Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

International Journal of Solids and Structures 56–57 (2015) 142–153

Contents lists available at ScienceDirect

International Journal of Solids and Structures


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

Subsequent yielding of polycrystalline aluminum after cyclic


tension–compression analyzed by experiments and simulations
Guijuan Hu a,b, Shihong Huang a, Damin Lu a, Xianci Zhong c, Zhenhuan Li d, Wolfgang Brocks e,
Keshi Zhang a,⇑
a
Key Laboratory of Disaster Prevention and Structural Safety, Guangxi University, Nanning 530004, China
b
Guangxi Polytechnic of Construction, Nanning 530004, China
c
College of Mathematics and Information Science, Guangxi University, Nanning 530004, China
d
Department of Mechanics, Huazhong University of Science and Technology, Wuhan 430074, China
e
Institute for Materials Science, Christian Albrecht University, Kiel, Germany

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

Article history: Subsequent yielding of polycrystalline aluminum after cyclic tension–compression is studied by both
Received 2 February 2014 experiments and finite element simulations applying crystal plasticity. The directional hardening induced
Received in revised form 14 October 2014 by pre-deformation is particularly emphasized. By means of a sub-model method, scale-bridging analyses
Available online 5 December 2014
of a specimen containing a ring-section constructed of a number of grains re-loaded in different ratios of
axial tension and torsion are carried out, after cyclic tension–compression pre-deformation. Both the
Keywords: Chaboche rate-dependent constitutive relation and the crystal plasticity theory extended by introducing
Subsequent yield surface
a back stress, which relates the mechanical behavior of polycrystalline aluminum at macroscopic scale
Sub-model method
Cyclic plasticity
and microscopic scale, are employed. The influences of different unloading loci, pre-loading directions
Polycrystalline aluminum and yield definitions on subsequent yield surfaces are investigated in detail by comparing computational
and experimental results. The results show that the shape of subsequent yield surfaces and the ‘‘sharp
corner’’ appearing at the front end of a yield surface are closely related to the tensile or compressive
pre-loading direction and different yield point definitions. The subsequent yielding by re-loading
obviously shows deformation-induced anisotropic hardening. The main characteristics of subsequent
yield surfaces observed in experiments can be satisfactorily captured by the present crystal plasticity
model with the introduction of a back stress.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction relationship can be established and accurate computational model-


ing can be performed.
The adoption of a yield surface is a basic concept in the theory of Since the shape of a yield surface strongly depends on the pre-
plasticity and hence is of long-term interest within the mechanics loading direction, the orientation of a ‘‘sharp corner’’ changes with
community. The initial yield surface depends on the material’s ini- accumulated plastic strain when the applied load changes from
tial microstructure, while the subsequent yield surfaces strongly tension to compression in the process of cyclic loading and unload-
depend on the pre-deformation path, in addition. Generally speak- ing (Zhang et al., 2011). Therefore, it is essential to any constitutive
ing, due to the deformation-induced microstructural changes, the theory of cyclic plasticity how to exactly characterize the aniso-
subsequent yield surfaces evolve and show a complex shape under tropic evolution of subsequent yield surfaces under cyclic loading.
different pre-deformation paths, such as expansion and contrac- In the past decades, a series of studies on the evolution of
tion (i.e. isotropic hardening or softening), translation (i.e. kine- subsequent yield surfaces has been performed. Most of them paid
matic hardening due to the evolution of back-stress), distortion their attention to monotonous proportional and non-proportional
(i.e. anisotropic hardening). Obviously, it is important to rationally pre-loading paths (Helling et al., 1986; Phillips and Lu, 1984; Wu
describe the evolution of subsequent yield surfaces under different and Yeh, 1991; Khan et al., 2009; Khan et al., 2010a,b; Sung
pre-deformation paths. On this basis, a rational constitutive et al., 2011), some also considering arbitrarily textured solids with
varied lattice structure (Graff et al., 2007), while only a few focused
⇑ Corresponding author. on cyclic pre-loading paths. For example, Kowalewski and
E-mail address: zhangks@gxu.edu.cn (K. Zhang).
Sliwowski (1997) studied the evolution of the yield surface of

http://dx.doi.org/10.1016/j.ijsolstr.2014.11.022
0020-7683/Ó 2014 Elsevier Ltd. All rights reserved.
G. Hu et al. / International Journal of Solids and Structures 56–57 (2015) 142–153 143

18G2A steel under cyclic loading with different strain amplitudes. ends of each tubular specimen are filled with additional steel
They showed that the subsequent yield surfaces shrank compared inserts, as shown Fig. 2.
to the initial yield surface due to cyclic softening. Ishikawa and All experiments are performed using an MTS809 axial-torsion
Sasaki (1988) and Ishikawa (1997) studied the yield surfaces of servo hydraulic fatigue testing machine. Axial and torsional dis-
304 stainless steel being initially anisotropic. Their results placements are measured by an axial–torsional extensometer of
exhibited that for proportional loading paths, the shapes of subse- gauge length 25 mm. The measurement ranges for axial strain,
quent yield surfaces were ellipses; however, for non-proportional E, and torsional strain (or shear angle), C, are ±0.1° and ±5°, respec-
cyclic pre-loading, the evolution of subsequent yield surfaces tively. Here, the uppercase Greek letters stand for the macroscopic
became very complicated, including translation, distortion and strain measures.
rotation. Besides, under cyclic pre-loading, it was repeatedly A multiple specimen technique has been applied to determine
observed in experiments that the shapes of subsequent yield subsequent yield surfaces. First, all specimens are subjected to
surfaces were strongly affected by several factors, such as cyclic strain controlled symmetrical tension–compression with an ampli-
hardening and softening, loading amplitudes, Bauschinger effect, tude of 0.3% and 30 cycles. Then the specimen is unloaded at differ-
creep, ratcheting and so on. To the best of our knowledge, deep ent specified positions, A, B, C or D, of the final hysteresis loop until
understanding and rational description of the evolution of subse- the end points, OA1, OA2, OB, OC or OD, as shown in Fig. 3(a) under
quent yield surfaces under complex pre-loading paths are still stress control. The two unloading end points OA1 and OA2 corre-
lacking in the literature. spond to the unloading starting point A, where OA1 is within the
In the earlier investigations of subsequent yield surfaces, sev- elastic domain and OA2 is the tensile-compressive turning point.
eral models were proposed to describe anisotropic deformation Besides, all unloading end points except OA2 must be located in
(Phillips and Tang, 1972; Eisenberg and Yen, 1984; Kurtyka, the elastic region; otherwise, local reverse loading may change
1988; Kurtyka and Zyczkowski, 1996; Francois, 2001). These mod- the yield point due to additional plastic strain. In order to obtain
els can depict yield surface distortion by geometric considerations, the subsequent yield surfaces corresponding to OA1, OA2, OB, OC
but their evolution equations lack definite physical meaning. and OD points, the specimens are reloaded from OA1, OA2, OB, OC
Moreover, these models usually include a group of material param- and OD by combined tension–torsion. If the proportion of the ten-
eters very difficult to be determined, especially when the material sion and torsion components keeps constant, the reloading paths
is subjected to complex loading or deformation paths. Generally are in the radial direction in the tension–torsion stress space with
speaking, the evolutions of the yield surface under complex pre- OA1, OA2, OB, OC and OD points being the origins, as shown in
loading paths are very complicated, and it seems to be impossible Fig. 3(b). The respective effective stress–strain curves,
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
to describe them rationally unless the details of material micro- 2 2
Reff ¼ R2 þ 3T 2 , Epeff ¼ ðEp Þ þ ðCp Þ =3, (here T is macroscopic
structure evolution are fully captured. Therefore, micromechanical
models, which are particularly suited for describing the material torsional stress) for two reloading paths are plotted in Fig. 3(c),
microstructure evolution, have become increasingly important in where K and K0 are initial and unloading stiffness, respectively,
modeling the material deformation under complex loading paths. which can be used to calculate the incremental residual strain.
As is well-known, the crystal plasticity models are based on aniso- Again, uppercase Greek letters stand for macroscopic stress and
tropic crystal slip and thus can fully capture the microstructural strain measures, respectively. For the polycrystalline sub-model
evolution. In other words, the crystal plasticity theory provides a described below, they represent homogenized values of stresses
promising approach for accurately characterizing anisotropic evo- and strains resulting from the crystal plasticity calculations. The
lution of yield surfaces. stress at a given (plastic) offset strain is defined as the subsequent
In the present contribution, the evolution of subsequent yield yield stress. The choice of the offset strain has a great influence on
surfaces of polycrystalline aluminum after cyclic tension–com- the experimentally determined yield stress. For a strain hardening
pression is studied in detail by means of experiments and FEM material, the greater the target strain is specified the higher is the
modeling. The paper is organized as follows. In Section 2, the measured subsequent yield stress.
methods to determine the subsequent yield surfaces under cyclic The procedure to determine the yield point related to the offset
strain controlled loading experimentally by multiple identical strain is as follows. Assume that the cumulative effective plastic
specimens and numerically by crystal plasticity simulations of strain at the unloading point is Epeff . After reloading beyond the
various sub-models of a polycrystalline ring section containing yield point, a residual plastic strain increment, DEpeff ¼
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
different numbers of grains are introduced. In Section 3, the influ- 2 2
ðDEp Þ þ ðDCp Þ =3, emerges with DEp being an axial strain incre-
ences of the value of offsetting strain applied to define the mate-
ment and DCp a shear angle increment. Specifying an offset value
rial yield on the yield surface shape and its evolution are
discussed by using experiment and simulation. The crystal plas- of DEpoffset ¼ DEpeff ¼ 2  104 , for instance, allows for determining
ticity model, which takes a back-stress based nonlinear kinematic the initiation of subsequent yielding.
hardening into account, is identified and validated with the Finally, the subsequent yield stress of the material, Ry , by reload-
experimental results. Finally, Section 4 ends the paper with some ing from the unloading point, Runload
eff ¼ Runload , can be expressed as
main conclusions.
Ry ¼ Runload þ f ðDEpoffset ; hÞ ð1Þ
where h stands for the loading direction angle in Fig. 3(b). The mate-
2. Experiment and modeling rial will re-yield when the macroscopic effective plastic strain incre-
ment due to reloading reaches the offset strain. The subsequent
2.1. Experiments yield stress is dependent on the loading path angle, h.

In order to study the subsequent yield surfaces of metals, a 2.2. Modeling


polycrystalline aluminum with purity 99.89% is used here. Its
chemical composition and mechanical properties are given in The macroscopic mechanical properties of polycrystalline mate-
Table 1. Approximately, the material has an equiaxed grain struc- rial are governed by their micro-structures and their evolution with
ture. Its micrograph is shown in Fig. 1. Thin-walled tube specimens deformation, including the grain size, grain orientation, anisotropy
are adopted in our tests. In order to ensure reliable clamping, both and heterogeneity of deformation within grains. Obviously, the
144 G. Hu et al. / International Journal of Solids and Structures 56–57 (2015) 142–153

Table 1
Chemical composition in weight % and mechanical properties of commercially pure aluminum.

Chemical composition Mechanical properties


Al/% Cu/% Mg/% Si/% Mn/% Zn/% E/GPa G/GPa r0:2 /MPa ru /MPa ef
99.89 0.02 0.03 0.03 0.02 0.01 60 25 20 81 24%

60 (a)
A
40 OA1

20

/MPa
OB OA2
0 OC

Stress
-20
OD
B
-40
C
D
-60

-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4


Strain E(%)
Fig. 1. Metallographic photo of the polycrystalline aluminum.
(b)
1/2 Reloading for searching
3 Subsequent yield surface
165 o
45 50 4 90
5 3
16

6
10.5

7 2
o o
180 Unloading
1
R25

8 10 Σ
O
Fig. 2. Specimen geometry. OA1,OA2,OB,OC,OD A,B,C,D
Cyclic pre-loading
micro-structure of a material significantly influences its subsequent
yield behavior. In order to reveal the underlying microscopic mech-
anisms, the crystal plasticity theory, which treats well the deforma-
tion at the grain level, is employed here.
To reduce the total computing time and simulate precisely the
micro-inhomogeneity of deformation, a sub-model method is
(c)
developed, which considers a global test specimen and polycrystal- Reloading path 1
K'
line aggregate model. The global specimen is analyzed by the mac-
A,B,C,D
roscopic Chaboche model, and then the displacement boundary
conditions are applied to the sub-model consisting of a polycrystal- Reloading path 2
Effective stress

line aggregate. By applying this method, the ‘‘scale-bridging’’ anal- Unloading


ysis from the macro-specimen scale to the micro-grain scale is Yield definition
implemented.
Pre-loading Offset strain
2.2.1. The rate-dependent Chaboche constitutive model
The standard equations of the Chaboche model (Chaboche, OA1,OA2,OB,OC,OD
1989) as adopted in this study are briefly summarized. The strain K
rate e_ can be partitioned into e_ ¼ e_ e þ e_ p . According to normality
assumption, the plastic strain rate, e_ p , can be written as
rffiffiffi  n rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  k0
3 3 Fy r0  x0 2 p p F
e_ p ¼ _ ¼
pn ; p_ ¼ e_ : e_ ¼ y ð2Þ Effective strain
2 2 K J 2 ðr0  x0 Þ 3 K
Fig. 3. Reloading paths at different unloading points after reaching a stable
where, p_ is the effective plastic strain rate, r0 and x0 are the devia- tension–compression cycle. (a) Schematic of unloading at different strain positions,
tors of stress and back-stress tensors respectively, J 2 is defined by (b) reloading paths for searching subsequent yield points, (c) unloading points and
qffiffiffiffiffiffiffiffiffiffiffiffi reloading positions.
J2 ðyÞ ¼ 32 y : y, k0 is rate-sensitivity parameter, K is a material
parameter reflecting the viscosity of the material, F y is a yield func- where, r0 ðpÞ
_ is used to describe the initial yield surface and R
tion which define the elasticity domain, and it is expressed as describes the isotropic surface expansion,
F y ¼ J 2 ðr0  x0 Þ  r0 ðpÞ
_ R¼0 ð3Þ R ¼ Q ½1  expðb  pÞ ð4Þ
G. Hu et al. / International Journal of Solids and Structures 56–57 (2015) 142–153 145

The shift of the yield surface, i.e. the kinematic hardening, is where q is a constant, whose value can be chosen in the range from
described by evolution equations of the deviatoric back stress, 0 to 1; and Chang and Asaro (1981) suggested
 
x_ ðkÞ ¼ C k ðak e_ p  xðkÞ pÞ
_ ð5Þ 2 h0 c
hðcÞ ¼ h0 sec h ; ð11Þ
ss  s0
X
M
x¼ xðkÞ ð6Þ where h0 is the initial hardening rate, s0 is the critical resolved shear
k¼1 stress and ss is the saturation value, which can be determined
according to the experimental data.
Here, xðkÞ is the component of the back-stress x; x_ ðkÞ is the corre-
Following Walker (1981) and Chaboche (1991), the evolution of
sponding rate; ak and C k are material parameters for describing
x(a) is expressed as (Zhang et al., 2011, 2013):
the material hardening characteristics. M is the number of compo-
ðaÞ
nents of the back stress, where M equals 2 in this paper. x_ ðaÞ ¼ ac_ ðaÞ  c ½1  e1 ð1  expðe2 cÞÞxðaÞ jc_ ðaÞ j  dx ð12Þ
In order to solve the model calculation of the incremental stress
and strain, the constitutive rate relation, assuming that the elastic where a is a linear hardening parameter, c and d are nonlinear hard-
deformation is small, can be expressed as: ening parameters, respectively, e1 and e2 are used to describe the
saturation of the cyclic hardening. The above material parameters
<4> <4>
are determined by the cyclic test data.
r_ J ¼ C : e_ e ¼ C : ðe_  e_ p Þ ð7Þ
<4>
In the above crystal plasticity model, the material hardening on
where r_ J is the Jaumann stress rate, C is the fourth-order tensor of each a-slip system is divided into two parts. One is the isotropic
the tangent elasticity of the material which consists of two indepen- hardening represented by g(a) and its evolution. Any slipping no
dent elastic constants for isotropic material: E and v, Young’s mod- matter if in positive or opposite direction always leads to an
ulus and Poisson’s ratio. This equation is also suitable for an increase of g(a), which means that the elastic domain will expand
anisotropic material, but its components of the tensor of the tan- until it saturates. The other is the directional hardening repre-
gent elasticity are different and the orientation of the material’s axis sented by the back stress, xðaÞ , of the a-slip system and its evolu-
should be considered. tion. For the evolution of the back stress in positive direction, the
Although above model is rate-dependent, it can describe the initial threshold value of the slip system is increased in the same
rate-independent behavior approximately if the value of parameter direction but reduced in opposite direction, and vice versa. There-
k0 is set large enough. This model can be adopted by using the soft- fore, the directional hardening of the back stress at the slip system
ware ABAQUS. Chaboche’s model is applied in the present paper to results in the Bauschinger effect, and the response to stretching is
simulate the macroscopic deformation of the specimen under axial directional dependent for the pre-deformed material. Thus, aniso-
and rotational loading, and the results are applied to the sub-mod- tropic hardening of the material will be represented by the evolu-
el’s calculation. tion of back stresses at the slip systems of many grains.
According to the geometric analysis (cf. Hill and Rice (1972),
2.2.2. A model of crystal plasticity considering the Bauschinger effect Asaro and Rice (1977), Needleman et al. (1985)), the plastic strain
In the present tests, the specimens have been frequently rate, e_ p can be expressed as:
unloaded and re-loaded. Therefore, a strong Bauschinger effect is n
observed. In this situation, the material hardening is neither simply e_ p ¼ R PðaÞ c_ ðaÞ ð13Þ
a¼1
isotropic nor simply kinematic. The conventional model of crystal
plasticity, however, is established on the basis of isotropic harden- where PðaÞ is the Schmid tensor. Thus, the constitutive rate relation
ing, which is described by considering critical resolved shear stress <4> <4>

and slip resistance. The slip resistances under tension and com- has a similar type as Eq. (7), r_ J ¼ C : e_ e ¼ C : ðe_  e_ p Þ, and it can
pression are assumed to be the same, so this hardening model can- be used to solve the incremental stress and strain calculation. The
not predict the Bauschinger effect. In order to capture the difference to isotropic material is that the calculation must consider
Bauschinger effect under cyclic loading, the conventional crystal- the orientation of the material’s axis and its rotation, the strain rate
line plastic theory has been improved by Zhang et al. (2011) as is is related to the slipping of the crystal lattice, furthermore the
briefly outlined here. Different from the model proposed by tensor of tangent elasticity is different. For pure Al, the tensor of
Hutchinson (1976), a back-stress xðaÞ is introduced in the viscoplas-
tic constitutive relation as follows:
 
sðaÞ  xðaÞ k
c_ ðaÞ ¼ c_ 0 sgnðsðaÞ  xðaÞ Þ ðaÞ

 ð8Þ
g

where c_ 0 is a reference shear strain rate on slip system, k is the rate-


sensitivity parameter, sðaÞ is the resolved shear stress on the ath slip
system of the single crystal, g ðaÞ is the threshold value at which the
slip system a is activated, the back-stress, xðaÞ , characterizes the
nonlinear kinematic hardening on the a-th slip system.
The evolution of g(a) is described as follows (Pan and Rice, θ
1983):
X
n
RX
n
g_ ðaÞ ðcÞ ¼ hab ðcÞjc_ ðbÞ j; c ¼ jdcðbÞ j ð9Þ
b b

where hab ðcÞ are the slip-plane hardening moduli, n is the number
of crystal slip systems. Moreover, Hutchinson (1970) suggested
hab ðcÞ ¼ hðcÞ½q þ ð1  qÞdab ; ð10Þ
Fig. 4. The ring shaped sub-model from the center section of the specimen.
146 G. Hu et al. / International Journal of Solids and Structures 56–57 (2015) 142–153

tangent elasticity is determined by three independent parameters: grains is adopted as a sub-model. The sub-model method provided
C 11 ; C 12 ; and C 44 . by the ABAQUS software can connect the ring section element,
The above constitutive relation for the model of plasticity has which can depict the microstructure properties of the polycrystal-
been implemented as user-supplied subroutine UMAT for the ABA- line material such as local anisotropy and plastic slipping, with the
QUS/Standard module (Zhang et al., 2011, 2013), the details for the global specimen model that depicts the overall response of materi-
model and the integral algorithm can be referred for readers. als at the specimen scale.
Assuming that the stress and deformation of the thin-walled
2.2.3. Procedure of calculation tubular specimen are uniform and homogenous over the gauge
Due to heavy computational costs, it is not a good choice to length, a ring with 1.0 mm radial thickness and 0.5 mm longitudi-
model a macroscopic test specimen at grain scale. Therefore, a rea- nal thickness is cut out from the center region of the specimen and
sonable ring section element which is constructed of a number of is regarded as a substructure, as shown in Fig. 4. Using this

(a) (b)
Fig. 5. The ring section sub-model as a 3D polycrystalline aggregate. (a) Grain core random distribution, (b) grain distribution in sub-model.

Experiment Experiment
60 60
Chaboche Simulation Crystal Plasticity Simulation

40 40

20 20
/MPa
/MPa

0 0
Stress
Stress

-20 -20

-40 -40

-60 Strain amplitude 0.3%


-60 Strain amplitude 0.3%

-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4
Strain E(%) Strain E(%)
(a) (b)
Fig. 6. Simulated stress–strain curves under cyclic tension–compression loading in comparison to experimental results.

Table 2
Material parameters of Chaboche model for pure Al.
0
E/MPa m b Q/MPa a1 /MPa C1/MPa a2/MPa C2/MPa r0 /MPa K/MPa k
56334 0.33 1.0 10 14 2500 20 50 15 10 200

Table 3
The material constants of single crystal aluminum.

C11/GPa C12/GPa C44/GPa s0/MPa ss/MPa h0/MPa a/GPa c/GPa e1 e2 d c_ 0 /s1 k

83.37 41.06 27 8.0 9.0 85 25 3.0 0.1 1.0 0 0.001 200


G. Hu et al. / International Journal of Solids and Structures 56–57 (2015) 142–153 147

sub-model the macroscopic axial strain increment, DE, and the 1X 1X


DE ¼ V ie Dezz ; DC ¼ V ie ð2Dezh Þ ð14Þ
shear increment, DC, of the sub-model can be calculated from V V
homogenization
where, V is the volume of the sub-model, Vie is the ieth element vol-
ume. In the same way, the axial stress increment, DR, and the shear
increment, DT, of the sub-model can be calculated by
Crystal Plasticity Simulation
60
Experiment
1X 1X
DR ¼ V ie Drzz ; DT ¼ V ie ðDrzh Þ ð15Þ
40 V V
/MPa

20 Thus, the response of DRij  DEij of the sub-model reflecting the


polycrystal behavior can be obtained. Furthermore, the axial plastic
0 strain increment, DEp , and the shear increment, DCp , for the speci-
Stress

men can also be calculated. The macroscopic strain increments


-20 can also be determined directly from the global FE model instead
by the homogenization calculation, which come up to be almost
-40 the same, numerically.
Therefore, the procedure of the scale-bridging calculation can
Strain amplitude
-60 be presented as following. First, the classic Chaboche constitutive
0.1%-0.3%-0.5%-0.8%
model is employed to simulate the macroscopic response of the
-1.0 -0.8 -0.6 -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 specimen. Then, the deformation field in the specimen is applied
Strain E(%)
as boundary condition to the ring-section sub-model. Finally, the
Fig. 7. Comparisons of steady hysteresis loops of cyclic tension–compression crystal plasticity theory is adopted to calculate the internal stresses
loading with different strain amplitude. and deformations of the material in the sub-model.
Considering that grains are randomly distributed in the sub-
model, the Voronoi method is adopted to generate polycrystalline
60 aggregates. First, the finite element mesh of the sub-model is com-
piled, which contains the information of nodes and elements. Then,
50 the coordinates of the grain cores in the sub-model are randomly
generated as shown in Fig. 5(a). Each element in the sub-model
is assigned to the nearest grain core by calculating the distance
40
/MPa

between the centroid of the element and the grain core, for the
50 grains
300 grains grain number is set previously the grain cores can randomly
30 200 grains arranged in the sub-model, by this way every element can find
Stress

500 grains the grain it belongs to, as shown in Fig. 5(b). In order to capture
1500 grains inhomogeneous deformation in grains, FE meshes should be
20
1000 grains refined so that each grain contains a sufficient number of finite
800 grains elements.
10 Finally, the orientation of each grain is described by an orthog-
onal crystallographic coordinate system, which is defined by a ran-
0 0.00 domly generated base vector and a second perpendicular vector.
0.05 0.10 0.15 0.20 0.25 0.30
For real materials, a certain texture will develop during processing
Strain E(%) and forming. The material considered in this study has approxi-
mately equiaxed grains (cf. Fig. 1). As a first step, this effect of
Fig. 8. Comparison of stress–strain curves under uniaxial tensile loading simulated
by the sub-model with varying number of grains. the initial texture on the shape and the orientation of the grains
has not been taken into account.

60 60
(a) (b)
Tresca surface Mises surface
40 50
/MPa

20 40
Τ /MPa

Axial tension
Effective stress

0 30 Pure torsion
1/2
3

-20 20

-40 -5
10
ΔΕpoffset=2x10
-3
ΔΕpoffset=2x10
-4 ΔΕpoffset=2x10
-60 0
-60 -40 -20 0 20 40 60 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
/MPa Effective Strain Eeq(%)

Fig. 9. Crystal plasticity simulations of monotonic loading; (a) initial yield surfaces in the stress space; (b) effective stress–strain curve.
148 G. Hu et al. / International Journal of Solids and Structures 56–57 (2015) 142–153

2.2.4. The determination of the parameters of the Chaboche and the only slightly with the number of cycles. In other words, the cyclic
crystal plasticity models hardening/softening behavior of the polycrystalline aluminum is
The constitutive parameters of the Chaboche model are deter- insignificant, and exhibits cyclic stable property. For this reason,
mined from the cyclic tension–compression test with an axial the parameters of the Chaboche model can be determined by
strain amplitude of 0.3%. As shown in Fig. 6(a), the experimentally matching the stress–strain saturation loop predicted by the model
measured cyclic stress–strain curves rapidly reach the saturated with the experimentally measured cyclic saturation loop. The
state within 30 tension–compression cycles. In addition, the max- value of the viscosity exponent is set as k0 = 200, so that the mate-
imum peak stress during each tension–compression cycle changes rial behavior is approximately rate-independent. All parameters

60
(a) Crystal-plasticity model A-O A1

40 -5
ΔΕ poffset=2x10
-5
A ΔΕ poffset=5x10
20

Τ /MPa
O A1 ΔΕ poffset=1x10
-4
unload-point p -4
Ε 0
OA1 ΔΕ offset=2x10
Symmetric points -4
1/2 ΔΕ poffset=6x10
3
-3
-20 ΔΕ poffset=1x10

-40

-60
-60 -40 -20 0 20 40 60
/MPa

60
(b) Crystal-plasticity model A-OA2

40

A ΔΕpoffset=2x10
-5
20 -5
ΔΕpoffset=5x10
3 1/2 Τ /MPa

O A2 Ε Unload-point ΔΕpoffset=1x10
-4

0 -4
Symmetric points OA2 ΔΕpoffset=2x10
p -4
ΔΕoffset=6x10
-20 -3
ΔΕpoffset=1x10

-40

-60
-60 -40 -20 0 20 40 60
/MPa
60
(c) Crystal-plasticity model B-OB

40

-5
ΔΕpoffset=2x10
20
3 1/2 Τ /MPa

p -5
ΔΕoffset=5x10
Unload-point -4
OB Ε 0
ΔΕpoffset=1x10
OB ΔΕpoffset=2x10
-4

-4
B -20 ΔΕpoffset=6x10
-3
ΔΕpoffset=1x10

-40 Symmetric points

-60
-60 -40 -20 0 20 40 60
/MPa
Fig. 10. Subsequent yield surfaces in transition from tension to compression after reaching a stable cycle of tension–compression loading as obtained by crystal plasticity
simulations. (a) Unloading at point A to OA1; (b) unloading at point A to OA2; (c) unloading at point B to OB; (d) unloading at point C to OC; (e) unloading at point D to OD.
G. Hu et al. / International Journal of Solids and Structures 56–57 (2015) 142–153 149

are listed in Table 2. In Fig. 6(a), the stable hysteresis loops be large enough to achieve a convergent homogeneous solution
obtained from the simulation and the experiment are plotted for of the stress and strain fields. However, as the computational costs
comparison. It is obvious that the macroscopic mechanical increase heavily with increasing number of grains, a proper grain
response predicted by the Chaboche model with the parameters number should be adopted in all simulations. Sub-models which
listed in Table 2 is in good agreement with experimental results. contain 50, 200, 300, 500, 800, 1000 and 1500 grains, respectively,
In order to determine the model parameters of crystal plasticity have been computed by crystal plasticity FEM. The computation-
theory, a representative volume element containing 4248 elements ally obtained stress–strain curves of sub-models under uniaxial
and 50 grains is first computed. By fitting the computationally tension are given in Fig. 8. It is seen that the computational results
obtained cyclic stress–strain curve to the experimentally measured become insensitive to the grain number once it is larger than 200.
one, the parameters of the crystal plasticity model are incipiently This convergence behavior of the homogenized mechanical
estimated. Then, by performing a further refined modeling of the response of the sub-model is attributed to the randomness of the
submodel, now containing 35,256 elements and 200 grains and grain orientation and the grain size. Therefore, a sub-model which
fine-tuning the initially estimated values, more accurate parame- contains 200 grains, a total of 35,256 elements and 44,296 nodes is
ters are obtained as listed in Table 3. In order to validate the pre- adopted in the following computations.
cision of the model parameters, the computationally obtained If the microstructural characterization of the material in the
cyclic stress–strain curves and experimentally measured ones are polycrystal sub-model is required to match with the real grain size,
plotted in Fig. 6(b) for the applied cyclic strain amplitude of 0.3%. the centroid positions of all crystal nuclei of the polycrystalline
The computed cyclic stress–strain loops cover the experimentally aggregate needs to be calculated from the spatial distribution of
measured ones, indicating that the model parameters of crystal the real grains. Due to the limited capacity of the computer, any
plasticity theory in Table 3 are sufficiently accurate. Crystal plastic- analysis using the real grain size is impossible. The study of yield
ity simulations for cyclic amplitudes of 0.1%, 0.3%, 0.5% and 0.8%, surfaces in this article is mainly concerned on the overall macro-
respectively, are shown in Fig. 7. scopic response of materials, which can be conducted by homogeni-
zation of a ‘‘mesoscopically’’ heterogeneous material. The attention
2.2.5. The effect of the grain numbers of the sub-model on the is paid mainly to the question whether the macroscopic plastic
macroscopic response behavior of the material can be reflected compared to the experi-
As mentioned above, the sub-model is cut out from the center mental observation or not. Thus, it is not relevant if the average grain
region of the specimen. The number of grains in this ring should size of model differs from the real size of the grains in the material.

60
(d) Crystal-plasticity model C-OC

40
-5
ΔΕpoffset=2x10
20 ΔΕpoffset=5x10
-5
3 1/2 Τ /MPa

-4
OC Ε Unload-point ΔΕpoffset=1x10
0 -4
OC ΔΕpoffset=2x10
-4
ΔΕpoffset=6x10
-20
-3
C ΔΕpoffset=1x10

-40 Symmetric points

-60
-60 -40 -20 0 20 40 60
/MPa

60
(e) Crystal-plasticity model D-OD

40
-5
ΔΕpoffset=2x10
20 -5
ΔΕpoffset=5x10
3 1/2 Τ /MPa

Ε unload-point ΔΕpoffset=1x10
-4
0
-4
OD ΔΕpoffset=2x10
OD -4
-20 ΔΕpoffset=6x10
-3
D ΔΕpoffset=1x10
-40 Symmetric points

-60
-60 -40 -20 0 20 40 60
/MPa
Fig. 10 (continued)
150 G. Hu et al. / International Journal of Solids and Structures 56–57 (2015) 142–153

p -4
60 (a) ΔΕoffset=1x10 Crystal-plasticity model

40

20 A-OA1

3 1/2 Τ /MPa
A-OA2
B-OB
0
C-OC
D-OD
-20
Symmetric points
-40
A
O A1
-60

OB O A2 Ε -60 -40 -20 0 20 40 60


OC /MPa

ΔΕpoffset=10
-4
OD B 60 (b) Test
C
D 40

A-OA1
20
A-OA2
3 1/2 Τ /MPa

B-OB
0 C-OC
D-OD
Symmetric points
-20

-40

-60

-60 -40 -20 0 20 40 60


/MPa

Fig. 11. Subsequent yield surfaces after reaching a stable tension–compression cycle in transition from tension to compression as obtained (a) by crystal plasticity
simulations and (b) by tests.

3. Results and discussions to the unloading points B, C and D, respectively. All subsequent
yield surfaces in Fig. 10 differ noticeably from the initial yield sur-
Using the above-mentioned model, through exerting different face. The largest deviations are found at the sharp corner in pre-
combined proportional axial and torsional loading, the different deformation direction and the relatively flat bottom at its opposite
initial yield surfaces have been determined with the specified off- direction (Bauschinger effect). A comparison between Fig. 10(a)
set strains as shown in Fig. 9(a). They represent the material’s ini- and (b) shows that the subsequent yield surface after unloading
tial state, that is without residual stresses or strains, with equiaxed to the end point OA2 is slightly different from that after unloading
crystal structure and a random orientation of grains. The surfaces to OA1 due to the reverse loading at OA2. Moreover, it can be clearly
are elliptic and their sizes depend on the specified offset strain. seen from Fig. 10 that the shape of a subsequent yield surface is
Mises’ and Tresca’s descriptions are equal when the material is closely related to the direction of plastic deformation induced by
uniaxially loaded, but differ in pure shear. The initial yield surfaces cyclic preloading, and the curvature of the subsequent yield sur-
obtained from crystal plasticity calculations are found to be very faces in the loading direction is larger than that in the opposite
close to Tresca’s description. The corresponding effective stress– direction. In the cyclic loading process, tensile and compressive
strain curves for tension and the torsion are plotted in Fig. 9(b). stress state changes alternately, then the ‘‘sharp corner’’ of yield
They differ because the definition of the effective stress is based surface should change with it. As shown in Fig. 10(a) and (b), the
on the Mises description whereas the real yield surface is similar ‘‘sharp corners’’ of the yield surfaces point to the tensile direction
to Tresca’s description. for tensile pre-deformation, while Fig. 10(c)–(e) show that the
Initial and subsequent yield surfaces differ. After plastic pre- ‘‘sharp corner’’ of the yield surface a point to the compressive
deformation by tensile deformation, for example, conspicuous direction for the compressive pre-deformation. Obviously, if the
anisotropic hardening can be observed. material undergoes plastic pre-deformation in a certain direction,
Adopting the same loading paths as in the experiments (Fig. 3) the strain hardening properties at other loading directions strongly
reloading the material in different ratios of tension (or compres- depend on the direction of accumulated plastic strain, resulting in
sion) and torsion starting from the unloading point, a series of crys- hardening anisotropy and visible change of the subsequent yield
tal plasticity analyses have been performed to obtain the surfaces. As mentioned above, the shape of subsequent yield sur-
subsequent yield surfaces of polycrystalline aluminum. They are face is closely related to the offset strain. In addition, it can be seen
plotted for five different cases in Fig. 10, where Fig. 10(a) and (b) in Fig. 10 that the subsequent yield surfaces are severely distorted
correspond to the same unloading point A but different unloading for small offset strains. With increasing the offset strain, the subse-
end points OA1 and OA2, respectively, and Fig. 10(c)–(e) correspond quent yield surfaces tend to become elliptic.
G. Hu et al. / International Journal of Solids and Structures 56–57 (2015) 142–153 151

60
(a) Subsequent yield surface A-O A1

40
A
OA1
20

Τ /MPa
OB OA2 Ε 0
OC Symmetric points

1/2
3
-20 initial
OD B
C -40 ΔΕ p ΔΕ poffset= 1x10
-4 -3
D offset= 1x10
Test Test
-60 Crystal-plasticity Crystal-plasticity
-60 -40 -20 0 20 40 60
/MPa

60 60
(b) Subsequent yield surface A-OA2 (c) Subsequent yield surface B-OB

40 40

20 20
31/2 Τ /MPa

Τ /MPa

0 0
Symmetric points Symmetric points
1/2

initial
3

-20 -20
initial

-40 ΔΕp ΔΕpoffset=1x10


-4 -3
-40 ΔΕp =1x10-4
offset
ΔΕ poffset=1x10-3 offset=1x10
Test Test Test Test
Crystal-plasticity Crystal-plasticity Crystal-plasticity Crystal-plasticity
-60 -60
-60 -40 -20 0 20 40 60 -60 -40 -20 0 20 40 60
෤/MPa /MPa
60 60
Subsequent yield surface
(d) Subsequent yield surface C-OC (e) D-OD

40 40

20 20
31/2 Τ /MPa

Τ /MPa

0 0
Symmetric points Symmetric points
1/2
3

-20 -20
initial initial

-40 -40 ΔΕp -4


ΔΕpoffset=1x10-4 ΔΕ poffset=1x10-3 offset=1x10 ΔΕpoffset=1x10
-3

Test Test Test Test


Crystal-plasticity Crystal-plasticity -60
Crystal-plasticity Crystal-plasticity
-60
-60 -40 -20 0 20 40 60 -60 -40 -20 0 20 40 60

෤ /MPa /MPa
Fig. 12. Subsequent yield surfaces after reaching a stable tension–compression cycle as obtained from tests and simulations. (a) Unloading at point A to OA1; (b) unloading at
point A to OA2; (c) unloading at point B to OB; (d) unloading at point C to OC; (e) unloading at point D to OD.

When the subsequent yield surface is determined with a smal- elements is different and random for different loading path. The
ler offset strain, the new plastic deformation is produced only in yield point determined by some specified loading path with smal-
very few elements or grains of the sub-model, and the number of ler offset strain may deviate abnormally. It may lead to scatter and
152 G. Hu et al. / International Journal of Solids and Structures 56–57 (2015) 142–153

concavity of the local yield surface. This phenomenon was also microstructure transformation satisfactorily characterizes the
observed and described by Graff et al. (2007). As shown in anisotropic hardening properties of materials.
Fig. 10, the subsequent yield defined by offset strains less or equal Subsequent yield surfaces for different unloading points A, B, C,
to 1  104 does noticeably deviate when re-loading from the D and the end points OA1, OA2, OB, OC,OD are plotted in Fig. 12(a)–
unloading point along the 120° direction. In classical plasticity the- (e), and the effects of two offset strains DEpoffset ¼ 104 and
ory of homogeneous materials, the yield surfaces are postulated as
DEpoffset ¼ 103 are considered. It is obvious that there are some dif-
convex based on Drucker postulate. If the offset strain is chosen
ferences in the sizes of computationally obtained and experimen-
higher than 1  104 , for example 2  104, 6  104 or 104 plas-
tally measured subsequent yield surfaces, although their shapes
tic deformation occurs in much more elements or grains of the sub-
are close to each other. For the smaller offset strain
model, leading to a plastic behavior close to a homogeneous mate-
DEpoffset ¼ 104 , the yield surfaces have a greater curvature in the
rial, and the phenomenon of concavity does no longer appear.
For the present preloading and unloading cases, the subsequent preloading direction, showing marked anisotropic hardening char-
yield surfaces computationally obtained by CPFEM and experimen- acteristics. However, for the larger offset strain DEpoffset ¼ 103 , the
tally measured are plotted together in Fig. 11(a) and (b) for com- differences due to directional hardening have significantly
parison, where the same offset strain of DEpoffset ¼ 104 is chosen. decreased, and the shape and the size of subsequent yield surfaces
The subsequent yield surfaces computationally obtained by CPFEM predicted by the crystal plasticity model are in agreement with the
are close to those experimentally measured, although there are results measured by experiments.
slight differences in shape. In particular, with the change of the For the classical J2-type plasticity theory, the shape of the yield
preloading directions from tension to compression, the ‘‘sharp cor- surface is circular. it can move (kinematic hardening) and similarly
ner’’ direction of the subsequent yield surfaces changes simulta- expand (isotropic hardening) in stress space with the increasing
neously. Obviously, the crystal plasticity model incorporating the accumulation of plastic strain. Obviously, these classical combined
description of the slip deformation, nonlinear hardening and hardening models cannot depict the complicated evolution of

Fig. 13. The stress and strain contours of a 3D Voronoi polycrystalline aggregate after 30 tension–compression pre-cycles tension up to point A. (a) Mises effective stress; (b)
effective plastic strain (c) tensile stress; (d) tensile strain.
G. Hu et al. / International Journal of Solids and Structures 56–57 (2015) 142–153 153

subsequent yield surfaces as exhibited in the present experiments. the evolution of subsequent yield surfaces under different
However, the crystal plasticity model, which inherently considers pre-loading paths is very complicated, the main characteris-
the anisotropy caused by the crystalline microstructure and crystal tics can be satisfactorily captured by the present model.
plastic slip, can give a reasonable description to the directional
hardening of the subsequent yield surfaces induced by preloading. Acknowledgments
The above-mentioned analysis demonstrates that the choice of
the offset strain has a significant effect on the size and shape of the The authors acknowledge the support rendered by the National
subsequent yield surface. In fact, similar conclusions from previous Natural Science Foundation of China (Nos. 11072064, 11272094
experimental researches were reported (Phillips and Tang, 1972; and 11472085) and Key Project of Guangxi Science and Technology
Moon, 1976). In addition, the calculated non-convexity of yield sur- Lab Center (No. LGZX201101) and Project supported by Guangxi
face may arise from an improper specified offset strain (Graff et al., Education Department (No. 2013YB312). These financial supports
2007), which is very serious in plasticity theory. When the subse- are gratefully acknowledged.
quent yield surfaces are determined by means of a very small offset
strain, they might produce abnormal deformation patterns for some References
special loading paths because very few elements or grains of the sub-
Asaro, R.J., Rice, J.R., 1977. Strain localization in ductile single crystals. J. Mech. Phys.
model sustain new plastic deformation. Then the respective yield Solids. 25, 309–338.
point may significantly deviate, leading to local concavity. This phe- Chaboche, J.L., 1989. Constitutive equations for cyclic plasticity and cyclic
nomenon is possible in the sub-model of heterogeneous material. viscoplasticity. Int. J. Plast. 5 (3), 247–302.
Chaboche, J.L., 1991. On some modifications of kinematic hardening to improve the
But in the classical plasticity theory for homogeneous material, the description of ratchetting effects. Int. J. Plast. 7, 661–678.
subsequent yield surface certainly has to be convex based on Druc- Chang, Y.W., Asaro, R.J., 1981. An experimental study of shear localization in
ker’s postulate. For a real heterogeneous material, the stress and aluminum–copper single crystals. Acta Metall. 29, 241–257.
Eisenberg, M.A., Yen, C.F., 1984. The anisotropic deformation of yield surfaces. ASME
strain fields are inhomogeneous, which can be realized by simula-
J. Eng. Mater. Technol. 106, 355–360.
tions using the crystal plasticity model. As an example, the contours Francois, M., 2001. A plasticity model with yield surface distortion for non
of Mises effective stress, accumulated effective plasticity strain, proportional loading. Int. J. Plast. 17 (5), 703–717.
axial tensile stress and axial tensile strain in the sub-model after Graff, S., Brocks, W., Steglich, D., 2007. Yielding of magnesium: from single crystal to
polycrystalline aggregates. Int. J. Plast. 23 (12), 1957–1978.
30 tension–compression cycles are displayed in Fig. 13. Stresses Helling, D.E., Miller, A.K., Stout, M.G., 1986. An experimental investigation of the
and strains in a heterogeneous material are very non-uniform. yield loci of 1100-0 aluminum, 70:30 brass, and an overaged 2024 aluminum
alloy after various prestrains. J. Eng. Mater. Technol. 108, 313–320.
Hill, R., Rice, J.R., 1972. Constitutive analysis of elastic-plastic crystal at arbitrary
strain. J. Mech. Phys. Solids 20, 401–413.
4. Conclusions Hutchinson, J.W., 1970. Elastic–plastic behaviour of polycrystalline metals and
composites. Proc. R. Soc. Lond. A 319, 247–272.
A crystal plasticity model with an especial consideration of non- Hutchinson, J.W., 1976. Bounds and self-consistent estimates for creep of
polycrystalline materials. Proc. R. Soc. Lond. A348, 101–127.
linear kinematic hardening is applied to study the subsequent yield Ishikawa, H., 1997. Subsequent yield surface probed from its current center. Int. J.
surfaces of polycrystalline aluminum under cyclic tension–com- Plast. 13, 533–549.
pression preloading. To describe the subsequent directional hard- Ishikawa, H., Sasaki, K., 1988. Yield surfaces of SUS 304 under cyclic loading. J. Eng.
Mater. Technol. 110, 364–371.
ening, the back-stress is especially introduced. A careful Khan, A.S., Kazmi, R., Pande, A., Stoughton, T., 2009. Evolution of subsequent yield
comparison between the experimentally measured subsequent surfaces and elastic constants with finite plastic deformation. Part-I: a very low
yield surfaces and computationally obtained ones by CPFEM is work hardening aluminum alloy (Al6061-T6511). Int. J. Plast. 25 (9), 1611–
1625.
made. Main conclusions are as follows: Khan, A.S., Pandey, A., Stoughton, T., 2010a. Evolution of subsequent yield surfaces
and elastic constants with finite plastic deformation. Part II: a very high work
(1) When the strain at the unloading point changes from ten- hardening aluminum alloy (annealed 1100 Al). Int. J. Plast. 26 (10), 1421–1431.
Khan, A.S., Pandey, A., Stoughton, T., 2010b. Evolution of subsequent yield surfaces
sion to compression, the direction of ‘‘sharp corner’’ of
and elastic constants with finite plastic deformation. Part III: yield surface in
subsequent yield surface is changed. In this process, the tension–tension stress space (Al 6061-T 6511 and annealed 1100 Al). Int. J.
size of the subsequent yield surface is slightly changed Plast. 26 (10), 1432–1441.
Kowalewski, Z.L., Sliwowski, M., 1997. Effect of cyclic loading on the yield surface
but the shape of the subsequent yield surface is distorted
evolution of 18G2A low-alloy steel. Int. J. Mech. Sci. 39, 51–68.
severely and its position in stress space is also translated Kurtyka, T., 1988. Parameter identification of a distortional model of subsequent
observably. yield surfaces. Arch. Mech. 40, 433–454.
(2) A scale-bridging modeling method, which crosses the mac- Kurtyka, T., Zyczkowski, M., 1996. Evolution equations for distortional plastic
hardening. Int. J. Plast. 12 (2), 191–213.
roscopic specimen scale to the approximate grain scale in Moon, H., 1976. An experimental study of the outer yield surfaces for annealed
which the material’s local anisotropy and plastic slipping polycrystalline aluminium. Acta Mech. 24, 191–208.
can be represented, is proposed. The Chaboche rate-depen- Needleman, A., Asaro, R.J., Lemonds, J., Peirce, D., 1985. Finite element analysis of
crystalline solids. Comput. Method. Appl. M. 52, 689–708.
dent constitutive relation is employed to model the mechan- Pan, J., Rice, J.R., 1983. Rate sensitivity of plastic flow and implications for yield-
ical behavior of the specimen and the crystal plasticity surface vertices. Int. J. Solids Struct. 19, 973–987.
theory extended by introducing a back stress is adopted to Phillips, A., Lu, W.Y., 1984. An experimental investigation yield surfaces and loading
of surfaces of pure aluminum with stress-controlled and strain-controlled paths
capture the grain level deformation behavior of the sub- of loading. J. Eng. Mater. Technol. 106, 349–354.
model. A comparison between the experimental and compu- Phillips, A., Tang, J.L., 1972. The effect of loading path on the yield surface at
tational results shows that the present cross-scale modeling elevated temperature. Int. J. Solids Struct. 8, 464–474.
Sung, S.J., Liu, L.W., Hong, H.K., Wu, H.C., 2011. Evolution of yield surface in the 2D
method can successfully capture the cyclic characteristics,
and 3D stress spaces. Int. J. Solids Struct. 48, 1054–1069.
the Bauschinger effect and other cyclic hysteresis phenom- Walker, K.P., 1981. Research and development program for non-linear structural
ena observed in the present experiments. modeling with advanced time-temperature dependent constitutive
relationships. Report PWA-5700-50, NASA CR-165533.
(3) For different unloading starting and end points after reach-
Wu, H.C., Yeh, W.C., 1991. On the experimental determination of yield surfaces and
ing a stable tension–compression cycle, the shape, size and some results of annealed 304 stainless steel. Int. J. Plast. 7 (8), 803–826.
position of subsequent yield surfaces in the stress space dis- Zhang, K.S., Shi, Y.K., Xu, L.B., Wei, D.K., 2011. Anisotropy of yielding/hardening and
play obvious deformation-induced anisotropic hardening. micro-inhomogeneity of deforming/rotating for a polycrystalline metal under
cyclic tension–compression. Acta Metall. Sin. 10 (47), 1292–1300 (in Chinese).
The subsequent yield surfaces are significantly affected by Zhang, K.S., Shi, Y.K., Ju, J.W., 2013. Grain-level statistical plasticity analysis on
the offset strain used to define the material yield. Although strain cycle fatigue of a FCC metal. Mech. Mater. 64, 76–90.

You might also like