y. Liu, Doctor of Philosophy, Wayne State, 2003.
= 128 − 2T + 5.77 × 10−3 T 2 + 4 × 10−6 T 3 (15) [14] S.B. Yi, S. Zaefferer, H.-G. Brokmeier, Mater. Sci. Eng. A 424 (2006) 275–281.
[15] A. Jäger, P. Lukác, V. Gärtnerová, J. Bohlen, K.-U. Kainer, J. Alloys Compd. 378
(2004) 184–187.
[16] T. Imai, S. Dong, N. Saito, I. Shigemastu, Microstructure and mechanical prop-
erties of Mg–Al–Zn alloys processed by different-speeds-rolling, presented at
5. Summary and conclusions TMS, 2004.
[17] A.G. Beer, M.R. Barnett, Metall. Mater. Trans. A 38A (2007) 1856 (August).
[18] D. William, J. Callister, Material Science and Engineering: An Introduction, 6th
In this research work, the effect of change in deformation con- ed., John Wiley & Sons Inc., 2003.
ditions on the behavior and microstructure of AZ31 magnesium [19] A. Mwembela, 1997, Doctor of Philosophy, Concordia.
[20] A. Staroselskya, L. Anandb, Int. J. Plast. 19 (2003) 1843–1864.
alloy was studied by employing mechanical testing, including ten-
[21] A. Mwembela, E.B. Konopleva, H.J. McQueen, Scr. Mater. 37 (1997) 1789–1795.
sile and hardness testing. A further investigation was carried out to [22] M.M. Myshlyaev, H.J. McQueen, A. Mwembela, E. Konopleva, Mater. Sci. Eng. A
correlate the material behavior to the microstructure evolution by 337 (2002) 121–133.
examination the volume fraction and size of dynamically recrystal- [23] A. Takara, Y. Nishikawa, H. Watanabe, H. Somekawa, T. Mukai, K. Higashi, Mater.
Trans. JIM 45 (2004) 2377–2382.
lized grains.

You might also like