You are on page 1of 10

Materials Science & Engineering A 712 (2018) 704–713

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

Constitutive modeling of flow behavior and microstructure evolution of T


AA7075 in hot tensile deformation

Wenchao Xiaoa, Baoyu Wanga, , Yong Wub, Xiaoming Yanga
a
School of Mechanical Engineering, University of Science and Technology Beijing, Beijing 100083, China
b
School of Materials Science and Engineering, Harbin Institute of Technology, Harbin 150001, China

A R T I C L E I N F O A B S T R A C T

Keywords: Hot tensile deformation behavior of AA7075 was studied on a Gleeble-3500 thermal simulation machine. The
Constitutive model deformation temperatures were 300 °C, 350 °C, 400 °C, 450 °C, and strain rates were 0.01 s−1, 0.1 s−1, 1 s−1.
Hot tensile tests Electron backscatter diffraction (EBSD) technique was performed on the deformed specimens to investigate the
Flow behavior microstructure evolution. The results showed that grain sizes can be refined with increasing deformation
Microstructure evolution
amount, temperature, and decreasing strain rate. A constitutive model coupling with the evolution of re-
AA7075
crystallization, dislocation, grain size, and damage was established based on the experimental results. The
comparisons between model predictions and experimental results were evaluated in terms of statistical methods,
the minimum correlation coefficient value was 0.934, and the maximum average absolute relative error and root
mean square error were 3.96% and 2.97 MPa, respectively. The results indicated a good prediction accuracy of
the model in describing the flow behavior and microstructural evolution of AA7075.

1. Introduction 150 °C, and established a constitutive model with the consideration of
damage mechanics. Results showed that the model can precisely de-
AA7075, as a typical high strength 7XXX aluminum alloy, exhibits scribe the hot creep behaviors of 7075 aluminum alloy.
high tensile strength, which makes it potential to be used as structural Much work has been done on the study of microstructure evolution
components in a car body [1]. However, the low ductility of AA7075 at during hot deformation. Rokni et al. [8] proposed three dynamic re-
room temperature makes the material exhibit typical disadvantages in crystallization mechanisms (continuous, discontinuous, and geome-
manufacturing complex-shaped parts through conventional cold trical recrystallization) for 7075 aluminum alloy to describe the hot
stamping. Therefore, hot stamping process was proposed to improve the compression behavior under different thermal conditions. It was found
formability in recent years [2–4]. that the continuous recrystallization was the main dynamic re-
In hot stamping condition, blanks undergo a complex deformation crystallization mechanism for 7075 aluminum alloy during hot com-
process and microstructures change dynamically. Understanding the pression deformation. Liu et al. [9] conducted hot compression tests of
hot flow behavior and microstructure evolution of a material is im- AA7085 at temperatures from 250 °C to 450 °C and found that dynamic
portant to optimize the hot stamping process. Sun et al. [5] conducted recrystallization and dynamic recovery were co-responsible for the
hot compression tests for as-extruded 7075 aluminum alloy at tem- dynamic flow softening. Due to the high stacking fault energy, dynamic
peratures of 300–450 °C and strain rates of 0.01 s−1 – 10 s−1, the mi- recovery was more prone to occur before dynamic recrystallization
crostructure analysis showed that the main mechanism for softening of during hot deformation. Ma et al. [10] studied the damage evolution of
the flow behavior was dynamic recovery, and dynamic recrystallization 6111 aluminum alloy during hot tensile deformation, and found that
was the supplementary mechanism. Poletti et al. [6] studied the hot the micro-voids were first initiated inside the material. With the in-
compression behavior of AA6082 at temperatures from 300 °C to crease of deformation, the micro-voids continue to grow up to a critical
550 °C, and used the flow data to calculate a set of phenomenological value, then the specimen will be pulled to fracture. Ying et al. [11]
constitutive equations. A high value of stress exponent n was obtained observed the damage evolution of 7075-T6 during warm deformation,
in the equations, which was attributed to the interaction of particles and found that the damage evolution within the tensile material cor-
with dislocations during hot deformation. Li et al. [7] studied the hot responded to the process of void nucleation, growth, coalescence, and
creep behaviors of 7075 aluminum alloy at temperatures from 120 °C to subsequent fracture.


Corresponding author.
E-mail address: bywang@ustb.edu.cn (B. Wang).

https://doi.org/10.1016/j.msea.2017.12.028
Received 18 July 2017; Received in revised form 4 December 2017; Accepted 8 December 2017
Available online 11 December 2017
0921-5093/ © 2017 Elsevier B.V. All rights reserved.
W. Xiao et al. Materials Science & Engineering A 712 (2018) 704–713

Since great efforts have been made in computer simulating material


deformation in recent years, a model that can correctly describe the
flow behavior and microstructure evolution of a material is crucial to
obtaining accurate simulation. Many researchers have established
plenty of phenomenological equations for different materials. Lin et al.
[12] proposed a phenomenological model to describe the flow behavior
of 7075 aluminum alloy under hot compression tests; in this model,
flow stress is mainly affected by strain, strain rate, and deformation
temperature. Shi et al. [13] established an Arrhenius model to char-
acterize the hot compression behavior of 6005A aluminum alloy at the
temperatures from 300 °C to 500 °C. Zhang et al. [14] tested the dy-
namic tensile behaviors of 7075 aluminum alloy at high strain rates and
established a modified Johnson–Cook model to reflect the hardening
effect of strain and strain rate. Although materials’ flow behaviors can
be correctly described by phenomenological models, the interaction
between flow behavior and microstructure evolution is not considered
in these models. Hence, physically based constitutive models were
proposed in recent years to capture the microstructure evolution during
deformation.
Much work has been done in constitutive modeling for different
Fig. 1. (a) Dimensions of the hot tensile specimen (unit: mm) and (b) process route for the
materials in the analyses of hot deformation behavior and micro-
hot tensile tests.
structure evolution. Lin et al. [15] developed a constitutive model
based on dislocation evolution to simulate a C-Mn steel in multi-pass
hot rolling, and results showed that the model was suitable for pre- simulation machine to characterize the flow behavior of AA7075.
dicting microstructure evolution during and after hot deformation. Wu Fig. 1a shows the dimensions of the specimen for thermal simulation,
et al. [16] conducted hot compression tests for 7050 aluminum alloy and Fig. 1b shows the process route. The tensile direction was parallel
and proposed a set of constitutive equations, which described the hot to rolling direction, and the specimens were deformed at different
flow behavior as a function of temperature, work hardening, and dy- temperatures of 300 °C − 450 °C and different strain rates of 0.01 s−1
namic recrystallization. Mohamed et al. [17] considered the dislocation − 1 s−1.
density and damage evolution in a visco-plastic constitutive model for EBSD was performed on the deformed specimens to observe the
AA6082, results showed that the model predicted the flow stress at high microstructures. Samples were taken from the deformation zone in the
temperature well. Ji et al. [18] established a constitutive model based middle of the specimen, as shown in Fig. 2. Samples were mechanically
on the evolution of microstructures, such as dislocation, dynamic re- polished and then electrochemical polished in 10% perchloric acid in
crystallization, and grain size, to characterize the flow behavior of alcohol solution at voltage of 20 V for 50 s at room temperature. Ob-
5Cr21Mn9Ni4N steel during compression deformation. Furthermore, servations were carried out on a Zeiss ULTRA 55-type field emission
Ma et al. [10], Zheng et al. [19], Yang et al. [20] used similar models to scanning electron microscopy operated at 20 kV. HKL Channel 5 soft-
characterize different materials under different conditions. Results in ware was used to characterize the microstructures and provide mis-
these studies showed that the experimental stress and microstructure orientation information.
evolution were fitted well by the physically based constitutive models,
demonstrating that the models were capable to accurately predict the 2.2. Experimental results
flow behavior and microstructure evolution during deformation.
The above discussion leads to the conclusion that a physically based The true stress-strain data were calculated using Eqs. (1) and (2).
constitutive model for AA7075 in hot tensile deformation needs to be
l l + Δl
developed. In this study, hot tensile tests for AA7075 were conducted, ε = ln = ln( 0 )
EBSD was performed on the deformed specimens to observe the mi- l0 l0 (1)
crostructure evolution. Then, a constitutive model coupled with the
F Fl F (l 0 + Δl)
evolution of microstructures will be proposed. Finally, the validated σ= = =
A A0 l 0 A0 l 0 (2)
constitutive model will be used to analyze the flow behavior and mi-
crostructure evolution of AA7075. Where ε and σ represent true strain and true stress respectively, l and A
are the instantaneous length and cross-sectional area of uniform tem-
2. Experimental details perature zone, and subscript of “0 ” means their initial values. F is the
tensile force and Δl is the clamp displacement obtained from mea-
2.1. Uniaxial hot tensile tests surement. A correction on the true stress calculation to eliminate the
influence of necking was conducted according to the method proposed
The material was obtained from the Southwest Aluminum (Group) by Li et al. [21].
Co., Ltd. The thickness of AA7075 is 2 mm, Table 1 shows its chemical Fig. 3 shows the true stress-strain data of AA7075 at different
composition.
Uniaxial hot tensile tests were conducted on a Gleeble-3500 thermal

Table 1
Chemical composition of the AA7075 provided by the Southwest Aluminum (Group) Co.,
Ltd.

Element Si Fe Cu Mn Mg Cr Zn Ti Al

Wt% 0.07 0.22 1.4 0.04 2.2 0.19 5.4 0.02 Balance
Fig. 2. EBSD sampling location in the deformed specimen (unit: mm).

705
W. Xiao et al. Materials Science & Engineering A 712 (2018) 704–713

Fig. 3. True stress-strain curves of AA7075 at (a) different temperatures and (b) different strain rates.

deformation temperatures and strain rates. The flow stress increases


with increasing strain rate and decreasing temperature. During de-
formation at strain rate 0.1 s−1, flow stress increases slightly at a lower
temperature such as 300 °C, while it decreases slightly at a higher
temperature such as 450 °C. The reason is that the effect of strain
hardening is more obvious at a lower temperature [22]. The ductility
increases with increasing deformation temperature from 300 °C to
400 °C, however, there is a significant decrease of ductility for AA7075
when deformation is conducted at 450 °C. In Fig. 3b, it can be found
that the effect of strain hardening is more obvious at a higher strain
rate, and ductility increases with increasing strain rate.
Fig. 4 shows the effect of strains on the microstructures of AA7075
at temperature 400 °C and strain rate 0.1 s−1. Since the material was
first heat treated at 480 °C for 600 s, grain growth was induced by the
thermal effect. At the strain 0, average grain size was 28.1 µm. With the
increase of deformation, the accumulation of dislocations increases, and
the generation of low-angle grain boundaries is promoted. At the strain
0.3, grain refinement was observed in Fig. 4b, average grain size was
21.6 µm. Grains will be further refined with the increase of plastic
strain, the average grain size was 17.8 µm at the strain 0.5. The number
of newly generated grains was small under large plastic strain. From the
above analysis, it can be found that dynamic recrystallization is prone
to occur at a larger strain.
Fig. 5 and Fig. 4c show the effect of temperatures on the micro-
structures of AA7075 at strain 0.5 and strain rate 0.1 s−1. At a lower
temperature such as 300 °C, recrystallization mainly occurs in the grain
boundaries, and dynamic recovery is the main mechanism to annihilate
dislocations. With the increase of temperature, the generation of new
grains was more obvious. The number of recrystallized grains increased
with increasing temperature, indicating that recrystallization is prone
to occur at a higher temperature. Finer grain size was observed at a
higher deformation temperature, and average grain size at different
temperatures were 23.0 µm − 300 °C, 20.1 µm − 350 °C, 17.8 µm −
400 °C, and 15.3 µm − 450 °C.
Fig. 6 and Fig. 4c show the effect of strain rates on the micro-
structures of AA7075 at strain 0.5 and deformation temperature 400 °C.
It is shown that the recrystallization volume fractions at strain rates of
0.01 s−1 and 0.1 s−1 were obviously higher that that at strain rate of
1 s−1. Average grain size at different strain rates were 17.3 µm −
0.01 s−1, 17.8 µm − 0.1 s−1, and 21.3 µm − 1 s−1.
Fig. 7 shows the grain boundary distribution of AA7075. The black Fig. 4. Microstructures of AA7075 at different strains (temperature 400 °C and strain rate
line describes a misorientation angle higher than 15°, indicating high- 0.1 s−1): (a) Strain 0; (b) Strain 0.3; (c) Strain 0.5.
angle grain boundaries (HAGBs), while blue line, green line, and red
line describe angles of 10–15°, 5–10°, and 2–5°, respectively, indicating
[23], and some LAGBs may be transformed into HAGBs due to static
low-angle grain boundaries (LAGBs). At strain 0 (Fig. 7a), grain
recrystallization [24] and dynamic recrystallization. At strain 0.3
boundaries were mainly HAGBs due to the solution heat treatment, and
(Fig. 7b), the original grain was divided into several sub-grains by the
the average misorientation angle was 34.3°. With the increase of strain,
sub-grain boundaries, and equiaxed grains can be found. The fraction of
a plenty of LAGBs will be formed in the material due to deformation
LAGBs increased, and the average misorientation angle was decreased

706
W. Xiao et al. Materials Science & Engineering A 712 (2018) 704–713

Fig. 6. Microstructures of AA7075 at different strain rates (temperature 400 °C and strain
0.5): (a) Strain rate 0.01 s−1; (b) Strain rate 1 s−1.

plastic strain, respectively.


During high temperature deformation, the overall flow stress can be
decomposed into three components: initial yield stress, viscoplastic
stress, and hardening stress [25]:
σ = σ0 + σv + H (4)
where σ0 is the initial yield stress, σv is the viscoplastic stress, H is the
isotropic hardening stress.
Flow rule of the material is described as:
σ − σ0 − H n
εṗ = ( )
Fig. 5. Microstructures of AA7075 at different temperatures (strain 0.5 and strain rate K (5)
0.1 s−1): (a) 300 °C; (b) 350 °C; (c) 450 °C. where εṗ is the plastic strain rate, K is a temperature dependent para-
meter, n is the viscoplastic exponent.
to 29.2°. At strain 0.5 (Fig. 7c), the average misorientation angle was The isotropic hardening stress H is related to dislocation density,
32.9°. The reason for the increase in average misorientation angle from and can be described as:
strain 0.3 to strain 0.5 is that, the annihilation effect of recrystallization
H = Bρ 0.5 (6)
on the LAGBs was greater than the formation effect of deformation.
Fig. 7c and d compare the grain boundary distribution at 400 °C and where B is a temperature dependent parameter, ρ is the normalized
350 °C. The fraction of LAGBs at 350 °C was obviously higher than that dislocation density, which is formulated as ρ = 1 − ρ0 / ρ [26,27], ρ0 is
at 400 °C, the reason is that recrystallization fraction is higher at the initial dislocation density, and ρ is the dislocation density after
400 °C; more LAGBs were transformed into HAGBs. Fig. 7c and e deformation. The value of ρ ranges from 0 to 1.
compare the grain boundary distribution of strain rate 0.1 s−1 and
strain rate 0.01 s−1. It can be found that the fraction of HAGBs at 3.1. Dislocation density evolution
0.01 s−1 was higher than that at 0.1 s−1, and average misorientation
angle at 0.01 s−1 was 34.0°, which is slightly higher than that at The dislocation density evolution is caused by three mechanisms:
0.1 s−1. the first is the creation and accumulation of dislocations caused by
plastic deformation, the second is the annihilation caused by static and
3. Constitutive modeling dynamic recovery, and the third is the annihilation caused by micro-
structure evolution such as recrystallization.
According to Hooke's law, the flow stress of a material is calculated Yang et al. [20] described the evolution of dislocation density as:
by [25]:
ρ ̇ = A (1 − ρ ) εṗ − K1 ρ δ1 (7)
σ = E (εT − εp) (3)
where A , K1, δ1 are material constants. The equation proposed by Yang
where σ is the overall flow stress, E is Young's modulus, which is a described the evolution of dislocation density due to plastic deforma-
temperature dependent parameter. εT and εp represent total strain and tion and recovery. However, the effect of microstructure on dislocations

707
W. Xiao et al. Materials Science & Engineering A 712 (2018) 704–713

Fig. 7. Grain boundary distribution of AA7075 at different conditions: (a) Strain 0 (400 °C and 0.1 s−1); (b) Strain 0.3 (400 °C and 0.1 s−1); (c) Strain 0.5 (400 °C and 0.1 s−1); (d) Strain
0.5 (350 °C and 0.1 s−1); (e) Strain 0.5 (400 °C and 0.01 s−1).

was not considered. occurs [28], and can be written as a function of strain rate [29]:
Under hot deformation, recrystallization is related to the dislocation
q
density. The evolution rate of recrystallization can be modeled as [25]: ρc = q3 εṗ 4 (10)
q
S ̇ = q1 [xρ − ρc (1 − S )](1 − S )q2 εṗ 6 (8) where q3 and q4 are material constants.
Considering the annihilation of dislocations caused by re-
where S is the recrystallized volume fraction, q1 and q2 are material
crystallization, the evolution rate of dislocation density can be de-
constants.
scribed as:
The onset of recrystallization requires an incubation period, which
is given as:
ρ ̇ = A (1 − ρ ) εṗ − K1 ρ δ1 − [K2 ρ /(1 − S )] S ̇ (11)
x ̇ = q5 (1 − x ) ρ (9)
where K2 is a material constant.
where x is the incubation fraction, q5 is a material constant.
ρc is a critical value of dislocation density when recrystallization

708
W. Xiao et al. Materials Science & Engineering A 712 (2018) 704–713

Fig. 7. (continued)

Table 2 d2̇ = ω2 εṗ d−γ2 (13)


Temperature dependent material parameters and their corresponding formulation.
where ω2 and γ2 are material constants.
σ0 = σ0,0 exp(Qσ 0/ RT ) q6 = q6,0 exp(Qq6/ RT ) γ3 = γ3,0 exp(Q γ3/ RT ) For high strain rates, the effect of static grain growth is less im-
portant compared with recrystallization induced grain growth. During
K = K 0 exp(QK / RT ) A = A0 exp(Q A/ RT ) γ4 = γ4,0 exp(Q γ4/ RT )
hot deformation, recrystallization eliminates dislocations and creates
n = n 0 exp(Qn/ RT ) K1 = K1,0 exp(Q k1/ RT ) γ5 = γ5,0 exp(Q γ5/ RT )
new grains, which may result in a decrease in the average grain size.
B = B0 exp(QB / RT ) K2 = K2,0 exp(Q k2/ RT ) η1 = η1,0 exp(Q η1/ RT )
The effect of recrystallization on grain size evolution can be written as:
q1 = q1,0 exp(−Qq1/ RT ) ω1 = ω1,0 exp(Qω1/ RT ) η2 = Cη2 + η2,0 exp(−Q η2/ RT )
γ
q2 = q2,0 exp(Qq2/ RT ) ω2 = ω2,0 exp(Qω2/ RT ) η3 = Cη3 + η3,0 exp(Q η3/ RT ) d3̇ = −ω3 S ̇ 3 d γ4 (14)
q3 = q3,0 exp(Qq3/ RT ) ω3 = ω3,0 exp(Qω3/ RT ) η4 = η4,0 exp(Q η4/ RT )
where ω3 , γ3 , and γ4 are material constants.
q4 = q4,0 exp(Qq4/ RT ) γ1 = γ1,0 exp(Q γ1/ RT ) E = E0 exp(QE / RT )
Combining Eqs. (12)–(14), the grain size evolution can be modeled
q5 = q5,0 exp(Qq5/ RT ) γ2 = γ2,0 exp(Q γ2/ RT )
as:

d ̇ = ω1 d −γ1 + ω2 εṗ d −γ2 − ω3 S ̇ 3 d γ4


γ
(15)
3.2. Grain size evolution
where ω1 is a material constant, representing the combined effect of Mb
During hot deformation, grain size evolution significantly influences and σsurf . Normalized grain size d = d/ d 0 was used to simplify the de-
the flow rule of the material [25]. Then, its effects need to be coupled in termination process of material constants. d 0 is the initial grain size
the constitutive model. after solution heat treatment, d is the grain size after deformation.
Static grain growth is related to grain boundary, which is influenced Considering the effect of grain size on flow stress, flow rule of the
by the thermal effect. The static grain growth rate is given as [30]: material (Eq. (5)) is modified as:
σ − σ0 − H n −μ
d1̇ = Mb σsurf d−γ1 εṗ = ( ) (d )
(12) K (16)
where μ is a material constant.
where Mb is the grain boundary mobility, σsurf is the grain boundary
energy per unit area, γ1 is a material constant, d is the average grain
size. 3.3. Damage evolution
Deformation induced grain growth is the dynamic grain growth,
which is modeled on a grain-boundary diffusion mechanism [31]. The The damage evolution is influenced by the damage nucleation,
dynamic grain growth is independent of static grain growth, and it can growth and coalescence, and healing [32]:
be written as: D = DN + DG + DH (17)

709
W. Xiao et al. Materials Science & Engineering A 712 (2018) 704–713

Table 3 Damage healing may occur during hot deformation. For example,
Determined values of material constants for the constitutive model. when a recrystallization front passes over a micro-void, damage may be
eliminated completely. However, the healing effects on damage are
Material constant Optimal value Material constant Optimal value
limited [34].
σ0,0 (MPa) 3.806E−01 Qω1(J/mol) 1.871E+03
̇ = −η4 D
DH (20)
Qσ 0 (J/mol) 1.766E+04 ω2,0 (μm) 2.179E−05
K 0 (MPa) 1.413E+00 Qω2 (J/mol) 2.570E+02 where η4 is a material constant.
QK (J/mol) 1.999E+04 ω3,0 (μm) 2.086E+00 Then, the damage evolution can be formulated as:
n 0 (dimensionless) 8.563E−02 Qω3 (J/mol) 6.070E+02
Qn (J/mol) 2.262E+04 γ1,0 (dimensionless) 2.658E+00 Ḋ = η1 (1 − D) εpḋ 1 + η2 d −γ5 exp(η3 εp) εpḋ 2 − η4 D (21)
μ (dimensionless) 5.010E−01 Q γ1(J/mol) 8.120E+02
B0 (MPa) 4.863E+00 γ2,0 (dimensionless) 2.060E+00 Considering the reduction in cross section area caused by damage,
QB (J/mol) 1.381E+04 Q γ2 (J/mol) 7.595E+01 the relationship between effective flow stress and plastic strain need to
q1,0 (dimensionless) 3.579E+00 γ3,0 (dimensionless) 9.252E−01 be modified as:
Qq1(J/mol) 1.203E+04 Q γ3 (J/mol) 9.290E+02
σ /(1 − D) − H − σ0 ⎞n −μ
q2,0 (dimensionless) 4.820E−01 γ4,0 (dimensionless) 1.460E−02 εṗ = ⎛ (d )
⎝ K ⎠ (22)
Qq2 (J/mol) 4.403E+03 Q γ4 (J/mol) 1.187E+04
q3,0 (dimensionless) 6.017E−06 γ5,0 (dimensionless) 2.736E−02 Furthermore, flow stress of a material need to be calculated by:
Qq3 (J/mol) 3.621E+04 Q γ5 (J/mol) 6.727E+03
σ = E (1 − D)(εT − εp) (23)
q4,0 (dimensionless) 1.302E−01 η1,0 (dimensionless) 2.460E−02
Qq4 (J/mol) 1.386E+03 Q η1 (J/mol) 9.270E+02 The damage parameter is isotropic [35], and it is assumed that the
q5,0 (dimensionless) 2.120E+01 Cη2 (dimensionless) 1.218E−02 damage factor is 0 before deformation, it accumulates with the increase
Qq5 (J/mol) 3.730E+03 η2,0 (dimensionless) 8.549E+10
of deformation. When the material is failed, the damage factor D is
q6,0 (dimensionless) 5.341E−02 Q η2 (J/mol) 1.706E+05
considered as 1.
Qq6 (J/mol) 5.446E+03 Cη3 (dimensionless) 2.662E+00
A0 (dimensionless) 2.197E−02 η3,0 (dimensionless) 7.321E−01
3.4. Determination of material constants
Q A (J/mol) 2.745E+04 Q η3 (J/mol) 1.240E+04
K1,0 (dimensionless) 2.821E−02 η4,0 (dimensionless) 9.249E−04
During the whole deformation process, the constitutive equations
Q k1(J/mol) 2.531E+04 Q η4 (J/mol) 4.069E+03
K2,0 (dimensionless) 1.683E−02 d1(dimensionless) 9.146E−01
are highly coupled together. Temperature dependent material para-
Q k2 (J/mol) 2.185E+04 d2 (dimensionless) 9.161E−01 meters are listed in Table 2.
δ1(dimensionless) 1.430E+00 E0 (MPa) 1.115E+03 In Table 2, R is the gas constant, with a value of 8.314 J/(mol K). T
ω1,0 (μm) 1.113E−07 QE (J/mol) 7.316E+03 is the absolute temperature. The symbol Q with a subscript denotes the
activation energy for the corresponding material constants.
Material constants of the constitutive model were calibrated using
where D is the damage parameter, DN represents the damage nuclea- the true stress-strain curves and the grain sizes evolution. A genetic
tion, DG represents the damage growth and coalescence, DH represents algorithm toolbox in software MATLAB was used, and details of the
the damage healing. optimization process have been reported by Cao and Bai [36,37]. The
Damage growth may occur due to diffusion of vacancies into the determined values of material constants are listed in Table 3.
cavity, or the coalescence of voids [33]. The damage growth rate in-
creases with increasing strain rate [20], and can be formulated as: 4. Results and discussions
̇
Dgrowth = η1 (1 − D) εpḋ 1 (18) Fig. 8 compared the true stress-strain data between computed and
where η1 and d1 are material constants. experimental results. Most of the experimental data points lie close to
The damage nucleation rate increases with increasing strain rate the computed results, indicating a good prediction of the constitutive
and deformation amount, it is given by [17]: model.
Predictability of the proposed constitutive model can be evaluated
̇
Dnucleation = η2 d −γ5 exp(η3 εp) εpḋ 2 (19) by statistical methods such as correlation coefficient (R ), average ab-
solute relative error ( AARE ), and root mean square error (RMSE )
where η2 , η3 , γ5 , and d2 are material constants. [38,39]. They are expressed as:

Fig. 8. Comparisons of the experimental (symbols) and computed (solid curves) stress-strain data at (a) different temperatures and (b) different strain rates.

710
W. Xiao et al. Materials Science & Engineering A 712 (2018) 704–713

Fig. 9. Correlation between the experimental and computed stress at (a) different temperatures and (b) different strain rates.

Fig. 10. Microstructure evolution of AA7075: (a) Normalized grain size; (b) Dislocation density; (c) Damage factor.

711
W. Xiao et al. Materials Science & Engineering A 712 (2018) 704–713

N
∑i = 1 (Ei − E )(Pi − P ) References
R=
N N
∑i = 1 (Ei − E )2 ∑i = 1 (Pi − P )2 (24) [1] J. Hirsch, Recent development in aluminium for automotive applications, T.
Nonferr. Metal. Soc. 24 (2014) 1995–2002.
[2] R.P. Garrett, J. Lin, T.A. Dean, An investigation of the effects of solution heat
N
1 Ei − Pi treatment on mechanical properties for AA 6xxx alloys: experimentation and
AARE =
N
∑ Ei
× 100% modelling, Int. J. Plast. 21 (2005) 1640–1657.
i=1 (25) [3] O.E. Fakir, L. Wang, D. Balint, J.P. Dear, J. Lin, T.A. Dean, Numerical study of the
solution heat treatment, forming, and in-die quenching (HFQ) process on AA5754,
Int. J. Mach. Tool. Manuf. 87 (2014) 39–48.
N
1 [4] X. Fan, Z. He, S. Yuan, K. Zheng, Experimental investigation on hot for-
RMSE =
N
∑ (Ei − Pi)2 ming–quenching integrated process of 6A02 aluminum alloy sheet, Mater. Sci. Eng.
i=1 (26) A 573 (2013) 154–160.
[5] Z.C. Sun, L.S. Zheng, H. Yang, Softening mechanism and microstructure evolution
where Ei represents the experimental stress and Pi represents the of as-extruded 7075 aluminum alloy during hot deformation, Mater. Charact. 90
(2014) 71–80.
computed stress predicted by the proposed constitutive model. E and P [6] C. Poletti, T. Wójcik, C. Sommitsch, Hot deformation of AA6082 containing fine
are the mean values. N is the total number of the data points. Fig. 9 intermetallic particles, Metall. Mater. Trans. A 44 (2013) 1577–1586.
shows the correlation between experimental and computed flow [7] L.T. Li, Y.C. Lin, H.M. Zhou, Y.Q. Jiang, Modeling the high-temperature creep be-
haviors of 7075 and 2124 aluminum alloys by continuum damage mechanics
stresses for different temperatures. The minimum R value is 0.934, and model, Comp. Mater. Sci. 73 (2013) 72–78.
the maximum statistical parameters AARE and RMSE are 3.96% and [8] M.R. Rokni, A. Zarei-Hanzaki, A.A. Roostaei, H.R. Abedi, An investigation into the
2.97 MPa, respectively. The correlation analysis indicates a good pre- hot deformation characteristics of 7075 aluminum alloy, Mater. Des. 32 (2011)
2339–2344.
diction accuracy of the model. [9] W. Liu, H. Zhao, D. Li, Z. Zhang, G. Huang, Q. Liu, Hot deformation behavior of
Fig. 10 shows the microstructure evolution of AA7075 at different AA7085 aluminum alloy during isothermal compression at elevated temperature,
temperatures and strain rates. Fig. 10a shows the computed and ex- Mater. Sci. Eng. A 596 (2015) 176–182.
[10] W. Ma, B. Wang, J. Bian, X. Tang, L. Yang, Y. Huo, A new damage constitutive
perimental normalized grain size evolution. It is shown that the com-
model for thermal deformation of AA6111 sheet, Metall. Mater. Trans. A 46 (2015)
puted results fitted the experimental data points well. The grain size 2748–2757.
was refined with the occurrence of recrystallization after a period of [11] L. Ying, W.Q. Liu, D.T. Wang, P. Hu, Q. Wang, Experimental and simulation of
incubation, and the strain that triggers dynamic recrystallization in- damage evolution behavior for 7075-T6 aluminum alloy in warm forming, Chin. J.
Nonferrous Met. 26 (2016) 1383–1390.
creases with increasing strain rate. Fig. 10b shows the computed nor- [12] Y.C. Lin, L.T. Li, Y.X. Fu, Y.Q. Jiang, Hot compressive deformation behavior of 7075
malized dislocation density evolution. The dislocation density increases Al alloy under elevated temperature, J. Mater. Sci. 47 (2012) 1306–1318.
with increasing strain rate and decreasing temperature. Under the effect [13] L. Shi, H. Yang, L.G. Guo, J. Zhang, Constitutive modeling of deformation in high
temperature of a forging 6005A aluminum alloy, Mater. Des. 54 (2014) 576–581.
of dynamic recovery and recrystallization, the dislocation density in- [14] D.N. Zhang, Q.Q. Shangguan, C.J. Xie, F. Liu, A modified Johnson–Cook model of
creases to a peak value, and then keeps a constant value. Fig. 10c shows dynamic tensile behaviors for 7075-T6 aluminum alloy, J. Alloy. Compd. 619
the computed damage evolution. The damage factor is zero at the be- (2015) 186–194.
[15] J. Lin, T.A. Dean, Modelling of microstructure evolution in hot forming using
ginning of deformation, indicating the material is undamaged. Then, it unified constitutive equations, J. Mater. Process. Technol. 167 (2005) 354–362.
grows approximately linearly with the increase of strain. At last, with [16] B. Wu, M.Q. Li, D.W. Ma, The flow behavior and constitutive equations in iso-
the nucleation and growth of microcrack [40], the damage factor grows thermal compression of 7050 aluminum alloy, Mater. Sci. Eng. A 542 (2012) 79–87.
[17] M.S. Mohamed, A.D. Foster, J. Lin, D.S. Balint, T.A. Dean, Investigation of de-
at a very fast rate to the ultimate tensile strain. At temperature 400 °C, formation and failure features in hot stamping of AA6082: experimentation and
the evolution rate of damage increases with decreasing strain rate. At modelling, Int. J. Mach. Tool. Manuf. 53 (2012) 27–38.
strain rate 0.1 s−1, the evolution rate of damage was lowest at 400 °C. [18] H. Ji, J. Liu, B. Wang, X. Tang, J. Lin, Y. Huo, Microstructure evolution and con-
stitutive equations for the high-temperature deformation of 5Cr21Mn9Ni4N heat-
resistant steel, J. Alloy. Compd. 693 (2017) 674–687.
[19] K. Zheng, J. Lee, J. Lin, T.A. Dean, A buckling model for flange wrinkling in hot
5. Conclusions deep drawing aluminium alloys with macro-textured tool surfaces, Int. J. Mach.
Tool. Manuf. 114 (2017) 21–34.
[20] L. Yang, B. Wang, G. Liu, H. Zhao, W. Xiao, Behavior and modeling of flow softening
(1) The flow behavior and microstructure evolution of AA7075 at
and ductile damage evolution in hot forming of TA15 alloy sheets, Mater. Des. 85
uniaxial hot tensile tests was analyzed. EBSD results showed that, (2015) 135–148.
average grain size can be refined with increasing deformation [21] N. Li, C. Sun, N. Guo, M. Mohamed, J. Lin, T. Matsumoto, C. Liu, Experimental
amount, temperature, and decreasing strain rate. investigation of boron steel at hot stamping conditions, J. Mater. Process. Technol.
228 (2016) 2–10.
(2) A plenty of LAGBs were formed in the material due to deformation, [22] J. Lin, M. Mohamed, D. Balint, T. Dean, The development of continuum damage
and some LAGBs were transformed into HAGBs due to subgrain mechanics-based theories for predicting forming limit diagrams for hot stamping
growth and dynamic recrystallization. The average misorientation applications, Int. J. Damage Mech. 23 (2014) 684–701.
[23] X. Tang, B. Wang, H. Ji, X. Fu, W. Xiao, Behavior and modeling of microstructure
angle increases with increasing temperature and decreasing strain evolution during metadynamic recrystallization of a Ni-based superalloy, Mater.
rate. Sci. Eng. A 675 (2016) 192–203.
(3) A set of constitutive equations for AA7075 was established, which [24] B. Radhakrishnan, G.B. Sarma, T. Zacharia, Modeling the kinetics and micro-
structural evolution during static recrystallization—Monte Carlo simulation of re-
couples with temperature, strain rate, strain, and the evolution of crystallization, Acta Mater. 46 (1998) 4415–4433.
microstructures. Genetic algorithm was used to determine the va- [25] J. Lin, Y. Liu, D.C.J. Farrugia, M. Zhou, Development of dislocation-based unified
lues of the material constants based on the hot tensile results. material model for simulating microstructure evolution in multipass hot rolling,
Philos. Mag. 85 (2005) 1967–1987.
(4) The predicted stress-strain data and average grain size were com- [26] J. Lin, Y. Liu, A set of unified constitutive equations for modelling microstructure
pared with the experimental results. Statistical results indicate a evolution in hot deformation, J. Mater. Process. Technol. 143–144 (2003) 281–285.
good prediction accuracy of the model in describing the flow stress [27] Y. Huo, Q. Bai, B. Wang, J. Lin, J. Zhou, A new application of unified constitutive
equations for cross wedge rolling of a high-speed railway axle steel, J. Mater.
and microstructural evolution of AA7075.
Process. Technol. 223 (2015) 274–283.
[28] A. Momeni, G.R. Ebrahimi, M. Jahazi, P. Bocher, Microstructure evolution at the
onset of discontinuous dynamic recrystallization: a physics-based model of subgrain
Acknowledgements critical size, J. Alloy. Compd. 587 (2014) 199–210.
[29] W. Roberts, B. Ahlblom, A nucleation criterion for dynamic recrystallization during
hot working, Acta Mater. 26 (1978) 801–813.
This work was supported by the Joint Funds of the National Natural [30] R. Sandström, R. Lagneborg, A model for hot working occurring by recrystalliza-
Science Foundation of China (Grant number U1564202). The authors tion, Acta Mater. 23 (1975) 387–398.
are grateful to Professor Jianguo Lin of Imperial College London for [31] B.N. Kim, K. Hiraga, Y. Sakka, B.W. Ahn, A grain-boundary diffusion model of
dynamic grain growth during superplastic deformation, Acta Mater. 47 (1999)
guidance in genetic algorithm-based optimization.

712
W. Xiao et al. Materials Science & Engineering A 712 (2018) 704–713

3433–3439. [37] Q. Bai, M. Mohamed, Z. Shi, J. Lin, T. Dean, Application of a continuum damage
[32] G.Z. Voyiadjis, A. Shojaei, G. Li, P.I. Kattan, Continuum damage-healing mechanics mechanics (CDM)-based model for predicting formability of warm formed alumi-
with introduction to new healing variables, Int. J. Damage Mech. 21 (2012) nium alloy, Int. J. Adv. Manuf. Technol. 88 (2017) 3437–3446.
391–414. [38] N. Kotkunde, A.D. Deole, A.K. Gupta, S.K. Singh, Comparative study of constitutive
[33] M.A. Khaleel, H.M. Zbib, E.A. Nyberg, Constitutive modeling of deformation and modeling for Ti–6Al–4V alloy at low strain rates and elevated temperatures, Mater.
damage in superplastic materials, Int. J. Plast. 17 (2001) 277–296. Des. 55 (2014) 999–1005.
[34] D. Wei, J. Han, Z.Y. Jiang, C. Lu, A.K. Tieu, A study on crack healing in 1045 steel, [39] J. Zhou, B. Wang, M. Huang, Two constitutive descriptions of boron steel 22MnB5
J. Mater. Process. Technol. 177 (2006) 233–237. at high temperature, Mater. Des. 63 (2014) 738–748.
[35] G.Z. Voyiadjis, P.I. Kattan, A comparative study of damage variables in continuum [40] T.R. Bieler, P. Eisenlohr, F. Roters, D. Kumar, D.E. Mason, M.A. Crimp, D. Raabe,
damage mechanics, Int. J. Damage Mech. 18 (2008) 315–340. The role of heterogeneous deformation on damage nucleation at grain boundaries in
[36] J. Cao, J. Lin, A study on formulation of objective functions for determining ma- single phase metals, Int. J. Plast. 25 (2009) 1655–1683.
terial models, Int. J. Mech. Sci. 50 (2008) 193–204.

713

You might also like