Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

International Journal of Plasticity 114 (2019) 196–214

Contents lists available at ScienceDirect

International Journal of Plasticity


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

Cyclic hardening/softening behavior of 316L stainless steel at


T
elevated temperature including strain-rate and strain-range
dependence: Experimental and damage-coupled constitutive
modeling
Xue-fang Xiea, Wenchun Jianga,b,∗, Jingkai Chena, Xiancheng Zhangc,
Shan-Tung Tuc,∗∗
a
State Key Laboratory of Heavy Oil Processing, College of Chemical Engineering, China University of Petroleum (East China), Qingdao, 266580, PR
China
b
Key Laboratory of Unconventional Oil & Gas Development, School of Petroleum Engineering, China University of Petroleum (East China), Qingdao,
266580, PR China
c
Key Laboratory of Pressure Systems and Safety (MOE), School of Mechanical and Power Engineering, East China University of Science and
Technology, Shanghai, 200237, PR China

ARTICLE INFO ABSTRACT

Keywords: In this study, the cyclic mechanical characters of 316L stainless steel at elevated temperature are
316L stainless steel extensively investigated by the experimental and cyclic constitutive models. The experiments
Elevated temperature include the monotonic tensile tests with different loading rates and the low cycle fatigue tests
Cyclic hardening/softening considering the effect of strain amplitudes, strain rates and loading sequences. The evolution of
Damage-coupled cyclic constitutive model
cyclic stress amplitudes, hysteresis loops and elastic modulus under various loading conditions
are comprehensively analyzed. The experimental results show that the 316L steel at elevated
temperature performs a typical three-stage cyclic mechanical response, i.e., initial hardening,
subsequent saturation and final accelerated softening. The cyclic softening in both stiffness and
flow stress is mainly caused by the nucleation of micro-voids or micro-cracks, and the subsequent
coalesce and propagation. Furthermore, although the nearly rate-independent mechanical be-
havior is observed at monotonic tensile and first several fatigue cycles due to the DSA effect, the
cyclic hardening/softening behavior shows a significant strain-rate and loading history depen-
dence. Finally, inspired by the experimental observations and analyses, a damage-coupled cyclic
elastic-viscoplastic constitutive model involving strain-range, strain-rate and loading history
dependence is proposed to predict the complex cyclic behaviors of the material at elevated
temperature. A hardening factor is incorporated into the Chaboche kinematic hardening equa-
tions to model the kinematic-induced hardening behavior. And the plastic strain memory surface
and the maximum plastic strain rate are introduced to model the strain-range, strain-rate and
loading history dependence of cyclic behavior. The proposed model is proved to effectively de-
scribe the complex evolution of not only cyclic stress amplitude but also hysteresis loops for the
316L steel at elevated temperature.


Corresponding author. State Key Laboratory of Heavy Oil Processing, College of Chemical Engineering, China University of Petroleum (East
China), Qingdao, 266580, PR China.
∗∗
Corresponding author.
E-mail addresses: jiangwenchun@126.com (W. Jiang), sttu@ecust.edu.cn (S.-T. Tu).

https://doi.org/10.1016/j.ijplas.2018.11.001
Received 20 September 2018; Received in revised form 30 October 2018; Accepted 2 November 2018
Available online 05 December 2018
0749-6419/ © 2018 Elsevier Ltd. All rights reserved.
X.-f. Xie et al. International Journal of Plasticity 114 (2019) 196–214

1. Introduction

The 316L austenitic stainless steel has been extensively utilized for high temperature applications due to its excellent corrosion
resistance and mechanical properties at high temperature (Hong et al., 2003; Kang et al., 2006; Maziasz and Busby, 2012; Polák et al.,
2014). However, the structural integrity is threatened severely by the fatigue damage due to the frequent start-up and shut-down
operation, or the cyclic loading in service. Hence, a comprehensive understanding about the cyclic behavior at elevated temperature
is crucial for safety design, and it is even more desirable to develop a constitutive model to predict the complex cyclic performance
accurately for life assessment.
The cyclic performance of 316L stainless steel at room temperature has been extensively studied, including the cyclic deformation
response and its relationship with the microstructural evolution (Kang et al., 2010; Chang and Zhang, 2012a, b; Facheris et al., 2013;
Pham et al., 2013) and the damage evolution in multiaxial cyclic loading (Mazánová et al., 2017). But with the elevation of tem-
perature, the cyclic mechanical properties of material are changed significantly due to the variation of fatigue mechanism. Pham and
Holdsworth (2013) concluded that at 300 °C, when the dynamic strain aging (DSA) is quite active, a form of ladder-free deformation
bands in microstructure and secondary cyclic hardening in macroscopic stress are observed under cyclic loading, which result in a
dramatic reduction of life. The DSA comes from the interaction between the diffusion of solute atoms and the mobile dislocation
during plastic deformation (Hong and Lee, 2004), and is quite sensitive to the temperature and loading rate. The DSA effect can not
only weaken the rate-dependent properties of material but also strengthen the cyclic hardening behavior (Kang et al., 2006; Yu et al.,
2012). With the further elevation of temperature, the fatigue failure mechanism becomes more complex due to the effect of creep,
oxidation, DSA and their interaction (Srinivasan et al., 1999, 2003). The cyclic performance of 316L stainless steel beyond 600 °C has
been studied by Srinivasan et al. (1999 and 2003), Hong and Lee (2004), and Hormozi et al. (2015). However, aforementioned
literature mainly focuses on the macroscopic stress amplitude and the final fractography, while the evolutional hysteresis loops, the
strain-rate dependence of cyclic hardening/softening and the cyclic constitutive model require an comprehensive investigation.
The constitutive relations of materials under cyclic loads have been investigated since several decades ago. Recently, the cyclic
crystal plasticity models based on the underlying physical micromechanism are attracting broad attentions, including Yu et al.
(2014), Li et al. (2016), Castelluccio and Mcdowell (2017), Guan et al. (2016), Rodas et al. (2017), and et al. But the modeling process
and parameter identification are quite complex, especially when applied to the engineering structures. Alternatively, a cyclic con-
stitutive model can be developed based on the internal stress state variables theories, which are usually simple but contain some sense
of physical meaning. The kinematic and isotropic components, which are respectively associated with the long- and short-range
dislocation interactions (Zhao and Xuan, 2011; Pham et al., 2013; Zhou et al., 2018), are generally incorporated into those models to
represent the anisotropic resistance to inelastic deformation and the domains of elastic deformation, respectively. And the devel-
opments of kinematic hardening laws are attracting more attention, as the isotropic hardening has been proved to be not a major
cyclic plasticity property for many engineering material (Jiang and Zhang, 2008; Zhang and Jiang, 2008; Krishna et al., 2009; Chang
and Zhang, 2012a, b; Pham et al., 2013; Ahmed et al., 2016; Zhou et al., 2018; Chen et al., 2018). Armstrong and Frederick (1966)
introduced a dynamic recovery term into a linear kinematic hardening formulation initially proposed by Prager (1949), and
Chaboche et al. (1979) further decomposed the back stress into several components to improve the description of transient hardening
behavior. After that, the dynamic recovery term has been further employed and developed to better predict the accumulation of the
inelastic strain under the ratcheting condition by Chaboche (1991), Ohno and Wang (1993a, b), Abdel-Karim and Ohno (2000), Kang
et al. (2004, 2005), Abdel-Karim (2005, 2009, 2010), and so on. Meanwhile, a static recovery term, which is originally proposed by
Malinin and Khadjinsky (1972), has been added into the nonlinear kinematic hardening law to describe the thermally activated
mechanism behavior such as creep deformation or stress relaxation (Yoshida, 1990; Yaguchi and Takahashi, 2005; Zhang and Xuan,
2017; Chen et al., 2018). And some other complex cyclic features have been involved into the corresponding extended cyclic con-
stitutive models, including the plastic strain memory effect (Chaboche et al., 1979; Ohno, 1982; Kang et al., 2002; Zhou et al., 2018),
thermal effect (Morin et al., 2011; Yu et al., 2015; Wang et al., 2017), and so on. Among those features, the incorporation of the DSA
effect has always been a challenge due to its complex association with many factors, including the initial microstructure, environ-
mental temperature, loading rate, viscoplastic proprieties and cyclic hardening behavior. Yaguchi and Takahashi (2000) defined an
aging stress as the function of the temperature and loading rate to reflect the effect of DSA, but a complex implicit algorithm was
introduced in the plastic flow formulation. Yu et al. (2012) presented the DSA effect into the saturated values of isotropic stress by an
exponential relationship with temperature in the framework of a thermo-viscoplastic modeling, but the rate dependence of cyclic
hardening behavior through DSA effect was not discussed. The further modeling investigation of DSA effect was performed by
Chaboche et al. (2013), Huang et al. (2014, 2016), and so on. However, the theoretical and experimental relationship between
loading rate, DSA effects and cyclic hardening behavior is still not well clarified and require an intensive investigation.
Furthermore, the aforementioned models have a common barrier in simulating the cyclic softening behavior along with cyclic
hardening due to that their kinematic or isotropic variables progressively approach to constant values. Thus, a memory evanescence
term was introduced by Nouailhas et al. (1985) to model the cyclic softening behavior for the cold-worked material, and was further
developed by Krishna et al. (2009), Ahmed et al. (2016) and so on. Recently, Xu et al. (2016) incorporated a logistic function into the
asymptotic values of the kinematic hardening and isotropic hardening parameters to trace the evolution of softening for a low yield
point BLY160 steel. Zhou et al. (2018) introduced a kinematic hardening coefficient into the nonlinear kinematic hardening rule to
incorporate the hardening/softening effect in the back stress for the 316L steel at room temperature. However, for the material at
elevated temperature, the nucleation of micro-voids or micro-cracks, and the subsequent coalesce and propagation play a crucial role
in the cyclic softening, which can not be represented reasonably by the above models. Since the continuum damage mechanics
initially proposed by Kachanov (1958) and Rabotnov (1968) introduce a continuum damage variable as a measurement of micro-

197
X.-f. Xie et al. International Journal of Plasticity 114 (2019) 196–214

Table 1
Chemical compositions of 316L stainless steel (%).
Element C Si Mn P S Ni Cr Mo N

Content 0.025 0.41 1.41 0.025 0.025 10.22 16.16 2.09 0.043

cracks and micro-voids in the material, it has been extensively employed in the long period failure problems by coupling damage
variable into different constitutive models, including the fatigue, the ratcheting or the creep-fatigue interaction (e.g. Voyiadjis and
Dorgan, 2007; Kang et al., 2009; Zhu and Sun, 2013; Xu et al., 2017). Kang et al. (2009) constructed a visco-plastic cyclic constitutive
model by coupling an isotropic damage variable based on thermodynamics to simulate the whole-life ratcheting behavior. The
simulated results are in a fairly good agreement with the experiments. However, the damage evolution model for 316L steel at
elevated temperature and its experimental validation require further investigation.
In this paper, a systematic set of experiments are performed to investigate the cyclic behavior of 316L stainless steel at 650 °C, and
the corresponding constitutive model is proposed. First, the monotonic tensile tests with different loading rates and the low cycle
fatigue (LCF) tests with various strain ranges, strain rates and loading sequences are conducted to characterize the mechanical
performance of the material. Then, the strain-range, strain-rate and loading history dependence on cyclic hardening/softening be-
havior is discussed in detail. Finally, a damage-coupled cyclic elastic-viscoplastic constitutive model with novel hardening laws and
damage evolution equation is proposed to predict the development of both stress amplitude and hysteresis loops under complex cyclic
loading. The capability of the proposed model is verified by the experimental results.

2. Experimental

2.1. Experimental details

The chemical compositions of the 316L steel are listed in Table 1, which has been solution treated at 1070–1100 °C and then water
quenched. The detailed sample dimensions are shown in Fig. 1.
The low cycle fatigue (LCF) tests were conducted on a MTS servo-hydraulic fatigue machine. A high-temperature axial ex-
tensometer with a clip gauge of 12 mm was used to monitor and control the strain, and the magnitudes of axial loads imposed on the
specimen were measured by a force sensor. A furnace with two pairs of silicon-carbide heating elements was used to heat the
specimens. In order to achieve a stabilized temperature throughout the gauge section of the specimen, two K-type thermocouples
were attached directly at two sides of the gauge length. The temperature fluctuation was controlled within ± 3 °C at steady state.
Besides, the specimen needed to remain prescribed temperature for 30 min before mechanical loading.
The monotonic tensile and LCF tests were conducted at 650 °C. Three levels of strain rate were considered at tensile tests, i.e.
1 × 10−3s−1, 1 × 10−4s−1 and 1 × 10−5s−1. And the triangular strain-controlled cyclic loading wave was imposed with the fully-
reversed strain amplitudes of ± 0.4%, ± 0.5%, ± 0.6% and ± 0.7% at a constant strain rate of 5 × 10−3 s−1. Besides, different strain
rates within 5 × 10−4 to 1 × 10−2 s−1 were imposed to a constant strain amplitude ± 0.6% to investigate the rate-dependent of
cyclic properties. To investigate the strain rate history dependence of the material, the multiple steps tests with different strain rates
were also carried out. Note that to compare the difference of cyclic response of 316L steel at ambient and elevated temperature, the
LCF tests under the strain amplitudes of ± 0.4%, ± 0.5%, ± 0.6% and ± 0.7% at a constant strain rate of 5 × 10-3 s−1 were also
performed at room temperature.

Fig. 1. The detailed dimension of fatigue sample.

198
X.-f. Xie et al. International Journal of Plasticity 114 (2019) 196–214

Fig. 2. Comparison of strain-stress relationship with different strain rates.

2.2. Experimental results

2.2.1. Monotonic behavior


A comparison of strain-stress relationship with different strain rates is illustrated in Fig. 2. It is clear that the curves at elastic
domains are identical, and only small differences are observed at the low plastic strain region. Hence, the 316L steel at elevated
temperature presents approximately rate-independent tensile proprieties, which is induced by the DSA effect as reported by Kang
et al. (2006) and Yu et al. (2012).

2.2.2. Cyclic stress response under a constant strain rate


The cyclic stress response of 316L steel for different strain amplitudes with a constant strain rate of 5 × 10−3/s at room and
elevated temperature is shown in Fig. 3. The tensile stress amplitude is plotted with respect to the number of cycles. It is easily found
that, on the one hand, a higher temperature indicates a smaller stress response and a shorter fatigue life for a same strain amplitude,
on the other hand, for a given temperature, the stress amplitude increases gradually but the fatigue life decreases significantly with
the increase of strain amplitude. In addition, at elevated temperature, the cyclic stress response during the whole-life fatigue exhibits
three typical stages regardless of the strain amplitude, i.e., the initial cyclic hardening, subsequent saturation and accelerated cyclic
softening until the final rupture. The higher hardening/softening rate and shorter saturation period are observed for the larger strain
amplitude. However, at room temperature, no cyclic saturation stage but a very remarkable softening behavior followed by initial
hardening is observed, and even a special secondary cyclic hardening phenomenon occurs for the larger strain amplitudes. More than
that, the primary hardening behavior at 650 °C is much stronger than that at ambient temperature by increasing hardening rate and
prolonging hardening period. The different mechanical behavior of the material at various temperature is caused by the different
fatigue mechanism. Considering that the cyclic behavior and corresponding mechanism of 316L at room temperature has been

Fig. 3. Cyclic stress response of 316L steel at ambient and elevated temperature.

199
X.-f. Xie et al. International Journal of Plasticity 114 (2019) 196–214

Fig. 4. The cyclic hardening (a) and softening (b) for a typical specimen with the strain amplitude ± 0.6% at the rate 5 × 10−3/s.

extensively researched by Kang et al. (2010), Pham et al. (2013), and so on, the elevated temperature cyclic characters of the material
is investigated in the following sections.

2.2.3. Cyclic hardening/softening behavior


To further illustrate the variation of cyclic performance of the material during the cyclic hardening/softening process, the hys-
teresis loops are investigated meticulously in this section. For comparison purpose, the hysteresis loops at different cycles are moved
to their minimum points, as shown in Fig. 4(a), which describes the cyclic hardening of a typical specimen with the strain ampli-
tude ± 0.6% at the rate 5 × 10−3/s. Obviously, the stress amplitude increases continuously as the plastic strain accumulates, which
is consistent with the character of cyclic hardening. Note that the elastic comprehensive modulus Ec and elastic domains 2R are nearly
identical for different cycles, which is also clearly presented in cyclic softening phase (see Fig. 4(b)) and saturated hysteresis loops for
different strain amplitudes (see Fig. 5). If the offset plastic strain is considered as 0.01%, the corresponding values of the elastic
domains at different loading condition fluctuate around 200 MPa. This indicates that the cyclic hardening/softening and strain-range
dependence of the material are mainly caused by their kinematic components, and the isotropic stress can be modeled as a tem-
perature-dependent constant. Furthermore, there is another interesting observation in Fig. 4(b) that not only the stress amplitude but
also elastic tensile modulus Et decreases significantly during cyclic softening, which is neglected in many traditional cyclic con-
stitutive models. Note that the hysteresis loops during cyclic softening phase are compared by collocating their maximum points. This
is because the mechanical performance of the material at tensile and compressive state is significantly different as the fatigue damage
accumulates (see Fig. 7), and the tensile properties get more concerned in this paper due to the fact that the fatigue fracture always
occurs at tensile stage. In addition, as shown in Fig. 5, the hardening behavior of the material performs a significant strain-range
dependence.

2.2.4. Elastic modulus and fatigue damage


As observed in Fig. 4(b), the elastic modulus decreases significantly at cyclic softening phase. To further illustrate the evolutional

200
X.-f. Xie et al. International Journal of Plasticity 114 (2019) 196–214

Fig. 5. The saturated hysteresis loops for different strain amplitudes.

Fig. 6. The evolutional elastic modulus under different strain amplitudes with a constant rate of 5 × 10−3/s.

Fig. 7. The morphology of the ruptured fatigue specimens with strain amplitude of ± 0.6% and strain rate of 5 × 10−3/s at room temperature
(lower) and 650 °C (upper).

law of elastic modulus, Fig. 6 describes the quantitative elastic modulus determined by the least square method during the whole-life
fatigue under different strain amplitudes. It is apparent that the evolution of both Et and Ec goes through two stages regardless of the
fatigue amplitude, i.e. the initial steady stage and the following softening stage. And the longer steady period and slower decreasing
rate are observed for lower strain amplitudes. In addition, the steady values of Et and Ec for different strain amplitudes fluctuate at
150 GPa, but Et is slightly bigger than that of Ec at the softening phase due to the close of some micro-cracks which are perpendicular
to the loading direction when compressing. This indicates that the elastic modulus can be modeled as a strain-range independent and
isotropic material parameter by ignoring the effect of micro-cracks closure.

201
X.-f. Xie et al. International Journal of Plasticity 114 (2019) 196–214

Fig. 8. The relationship between tensile stress amplitude and fatigue damage for the strain amplitude ± 0.6% with strain rate 5 × 10−3/s.

Furthermore, Fig. 7 describes the morphology of ruptured fatigue specimens with strain amplitude of ± 0.6% and strain rate of
5 × 10−3/s at room temperature and 650 °C. It is clear that lots of distributed small cracks are observed throughout the gauge section
of the specimen at elevated temperature, while only one dominant crack leads to the final rupture at room temperature. This indicates
that, under the effect of elevated temperature, the nucleation of distributed micro-voids or micro-cracks and their subsequent coa-
lesce and propagation play a vital role on the degradation of elastic modulus by reducing the effective loading areas, which arouses a
macroscopic cyclic softening behavior. However, according to the microstructural observation by Pham et al. (2013), the cyclic
softening of 316L steel at room temperature is due to the rearrangement of dislocation by the secondary slip, which results in the
formation of dislocation walls/channels. This indicates that the cyclic softening at ambient and elevated temperature is respectively
induced by dynamic recovery and fatigue damage, and a comprehensive elevated temperature cyclic constitutive model must take the
role of fatigue damage into consideration.
Then, inspired by Lemaître and Desmorat (2005) and Kang et al. (2009), an isotropic fatigue damage variable D is defined by the
reduction of Et, as following:
E0 Eit Eit
D= =1
E0 E0 (1)

Eti
where E0 is the initial steady elastic modulus, and is the elastic tensile modulus at the ith cycle. Then, the relationship between the
cyclic stress amplitude and the fatigue damage at 650 °C is illustrated by performing a typical test with strain amplitude ± 0.6% and
strain rate 5 × 10−3/s, shown in Fig. 8. It is seen that the fatigue damage is negligible during the cyclic hardening. As the micro-
structure observation by Pham et al. (2013), the cyclic hardening behavior is aroused from the increase of total dislocation density.
Note that this evolution of microstructure is insufficient to introduce a significant evolution of elastic modulus, hence the values of
fatigue damage determined based on Eq. (1) is approximately zero in this stage. But for the subsequent long-time softening phase, the
stress amplitude decreases gradually with the increase of fatigue damage. It is also worth noting that the drastic reduction of cyclic
stress amplitude is observed after 820 cycles, which is due to the propagation of a macro-crack.

2.2.5. Cyclic behavior with different strain rates


Fig. 9 illustrates the evolution of tensile stress amplitude for a constant strain amplitude ± 0.6% with the rate range from
5 × 10−4/s to 1 × 10−2/s. It is obvious that the typical three-stage cyclic response (hardening, saturation and softening) is still
observed for all investigated strain rates. Little difference is observed at first 10 cycles, but the saturated cyclic stress amplitude is
significant rate-dependent, that is, the greater strain rate indicates a higher stress amplitude for the rate range from 5 × 10−4/s to
5 × 10−3/s. But the saturated stress of 1 × 10−2/s is slightly lower than that of 5 × 10−3/s. Besides, comparing the evolutional
curves at the strain rate of 5 × 10−3/s and 1 × 10−2/s, the cyclic hardening phase is very close, but the fatigue lifetime at strain rate
5 × 10−3/s (820 cycles) is far less than that at strain rate 1 × 10−2/s (1261 cycles), which indicates that the cyclic softening also
performs strain-rate dependence, i.e., a greater strain rate indicates a slower softening rate, which leads to a longer lifetime, for whole
investigated strain rate. As stated in section 2.2.3, at elevated temperature, the nucleation of micro-voids or micro-cracks, and the
subsequent coalesce and propagation leads to the cyclic softening behavior. Hence, it can be accepted that the time for the specimen
at high stress level in one cycle is longer than that at a slower loading rate, which arouses the sufficient growth of micro-defects, thus
a faster cyclic softening behavior is observed.
To further reveal the strain-rate sensitivity of cyclic behavior, the hysteresis loops at different cycles for various strain rates is
described in Fig. 10. On one hand, for the three larger rates of 2.5 × 10−3/s, 5 × 10−3/s and 1 × 10−2/s, the drastically serrated
stresses are observed at the first cycle. And the serrated stresses disappear completely after the 10th cycles for the rates 2.5 × 10−3/s
and 5 × 10−3/s, but they are presented during the whole life for the rate 1 × 10−2/s. As reported by Yu et al. (2012) and Pham et al.
(2013), the serrated stress is regarded as the manifestation of the DSA. This indicates that the DSA depends heavily on the strain rate

202
X.-f. Xie et al. International Journal of Plasticity 114 (2019) 196–214

Fig. 9. The tensile stress amplitude for a constant strain amplitude ± 0.6% with various strain rates.

and is more pronounced at higher rates. The variation of cyclic stress amplitude under different strain rates also indicates the DSA can
strengthen the cyclic hardening behavior, but the effect will be saturated when the strain rate is bigger than 5 × 10−3/s. On the other
hand, the hysteresis loops for various rates are almost identical at the 1st and 10th cycles, which indicates the rate-dependent
properties of the material are significantly weakened by the DSA effects. But they subsequently diverge at the 100th cycle. The deeper
analysis about the rate dependence for saturated hysteresis loops is illustrated in Fig. 11, which moves the loops to their minimum
points. As shown in Fig. 11, the elastic domains and elastic modulus are nearly identical regardless of the strain rates, which indicates
that they can be modeled as a rate independence material parameter, but the kinematic stress is very sensitive to the strain rate, i.e.,
the kinematic stress increases with the increase of strain rate, except for the rate of 1 × 10−2/s.

2.2.6. Multiple-step tests with different strain rates


To investigate the dependence of loading history, a multiple loading tests with a constant strain amplitude ± 0.6% but different
strain rates was performed. The loading scheme is divided into three phases: a prior lower rate (1 × 10−3/s) followed by a higher rate
(5 × 10−3/s), and then return to the low rate (1 × 10−3/s), and each loading level maintains 30 cycles. The experimental result is
shown in Fig. 12. The stress amplitude increases continuously and reaches a plateau in the prior phase, then the additional hardening
is observed in the second phase, and the saturated value is nearly same with that of the test with a constant strain rate of 5 × 10−3/s.
It is indicated that the prior loading history in a lower strain rate has a small effect on the saturated stress amplitude. But when the
strain rate return to the prior rate, the amplitude stress cannot return to the prior saturated level. This is because that the material has
been hardened in the second phase. A similar phenomenon has been reported by Xu et al. (2016) when investigating the dependence
of loading history with different strain amplitude for the low yield point steel BLY160.

3. Cyclic constitutive model

3.1. Framework of model

As mentioned before, although the approximate rate-independent mechanical behavior is observed at monotonic tensile and first
several cycles due to the DSA effect for 316L steel at elevated temperature, the cyclic hardening and softening behavior exhibits a
significant rate-dependent propriety. Furthermore, the role of fatigue damage in the cyclic softening should not be ignored in the
comprehensive constitutive model. Hence, a modified damage-coupled cyclic model is proposed based on the framework of the
elastic-viscoplastic model.
First, the total strain is decomposed into the elastic and plastic strain under the assumption of small deformation:
= e + p (2)

where the ɛe and ɛp represent the elastic and plastic strain tensors, respectively. Then, the Cauchy stress in classical constitutive model
is replaced by the effective stress based on the strain equivalent principle (Lemaître and Desmorat, 2005). Thus, the elastic law
coupled with the isotropic damage variable D is expressed as:

= (1 D ) C: e
(3)

where C is the fourth-order elastic tensor determined by Young's modulus E and Poission's radio μ. And the Von Mises yield rule is
applied to define yield function, as:

203
X.-f. Xie et al. International Journal of Plasticity 114 (2019) 196–214

Fig. 10. The hysteresis loops at different cycles for various strain rates: (a) the first cycle, (b)10th cycle and (c)100th cycle.

f=
3 s
2 1 D
X
: 1
s
D
X R
(4)

where s and X are the deviatoric and the back stress tensors, respectively. R is the isotropic (effective) stress representing the radius of
the yield surface. This means that the elastic domains in a typical hysteresis loops are the twice of the isotropic stress, as shown in
Fig. 5. The flow rule is defined as:

204
X.-f. Xie et al. International Journal of Plasticity 114 (2019) 196–214

Fig. 11. The rate sensitivity of cyclic behavior for saturated hysteresis loops.

Fig. 12. The evolution of stress amplitude under multiple loading steps.

p f 3 d s /(1 D) X
d =d =
2 1 D s /(1 D) X (5)

d f n
dp = =< >
1 D K (6)

where the λ and p are plastic multiplier and accumulated equivalent plastic strain, respectively. K and n are the rate-dependent
material parameters. Considering that the near rate-independent behavior is observed during the monotonic tensile phase and first
several fatigue cycles, a smaller K and a bigger n are adopted in this study (K = 10, n = 50).

3.2. Cyclic hardening

The mixture hardening model including kinematic and isotropic hardening equations is generally applied to model different cyclic
characters, for example, the cyclic hardening (Lee et al., 2014), strain-range dependence (Kang et al., 2002; Xu et al., 2016; Zhang
and Xuan, 2017), strain history (Xu et al., 2016) and etc. However, it has been found that, for the 316L steel at elevated temperature,
the isotropic phase is nearly unchanged during cyclic hardening and softening (Fig. 4) and can be regarded as strain-range in-
dependent (Fig. 5) and rate-independent (Fig. 11). Hence, in this study, the isotropic stress is modeled as a constant parameter, while
the kinematic hardening equations coupled with fatigue damage are proposed to characterize the cyclic hardening and the strain-
rate, strain-range and loading history dependence.
The nonlinear kinematic hardening Chaboche model, which contains a linear hardening term, a dynamic recovery term and a
static recovery term, is expressed as:

205
X.-f. Xie et al. International Journal of Plasticity 114 (2019) 196–214

N
X= Xk
k=1 (7)

dXk = Ck d p
k Xdp ak [J ](X k ) vk 1Xk (8)

where k is the number of back stress, and Ck, γk, ak and vk are material parameters. Ck is the initial kinematic hardening modulus, and
γk determines the rate at which the kinematic hardening modulus decreases with the increase of plastic deformation. The third term
(the static recovery term) is built for modeling the thermal recovery effect, such as the creep behavior at a constant stress level or the
stress relaxation phenomenon at a constant strain level, which is highlighted in Zhang and Xuan (2017) when modeling the creep-
fatigue behavior of 9–12% Cr steel, but not considered in this study, thus it is not discussed in our model. Then, according to Kang
et al. (2009), when the temperature and field variable dependencies are omitted, the coupled-damage Chaboche model is expressed
as:
dXk = Ck (1 D) d p
k Xdp (9)

However, the cyclic hardening induced by kinematic component cannot be reproduced by this model, because the constant
hardening modulus is included. Therefore, inspired by Zhou et al. (2018), a new nonlinear kinematic rule is proposed by introducing
a hardening factor δ, as following:
dXk = Cks (p)(1 D) d p
k Xdp (10)

d = (1 ) dp with 0 = (11)

Here, Cks is the hardening modulus of saturated cycles. δ is defined as the exponential function of equivalent plastic strain with an
initial value φ which reflects hardening behavior for the first cycle, and ω determines the hardening rate, as depicted in Eq. (11).

3.3. Strain-range, strain-rate and loading history dependence

It has been stated that the strain-range dependence of cyclic behavior for 316L steel at elevated temperature is presented at the
nonlinear hardening range for the larger strain, but not the elastic domains or transitioning hardening areas (see Fig. 5). And
according to the Chaboche (1986), three back stresses are proposed to recover the hardening behavior at different length scales, i.e.,
the 1st, 2nd and 3rd back stress take change of the elastic-to-plastic transitioning hardening at small strains, the nonlinear hardening
at the middle plastic strains and nearly linear hardening behavior for large plastic strains, respectively, as shown in Fig. 13. Hence,
this paper develops the saturated hardening modulus of the 2nd back stress as the function of q, as following:

C2s = A1 + B1 (1 e C1 q
) (12)

where A1, B1 and C1 are material parameters. And q is determined by the plastic strain memory surface proposed by Chaboche et al.
(1979). The equations of the plastic strain memorization are shown as following:

G=

d = (1
2
3
( p )
: ( p

) H (G ) n: n n dp
) q2 0
(13)

(14)

Fig. 13. The comparison between experimental data and fitting lines for the strain amplitude of ± 0.6%.

206
X.-f. Xie et al. International Journal of Plasticity 114 (2019) 196–214

Table 2
Formulations of the proposed cyclic constitutive model.
Framework of the model:

= e

d
+

p =
p;

3 d
= (1
s / (1
D) C :
D)
e;

X
f=

; dp =
3
( s
2 1 D
d
=<
X
f n
>
:
) ( 1
s
D
X ) R;

2 1 D s / (1 D) X 1 D K
Kinematic hardening laws:
dXk = Cks (p, pmax )(1 D ) d p
k Xdp k = 1,3 ;
2 Xdp ;
dX2 = C2s (q) (p)(1 D) d p

d = (1 ) dp m (pmax )
with 0 =
= m (pmax ) > m (pmax )
Strain-range dependence:
C2s = A1 + B1 (1 e C1 q) ;

Strain-rate and loading history dependence:


G=
2 p
3
( )
: ( p ) q2 0 ; dq = H (G ) n: n dp

m (pmax ) = 1 H (pc pmax )(A2 + B2 (1 e C2 p max ))

Damage evolution laws:


Rv (dp)r + 1 2
dD = ; = A3 + B3 (1 e C3 pmax ) ; Rv = (1 + µ) + 3(1 2µ) m
(r + 1) (Dc D ) 3 e

Fig. 14. The comparison between fitting results of damage parameter and experimental data.

Table 3
Material parameters of the proposed model.
Isotropic parameter:
E = 150.0 GPa, ν = 0.3, R = 100 MPa
Kinematic hardening rule:
C1s = 48705.1 MPa, γ1 = 478.9, γ2 = 3746.5, C3s = 20 MPa, γ3 = 5.0, φ = 0.33, ω = 8.0
Strain-range dependence:
η = 0.5, A1 = 153549.8 MPa, B1 = 281956.4 MPa, C1 = 292.6
Strain-rate dependence:
pc = 5 × 10−3/s, A2 = 0.225, B2 = 0.049, C2 = −378.5
Damage evolution rule:
Dc = 0.25, r = 1.3, β = 7.0, A3 = 115.4, B3 = 19.4, C3 = −152.4

dq = H (G ) n: n dp (15)

where ψ and q represent the center and radius of the plastic strain memory surface G, respectively. η is the evolution rate, and the
instantaneous stable memorization will be achieved after one cycle when η = 0.5. The H(x) donates the Heaviside function, which
means H(x) is equal to 1 when x is bigger than 0, otherwise, H(x) is 0. n and n are the normal vectors of the yield stress surface and
the plastic strain surface.
Furthermore, to reproduce the strain rate and loading history dependence of the cyclic hardening behavior, the kinematic
hardening laws is further developed in this section. Based on the fact that the prior loading history in a lower strain rate has a small
effect on the saturated stress amplitude and it is determined by the maximum strain rate, which has been proved in section 2.2.6, Eq.

207
X.-f. Xie et al. International Journal of Plasticity 114 (2019) 196–214

Fig. 15. The tensile stress amplitude for different strain amplitudes with a constant strain rate of 5 × 10−3/s.

(11) is further developed by introducing a limited maximum values m, which is defined as the function of the maximum plastic strain
rate pmax , as following:

d = (1 ) dp m
=m >m (16)

m=1 H (pc pmax ) (A2 + B2 (1 e C2 pmax


)) (17)
where the pc is the critical plastic strain rate and is introduced to describe the saturated effect of DSA on cyclic hardening when strain
rate is greater than pc .

3.4. Cyclic softening and damage evolution

Based on the continuum damage mechanism, a damage evolution equation should be coupled into the cyclic constitutive model to
characterize the evolutional law of the fatigue damage under cyclic loading. As the plastic strain dominates the fatigue life of material
within the regime of low cycle fatigue, Lemaître and Desmorat (2005) proposed a widely-used low cycle fatigue damage evolution
equation, as following:
Rv (dp)r + 1
dD =
(r + 1) (1 D ) (18)
where Ω, r and β are material parameters, and Rv is the effect of stress triaxiality, defined as:
2 m
Rv = (1 + µ) + 3(1 2µ )
3 e (19)
where μ is the Poison's ratio. σm and σe are the hydrostatic and equivalent stresses, respectively. The Eq. (19) indicates a complete
failure when the damage variable reaches 1. However, as shown in Fig. 8, the critical damage values corresponding to the initiation of
the macro-crack is far less than 1. And note that the subsequent propagation is within the domain of the fracture mechanics but
beyond the scope of cyclic constitutive model. Thus, inspired by Kang et al. (2009), the damage evolution equation is developed by
introducing a critical fatigue damage Dc. Furthermore, the damage coefficient Ω is defined as the function of pmax to present the
strain-rate dependence of cyclic softening, as following:
Rv (dp)r + 1
dD =
(r + 1) (Dc D) (20)

= A3 + B3 (1 e C3 pmax ) (21)
where A3, B3 and C3 are material parameters. All formulations of the proposed model are summarized in Table 2.

4. Model validation

4.1. Identification of parameters

A total of 24 parameters are involved in the proposed model including kinematic hardening, strain-range dependence, strain-rate

208
X.-f. Xie et al. International Journal of Plasticity 114 (2019) 196–214

Fig. 16. The hysteresis loops at different cycles for a typical strain amplitude of ± 0.6%: (a)1st, (b)10th, (c)100th, (d) 500th, and (e)700th cycle.

dependence and fatigue damage evolution. The steps of determining a full set of model parameters are elaborated in the following.
First, the saturated loops of different strain amplitudes under constant strain rate are applied to determine the parameters of
kinematic hardening rule. By ignoring the fatigue damage, the integration of Eq. (9) in a half loop is shown as following:

Cks
Xk = (1 e kp ) + Xk,0 e kp
k (22)

where Xk,0 is the back stress component with zero equivalent plastic strain. Fig. 13 gives a comparison between experimental data and
fitting lines by taking the strain range of ± 0.6% as example. Hence, the values of Cks and γk are firstly determined, and the different
values of C2s for various strain amplitudes are applied to identify the strain-range parameters A1, B1 and C1 in terms of Eq. (12). The
hardening parameters φ, ω and m are determined by the trail-and-error method with the accepted accuracy. For simplification, the pc

209
X.-f. Xie et al. International Journal of Plasticity 114 (2019) 196–214

Fig. 17. The saturated hysteresis loops for different strain amplitudes.

Fig. 18. The comparison of cyclic stress amplitudes under a constant amplitude with various strain rates from simulation and experiments.

is determined based on the test with strain rate of 5 × 10−3/s. Then, the rate-dependent parameters A2, B2 and C2 are identified
based on the determined values of m under different strain rates in terms of Eq. (17).
Then, the parameters in fatigue damage evolution model are identified. Based on the assumption that the fatigue damage D within
one loading cycle is constant, Eq. (20) can be written as:

dD Rv ( p) r + 1
=
dN (r + 1) (Dc D ) (23)

where Δp represents the variation of equivalent plastic strain in one cycle. Integrating Eq. (23) and combining the boundary con-
ditions (D = 0 for N = 0 and D = Dc for N=Nf), for the uniaxial condition (Rv = 1), it can be obtained as:

(r + 1) 1 +
( p)r + 1Nf = Dc
1+ (24)

The Eq. (24) is the damage-coupled Masson-coffin equation. Then, Eq. (23) can be also expressed as a function of the number of
cycles by integrating from N (N < Nf) to Nf and D to Dc:

1 1
(Dc D )1 + = ( p)r + 1 (Nf N)
1+ (r + 1) (25)

Substituting Eq. (24) into (25), it is described as:

210
X.-f. Xie et al. International Journal of Plasticity 114 (2019) 196–214

Fig. 19. The saturated hysteresis loops for the strain amplitude of ± 0.6% with various strain rates: (a)1 × 10−2/s, (b)2.5 × 10−3/s, (c)1 × 10−3/s
and (d)5 × 10−4/s.

Fig. 20. The comparison of stress amplitude evolution under multiple loading steps condition obtained by experiments and simulation.

1
N 1+
D = Dc Dc 1
Nf (26)

Therefore, the β is firstly calibrated by fitting Eq. (26) and experimental relationship between damage and cycles, as shown in Fig.
(14). Then the Ω and r are subsequently obtained by Eq. (24) based on the relationship between the plastic strain amplitude and

211
X.-f. Xie et al. International Journal of Plasticity 114 (2019) 196–214

fatigue life. Finally, the rate-dependent parameters A3, B3 and C3 are identified based on the determined values of Ω under different
strain rates in terms of Eq. (21). All the determined material parameters are listed in Table 3.

4.2. Simulation for the constant strain rate

The proposed damage-coupled cyclic plasticity constitutive model was implemented into the finite element code ABAQUS by a
subroutine UMAT based on an implicit integration algorithm. The symmetry boundary conditions and the CAX4 elements (4-nodes
axisymmetric) were applied in the finite element model. First, the calculated results of the evolutional tensile stress amplitude for a
constant strain rate of 5 × 10−3/s were compared with the experimental data, as shown in Fig. 15. It is easily observed that three
typical cyclic characters (hardening, saturation and softening) are simulated well by the proposed model, and a great agreement is
obtained for all investigated strain amplitudes. The comparison of the hysteresis loops shown in Fig. 16 indicates that although the
saturated stress phenomenon cannot be simulated by the proposed model for the first cycle, the overall trend of hysteresis loops from
the simulation agrees well with experimental data at both elastic stiffness and plastic flow. Furthermore, the saturated hysteresis
loops for different cyclic amplitudes are compared in Fig. 17. It is obvious that the experimental points are nearly all located on the
simulated lines, which indicates that the strain-range dependence of kinematic component is well recovered by the proposed model.

4.3. Simulation for different strain rates

The capability of reproducing strain-rate dependent cyclic behavior is validated in this section. The evolutions of cyclic stress
amplitudes and saturated hysteresis loops under a constant amplitude of ± 0.6% with various strain rates from 5 × 10−4/s to
1 × 10−2/s are compared in Figs. 18 and 19, respectively. Note that the hysteresis loops at strain rate 5 × 10−4/s has been validated
in Fig. 17. As shown in the figures, the simulated results meet the experimental data very well, which indicates that the strain-rate
dependence of cyclic behavior is fully considered in the proposed model. Besides, the multiples loading step tests with a constant
strain amplitude ± 0.6% but three different strain rates, i.e., a prior lower rate (1 × 10−3/s) followed by a higher rate (5 × 10−3/s),
and then return to the low rate (1 × 10−3/s), was reproduced by the proposed model, and the results were shown in Fig. 20. It is clear
that the effect of prior loading on the subsequent cyclic hardening behavior has been simulated well by the proposed model.

5. Conclusion

In this study, a systematic set of monotonic tensile and LCF tests has been performed to investigate the cyclic hardening/softening
behavior of 316L stainless steel at 650 °C under the influences of strain amplitude, strain rate and loading consequence. The role of
fatigue damage and DSA effect on the cyclic behavior has been extensively analyzed. Lastly, a damage-coupled constitutive model has
been proposed to predict the LCF responses up to failure at elevated temperature. The main conclusions are listed as follows:

(1) The cyclic response of the material at ambient and elevated temperature are different. The cyclic hardening phenomenon is more
pronounced at elevated temperature, and the cyclic softening behaviors at ambient and elevated temperature are induced by
dynamic recovery and fatigue damage, respectively.
(2) The fatigue damage variable is determined experimentally by the degradation of elastic tensile modulus and is incorporated into
the proposed cyclic elastic-viscoplastic constitutive model.
(3) Although the nearly rate-independent mechanical behavior is observed at monotonic tensile and first several cycles due to the
DSA effect for the material at elevated temperature, the cyclic hardening and softening behavior are strongly rate-dependent for
the remaining load cycles.
(4) The damage-coupled cyclic elastic-viscoplastic constitutive model including strain-range, strain-rate and loading history de-
pendence is proposed to effectively predict the complex cyclic behavior of the material at elevated temperature.

Acknowledgments

The authors gratefully acknowledge the support provided by National Natural Science Foundation of China (51575531), Chang
Jiang Scholars Program, Taishan Scholar Construction Funding (ts201511018) and National Key R&D Program of China
(2018YFC0808800).

References

Abdel-Karim, M., 2005. Numerical integration method for kinematic hardening rules with partial activation of dynamic recovery term. Int. J. Plast. 21, 1303–1321.
Abdel-Karim, M., 2009. Modified kinematic hardening rules for simulations of ratchetting. Int. J. Plast. 25, 1560–1587.
Abdel-Karim, M., 2010. An evaluation for several kinematic hardening rules on prediction of multiaxial stress-controlled ratchetting. Int. J. Plast. 26, 711–730.
Abdel-Karim, M., Ohno, N., 2000. Kinematic hardening model suitable for ratcheting with steady-state. Int. J. Plast. 16, 225–240.
Ahmed, R., Barrett, P.R., Hassan, T., 2016. Unified viscoplasticity modeling for isothermal low-cycle fatigue and fatigue-creep stress–strain responses of Haynes 230.
Int. J. Solid Struct. S88–89, 131–145.
Armstrong, P.J., Frederick, C.O., 1966. 2007. A Mathematical Representation of the Multiaxial Bauschinger Effect, vol. 24. CEGB Report, Berkeley Laboratories, R&D
Department, CA, pp. 11–26 (reappears in Materials at High Temperature).
Castelluccio, G.M., Mcdowell, D.L., 2017. Mesoscale cyclic crystal plasticity with dislocation substructures. Int. J. Plast. 98, 1–26.
Chaboche, J.L., 1986. Time-independent constitutive theories for cyclic plasticity. Int. J. Plast. 2, 149–188.

212
X.-f. Xie et al. International Journal of Plasticity 114 (2019) 196–214

Chaboche, J.L., 1991. On some modifications of kinematic hardening to improve the description of ratcheting effects. Int. J. Plast. 7, 661–678.
Chaboche, J.L., Dang-Van, K., Cordier, G., 1979. Modelization of the strain memory effect on the cyclic hardening of 316 stainless steel. In: SMIRT 5, Berlin.
Chaboche, J.L., Gaubert, A., Kanouté, P., Longuet, A., Azzouz, F., Mazière, M., 2013. Viscoplastic constitutive equations of combustion chamber materials including
cyclic hardening and dynamic strain aging. Int. J. Plast. 46, 1–22.
Chang, B., Zhang, Z., 2012a. Cyclic deformation behavior in a nitrogen-alloyed austenitic stainless steel in terms of the evolution of internal stress and microstructure.
Mater. Sci. Eng., A 556, 625–632.
Chang, B., Zhang, Z., 2012b. Low cycle fatigue behavior of a high nitrogen austenitic stainless steel under uniaxial and non-proportional loadings based on the
partition of hysteresis loops. Mater. Sci. Eng., A 547, 72–79.
Chen, W., Kitamura, T., Feng, M., 2018. Creep and fatigue behavior of 316L stainless steel at room temperature: experiments and a revisit of a unified viscoplasticity
model. Int. J. Fatig. 112, 70–77.
Facheris, G., Pham, M.S., Janssens, K.G.F., Holdsworth, S.R., 2013. Microscopic analysis of the influence of ratcheting on the evolution of dislocation structures
observed in AISI 316l stainless steel during low cycle fatigue. Mater. Sci. Eng., A 587, 1–11.
Guan, Y., Chen, B., Zou, J., Britton, B., Jiang, J., Dunne, F.P.E., 2016. Crystal plasticity modelling and HR-DIC measurement of slip activation and strain localisation in
single and oligo-crystal Ni alloys under fatigue. Int. J. Plast. 88, 70–88.
Hong, S.G., Lee, S.B., 2004. The tensile and low-cycle fatigue behavior of cold worked 316L stainless steel: influence of dynamic strain aging. Int. J. Fatig. 26, 899–910.
Hong, S.G., Yoon, S., Lee, S.B., 2003. The effect of temperature on low-cycle fatigue behavior of prior cold worked 316l stainless steel. Int. J. Fatig. 25, 1293–1300.
Hormozi, R., Biglari, F., Nikbin, K., 2015. Experimental and numerical creep–fatigue study of type 316 stainless steel failure under high temperature LCF loading
condition with different hold time. Eng. Fract. Mech. 141, 19–43.
Huang, Z.Y., Chaboche, J.-L., Wang, Q.Y., Wagner, D., Bathias, C., 2014. Effect of dynamic strain aging on isotropic hardening in low cycle fatigue for carbon
manganese steel. Mater. Sci. Eng., A 589, 34–40.
Huang, Z.Y., Wagner, D., Wang, Q.Y., Khan, M.K., Chaboche, J., 2016. A low cycle fatigue model for low carbon manganese steel including the effect of dynamic strain
aging. Mater. Sci. Eng., A 654, 77–84.
Jiang, Y., Zhang, J., 2008. Benchmark experiments and characteristic cyclic plasticity deformation. Int. J. Plast. 24, 1481–1515.
Kachanov, L.M., 1958. On the creep fracture time. Izvestiya Akademii, Nauk USSR Otd. Tech. 8, 26–31 (in Russian).
Kang, G., Gao, Q., Yang, X., 2002. A visco-plastic constitutive model incorporated with cyclic hardening for uniaxial/multiaxial ratcheting of SS304 stainless steel at
room temperature. Mech. Mater. 34, 521–531.
Kang, G.Z., Gao, Q., Yang, X.J., 2004. Uniaxial and multiaxial ratcheting of SS304 stainless steel at room temperature: experiments and visco-plastic constitutive
model. Int. J. Non-linear Mech. 39, 843–857.
Kang, G.Z., Li, Y.G., Gao, Q., 2005. Non-proportionally multiaxial ratcheting of cyclic hardening materials at elevated temperatures: experiments and simulations.
Mech. Mater. 37, 1101–1118.
Kang, G., Kan, Q., Zhang, J., Sun, Y., 2006. Time-dependent ratchetting experiments of SS304 stainless steel. Int. J. Plast. 22, 858–894.
Kang, G., Liu, Y., Ding, J., Gao, Q., 2009. Uniaxial ratcheting and fatigue failure of tempered 42CrMo steel: damage evolution and damage-coupled visco-plastic
constitutive model. Int. J. Plast. 255, 838–860.
Kang, G., Dong, Y., Wang, H., Liu, Y., Cheng, X., 2010. Dislocation evolution in 316L stainless steel subjected to uniaxial ratchetting deformation. Mater. Sci. Eng., A
527, 5952–5961.
Krishna, S., Hassan, T., Naceur, I.B., Saï, K., Cailletaud, G., 2009. Macro versus micro-scale constitutive models in simulating proportional and nonproportional cyclic
and ratcheting responses of stainless steel 304. Int. J. Plast. 25, 1910–1949.
Lee, C.H., Do, V.N.V., Chang, K.H., 2014. Analysis of uniaxial ratcheting behavior and cyclic mean stress relaxation of a duplex stainless steel. Int. J. Plast. 62, 17–33.
Lemaître, J., Desmorat, R., 2005. Engineering Damage Mechanics: Ductile, Creep, Fatigue and Brittle Failures. Springer-Verlag, Heidelberg.
Li, Y., Pan, X., Wu, G., Wang, G., 2016. Shape-instability life scatter prediction of 40Cr steel: damage-coupled crystal plastic probabilistic finite element method. Int. J.
Plast. 79, 1–18.
Malinin, N.N., Khadjinsky, G.M., 1972. Theory of creep with anisotropic hardening. Int. J. Mech. Sci. 14, 235–246.
Mazánová, V., Škorík, V., Kruml, T., Polák, J., 2017. Cyclic response and early damage evolution in multiaxial cyclic loading of 316L austenitic steel. Int. J. Fatig. 100,
466–476.
Maziasz, P.J., Busby, J.T., 2012. Properties of austenitic steels for nuclear reactor applications. Comp. Nuc. Mater. 267–283.
Morin, C., Moumni, Z., Zaki, W., 2011. A constitutive model for shape memory alloys accounting for thermomechanical coupling. Int. J. Plast. 27, 748–767.
Nouailhas, D., Cailletaud, G., Policella, H., Marquis, D., Dufailly, J., Lieurade, H.P., Ribes, A., Bollinger, E., 1985. On the description of cyclic hardening and initial cold
working. Eng. Fract. Mech. 21, 887–895.
Ohno, N., 1982. A constitutive model of cyclic plasticity with a nonhardening strain region. J. Appl. Mech. 49, 721–727.
Ohno, N., Wang, J.D., 1993a. Kinematic hardening rules with critical state of dynamic recovery: I. Formulation and basic features for ratcheting behaviour. Int. J. Plast.
9, 375–389.
Ohno, N., Wang, J.D., 1993b. Kinematic hardening rules with critical state of dynamic recovery: II. Application to experiments of ratcheting behaviour. Int. J. Plast. 9,
390–403.
Pham, M.S., Holdsworth, S.R., 2013. Role of microstructural condition on fatigue damage development of AISI 316L at 20 and 300°C. Int. J. Fatig. 51, 36–48.
Pham, M.S., Holdsworth, S.R., Janssens, K.G.F., Mazza, E., 2013. Cyclic deformation response of AISI 316L at room temperature: mechanical behaviour, micro-
structural evolution, physically-based evolutionary constitutive modelling. Int. J. Plast. 47, 143–164.
Polák, J., Petráš, R., Heczko, M., Kuběna, I., Kruml, T., Chai, G., 2014. Low cycle fatigue behavior of Sanicro25 steel at room and at elevated temperature. Mater. Sci.
Eng., A 615, 175–182.
Prager, W., 1949. Recent developments in the mathematical theory of plasticity. J. Appl. Phys. 20, 235–241.
Rabotnov, Y.U.N., 1968. Creep rupture. In: Proceedings of the XII International Congress on Applied Mechanics. StanfordSpringer, pp. 342–349.
Rodas, E.A.E., Neu, R.W., Rodas, E.A.E., Neu, R.W., 2017. Crystal viscoplasticity model for the creep-fatigue interactions in single-crystal Ni-base superalloy CMSX-8.
Int. J. Plast. 100, 14–33.
Srinivasan, V.S., Valsan, M., Sandhya, R., Rao, K.B.S., Mannan, S.L., Sastry, D.H., 1999. High temperature time-dependent low cycle fatigue behaviour of a type 316L
(N) stainless steel. Int. J. Fatig. 21, 11–21.
Srinivasan, V.S., Valsan, M., Rao, K.B.S., Mannan, S.L., Raj, B., 2003. Low cycle fatigue and creep–fatigue interaction behavior of 316L(N) stainless steel and life
prediction by artificial neural network approach. Int. J. Fatig. 25, 1327–1338.
Voyiadjis, G.Z., Dorgan, R.J., 2007. Framework using functional forms of hardening internal state variables in modeling elasto-plastic-damage behavior. Int. J. Fatig.
23, 1826–1859.
Wang, J., Moumni, Z., Zhang, W., 2017. A thermomechanically coupled finite-strain constitutive model for cyclic pseudoelasticity of polycrystalline shape memory
alloys. Int. J. Plast. 97, 194–221.
Xu, L., Nie, X., Fan, J., Tao, M., Ding, R., 2016. Cyclic hardening and softening behavior of the low yield point steel BLY160: experimental response and constitutive
modeling. Int. J. Plast. 78, 44–63.
Xu, L., Zhao, L., Gao, Z., Han, Y., 2017. A novel creep-fatigue interaction damage model with the stress effect to simulate the creep-fatigue crack growth behavior. Int.
J. Mech. Sci. 130, 143–153.
Yaguchi, M., Takahashi, Y., 2000. A viscoplastic constitutive model incorporating dynamic strain aging effect during cyclic deformation conditions. Int. J. Plast. 16,
241–262.
Yaguchi, M., Takahashi, Y., 2005. Ratchetting of viscoplastic material with cyclic softening, part 2: application of constitutive models. Int. J. Plast. 21, 835–860.
Yoshida, F., 1990. Uniaxial and biaxial creep-ratcheting behavior of SUS304 stainless steel at room temperature. Int. J. Pres. Ves. Pip. 44, 207–223.
Yu, D., Chen, X., Yu, W., Chen, G., 2012. Thermo-viscoplastic modeling incorporating dynamic strain aging effect on the uniaxial behavior of Z2CND18.12N stainless

213
X.-f. Xie et al. International Journal of Plasticity 114 (2019) 196–214

steel. Int. J. Plast. 37, 119–139.


Yu, C., Kang, G., Kan, Q., 2014. Crystal plasticity based constitutive model of NiTi shape memory alloy considering different mechanisms of inelastic deformation. Int.
J. Plast. 54, 132–162.
Yu, C., Kang, G., Kan, Q., Zhu, Y., 2015. Rate-dependent cyclic deformation of super-elastic NiTi shape memory alloy: thermo-mechanical coupled and physical
mechanism-based constitutive model. Int. J. Plast. 72, 60–90.
Zhang, J.X., Jiang, Y.Y., 2008. Constitutive modeling of cyclic plasticity deformation of a pure polycrystalline copper. Int. J. Plast. 24, 1890–1915.
Zhang, S.L., Xuan, F.Z., 2017. Interaction of cyclic softening and stress relaxation of 9-12% Cr steel under strain-controlled fatigue-creep condition: experimental and
modeling. Int. J. Plast. 98, 45–64.
Zhao, P., Xuan, F.Z., 2011. Ratchetting behavior of advanced 9-12% chromium ferrite steel under creep-fatigue loadings. Mech. Mater. 436, 299–312.
Zhou, J., Sun, Z., Kanouté, P., Retraint, D., 2018. Experimental analysis and constitutive modelling of cyclic behaviour of 316l steels including hardening/softening
and strain range memory effect in LCF regime. Int. J. Plast.
Zhu, H., Sun, L., 2013. A viscoelastic–viscoplastic damage constitutive model for asphalt mixtures based on thermodynamics. Int. J. Plast. 40, 81–100.

214

You might also like