Anda di halaman 1dari 19

International Journal of Plasticity 43 (2013) 1–19

Contents lists available at SciVerse ScienceDirect

International Journal of Plasticity


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

About the cyclic accumulation of the inelastic strain observed


in metals subjected to cyclic stress control
Lakhdar Taleb ⇑
INSA, GPM, CNRS UMR 6634, BP 08 avenue de l’université, 76801 St. Etienne du Rouvray Cedex, France

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

Article history: This work discusses the sources of the accumulated inelastic strain observed under unsym-
Received 11 August 2012 metrical stress cycling. Different metallic materials have been chosen for their large variety
Received in final revised form 11 October 2012 of elasto-visco-plastic behaviors: two austenitic stainless steels (304L and 316L), two ferritic
Available online 25 October 2012
steels (35NiCrMo16 and XC18), one aluminum alloy (Al 2017) and one copper–zinc alloy
(CuZn27). Each material has been tested at room temperature under two loading paths:
Keywords: one proportional (tension–compression) and the other non proportional following a trian-
Plastic collapse
gular path in the axial stress – shear strain space. The results demonstrate that ratcheting
Cyclic loading
Metallic materials
under multiaxial loading is clearly present even if it may be associated with other phenom-
Mechanical testing ena. Indeed, the multiaxial stress state governs the direction of the inelastic strain rate
Ratcheting through the normality rule. Under unsymmetrical tension–compression load control, the
cyclic growth of the inelastic strain is almost absent in the aluminum and copper–zinc alloys
under the loading considered while for the austenitic steels, the cyclic accumulation seems
mainly due to creep. Namely, 1D ratcheting seems rather very small for these four materials.
However, for the ferritic steels, a cyclic progressive deformation is observed and may be
associated with ratcheting even if the mode of control (load) may also have a contribution.
Ó 2012 Elsevier Ltd. All rights reserved.

1. Introduction

The industrial mechanical structures are often subjected to the combination of force and displacement cyclic loadings
which results on the interaction between primary and secondary stresses respectively. Such loading may lead to complex
behavior of the material with or without phase changes especially under non proportional paths (Abdel Karim and Ohno,
2000; Bari and Hassan, 2002; Bocher et al., 2001; Chaboche, 2008; Chaboche et al., 2012; Grostabussiat et al., 2004; Hassan
et al., 2008; Kang et al., 2005, 2006; Kang and Kan, 2007; Khan et al., 2007; McDowell, 1987; Moosbrugger, 1993; Ohno, 1997;
Ohno and Wang, 1993, 1994; Taleb and Hauet, 2009; Taleb and Cailletaud, 2010, 2011; Taleb et al., 2006; Wolff and Taleb,
2008. . .). Under prescribed cyclic displacement the material may exhibit cyclic hardening or softening depending not only
on the amplitude of the applied strain but also on the non proportionality character of the loading path. Under unsymmetrical
stress control, a strain accumulation cycle by cycle may be observed: our objective here is to try to better understand the
sources of this progressive strain. Such phenomenon is generally called ratcheting as it can be checked in the following ref-
erences of literature:

 After Kang and his co-workers (Kang and Gao, 2002; Kang et al., 2004, 2006; Kang, 2008). . ., we take the following sentence:
‘‘Under asymmetrical stress cycling, a cyclic accumulation of strain, namely ratcheting, occurs’’ (Kang and Gao, 2002);
 After Abdel-Karim and his co-workers (Abdel Karim and Ohno, 2000; Abdel Karim, 2009, 2011): ‘‘Ratchetting, defined as
accumulation of plastic strain increments under cyclic loading with nonzero mean stress’’ (Abdel Karim, 2009);

⇑ Tel.: +33 2 32 95 97 65; fax: +33 2 32 95 97 10.


E-mail addresses: ltaleb@insa-rouen.fr, lakhdar.taleb@insa-rouen.fr

0749-6419/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijplas.2012.10.009
2 L. Taleb / International Journal of Plasticity 43 (2013) 1–19

 After (Bari and Hassan, 2000): ‘‘. . . ratcheting response (accumulation of strains with cycles) are . . .’’
 After (Dafalias and Feigenbaum, 2011) ‘‘Cyclic plasticity addresses such response under a sequence of repeated stress
reversals and the ensuing technologically important phenomenon of plastic strain accumulation, called ratcheting’’.
 After (Gaudin and Feaugas, 2004): ‘‘The ratchetting process was extensively studied in fcc alloys with medium and high
stacking fault energy. This phenomenon is characterized by a translation of the hysteresis loop under non-symmetrical
stress loading’’
 After (Gao et al., 2006): ‘‘Ratchetting, namely cyclic plastic strain accumulation, is a phenomenon that can occur in a
structure subjected to a primary load with a secondary cyclic load’’
 After (Gupta et al., 2005): ‘‘The study of accumulation of strain due to asymmetrical stress cycling in structural materials
for piping application has been the subject of a number of investigations in recent years [1–3]. The phenomenon, known
as ratcheting, is potentially dangerous. . .’’
 After (Jiang and Zhang, 2008): ‘‘Cyclic strain ratcheting refers to the progressive and directional plastic strain accumula-
tion due to unsymmetrical stress cycling’’
 After (Vincent et al., 2004): ‘‘When a metal is submitted to repeated cycles in the plastic range, plastic strain can accu-
mulate in at least one direction. This is material ratchetting.’’
 After (Chaboche, 2008): ‘‘As pointed out in Chaboche and Nouailhas (1989), ratchetting corresponds with a progressive
accumulation of plastic strain, cycle-by-cycle, that resembles to a creep effect. For example, maximum strain evolution
shows a primary stage (decreasing ratchetting rate) followed by a stationary stage (constant ratchetting rate),
like in the creep test. However, such an accumulation must not be confused with creep (induced by time under a constant
stress), because due to plasticity mechanisms during the unloading–reloading (not specifically influenced by time).

One can notice that all these references agree with the definition of the ratcheting phenomenon as the accumulation
of the inelastic strain under unsymmetrical stress cycling. However, the definition given by Chaboche (2008) is more
accurate as it enables to avoid the confusion between ratcheting and creep. This specification contrasts with the defini-
tion of ratcheting by ‘‘cyclic creep’’ sometimes employed (Gaudin and Feaugas, 2004) as ratcheting is rate-independent
phenomenon. It is not always easy to discriminate between creep and ratcheting through the responses of ‘‘classical’’
stress controlled experiments. However, in earlier study performed on the 304L stainless steel at room temperature,
we have achieved such ‘‘separation’’ by the application of a creep test prior to cycling (Taleb and Cailletaud, 2011).
Through such work we have demonstrated that most of the accumulated plastic strain observed under unsymmetrical
stress cycling is due to creep. This result means that ratcheting is very small for the 304L steel (in the conditions con-
sidered) which contrasts with number of literature studies where the cyclic behavior of this material is assumed mainly
rate-independent (Hassan et al., 2008), (Krishna et al., 2009), (Abdel Karim, 2010) and so on. Furthermore, it is probable
that the bahavior of the 316L SS at room temperature is also time dependent if we consider the results of Portier et al.
(2000). Indeed, a cyclic accumulation of the inelastic strain was observed for this steel under tension–compression load
control despite the fact that the applied amplitude (140 MPa) is smaller than the yield stress (about 180 MPa, evaluated
from the stablized stress–strain loop of Fig. 4 of the same reference). It is obvious that in these conditions, the unloading
occurs mainly into the yield surface and then no significant ratcheting may be expected! Therefore, despite the large
number of works carried out, the cyclic behavior of metals remains not completely understood. Furthermore, if the
mechanism of biaxial ratcheting may be explained through the normality rule, the physical origin of the cyclic growth
of the strain under unsymmetrical tension–compression is still not well understood. The present paper is a new contri-
bution in this direction.
In order to better understand the origin of the accumulated inelastic strain under unsymmetrical stress cycling, dif-
ferent metallic materials have been chosen with a large variety of elasto-visco-plastic behaviors: two austenitic stain-
less steels (304L and 316L), two ferritic steels (35NiCrMo16 and XC18), an aluminum alloy (Al 2027) and, a copper–
zinc alloy (CuZn27). Each material has been tested at room temperature under two cyclic loading paths: one propor-
tional (tension–compression) and the other non proportional following a triangular path in the axial stress – shear
strain space.
The paper is divided into three main parts: after the introduction, the experimental procedure is detailed in the first sec-
tion. The second section is devoted to the presentation of the main experimental responses obtained. Section 3 includes a
discussion about the different sources leading to a cyclic accumulation of the inelastic strain. In this analysis, our aim is
to know if what we call ‘‘ratcheting’’ does not include the contributions of other known phenomena like the mode of control
(load instead of true stress), creep and, fatigue damage. Finally, some concluding remarks are given.

2. Experimental procedure

2.1. Materials

Six different metallic materials have been considered in this study: two austenitic stainless steels (304L and 316L), two
ferritic steels (35NiCrMo16 and XC18), an aluminum alloy 2017 and finally a copper–zinc alloy CuZn27. Their average chem-
ical compositions are given in Table 1a–f.
L. Taleb / International Journal of Plasticity 43 (2013) 1–19 3

Table 1
Average chemical composition (in% mass) of the different materials used in this study.

(a) 304L SS
C Si Mn Ni Cr N S P Fe
0.028 0.68 1.54 9.04 18.83 0.085 0.026 0.035 Balance
(b) 316L SS
C Si Mn Ni Cr Mo S P Fe
0.02 0.36 1.32 10.13 16.57 2.03 0.029 0.031 Balance
(c) 35NiCrMo16 (THYSSEN-SENARD)
C Si Mn P S Cr Ni Mo Cu Fe
0.35 0.29 0.48 0.21 0.011 1.81 3.78 0.27 0.17 Balance
(d) XC18
C Si Mn P S Cr Ni Mo Cu Fe
0.20 0.22 0.57 0.022 0.021 0.11 0.06 0.01 0.18 Balance
(e) Al 2017
Cu (%) Mg (%) Mn (%) Si (%) Fe (%) Al (%)
3.95 0.67 0.53 0.76 0.90 Balance
(f) CuZn27
Cu Pb Sn Fe Ni Al Mn Si Zn
63.0 1.50 1.00 1.25 2.50 1.75 2.25 1.00 Balance

2.2. Specimen and experimental device

All specimens have the same geometry given in Fig. 1: tubular shape with a gage length of 46 mm where a central part of
25 mm is used for the extensometry. In this zone, the outer and inner diameters are respectively equal to 20 mm and 17 mm
that enables to have relatively thin tubes.
The steel specimens have been machined from solid bars and then heat treated in order to have almost the same metal-
lurgical and mechanical initial states. The applied thermal cycle is composed of heating up 1050 °C followed by a hold time of
1 h at this temperature and, finally a slow cooling up to room temperature.
The aluminum alloy specimens have been machined from extruded solid bars parallel to the extrusion axis, they have also
heat treated in order to ensure almost the same initial state. The treatment includes a 1 h hold time at 500 °C followed by a
quenching up to room temperature.
Finally, the copper–zinc alloy specimens have been obtained by casting process performed by FAVI Company.
The tests were carried out using MTS axial–torsional servo-hydraulic machine with a capacity of ±250 KN/±75 mm in
tension–compression and ±2200 Nm/±45° in torsion. The axial displacement and the rotation in the central zone of the
specimen were measured by MTS axial–torsional extensometer with 25 mm gage length and a capacity of ±2.5 mm/±5°.
Hydraulic collet grips enable to rigidly maintain the heads of the specimen under adjustable pressure control.
The engineering stress and strain tensors are assumed constant in the gage length; they have the following general form
in a cylindrical coordinate system ðr; h; zÞ with z oriented along the specimen axis:
0 1 0 1
0 0 0 err 0 0
r¼B
@0 0 rzh C B C
A and; e ¼ @ 0 ehh ezh A
0 rzh rzz 0 ezh ezz

Fig. 1. Geometry of the specimens.


4 L. Taleb / International Journal of Plasticity 43 (2013) 1–19

The following components rij and eij are deducted from the measurements carried out:
F
rzz ¼ ð1Þ
S0

Mt r
rzh ¼ ð2Þ
I0

R0 a
2ezh ¼ czh  ð3Þ
l0

uz
ezz ¼ ð4Þ
l0
S0, l0, R0 and I0 are the cross section, the length, the outer radius and the polar moment of the gage length calculated before
deformation.
F is the axial force and Mt the torque measured by axial–torsional cell. uz is the axial displacement and a the angle (in
radians) measured by axial–torsional extensometer.
The axial true strain is defined by,
Z l
dl
eTzz ¼ ¼ Lnð1 þ ezz Þ ð5Þ
l0 l

Table 2
Main characteristics of the tests performed. Forces (Newtons); angles (°); stresses (MPa) and; strains (%).

Material r0:2% (conv. yield Tension–compression F max =F min Triangle-inverse triangle F max =amax Reference
czhmax
stress, MPa) rmax min
zz =rzz rmax pffiffi
zz = 3

304L 260 21750/13050 304L_50-200*


250/150
21750/1 304L_250-0.4
250/0.4
316L 210 21750/13050 316L_250-200
250/150
21750/1 316L _250-0.4
250/0.4
35NiCrMo16 900 104400/69600 35NiCrMo16_200-1000
1200/800
108750/73950 35NiCrMo16_200-1050
1250/850
121800/87000 35NiCrMo16_200-1200
1400/1000
78300/1 35NiCrMo16_900-0.4
900/0.4
78300/2 35NiCrMo16_900-0.8
900/0.8
XC18 300 32190/23490 XC18_50-320
370/270
21750/1 XC18_250-0.4
250/0.4
Al 2017 286 26100/17400 AU4G_50-250
300/200
30450/21750 AU4G_50-300
350/250
26100/1 AU4G_300-0.4
300/0.4
26100/2 AU4G_300-0.8
300/0.8
CuZn27 260 26100/17400 CuZn27_50-250
300/200
28710/21750 CuZn27_50-280
330/20010
26100/1 CuZn27_250-0.4
300/0.4
26100/2 CuZn27_280-0.8
300/0.8
*
Test performed in (Taleb and Cailletaud, 2011).
L. Taleb / International Journal of Plasticity 43 (2013) 1–19 5

If we assume that the volume of the gage length remains constant, we can calculate the axial true stress as follows,

rTzz ¼ rzz ð1 þ ezz Þ ð6Þ

2.3. Tests performed

Two cyclic loading paths, one proportional and the other non proportional, have been chosen in this study in order to
investigate the origin of the cyclic accumulation of the inelastic strain. The proportional loading path is a non symmetric
tension–compression under load control while the non proportional one is a combination of axial force and torsional rotation
following a triangular shape. The choice of the latter path is motivated by its ‘‘complexity’’ as several earlier studies show
that different constitutive equations fail in simulating such a test type (Hassan et al., 2008; Abdel Karim, 2010).
The list providing the main characteristics of all tests performed in the present study is given in Table 2.
In the tension–compression tests 100 cycles have been applied (Fig. 2a). For the austenitic steels where the behavior is
time dependent, classical ‘‘ratcheting’’ tests (without hold time at peak stresses) have been performed in addition to
‘‘creep-ratcheting’’ tests. In the latter (Fig. 2a), the maximum force of the first cycle is maintained constant during a certain
time in order to minimize the contribution of creep in the cyclic accumulation of the inelastic strain. In the multiaxial tests,
the non proportional loading is applied in two stages: 100 ‘‘triangular’’ cycles (Fig. 2b) followed by 100 ‘‘Inverse triangular’’
cycles (Fig. 2c) are applied successively on the same specimen. Note that no hold time at the maximum force is applied in the
second stage (Fig. 2c). All tests have been carried out at room temperature.

(a) α F
Fmax
Fmin Fm Fmax
F
Fa time
Fmin

(b) F
α Fmax
α max
2 time
3 Fmax
1
4
α
F α max
6 5 time
− α max
− α max

(c) α F time
α max
2
− Fmax 3 − Fmax
1
4 F α
6 α max
5
− α max time

− α max
Fig. 2. Loading paths applied in the tests performed. (a) Cyclic tension–compression under force control between Fmax and Fmin starting by creep test at Fmax
for the ‘‘creep-ratcheting’’ tests; no hold time is applied for the classical ‘‘ratcheting’’ tests. (b) Two first cycles of a path combining prescribed force between
0 and Fmax and torsional rotation between amax and amax following triangular shape, a creep test follows the first loading at Fmax. (c) Cyclic path combining
prescribed force between 0 and Fmax and torsional rotation between amax and amax following ‘‘inverse’’ triangular shape.
6 L. Taleb / International Journal of Plasticity 43 (2013) 1–19

Due to the presence of the Lüders phenomenon in the XC18 steel, the tension compression test started by a first pre-load-
ing under strain control ensuring to go over the plateau. However, the stress range of such pre-loading remains lower than
the stress range of the subsequent cycles performed at force control.

3. Results

3.1. Responses under tension–compression

Figs. 3–7 give the evolution of the engineering axial stress and strain in the tests performed. For the sake of better read-
ability, only the cycles 1–5, 10, 20 and every 10 up to 100 are reported.
The interactions between plasticity and creep/relaxation have been studied in literature for many years (Kawai and
Ohashi, 1987; Krempl, 1979; Murakami and Ohno, 1982; Ikegami and Niitsu, 1985; Yoshida, 1990). The results given in
Fig. 3 are taken from a recent paper (Taleb and Cailletaud, 2011) which constitutes one more contribution in order to better
understand such interactions. Even if a cyclic growth of the inelastic strain is observed in the classical ‘‘ratcheting’’ tests (see
the original paper), a plastic shakedown clearly takes place if a hold time is applied at the peak stress in the first cycle. This
result points out the fact that creep constitutes the most part of the accumulated inelastic strain exhibited by the 304L stain-
less steel subjected to cyclic load control at room temperature. The same tendency is demonstrated here (Fig. 4) for the 316L
stainless steel where creep is also very important at room temperature and the subsequent cyclic accumulation of the inelas-
tic strain is rather not significant.
Apart from austenitic stainless steels, the evolution of the axial strain observed when the first peak of the first cycle is
maintained constant is rather small. Therefore, one can neglect creep at room temperature for the ferritic steels, the alumi-
num alloy and the copper–zinc alloy under consideration. We think that this observation remains true for the test 35NiC-
rMo16_200-1200 (mean stress 200 MPa and amplitude 1200 MPa) which leads to the fracture of the specimen during the
loading in the first cycle. Indeed, such fracture is rather due to the fact that the maximum applied stress (1400 MPa) may
be close to the maximum resistance of the material as it can be remarked in the axial stress–strain diagram of Fig. 9. Further
analyses about creep tests will be given later.

Fig. 3. Evolution of the axial stress and strain for a ‘‘creep-ratcheting’’ test performed on the 304L austenitic stainless steel 304L_50-200: (a) axial stress vs
axial strain; (b) axial strain peaks vs the number of cycles after (Taleb and Cailletaud, 2011).

Fig. 4. Evolution of the axial stress and strain for both ‘‘ratcheting’’ and ‘‘creep-ratcheting’’ tests performed on the 316L austenitic stainless steel (316L_50-
200): (a) axial stress vs axial strain; (b) axial strain peaks vs the number of cycles.
L. Taleb / International Journal of Plasticity 43 (2013) 1–19 7

Fig. 5. Evolution of the axial stress and strain for the tests performed on the 35NiCrMo16 steel: (a–b) 35NiCrMo16_200-1000 and (c–d) 35NiCrMo16_200-
1050. (a–c) axial stress vs axial strain; (b–d) axial strain peaks, valleys and averages vs the number of cycles.

Fig. 6. Evolution of the axial stress and strain for the tests performed on the XC18 steel: XC18_50-320: (a) axial stress vs axial strain; (b) axial strain peaks,
valleys and averages vs the number of cycles.

For the 35NiCrMo16 steel, the cyclic accumulation of the inelastic strain depends on the applied loading; it is very small
for the test 35NiCrMo16_200-1000 (Fig. 5a–b) while it is significant for the test 35NiCrMo16_200-1050 (Fig. 5c–d). Fig. 5d
shows that after about 50 cycles of regular cyclic growth, the rate of the inelastic strain becomes more and more important.
The cyclic growth of the inelastic strain is very clear in the test XC18_50-320 steel where almost regular increments are
observed (Fig. 6).
Both tests carried out on the aluminum alloy show very small cyclic accumulation of the inelastic strain during the few
first cycles. However, such accumulation vanishes very quickly (Fig. 7b–d) leading to macroscopic elastic shakedown
(Fig. 7a–c).
8 L. Taleb / International Journal of Plasticity 43 (2013) 1–19

Fig. 7. Evolution of the axial stress and strain for the tests performed on the Al 2017 alloy: (a, b) AU4G_50-250 and (c, d) AU4G_50-300. (a, c) axial stress vs
axial strain; (b, d) axial strain peaks, valleys and averages vs the number of cycles.

The copper–zinc alloy exhibits ‘‘unusual’’ behavior compared to the other metals considered here. After checking that the
initial yield limit is the same in tension and compression, if we compare the apparent sizes of the elastic domain during the
loading (twice the yield limit) and unloading in the first cycle (Fig. 8a, c), one can observe significant isotropic softening.
However, in the following cycles, the material exhibits an isotropic hardening which leads to the quasi absence of the global
cyclic accumulation of the inelastic strain. Such evolution cannot be described through the evolution of the peaks or the val-
leys of the axial strain. Indeed, if we analyze the evolution of the axial strain peaks (Fig. 8b, d), one can conclude about neg-
ative accumulation of the inelastic strain while the evolution of the axial strain valleys leads to the opposite observation.
Therefore, the evolution of the strain peaks considered alone cannot be always reliable to indicate the right sense of the evo-
lution of the cyclic behavior.
To assess the results obtained after the tension–compression tests given above, one can conclude that no significant cyclic
accumulation of the inelastic strain is observed for the aluminum and copper–zinc alloys while the accumulation observed
for the austenitic steels 304L and 316L seems mainly due to creep. These results mean that the question about ratcheting
remains open. Further investigations are necessary in order to better understand this phenomenon in these metals under
uniaxial loading at room temperature. However, the ferritic steels 35NiCrMo16 and XC18 exhibit cyclic accumulation of
the inelastic strain; we will discuss the origin of this growth later.

3.2. Responses under non proportional loading

The evolution of the engineering axial stress and strain in the tests performed under triangular followed by inverse tri-
angular paths are given in Figs. 10–15. One hundred cycles have been applied for each path: the first and the second 100
cycles of the test will be referred to as part 1 and part 2, respectively. For the sake of better readability, only the cycles
1–5, 10, 20 and every 10 up to 100, 101–105, 110, 120 and every 10 up to 200 are reported. Furthermore, from now in
the absence of other specification, by the term strain we mean the axial component of the strain tensor. Note that the test
35NiCrMo16_900-0.8% has not been conducted completely as the fracture of the specimen occurred after 66 cycles. The re-
sults obtained above suggest that the 304L and 316L steels are the only materials from the ones under consideration here
which exhibit creep phenomenon at room temperature. For this reason, only the tests 304L_250-0.4% and 316L_250-0.4%
started by a hold time (2 h) at the maximum force in the first cycle (end of the first segment of the loading path).
Without surprise the importance of creep for the 304L stainless steel at room temperature is confirmed after the test
304L_250-0.4% (Fig. 10); the strain reaches about 1% after a hold time of two hours at 250 MPa. However, contrary to the
L. Taleb / International Journal of Plasticity 43 (2013) 1–19 9

Fig. 8. Evolution of the axial stress and strain for the tests performed on the copper–zinc alloy: (a, b) CuZn27_50-250 and (c, d) CuZn27_50-280. (a, c) axial
stress vs axial strain; (b, d) axial strain peaks, valleys and averages vs the number of cycles.

Fig. 9. Axial stress vs axial strain for the test 35NiCrMo16_200-1200 where the fracture of the specimen occurred during the loading in the first cycle.

uniaxial case (Fig. 3), the cyclic accumulation of the inelastic strain is clearly observed here for both parts of the test. The
axial strain peaks obtained after the two parts of the test are 1.85% and 1.25% respectively (Fig. 10). Therefore, apart from
creep (1%), 70% of the axial strain accumulated after the first 100 cycles is erased after the second 100 cycles. One can note
that no steady state is obtained during the test; the rate of the accumulated strain remains significant in both parts.
The same may be concluded about the 316L stainless steel where the creep phenomenon seems more important under a
given stress. Indeed, in the test 316L_250-0.4% (Fig. 11), the strain goes over 3% after a hold time of two hours under
250 MPa.
The behavior of the 35NiCrMo16 steel is very sensitive to the applied shear strain. Indeed, if no significant accumulation
of the plastic strain is observed for the test 35NiCrMo16_900-0.4%, the fracture of the specimen occurred after 66 cycles in
the test 35NiCrMo16_900-0.8% (Fig. 12). The value of the axial strain (which is not very different from the equivalent plastic
strain) at which the fracture occurs here is relatively small (less than 1.5%!) compared to the fracture strain obtained under
10 L. Taleb / International Journal of Plasticity 43 (2013) 1–19

Fig. 10. Evolution of the axial stress and strain for the test performed on the 304L austenitic stainless steel 304L_250-0.4%: (a) axial stress vs axial strain; (b)
axial strain peaks, valleys and averages vs the number of cycles.

Fig. 11. Evolution of the axial stress and strain for the test performed on the 316L austenitic stainless steel 316L_250-0.4%: (a) axial stress vs axial strain; (b)
axial strain peaks, valleys and averages vs the number of cycles.

monotonic tensile loading which approaches 10% (Fig. 9). This result demonstrates the sensitivity of the material’s fatigue
life to non proportional loading.
As for the tension–compression case (Fig. 6), the XC18 steel exhibits the most important and ‘‘regular’’ cyclic evolution
compared to the other materials under consideration. Indeed, after part 1, the axial strain reaches about 4.3% but during part
2, the ‘‘negative’’ strain accumulation proved to be more important as the final axial strain is about 1.5% (Fig. 13). Further-
more, no steady state is obtained during the test; the rate of the accumulated strain remains significant at the end of both
parts.
The aluminum alloy Al 2017 shows different cyclic behavior. In the test AU4G_300-0.4%, the cyclic accumulation of the
axial strain is relatively important during the first 20 cycles of part 1; after that, a steady state is observed with an axial strain
around 1.3% (Fig. 14). During the next 100 cycles (part 2), the cyclic evolution is rather not significant as the final axial strain
is about 1.2%. The behavior in the test AU4G_300-0.8% is significantly different which prove the sensitivity of the material to
the prescribed torsional rotation. Indeed, we obtain 9.3% after part 1 but this amount is reduced to 5% by the negative accu-
mulation of the strain during part 2 (Fig. 14c–d). Furthermore, no steady state is obtained during the test as the rate of the
accumulated strain remains significant at the end of both parts (Fig. 14d).
The copper–zinc alloy CuZn27 shows moderated cyclic accumulation of the strain in both tests. In the test CuZn27_250–
0.4%, the final strains obtained after part 1 and part 2 are 0.2% and 0.06% respectively while in the test CuZn27_280–0.4%, we
obtain 0.5% and 0.08% respectively (Fig. 15). These results mean that the strain accumulated during part 1 is ‘‘quasi-exactly’’
erased by part 2. A steady state is reached where the stress and the strain come back to zero at the end of part 2 which leads
to close axial stress–strain loops in both tests (Fig. 15a, c). This result may be due to the absence of significant cyclic strain
hardening or softening in this material.
The results obtained above show once more the complexity of the cyclic behavior under stress/strain control. Such com-
plexity is amplified as each metallic material has its own behavior.
Under tension–compression load control, the aluminum and copper–zinc alloys considered here did not exhibit signifi-
cantly neither creep nor cyclic strain accumulation. Furthermore, the most part of the cyclic accumulated strain observed
L. Taleb / International Journal of Plasticity 43 (2013) 1–19 11

Fig. 12. Evolution of the axial stress and strain for the tests performed on the 35NiCrMo16 steel 35NiCrMo16_900-0.4% (a, b) and 35NiCrMo16_900-0.8%
(c, d): (a, c) axial stress vs axial strain; (b, d) axial strain peaks, valleys and averages vs the number of cycles.

Fig. 13. Evolution of the axial stress and strain for the test performed on the XC18 steel XC18_250-0.4%: (a) axial stress vs axial strain; (b) axial strain peaks,
valleys and averages vs the number of cycles.

for the 304L (Taleb and Cailletaud, 2011) and 316L steels seems due to creep. However, the 35NiCrMo16 and the XC18 steels
are not sensitive to creep but show a clear cyclic accumulation of the inelastic strain.
Under non proportional loading path (Triangle-Inverse Triangle), all metals considered here show a cyclic accumulation of
the inelastic strain.
The next section will be devoted to the analyses of these results with the objective to understand the mechanism of the
cyclic accumulated strain when it appears.

4. Discussion

The cyclic accumulation of the inelastic strain may have various sources: experimental conditions (like the control of the
load instead of the true stress), presence of creep, fatigue damage and/or, ratcheting. In the absence of creep and fatigue
12 L. Taleb / International Journal of Plasticity 43 (2013) 1–19

Fig. 14. Evolution of the axial stress and strain for the tests performed on the aluminum alloy AU4G_300-0.4% (a, b) and AU4G _300-0.8% (c, d): (a, c) axial
stress vs axial strain; (b, d) axial strain peaks, valleys and averages vs the number of cycles.

Fig. 15. Evolution of the axial stress and strain for the tests performed on the copper–zinc alloy CuZn27_250-0.4% (a, b) and CuZn27_280-0.4% (c, d): (a, c)
axial stress vs axial strain; (b, d) axial strain peaks, valleys and averages vs the number of cycles.
L. Taleb / International Journal of Plasticity 43 (2013) 1–19 13

damage phenomena, ratcheting is generally defined as the cyclic accumulation of the inelastic strain observed under non
symmetric stress control: see the introduction. In the following, we will try to point out qualitatively the importance of rat-
cheting and the contribution of other phenomena in the cyclic evolution under consideration.

4.1. Cyclic accumulation of the inelastic strain under tension–compression

We will consider here the two tests 35NiCrMo16_200-1050 and XC18_50-320 where significant cyclic growth of the
inelastic strain has been observed.

4.1.1. Contribution of the mode of control


Our tests have been performed under prescribed axial force which means that the applied true stress varies continuously
during the tests. This fact may influence the ratcheting rate as advocated in (Jiang and Zhang, 2008) and demonstrated in
(Yoshida et al., 1980). In order to see if such effect is important in our tests, we compare the peaks of engineering and true
stresses for the tests 35NiCrMo16_200-1050 (Fig. 16a) and XC18_50-320 (Fig. 16b).
The results of the test 35NiCrMo16_200-1050 show significant difference between the peaks of engineering and true
stresses: if the first remains obviously constant (1250 MPa), the second varies between about 1275 MPa in the first cycle
and 1280 MPa in the cycle 100. Even if this variation seems not significant, it may have a small contribution in the cyclic
evolution of the strain due to the small strain hardening of the materiel at this level of stress (see Fig. 9). For the test
XC18_50-320, the difference is more important: the peak of the engineering stress is 370 MPa while the peak of the true
stress varies between 385 MPa in the first cycle and 408 MPa in the cycle 100. Therefore the true stress increases continu-
ously during the cycles which may certainly have significant part in the cyclic growth of the inelastic strain.

4.1.2. Contribution of creep


Fig. 17 gives the results of the ‘‘creep tests’’ performed by the application of a given hold time at the peak stress of the first
cycle of the tests 35NiCrMo16_200-1050 and XC18_50-320.

Fig. 16. Comparison between the peaks of engineering and true stresses in the tests carried out: (a) 35NiCrMo16_200-1050; (b) XC18_50-320.

Fig. 17. Results of the creep tests carried out during the first cycle of the tests: (a) 35NiCrMo16_200-1050; (b) XC18_50-320.
14 L. Taleb / International Journal of Plasticity 43 (2013) 1–19

The results of Fig. 17a suggest that if creep seems exist, it remains not significant in the 35NiCrMo16 steel. Furthermore,
as the applied hold time leads to apparent stabilization of the creep strain, one can conclude that the contribution of this
phenomenon in the cyclic growth of the inelastic strain is rather negligible for the test 35NiCrMo16_200-1050. The same
conclusion may be adopted for the test XC18_50-320 where creep seems also not significant (Fig. 17b). Note that as specified
previously, the creep test on the XC18 steel is carried out after a pre-loading cycle performed under displacement control in
order to go over the plateau (see Fig. 17b).

4.1.3. Contribution of fatigue damage


According to Lemaître and Chaboche (1985), damage may be evaluated through the evolution of the elastic modulus (E)
using a variable D that expresses the coupling between elasticity and damage:

~
E
D¼1 ð7Þ
E
where E ~ is the effective elastic modulus.
It is obvious that in the absence of damage (D = 0), we have E ¼ E. ~ This approach was used for instance by Kang et al.
(2009) in order to evaluate fatigue damage of specimens subjected to cyclic loadings. However it is important to have in
mind the limitations of this method. Indeed, the decrease of the elastic modulus does not indicate necessarily the presence
of fatigue damage especially during the first cycles at small strain. Indeed, such evolution may be the consequence of a cou-
pling between elasticity, plasticity and internal stresses. One can refer to the works of Yoshida and his co-workers (Yoshida
et al., 2002) pointing out such interaction and its dependence on the pre-strain using mild and high strength steels. It is dem-
onstrated that the unloading stress–strain slope is smaller than the ‘‘initial’’ Young modulus from the first unloading, this
phenomenon is explained by the existence of a very small plasticity appearing in the course of unloading.
In order to have an idea on the validity of this approach in our case, let us evaluate the cyclic evolution of the Young mod-
ulus (or more accurately the cyclic evolution of the unloading stress–strain slope) for the tests 35NiCrMo16_200-1050 and
XC18_50-320 (Fig. 18).
~ is evaluated from the monotonic tensile curve of the material represented
For both tests, the initial value (cycle 1) of E ¼ E
by the loading part of the first cycle. For the other cycles, E ~ is estimated during the unloading of the cycle under question
(Figs. 5c, 6a).
In both tests, the engineering elastic modulus (ratio between the engineering axial stress and strain) decreases during the
cycles. However, the evolution of the true elastic modulus (ratio between the true axial stress and strain) is rather small after
the first cycle especially for the 35NiCrMo16 steel which brings the doubt about the presence of fatigue damage here. There-
fore, one can conclude that in our case it is difficult to assess fatigue damage through the cyclic variation of the stress–strain
slope during the unloading parts of the cycles. Microstructural investigations may be welcome in order to have a better idea
about the presence of damage fatigue.
Finally, more investigations seem necessary in order to evaluate the contribution of the mode of control (make tests un-
der controlled true stress) and the fatigue damage (investigate the microstructure of the specimens) in the cyclic accumu-
lation of the inelastic strain.

4.2. Cyclic accumulation of the inelastic strain in multiaxial tests

In the tests 304L_250-0.4%, 35NiCrMo16_900-0.8%, XC18_250-0.4% and AU4G_300-0.8%, significant cyclic accumulation
of the inelastic strain is observed: we try here to understand the origin of such accumulation. However, as the triangle path

Fig. 18. Evolution of the true and engineering elastic modulus vs the number of cycles: (a) 35NiCrMo16_200-1050; (b) XC18_50-320.
L. Taleb / International Journal of Plasticity 43 (2013) 1–19 15

applied is rather complex, let us start by giving a schematic representation of the behavior during the first and the 100th
cycles of this path.

4.2.1. Approximation of the different positions of the yield surface during a cycle
Our goal is to have an idea about the different positions taken by the yield surface during a cycle of the triangular path. For
this objective, we consider the test 304L_250-0.4% where a hold time at 250 MPa in the first cycle has been applied in order
pffiffiffi
to minimize creep during cycling. We assume that the yield surface keeps constant size and circular shape in the (rz ; 3rzh )
space. Of course these assumptions are not exact as in addition to isotropic hardening, numbers of studies of literature like
Aubin and Degallaix (2006) demonstrate the distortion of the yield surface under non proportional loading paths. However,
our objective here is only to have an idea about how the yield surface moves. The position may be approximated through the
pffiffiffi pffiffiffi
knowledge of two points: (rz ; 3rzh ) gives the position of the current stress state and (e_ pz ; c_ pzh = 3) enables to have an idea
about the orientation of the inelastic strain rate following the normality rule. For a given instant in the cycle, these quantities
pffiffiffi pffiffiffi
may be deduced from rz = f(t), 3rzh ¼ f ðtÞ, epz ¼ f ðtÞ and, cpzh = 3 ¼ f ðtÞ: These evolutions are given in Fig. 19 for the first and
pffiffiffi pffiffiffi
the 100th cycles of the test 304L_250-0.4% leading to the values of rz ; 3rzh ) and the signs of (e_ pz ; c_ pzh = 3) at the end of each
segment of these cycles (Table 3). In order to have only the evolution due to a given cycle, epz has been shifted to zero at the
beginning of the cycle under question. The data of Table 3 enables us to represent schematically the different positions taken
by the yield surface during the first cycle (Fig. 20).
p
At the end of the first segment (Fig. 20a), the applied stress is purely axial which leads to e_ with no shear component.
After the first segment, the axial stress decreases while the shear strain increases; this variation starts in elasticity as indi-
cated in Fig. 20a by the discontinuous arrow. When the stress state reaches the yield surface the inelastic strain rate has two
positive axial and shear components but at the end of segment 2, the axial component becomes small and negative: most of
the flow occurs in the shear direction at this moment (Fig. 20a). In segment 3, shear strain comes back to zero while the axial
stress is kept equal to zero. The behavior starts inside the yield surface (discontinuous vertical arrow) and ends with a pure
shear strain rate (Fig. 20b). Segment 4 is the same as segment 1: the axial stress goes from zero to the maximum when the
shear strain is kept equal to zero. The behavior here starts slightly in elasticity (discontinuous arrow) and at the end of this
segment, significant inelastic strain rate is observed: positive for the axial component and negative for the shear component
(Fig. 20c). In segment 5, the axial stress comes back to zero while the shear strain varies from zero to its maximum (negative)

pffiffiffi pffiffiffi
Fig. 19. Evolution of the stress and plastic strain components versus time: (a) rz and 3rzh and b) epz and, cpzh = 3 for the first cycle of the test 304L_250-
0.4%. (c) and (d): same evolutions but for the 100th cycle. epz is shifted to zero in order to have only the evolution during the cycle under question. Time is in
seconds and stresses in MPa.
16 L. Taleb / International Journal of Plasticity 43 (2013) 1–19

Tablepffiffiffi3 pffiffiffi
(rz ; 3rzh ) and signs of (e_ pz ; c_ pzh = 3) at the end of each segment (S) of the cycles 1 and 100 of the test 304L_250-0.4%.

End of S1 End of S2 End of S3 End of S4 End of S5 End of S6


pffiffiffi
Cycle 1 rz = 3rzh 250/0 0/290 0/255 250/157 0/300 0/236
Sign of e_ pz =c_ pzh +/0 /+ 0/ +/ / /+
pffiffiffi
Cycle 100 rz = 3rzh 250/164 0/320 0/200 250/160 0/315 0/204
Sign of e_ pz =c_ pzh 0/+ /+ +/ 0/ / /+

p
Fig. 20. Schematic positions of the yield surface and direction of the inelastic strain rate (e_ ) at different instants of the first cycle: (a) ends of segments 1
and 2; (b) ends of segments 2 and 3; (c) ends of segments 3 and 4; (d) ends of segments 4 and 5; (e) ends of segments 5 and 6; (f) ends of all segments.
Stresses are in MPa.

value; the behavior during this part of the cycle seems entirely elastoplastic leading to purely shear inelastic strain rate at the
end of the segment (Fig. 20d). In segment 6, the shear strain comes back to zero while the axial stress is maintained equal to
zero. The behavior starts in elasticity (discontinuous vertical arrow) and when the stress state reaches the yield surface, the
variation of the inelastic strain rate is mainly in the shear direction even if small negative inelastic axial strain rate is ob-
served. These results suggest that the cyclic accumulation of the inelastic strain occurs mainly during the two segments 4
and 5. If we examine the results related to the 100th cycle (Fig. 19c, d), one can remark that the inelastic axial strain is
the same between the beginning and the end of the cycle. A steady state is obtained after the development of a cyclic strain
hardening in the material in both directions. Indeed, if we compare the 100th and the first cycles: the amplitude of variation
of the inelastic axial strain is smaller while the shear stress amplitude is larger.

4.2.2. Contribution of the mode of control


A comparison between the axial true stress and the axial engineering stress is given in Fig. 21.
The results of the tests 304L_250-0.4%, 35NiCrMo16_900-0.8% and XC18_250-0.4% do not show any significant relative
difference between the engineering and true stresses. The evolution is different for the test AU4G_300-0.8% where the dif-
ference increases during the first 100 cycles while it decreases during the following 100 cycles. However, one can assume
that such variation increases globally the cyclic accumulation of the strain in the test AU4G_300-0.8%.
L. Taleb / International Journal of Plasticity 43 (2013) 1–19 17

Fig. 21. Comparison between the peaks of the engineering and the true stresses in the tests: (a) 304L_250-0.4%; (b) 35NiCrMo16_900-0.8%; (c) XC18_250-
0.4% and, (d) AU4G_300-0.8%.

4.2.3. Contribution of creep


As demonstrated earlier, among the metals considered here, the 304L and 316L steels seem the ones where creep appears
significantly at room temperature. For that reason, the tests 304L_250-0.4% and 316L_250-0.4% have been started by a hold
time at the peak axial stress of the first segment of cycle 1 (see Fig. 2b). The results confirm the importance of this phenom-
enon but the considered hold time (2 h) is plainly not sufficient to vanish creep as the strain rate remains significant at the
end of the 2 h (Fig. 22). Therefore, one can assume that creep has probably a small contribution in the cyclic growth of the
inelastic strain during the subsequent applied cycles.
As the loading is multiaxial, the evaluation of the pseudo-damage through the variations of the elastic modulus (E) as for
the uniaxial case considered above is not easy here. Indeed, as the loading combine axial and shear stresses, E is the ratio
between the axial stress and axial strain variations provided that these variations occur into the elastic domain. However
this condition is not always verified here (see Figs. 19 and 20) which makes unfortunately such analysis difficult.

Fig. 22. Creep tests performed during the first cycle of the tests: (a) 304L_250-0.4%; (b) 316L_250-0.4%.
18 L. Taleb / International Journal of Plasticity 43 (2013) 1–19

After these results of the multiaxial tests, it is clear that even if the mode of control and other phenomena like creep may
contribute in the cyclic growth of the inelastic strain, the latter is mainly due to the multiaxial character of the loading where
the direction of the inelastic strain rate is governed by the normality rule.

5. Conclusion

In this paper, the cyclic behavior of different classes of metallic materials has been investigated at room temperature. The
tests have been conducted on specimens made up of two austenitic stainless steels 304L and 316L, two ferritic steels XC18
and 35NiCrMo16, an aluminum alloy Al 2017 and a copper–zinc alloy CuZn27. Two loading paths have been applied for each
material: the first is proportional (tension–compression load control) while the second is non proportional following a tri-
angular path in the axial stress – shear strain space. The contributions of certain phenomena like creep and ratcheting in the
cyclic accumulation of the inelastic strain have been scrutinized here.
The obtained results lead to two preliminary observations:

i. the right sense of the cyclic accumulation of the inelastic strain cannot be always described by the evolution of the
strain peaks; the use of the average value may be more suitable;
ii. Contrary to austenitic steels, creep is rather negligible at room temperature for the ferritic steels, the aluminum and
the copper–zinc alloys under consideration.

The other main observations may be summarized in the following concluding remarks:

5.1. Under tension–compression

 The cyclic accumulation of the inelastic strain observed for the austenitic steels 304L and 316L at room temperature
seems mainly due to creep.
 No significant cyclic accumulation of the inelastic strain is observed for the aluminum and copper–zinc alloys under the
loadings applied in the present study.
 The ferritic steels 35NiCrMo16 and XC18 exhibit cyclic accumulation of the inelastic strain where the contribution of the
mode of control (load) may be present. The macroscopic assessment of the presence of fatigue damage through the vari-
ations of the elastic modulus does not lead to indisputable conclusions, microstructural investigations of the specimens
will be necessary.

These results mean that more investigations are necessary to understand ratcheting for these metals under tension–com-
pression stress control at room temperature.

5.2. Under multiaxial loading

All metals considered here show cyclic accumulation of the inelastic strain under non proportional loading path. Further-
more, the test responses demonstrate that even if some phenomena like the mode of control and creep may have a contri-
bution in the cyclic growth observed, the latter is mainly due to the multiaxial character of the loading where the direction of
the inelastic strain rate is governed by the normality rule.
In order to understand the physical mechanisms responsible of the macroscopic observations, some microstructural anal-
yses will be performed in the future on the specimens tested in this study. Numerical investigations will also be carried out
using the multimechanism model (Taleb and Cailletaud, 2010) after the identification of its parameters for the different
materials considered here. The extension of this model to take into account the coupling between viscoplasticity and damage
will also be regarded in the future.

Acknowledgements

The authors acknowledge the ‘‘Région Haute Normandie’’ and the European Community, through the FEDER program for
their grateful financial support and the company FAVI for providing the specimens of copper–zinc alloy CuZn27.

References

Abdel Karim, M., 2011. Effect of elastic modulus variation during plastic deformation on uniaxial and multiaxial ratchetting simulations. Eur. J. Mech. A.
Solids 30 (2011), 11–21.
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., 2009. Modified kinematic hardening rules for simulations of ratchetting. Int. J. Plast. 25 (2009), 1560–1587.
Abdel Karim, M., Ohno, N., 2000. Kinematic hardening model suitable for ratcheting with steady-state. Int. J. Plast. 16, 225–240.
Aubin, V., Degallaix, S., 2006, ‘‘Contribution of yield surface measurement to plastic behaviour modeling’’, Plasticity’06, In: twelfth International Symposium
on Plasticity and its Current Applications, 17–22 July, Halifax, Nova Scotia, Canada.
Bari, S., Hassan, T., 2000. Anatomy of coupled constitutive models for ratchetting simulation. Int. J. Plast. 16, 381–409.
L. Taleb / International Journal of Plasticity 43 (2013) 1–19 19

Bari, S., Hassan, T., 2002. An advancement in cyclic plasticity modeling for multiaxial ratcheting simulation. Int. J. Plast. 18, 873–894.
Bocher, L., Delobelle, P., Robinet, P., Feaugas, X., 2001. Mechanical and microstructural investigations of an austenitic stainless steel under non-proportional
loadings in tension torsion internal and external pressure. Int. J. Plast. 17, 1491–1530.
Chaboche, J.L., 2008. A review of some plasticity and viscoplasticity constitutive theories. Int. J. Plast. 24, 1642–1693.
Chaboche, J.L., Kanouté, P., Azzouz, F., 2012. Cyclic inelastic constitutive equations and their impact in the fatigue life predictions. Int. J. Plast. 35, 44–66.
Chaboche, J.L., Nouailhas, D., 1989. A unified constitutive model for cyclic viscoplasticity and its applications to various stainless steels. J. Eng. Mater.
Technol. 111, 424–430.
Dafalias, Y.F., Feigenbaum, H.P., 2011. Biaxial ratchetting with novel variations of kinematic hardening. Int. J. Plast. 27 (2011), 479–491.
Gaudin, C., Feaugas, X., 2004. Cyclic creep process in AISI 316L stainless steel in terms of dislocation patterns and internal stresses. Acta Mater. 52, 3097–
3110.
Gao, B., Chen, X., Chen, G., 2006. Cyclic creep process in AISI 316L stainless steel in terms of dislocation patterns and internal stresses. Int. J. Press. Vessels
Pip. 83 (2006), 96–106.
Gupta, C., Chakravartty, J.K., Reddy, G.R., Banerjee, S., 2005. Uniaxial cyclic deformation behaviour of SA 333 Gr 6 piping steel at room temperature. Int. J.
Press. Vessels Pip. 82 (2005), 459–469.
Hassan, T., Taleb, L., Krishna, S., 2008. Influences of non proportional loading paths on ratcheting responses and simulations by two recent cyclic plasticity
models. Int. J. Plast. 24, 1863–1889.
Ikegami, K., Niitsu, Y., 1985. Effect of creep prestrain on subsequent plastic deformation. Int. J. Plast. 1, 331–345.
Jiang, Y., Zhang, J., 2008. Benchmark experiments and characteristic cyclic plasticity deformation. Int. J. Plast. 24, 1481–1515.
Kang, G.Z., 2008. Ratchetting: recent progresses in phenomenon observation, constitutive modeling and application. Int. J. Fatigue 30 (2008), 1448–1472.
Kang, G.Z., Gao, Q., 2002. Uniaxial and non-proportionally multiaxial ratcheting of U71Mn rail steel: experiments and simulations. Mech. Mater. 34, 809–
820.
Kang, G.Z., Kan, Q., 2007. Constitutive modeling for uniaxial time-dependent ratcheting of SS304 stainless steel. Mech. Mater. 39, 488–497.
Kang, G.Z., Kan, Q., Zhang, J., Sun, Y.F., 2006. Time-dependent ratchetting experiments of SS304 stainless steel. Int. J. Plast. 22, 858–894.
Kang, G.Z., Gao, Q., Yang, X.J., 2004. Uniaxial and nonproportional multiaxial ratchetting of SS304 stainless steel at room temperature: Experiments and
simulations. 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 and its constitutive
modeling. Mech. Mater. 37, 1101–1118.
Kang, G.Z., Liu, Y., Ding, J., Gao, Q., 2009. Uniaxial ratcheting and fatigue failure of tempered 42CrMo steel: damage evolution and damage-coupled. Int. J.
Plast. 25, 838–860.
Kawai, M., Ohashi, Y., 1987. Couple effect between creep and plasticity of type 316 stainless steel at elevated temperature. In: Desai, C.S. (Ed.), Second Int.
Conf. on Constitutive Laws for Engineering Materials: Theory and Applications, Elsevier, Tucson, pp. 967–974.
Khan, A.S., Chen, X., Abdel-Karim, M., 2007. Cyclic multiaxial and shear finite deformation response of OFHC: Part I, experimental results. Int. J. Plast. 23,
1285–1306.
Krempl, E., 1979. An experimental study of room temperature rate-sensitivity, creep and relaxation of AISI Type 304 stainless steel. J. Mech. Phys. Solids 27,
363–375.
Krishna, S., Hassan, T., Ben Naceur, I., Sai, K., Cailletaud, G., 2009. Macro versus micro-scale constitutive models in simulating proportional and non
proportional cyclic and ratcheting responses of stainless steel 304. Int. J. Plast. 25, 1910–1949.
Lemaître, J., Chaboche, J.L., 1985. Mécanique des Matériaux Solides. Dunod, Paris.
McDowell, D.L., 1987. An evaluation of recent developments in hardening and flow rules for rate-independent, nonproportional cyclic plasticity. ASME J.
Appl. Mech. 54, 323–334.
Moosbrugger, J.C., 1993. Experimental parameter estimation for nonproportional cyclic viscoplasticity: nonlinear kinematic hardening rules for two
waspaloy microstructures at 650 °C. Int. J. Plast. 9, 345–373.
Murakami, S., Ohno, N., 1982. A constitutive equation of creep based on the concept of a creep-hardening surface. Int. J. Solids Struct. 18, 597–609.
Ohno, N., 1997. Recent progress in constitutive modeling for ratchetting. Mater. Sci. Res. Int. 3, 1–9.
Ohno, N., Wang, J.D., 1993. Kinematic hardening rules with critical state of dynamic recovery, part I: formulations and basic features for ratcheting behavior.
Int. J. Plast. 9, 375–403.
Ohno, N., Wang, J.D., 1994. Kinematic hardening rules for simulation of ratchetting behavior. Eur. J. Mech. A. Solids 13, 519–531.
Portier, L., Calloch, S., Marquis, D., Geyer, P., 2000. Ratchetting under tension-torsion loadings: experiments and modelling. Int. J. Plast. 16, 303–335.
Grostabussiat, S., Taleb, L., Jullien, J.F., 2004. Experimental results on classical plasticity of steels subjected to structural transformations. Int. J. Plast. 20 (8-9),
1371–1386.
Taleb, L., Cailletaud, G., Blaj, L., 2006. Numerical simulation of complex ratcheting tests with a multi-mechanism model type. Int. J. Plast. 22, 724–753.
Taleb, L., Cailletaud, G., 2011. Cyclic accumulation of the inelastic strain in the 304L SS under stress control at room temperature: ratcheting or creep? Int. J.
Plast. 27, 1936–1958.
Taleb, L., Cailletaud, G., 2010. An updated version of the multimechanism model for cyclic plasticity. Int. J. Plast. 26, 859–874.
Taleb, L., Hauet, L., 2009. Multiscale experimental investigations about the cyclic behavior of the 304L SS. Int. J. Plast. 25, 1359–1385.
Vincent, L., Calloch, S., Marquis, D., 2004. A general cyclic plasticity model taking into account yield surface distortion for multiaxial ratcheting. Int. J. Plast.
20, 1817–1850.
Wolff, M., Taleb, L., 2008. Consistency of two multi-mechanism models in isothermal plasticity. Int. J. Plast. 24, 2059–2083.
Yoshida, F., 1990. Uniaxial and biaxial creep-ratcheting behavior of SUS304 stainless steel at room temperature. Int. J. Press. Piping 44, 207–223.
Yoshida, F., Murata, K., Shiratori, E., 1980. Constitutive equation of cyclic creep under increasing stress condition. Bull. JSME 177 (23), 337–344.
Yoshida, F., Uemori, T., Fujiwara, K., 2002. Elastic-plastic behavior of steel sheets in-plane cyclic tension-compression at large strain. Int. J. Plast. 18, 633–
659.

Anda mungkin juga menyukai