You are on page 1of 141

ABSTRACT

BARI, MD. SHAFIQUL Constitutive Modeling for Cyclic Plasticity and Ratcheting.
(under the supervision of Dr. Tasnim Hassan)

This study critically evaluates the performance of a number of constitutive models in


predicting ratcheting responses of carbon steel for a set of uniaxial and biaxial loading
histories. Two types of modeling schemes, coupled and uncoupled, are evaluated. The
coupled models from Prager, Armstrong-Frederick, Chaboche, Ohno-Wang, and
Guionnet are examined. The Prager and the Armstrong-Frederick models perform
inadequately. The Chaboche and Ohno-Wang models perform well for uniaxial
ratcheting responses, but overpredict the biaxial ratcheting. The Guionnet model
simulates one set of biaxial ratcheting response well, but fails in others. Performances of
several kinematic hardening rules, when used with the uncoupled Dafalias-Popov model
are also evaluated. The Armstrong-Frederick rule simulates one set of biaxial response
reasonably. The Voyiadjis-Sivakumar, Phillips, Tseng-Lee, Kaneko and Xia-Ellyin rules
fail to simulate the biaxial ratcheting responses. The Chaboche rule, with three
decomposed terms, performs reasonably for the whole set of responses. The Ohno-Wang
rule also performs reasonably, except for one biaxial response. This study indicates a
strong influence of the kinematic hardening rule and its parameter determination scheme
on multiaxial ratcheting simulations. The coupled models by McDowell, Jiang-Sehitoglu,
Voyiadjis-Basuroychowdhury and AbdelKarim-Ohno, where additional multiaxial
parameters are included in the hardening rules, are also investigated. None of these
models perform consistently for the whole set of experiments. A modified kinematic
hardening rule using the idea of Delobelle and his co-workers in the framework of the
Chaboche model is proposed. This new rule performs impressively for all of the
ratcheting responses considered. Several models for anisotropic deformation of the yield
surface are scrutinized. Most of these models use complex and numerically extensive
higher order tensors for the yield surface formulations and thus become less attractive for
implementation with a cyclic plasticity model. This study demonstrates the methodology
and promise in incorporating the equi-plastic-strain surface proposed by Shiratori and his
co-workers into the Dafalias-Popov model for general multiaxial ratcheting simulations.
CONSTITUTIVE MODELING FOR CYCLIC
PLASTICITY AND RATCHETING

BY

SHAFIQUL BARI

A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE


REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILISOPHY

DEPARTMENT OF CIVIL ENGINEERING


NORTH CAROLINA STATE UNIVERSITY

UNDER THE GUIDANCE OF

DR. TASNIM HASSAN

RALEIGH
JANUARY 2001

APPROVED BY:

CHAIR OF ADVISORY COMMITTEE CO-CHAIR OF ADVISORY COMMITTEE


BIOGRAPHY

Shafiqul Bari was born on January 11th, 1969 in Faridpur, Bangladesh. After
graduating from the Bangladesh University of Engineering and Technology in 1993, with
a degree in Civil Engineering, he joined the University as a lecturer in the Department of
Civil Engineering. He obtained the Master of Science degree from the same university in
1996. In the fall of 1996, he enrolled at the North Carolina State University to pursue his
doctoral studies in Engineering Mechanics. After completing his doctoral study, he will
work for a company in the automobile industry at Michigan, USA.

ii
ACKNOWLEDGEMENTS

I sincerely thank Dr. Tasnim Hassan for his guidance and support during my
research. His time and effort in guiding the direction of this research was indispensable in
the completion of this thesis. He was always there to support me in every aspect of my
stay at North Carolina State university.

I would like to express my thanks to Prof. Vernon Matzen for his constant support
and suggestions during the course of this thesis. Working with him in conducting
experiments on nuclear power plant piping components for monotonic and cyclic loading
was an enlightening experience for me.

I would also like to express my gratitude to Prof. Kerry Havner for his invaluable
suggestions and counsel. His comments have been extremely valuable in the preparation
of this thesis.

I would also like to thank Dr. Mohammed Zikry for his time, encouragement and
suggestions.

Special Thanks should go to Dr. James Nau for his support and encouragement
during my stay at NC State University.

I like to acknowledge gratefully the financial support provided by the Center for
Nuclear Power Plant Structures, Equipment and Piping and the Department of Civil
Engineering to conduct this research.

I would like to express my heartfelt gratitude to the professors, staff and students
at the Center for Nuclear Power Plant Structures, Equipment and Piping for helping me in
numerous occasions during the course of this thesis.

A very special note of thanks and appreciation goes to my wife, Tahsina Ahmed
for her constant emotional and moral support to keep me in focus.

iii
TABLE OF CONTENTS

List of Tables v

List of Figures vi

1 Introduction 1

1.1. Background and Motivation 1


1.2. Scope and Organization 3

2 Coupled Constitutive Models for Ratcheting Simulations 8

Paper I 9

3 Uncoupled Constitutive Models for Ratcheting Simulations 48

Paper II 49

4 Improvement in Coupled Constitutive Models for Ratcheting Simulations 76

Paper III 77

5 Ratcheting Simulations Using Formative Hardening Rule 106

5.1. Introduction 106


5.2. Plastic Potential Theory based on the equi-plastic-strain surface 115
5.2.1. Coordinate system 115
5.2.2. Flow rule 115
5.2.3. Model I 116
5.2.3. Model II 120
5.3. Conclusion and Discussions 124

6 Conclusions and Recommendations 126

6.1. Conclusions 126


6.2. Recommendations for Future Research 130

References 131

iv
LIST OF FIGURES

1.1 Loading histories; (a) Uniaxial stress cycles, (b) Axial strain cycle with 4
constant pressure, (c) biaxial bow-tie cycle, (d) biaxial reverse bow-tie
cycle.

1.2 Axial strain ratcheting response for uniaxial stress controlled cycles with a 5
mean stress (from Hassan and Kyriakedes [1992])

1.3 Circumferential strain ratcheting response for (a) constant pressure and (b) 6
bow-tie loading histories of Fig 1.1 (from Hassan et al. [1992] and Corona
et al. [1196])

5.1 Comparison of experimental and von-Mises yield surfaces at different 106


stress points

5.2 Yield surface deformation law from Phillips and Tang [1972] 107

5.3 A schematic diagram of deformation and translation of yield surface from 109
Eisenberg and Yen [1984]

5.4 Qualitative equi-plastic-strain surface in Ilyushin [1954] stress space 111

5.5 Subsequent equi-plastic-strain surfaces and plastic strain increment 113


directions

5.6 Equi-plastic-strain surface I in the moving stress space i of the Ilyushin 117
space

5.7 Biaxial ratcheting predictions by model I and experiments from Hassan 119
and Kyriakides [1992].

5.8 Equi-plastic-strain surface II in the moving stress space i of the 121


Ilyushin space

5.9 Biaxial ratcheting predictions by model II and experiments from Hassan 123
and Kyriakides [1992].

v
CHAPTER 1

INTRODUCTION

1.1. BACKGROUND AND MOTIVATION

Ratcheting, one of the low cycle fatigue responses, is defined as the accumulation
of plastic strain with cycles. There are many type of structures that are subjected to cyclic
loading where the stress state exceed the elastic limit of the materials used. For design
and analysis of these type of structures, accurate prediction of ratcheting response is
critical as ratcheting can lead to catastrophic failure of the structures. Even for structures
that are designed to be within the elastic limit, plastic zones may exist at discontinuities
or at the tip of cracks. The fatigue cracks can initiate at these plastic zones. Therefore,
better simulation model for cyclic plasticity response is important for the prediction of
the high cycle fatigue life as well.

The modeling of cyclic plasticity responses is quite complex. Experimental


studies demonstrate that the yield surfaces grow, translate as well as change shape with
plastic loading. Some metals harden, while many others soften, during plastic cyclic
loading. Moreover, the cyclic plasticity responses are history dependent. Most of the
existing constitutive models fail to simulate these complex phenomena representatively
as they assume idealized simple yield functions and hardening rules. In addition, most of
the cyclic plasticity models for ratcheting simulations are developed and verified using
data from limited or simple experiments. They are rarely tested against a wide variety of
ratcheting responses to verify the generality of these models. Consequently, most of these
constitutive models might predict a special class of ratcheting responses quite well, but
fail to predict ratcheting responses in structures.

Although most metals cyclically harden or soften up to a certain number of


cycles, they subsequently stabilize. Ratcheting strains, however, keep on accumulating
with cycles even after the material stabilizes. The kinematic hardening (translation of the

1
yield surface in stress space) is attributed to be the primary reason for ratcheting.
Therefore, in order to develop and verify a basic model for ratcheting simulation, it is
essential to study the ratcheting responses of stabilized materials. After achieving a
robust model for ratcheting responses of cyclically stable materials, it can be extended to
cyclically hardening and softening materials.

The modeling for ratcheting simulations under uniaxial loading is inherently


different from that under multiaxial loading. Also, the role of various parameters differs
considerably depending on the loading condition. Most of the existing models apparently
fail to appreciate this distinction. The parameter determination schemes followed by these
models are, sometimes, incomplete and vague. As a result, these models are not general
and rarely perform representatively for varied loading conditions. It is, therefore,
essential to scrutinize the parameter determination schemes of various models for
ratcheting simulation and evaluate their performance against a broad set of experimental
responses.

The modeling schemes for the cyclic plasticity and ratcheting simulations can be
classified into two types: coupled and uncoupled. In the coupled modeling scheme, the
plastic modulus calculation is coupled with the kinematic hardening rule through the
consistency condition of the yield surface (Prager [1956]). Consequently, in uniaxial
loading condition, the kinematic hardening rule is effectively used to calculate the plastic
modulus, as the evolution direction is constant in this case. In multiaxial loading
condition, however, the hardening rule affects both the plastic modulus and the evolution
direction. If hardening rule parameters are calibrated from uniaxial loading responses
only, as has been done in many existing coupled models, the hardening rule is basically
designed for better representation of the plastic modulus, which undermines its role in
describing the direction of the yield surface evolution. Because of this approach in
modeling, many popular coupled models perform well in one type of loading condition
while fail in others.

In the uncoupled modeling scheme, on the other hand, the plastic modulus may be
indirectly influenced by the kinematic hardening rule but its calculation does not depend

2
on it. Therefore, the hardening rule parameters can be calibrated from multiaxial loading
conditions without having any effect on the simulation by the model in uniaxial loading
condition. However, the determination of the kinematic hardening rule parameter can be
somewhat subjective in uncoupled modeling scheme, unless simple hardening rules are
used.

In view of all these concerns in cyclic plasticity modeling to simulate ratcheting


responses, efforts have been made in this study to critically examine and evaluate many
existing coupled and uncoupled models against a uniform set of experiments which
include both uniaxial and multiaxial loading conditions. For all of the models considered,
emphasis has been given on the theoretical basis, numerical soundness, parameter
determination scheme and their effects on ratcheting responses. Insight gained from this
process has been employed to propose improved ratcheting simulation models that
perform consistently better in both uniaxial and multiaxial loading conditions. Finally,
investigations are conducted to evaluate the prospect of incorporating the yield surface
shape change into the constitutive modeling of cyclic plasticity.

1.2. SCOPE AND ORGANIZATION

Ratcheting experiments

It is imperative to obtain a broad set of ratcheting experiment to evaluate the state-


of-art of the ratcheting models. A broad set of quasi-static ratcheting data, which include
uniaxial to complex biaxial ratcheting responses of stabilized carbon steels, have been
developed by Hassan and Kyriakides [1992], Hassan et al. [1992] and Corona et al.
[1996]. These data are used to evaluate the performance of all the models considered in
this study.

The loading histories of the experiment set are shown in Fig. 1.1. The history in
Fig. 1.1a, which results in ratcheting of axial strains, involves axial stress controlled
cycles with a mean stress. The ratcheting response from this uniaxial loading history is

3
x
xc
xa
xm
t m

x
(a) (b)


xc xc

m m

x x
(c) (d)

Fig 1.1 Loading histories; (a) Uniaxial stress cycles, (b) Axial strain cycle with constant
pressure, (c) biaxial bow-tie cycle, (d) biaxial reverse bow-tie cycle.

shown in Fig. 1.2. Two sets of uniaxial ratcheting experiments are considered in this
xm) vary with constant amplitude stresses (
study. In the first set, the mean stresses ( xa).
Whereas in the second set, the amplitude stresses vary while mean stresses are kept
constant. The plots of peak axial strains (xp) at each cycle against the number of cycles,
N are presented in the paper I (Fig. 8a and b).

The loading history in Fig. 1.1b involves axial strain cycles in the presence of a
constant internal pressure. This history results in ratcheting of the circumferential strain
as shown in Fig. 1.3a. Two sets of data from this biaxial loading history are also
considered in this study. The peak circumferential strains (p) at each cycle are plotted
against the number of cycles, N in paper I (Figs. 8c and d). For the set shown in Fig. 8c,
the axial strain amplitudes (xc) are varied with same internal pressure (
m), whereas for

4
the set shown in Fig. 8d, the internal pressure varies while the axial strain amplitudes are
unchanged.

x
xa
x
t

Fig. 1.2 Axial strain ratcheting response for uniaxial stress controlled cycles
with a mean stress (Hassan and Kyriakides [1992])

The so-called bow-tie and reverse bow-tie loading histories shown in Fig. 1.1c
and d also result in circumferential strain ratcheting. The ratcheting response from the
bow-tie loading history is shown in Fig 1.3b. It should be noted that the bow-tie and
reverse bow-tie loading histories are typically observed in piping components with
internal pressure. The peak circumferential strains (p) at each cycle for these histories
are plotted in the paper I (Fig. 8e and f) against the number of cycles, N.

Coupled models

The performance of five well-known coupled constitutive models are evaluated in


Chapter 2 in terms of their simulations for the set of uniaxial and biaxial ratcheting
responses on carbon steel. The models proposed by Prager [1956], Mroz [1967],
Armstrong and Frederick [1966], Chaboche [1986, 1991], Guionnet [1992] and Ohno and

5
(a) (b)

Fig. 1.3 Circumferential strain ratcheting response for (a) constant pressure and (b)
bow-tie loading histories of Fig 1.1 (from Hassan et al. [1992 and Corona et al. [1996])

Wang [1993] are discussed anatomically in this chapter. For some of the models,
elaborate parameter determination schemes have been proposed. A journal paper has
been written from this part of the study, which is enclosed in Chapter 2 as paper I.

Uncoupled models

The performances of eight different kinematic hardening rules, when engaged


with the modified Dafalias and Popov [1976] uncoupled model, are evaluated in Chapter

6
3 against the same set ratcheting responses. The ratcheting simulations from the
kinematic hardening rules by Armstrong and Frederick [1966], Phillips [1972,1979],
Tseng and Lee [1983], Voyiadjis and Sivakumar [1991,1994], Kaneko [1981,1984], Xia
and Ellyin [1994,1997], Chaboche [1991] and Ohno-Wang [1993] are evaluated in this
chapter. The journal paper written from this part of the study is enclosed Chapter 3 as
paper II.

Improved coupled models

Chapter 4 evaluates several coupled models that try to improve their multiaxial
ratcheting simulations by introducing additional maultiaxial parameter(s) or term(s) in the
kinematic hardening rule. The models studied are by McDowell [1995], Jiang and
Sehitoglu [1996], Voyiadjis and Basuroychowdhury [1998], and AbdelKarim and Ohno
[2000]. A modified kinematic hardening rule using the idea of Delobelle and his
coworkers [1995] in the framework of the Chaboche [1991] model is proposed in this
chapter. A journal paper has been written from this part of the study and is enclosed in
Chaper 4 as paper III.

Formative Hardening rules with anisotropic yield surfaces

In all the models considered in the Chapters 2, 3 and 4, the von-Mises yield
surface is considered as the yielding boundary and also, the plastic potential surface. But
in reality, the yield surface translates as well as deforms during plastic loading (Phillips
and Tang [1972], Phillips and Lee [1979]). The incompleteness, thus, introduced in the
modeling bars the models from being successful in the general multiaxial loading
conditions. Therefore, investigations have been conducted in Chapter 5 to incorporate the
anisotropically deformed yield surface in the cyclic plasticity modeling (formative
hardening rule).

Finally, Chapter 6 presents a summary of significant conclusions and


recommendations from this study.

7
CHAPTER 2

COUPLED CONSTITUTIVE MODELS FOR RATCHETING


SIMULATIONS

The cyclic plasticity modeling schemes can be classified into two types: coupled
and uncoupled. There are four basic aspects that are used to define the constitutive
relations in cyclic plasticity models: (i) the yield surface, (ii) the flow rule derived from
the normality condition of plastic strain increments, (iii) the hardening rule and (iv) the
consistency condition. In coupled models, following the classical model by Parger
[1956], the plastic modulus is calculated from the kinematic hardening rule and the
consistency condition. This study critically evaluates the performances of five
constitutive models in predicting ratcheting responses of carbon steels for a broad set of
uniaxial and biaxial loading histories. A journal paper "Anatomy of Coupled Constitutive
Models for Ratcheting Responses" is published in the International Journal of Plasticity
from this part of the study. This paper is enclosed next as paper I. Readers are referred to
the enclosed paper I for details of the study conducted in this chapter.

8
Paper I-International Journal of Plasticity 16 (2000) 381-409

ANATOMY OF COUPLED CONSTITUTIVE MODELS FOR RATCHETING


SIMULATION

Shafiqul Bari and Tasnim Hassan*


Center for Nuclear Power Plant Structures, Equipment and Piping
Department of Civil Engineering
North Carolina State University
Raleigh, NC 27695-7908
e-mail: thassan@eos.ncsu.edu

Keywords: Cyclic plasticity, Ratcheting, Constitutive Modeling and Simulation


Shortened Title: Constitutive Models for Ratcheting

ABSTRACT

This paper critically evaluates the performance of five constitutive models in predicting ratcheting

responses of carbon steel for a broad set of uniaxial and biaxial loading histories. The models proposed

by Prager, Armstrong-Frederick, Chaboche, Ohno-Wang, and Guionnet are examined. Reasons for

success and failure in simulating ratcheting by these models are elaborated. The bilinear Prager and the

nonlinear Armstrong-Frederick models are found to be inadequate in simulating ratcheting responses.

The Chaboche and Ohno-Wang models perform quite well in predicting uniaxial ratcheting responses;

however, they consistently overpredict the biaxial ratcheting responses. The Guionnet model simulates

one set of biaxial ratcheting responses very well, but fails to simulate uniaxial and other biaxial ratcheting

responses. Similar to many earlier studies, this study also indicates a strong influence of the kinematic

hardening rule or backstress direction on multiaxial ratcheting simulation. Incorporation of parameters

dependent on multiaxial ratcheting responses, while dormant for uniaxial responses, into Chaboche-type

kinematic hardening rules may be conducive to improve their multiaxial ratcheting simulations. The

uncoupling of the kinematic hardening rule from the plastic modulus calculation is another potentially

viable alternative. The best option to achieve a robust model for ratcheting simulations seems to be the

incorporation of yield surface shape change (formative hardening) in the cyclic plasticity model.

* The corresponding author.

9
Paper I-International Journal of Plasticity 16 (2000) 381-409

I. INTRODUCTION

As the data base and understanding of ratcheting response (accumulation of strains with cycles) are

growing, the number of efforts in developing constitutive models for ratcheting is also increasing

(Chaboche and his coworkers [1979,1986,1991,1994], Voyiadjis and his coworkers [1991,1998],

Guionnet [1992], Ohno and Wang [1993], Hassan and Kyriakides [1994a,b], Delobelle et al. [1995],

McDowell [1995], Jiang and Sehitoglu [1996a], Ohno [1997], Xia and Ellyin [1997] and others). In many

of these models the plastic modulus calculation is coupled with its kinematic hardening rule through the

yield surface consistency condition as in the classical model proposed by Prager [1956]. These models

are referred to as coupled models in this paper. Models proposed by Armstrong-Frederick [1966],

Chaboche [1986,1991,1994], Guionnet [1992], and Ohno and Wang [1993] belong to this class and are

studied along with the Prager model [1956] in this paper.

In another class of models, the plastic modulus might be indirectly influenced by the kinematic

hardening rule but its calculation is not coupled to the kinematic hardening rule through the consistency

condition. The models proposed by Mroz [1967], Dafalias and Popov [1976], Drucker and Palgen [1981],

Tseng and Lee [1983] and many others belong to this class. These models are referred to as uncoupled

models and will be discussed in another paper (Bari and Hassan, [1999a]).

Most of the models proposed so far are developed and verified using data from limited or simple

experiments. These models have not been tested against a wide variety of ratcheting responses to verify

the generality of these models. Consequently, most of these constitutive models might predict a special

class of ratcheting responses quite well, but fail to predict a broad class of ratcheting responses (Hassan

and Kyriakides [1994b], Corona et al. [1996]).

Most metals cyclically harden or soften up to a certain number of cycles and subsequently stabilize

or cease to change the size of the yield surface (Morrow [1965], Jhansale [1975], Tuegel [1987], Ishikawa

10
Paper I-International Journal of Plasticity 16 (2000) 381-409

and Sasaki [1988]). Ratcheting, though, keeps on occurring with cycles even after the material stabilizes

(Hassan and Kyrikides [1992], Hassan et al. [1992], Hassan and Kyriakides [1994a,b]). Hence, the

kinematic hardening (translation of the yield surface in stress space) is attributed to be the primary reason

for ratcheting. Thus, in order to develop and verify a model for ratcheting, it is essential to study the

ratcheting responses of stabilized materials. This, in effect, means that the parameters affecting the

isotropic hardening (i.e, yield surface size change) should not be included during the model development

for ratcheting. Also, all of the kinematic hardening (i.e, yield surface translation) parameters should be

determined using experiments performed on stabilized materials. After achieving a robust model for

ratcheting responses of cyclically stable materials, it can easily be extended to cyclically hardening and

softening materials following the approach demonstrated by Hassan and Kyriakides [1994 a,b].

A broad set of quasi-static ratcheting data which include uniaxial to complex biaxial ratcheting

responses of stabilized carbon steels have been developed by Hassan and Kyriakides [1992], Hassan et al.

[1992] and Corona et al. [1996]. These data are used in this study to evaluate the performance of the

models considered. The cyclic loading histories prescribed in the experiments are shown in Fig. 1. The

history in Fig. 1a, which results in uniaxial ratcheting, involves axial stress cycles with a mean stress. The

readers are referred to Hassan and Kyriakides [1992] for demonstration of a uniaxial ratcheting response

(Fig. 7 in the reference). Two sets of uniaxial experiments are considered in this study. In the first set, the

amplitude stresses (xa) in the experiments are the same while the mean stresses (xm) vary. Whereas in

the second set, the amplitude stresses (xa) vary while the mean stresses (xm) are kept constant. Data

from these experiments are presented in Figs. 8a and 8b, where the peak axial strains (xp) of each cycle

are plotted against the number of cycles, N.

The loading history in Fig. 1b involves axial strain cycles in the presence of a constant internal

pressure. This history results in circumferential ratcheting as demonstrated in Fig.1 of Hassan et al.

11
Paper I-International Journal of Plasticity 16 (2000) 381-409


xc xc
xc
x a a
xa
xm m
m m
t
x x x
(a) (b) (c) (d)

Fig. 1. Loading histories; (a) Uniaxial stress cycle, (b) Axial strain cycle with constant internal
pressure, (c) biaxial bow-tie cycle, (d) biaxial reverse bow-tie cycle.

[1992]. Two sets of data from this biaxial loading history are also considered in this study as shown in

Figs. 8c and 8d. In these plots, the peak circumferential strains (p) of each cycle are plotted against the

number of cycles, N. For the set shown in Fig. 8c, the axial strain amplitude (xc) of the experiments vary

with same circumferential stress (m) due to internal pressure, whereas in the set of Fig. 8d, the internal

pressures of the experiments are different with same axial strain amplitude.

The bow-tie and reverse bow-tie loading histories in Figs. 1c and 1d also result in circumferential

ratcheting as demonstrated in Figs. 10 and 11 of Corona et al. [1996]. The data set from these two biaxial

loading histories considered in this study are shown in Figs. 8e and 8f.

The above mentioned ratcheting data are used to evaluate the performance of several coupled cyclic

plasticity models. Also, parameters for each model are determined using experiments on stabilized

carbon steels. The same set of parameters for each model are used to simulate uniaxial and multiaxial

ratcheting responses. The reasons for the success and failure in simulating these responses by the models

are presented through anatomical discussion of the influences of modeling schemes and parameters on

ratcheting simulation. Although the Prager [1956] and Armstrong and Frederick [1966] models are not

capable of simulating ratcheting responses satisfactorily, these models are discussed in this study in order

to demonstrate the gradual development of different features of plasticity models over time.

12
Paper I-International Journal of Plasticity 16 (2000) 381-409

II. RATCHETING MODELS

The plasticity models with the assumption of rate-independent material behavior has the following

common features:

12
i. von-Mises yield criterion: f ( ) = 3--- ( s a ) ( s a ) = 0 , (1)
2
p 1 f f
ii. flow rule: d = ---- ------ d ------ , (2)
H

p
Where, is the stress tensor, is the plastic strain tensor, s is the deviatoric stress tensor, is the

current center of the yield surface, a is the current center of the yield surface in the deviatoric space, 0 is

the size of the yield surface (constant for a cyclically stable material), and H is the plastic modulus. Also,

indicates the MacCauley bracket and the inner product a b = a ij b ij .


iii. The third and most important feature for ratcheting simulation, the kinematic hardening rule is

given by:

p p
da = g ( , , a, d, d , etc ) . (3)

The Kinematic hardening rule dictates the evolution of the yield surface during a plastic loading

increment by translation in the stress space only. Different models discussed in this study basically differ

in the kinematic hardening rules they employ.

In coupled models, the plastic modulus H is evaluated using the consistency condition, f = 0 ,

kinematic hardening rule (Eq. 3), flow rule (Eq. 2), and yield criterion (Eq. 1). The models considered in

this study are discussed below in terms of their features and their influence on ratcheting simulation.

13
Paper I-International Journal of Plasticity 16 (2000) 381-409

II.1. Linear Kinematic Hardening Model

Prager [1956] proposed perhaps the most simple kinematic hardening rule,

da = C d p (4)

to simulate plastic response of materials. For uniaxial loading, as the yield surface in this model moves

linearly with plastic strain (see x trace in Fig. 2a), the ensuing hysteresis loop is bilinear as shown in Fig.

2a. It is clear from the figure that this model cannot represent the experimental hysteresis curve during the

initial nonlinear part. Also, for a prescribed uniaxial stress cycle with a mean stress, this model fails to

distinguish between shapes of the loading and reverse loading hysteresis curves and consequently

produces a closed loop with no ratcheting (see Fig. 2b and 2c). For all the biaxial histories of Fig. 1, the

ratcheting by Pragers model stabilizes (shakedown) after some initial overprediction of ratcheting as

shown in Figs. 2d, 2e and 2f.

II.2. Multilinear Model

Improvement to the linear kinematic hardening model was proposed by Mroz [1967] as a

multisurface model, where each surface represents a constant work hardening modulus in the stress

space. Besseling [1958] introduced a multilayer model without any notion of surfaces. Also, Ohno and

Wang [1993] introduced a piecewise linear kinematic hardening rule. In uniaxial loading, all these

models essentially divide the stress-strain curve into many linear segments. When a sufficient number of

segments are chosen, the hysteresis loop simulations by these models are very good as shown in Fig. 3a.

Unfortunately, like the linear kinematic hardening model, multilinear models also predict a closed loop

when subjected to a uniaxial stress cycle with a mean stress and hence produce no uniaxial ratcheting (see

Fig. 3b).

14
Paper I-International Journal of Plasticity 16 (2000) 381-409

x Experiment x CS 1026
(ksi) 40 (ksi) 40 xa = 38.57 ksi
xm = 2.57 ksi

20 20

x
0 0

-20 -20

-40 -40

-1 -0.5 0 0.5 1 -0.5 0 0.5 1

(a) x (%)
p
(b) x (%)
p

xp p
4 4
CS 1026 CS 1026
(%) xa = 32.0 ksi (%) xc m = 9.65 ksi
xm = 6.52 ksi
x m (1) xc = 0.40 %
3 3
xa x (2) xc = 0.50 %
xm (3) xc = 0.65 %
t
2 2 (3)
(3)
(2)
(1)
(2)
Experiment (1)
1 Prager model 1

0 0
0 10 20 30 40 50 0 10 20 30 40 50

(c) N (d)
(c) N

p p
4 4
xc CS 1026 xc CS 1018
(%) a = 2.36 ksi (%) a = 2.36 ksi
a xc = 0.50 % a xc = 0.50 %
3 m 3 m
(2)
x (1) m = 9.65 ksi x (2) (1) m = 9.65 ksi
(2) m = 14.54 ksi (2) m = 14.54 ksi
2 (2) 2 (2)
(1)
(1) (1)

1 1 (1)
Experiment
Prager model

0 0
0 10 20 30 40 50 0 10 20 30 40 50

(e) N (f) N

Fig 2. Predictions from Pragers model for (a) strain-controlled stable hysteresis loop, (b)
stress-controlled hysteresis loop, (c) axial strain at positive stress peaks of uniaxial cycles,
(d) Circumferential strain peaks of axial strain cycles with constant pressure, (e,f)
Circumferential strain peaks under biaxial bow-tie and reverse bow-tie cycles.

15
Paper I-International Journal of Plasticity 16 (2000) 381-409

x Experiment x CS 1026
(ksi) 40 (ksi) 40 xa = 32.0 ksi
xm = 9.14 ksi
x
20 20

0 0

-20 -20

-40 -40

-1 -0.5 0 0.5 1 -0.5 0 0.5 1

x (%)
p
x (%)
p

(a) (b)
Fig 3. Predictions from Multilinear model for (a) strain-controlled stable hysteresis loop, (b)
stress-controlled hysteresis loop.

For multiaxial responses, the original Mroz [1967] model or a modified version of the model can

not properly describe the ratcheting responses from non-proportional loadings (Mroz and Rodzik [1996],

Jiang and Sehitoglu [1996b]). The Besseling model [1958] as used in ANSYS [1998] over-predicts the

ratcheting rates for most of the biaxial loading histories considered (Yuan. et al. [1999]). Since the

multilinear models also fail to simulate any uniaxial ratcheting, they are not considered further in this

study. When a slight nonlinearity is introduced into the multilinear Ohno and Wang [1993] model, it

shows promise in simulating both uniaxial and biaxial ratcheting responses. This model is presented later

in this paper.

II.3. Nonlinear Kinematic Hardening Model

The most well-known nonlinear kinematic hardening model has been proposed by Armstrong and

Frederick [1966]. They introduce a kinematic hardening rule containing a recall term which

incorporates the fading memory effect of the strain path and essentially makes the rule nonlinear in

nature. The kinematic hardening rule in this model is given in the form:

16
Paper I-International Journal of Plasticity 16 (2000) 381-409

2 p
da = --- Cd a dp , (5)
3
12
p 2 p p
where dp = d = --- d d .
3
For uniaxial loading, this rule basically provides an exponential x trace (see Fig. 4a), which always

starts with a modulus given by:

H = C
+ x (6)
and stabilizes to a value of C/ after traversing some amount of plastic strains. In Eq. (6), the negative

sign is used for forward loading curve and the positive sign for reverse loading curve. Figure 4a shows

this models simulation of the stable hyteresis loop. It is apparent from the figure that the experimental

stress-strain curve is not necessarily exponential in nature and the attempt to simulate it by a single

exponential equation does not yield a good fit. Increasing the value of C would improve the simulation

during the initial nonlinear part, but the simulation for the rest of the curve would suffer. Another

limitation of this model is its inability to produce constant plastic modulus exhibited by experiments for a

high strain range, for which this model always stabilizes to zero plastic modulus.

For a uniaxial stress cycle with mean stress, the recall term in the Armstrong-Frederick kinematic

hardening rule produces change in shapes between forward and reverse loading paths. Therefore, the loop

does not close and results in ratcheting (see Fig. 4b). But the stress-strain loop produced by this model

deviates significantly from the experiment and the ratcheting strain is also overpredicted, as demonstrated

in Fig. 4b. For continued cycles between two fixed stress levels, this model simulates the same ratcheting

loops for all cycles and thus, produces a constant ratcheting rate (strain accumulation per cycle) which is

evident from Fig. 4c.

If the experimental hysteresis loops in uniaxial ratcheting tests from Hassan and Kyriakides [1992]

are examined, it is observed that the loading curves gradually stiffen while the unloading curves soften

with cycles. This results in the gradual decrease in the rate of ratcheting up to a strain level and

17
Paper I-International Journal of Plasticity 16 (2000) 381-409

x Experiment x CS 1026
(ksi) 40 (ksi) 40 xa = 32.0 ksi
xm = 6.52 ksi
x
20 20

0 0

-20 -20

AF model
-40 -40 Experiment

-1 -0.5 0 0.5 1 -0.5 0 0.5 1

(a) x (%)
p
(b) x (%)
p

xp p
4 4
CS 1026 CS 1026
(%) xa = 32.0 ksi (%) m = 9.65 ksi
xm = 6.52 ksi xc = 0.50 %

3 3

xc

m
2 x 2 x
xa
xm
t
1 1
Experiment
AF model

0 0
0 10 20 30 40 50 0 10 20 30 40 50

(c) N (d)
(c) N

p p
4 4
CS 1026 CS 1018
(%) a = 2.36 ksi (%) a = 2.36 ksi
xc = 0.50 % xc = 0.50 %
m = 9.65 ksi m = 9.65 ksi
3 3

xc
a xc
2 2
a
m
m
x
1 1 x
Experiment
AF model

0 0
0 10 20 30 40 50 0 10 20 30 40 50

(e) N (f) N

Fig 4. Predictions from Armstrong and Frederick model (AF) for (a) strain-controlled stable
hysteresis loop, (b) stress-controlled hysteresis loop, (c) axial strain at positive stress peaks of
uniaxial cycles, (d) Circumferential strain peaks of axial strain cycles with constant pressure,
(e,f) Circumferential strain peaks under biaxial bow-tie and reverse bow-tie cycles.

18
Paper I-International Journal of Plasticity 16 (2000) 381-409

subsequent stabilization to a constant rate of ratcheting. The plastic modulus H expressed in Eq. (6), on

the other hand, is a function of x only. Hence, for stress cycles between two fixed limits, the shapes of

forward and reverse loading curves are repeatedly reproduced by this model. This, in effect, results in a

constant rate of ratcheting simulation for all cycles. It should be pointed out that by increasing the value

of the parameter C and selecting a corresponding , the rate of uniaxial ratcheting can be reduced, but this

impairs the simulation of the strain-controlled stable hysteresis loop. Moreover, this model always

grossly overpredicts the circumferential ratcheting strains for all the biaxial loading histories in Fig. 1

(see Figs. 4d, 4e and 4f).

Conceptually, the Armstrong and Frederick model has been a leap in representing cyclic plasticity

responses of materials, but is not robust enough to predict the ratcheting responses of materials. Several

improved models which are based on the Armstrong-Frederick kinematic hardening rule have

demonstrated promise in simulating ratcheting responses. Elaborate discussions on the modeling

structures and performances of three of these models are presented below.

II.4. Chaboche Model

II.4.1. Model Formulation and Ratcheting Responses: Chaboche and his coworkers [1979, 1986]

proposed a decomposed nonlinear kinematic hardening rule in the form:

M
2
d a i ,
p p
da = da i = --- C d i a i dp where dp = d (7)
3 i
i=1

As can be observed in Eq. (7), the Chaboche kinematic hardening rule is a superposition of several

Armstrong-Frederick hardening rules. Each of these decomposed rules has its specific purpose.

A stable hysteresis curve can be divided into three critical segments where the Armstrong-Frederick

model fails: the initial high modulus at the onset of yielding, the constant modulus segment at a higher

strain range and the transient nonlinear segment (knee of the hysteresis curve, see Fig. 4a). Chaboche

19
Paper I-International Journal of Plasticity 16 (2000) 381-409

[1986] initially proposed to use three decomposed hardening rules (M = 3 in Eq. 7) to improve the

simulation of the hysteresis loops in these three segments. They suggested that the first rule (1) should

start hardening with a very large modulus and stabilizes very quickly. The second rule (2) should

simulate the transient nonlinear portion of the stable hysteresis curve. Finally, the third rule (3) should

be a linear hardening rule (3 = 0) to represent the subsequent linear part of the hysteresis curve at a high

strain range. If this scheme is followed (see section II.4.1.1 for detailed parameter determination method),

the simulation for a stable hysteresis loop improves as shown in Fig. 5a (compare to Fig. 4a). Also note in

Fig. 5a, the traces of the three decomposed rules 1, 2 and 3 and the resulting yield surface center x

x Experiment x CS 1026
(ksi) 40 (ksi) 40 xa = 32.0 ksi
xm = 6.52 ksi
x
20 20
2
1
0 3 0

-20 -20

C-H3 model
-40 -40 Experiment

-1 Lp
p
-0.5 0 0.5 1 -0.5 0 0.5 1
L
x (%)
p
x (%)
p

(a) (b)
xp
4
CS 1026 x
xa = 32.0 ksi xa
(%) xm
xm = 6.52 ksi
t
3

1 Experiment
3 = 0
3 = 9

0
0 10 20 30 40 50

N
(c)
Fig 5. Predictions from Chaboche model with 3 decomposed rules (C-H3) for (a) strain-
controlled stable hysteresis loop, (b) stress-controlled hysteresis loop, (c) axial strain at positive
stress peaks of uniaxial cycles.

20
Paper I-International Journal of Plasticity 16 (2000) 381-409

(=1+2+3). The hysteresis curve simulation still deviates slightly from the experimental curve. This

can be improved further by adding more kinematic rules as is discussed later.

Figure 5b shows the simulation for a stress-controlled hysteresis loop by the Chaboche model using

three decomposed rules. The simulated loop is a significant improvement from that of the original

Armstrong-Frederick model and traces the experimental loop very closely. But this model still

overpredicts the ratcheting strain, though, by a relatively smaller amount at the end of the first cycle. The

ratcheting strain simulations (by three decomposed rules with 3 = 0) are plotted against the number of

cycles in Fig. 5c. It is observed in Fig. 5c that this model overpredicts the ratcheting strain for some initial

cycles, but gradually approaches to complete shakedown of ratcheting, which is in contrast to the

experimental trend. This shakedown is mainly caused by the incorporation of the linear kinematic

hardening rule (third component, 3) along with other nonlinear ones.

To understand this shakedown phenomenon, consider two points a and b at the same stress level

as shown in Fig. 6, where the total backstress x and the decomposed backstresses 1, 2 and 3 are

plotted against the plastic strain (note that a becomes x in uniaxial case). For 3 =0 (or linear third

rule), the backstress increments dx, for the same plastic strain increment dxp at these two points are:

dxa = d1a+d2a+d3a = (2/3 C1 dxp - 11a dp) + (2/3 C2 dxp - 2 2a dp) + (2/3 C3 dxp ) (8)

dxb = d1b+d2b+d3b = (2/3 C1 dxp - 11b dp) + (2/3 C2 dxp - 2 2b dp) + (2/3 C3 dxp ) (9)

Since a and b represent the same stress level, xa= xb for a cyclically stable material, thus

1a+2a+3a = 1b+2b+3b (10)

Since 1a = 1b (as evident from Fig. 6b), by subtracting Eq. (9) from Eq. (8) one obtains,

dxa - dxb= d2a d2b = 2 (2b2a ) dp (11)

21
Paper I-International Journal of Plasticity 16 (2000) 381-409


30 30
x x
(ksi) (ksi)
a b
20 20

10 10

0 0

-10 -10
0 0.5 1 1.5 2 0 0.5 1 1.5 2 2.5 3

(a) xp (%) (a) xp (%)


30 30
1
(ksi)
x 1
(ksi) x
CS 1026 CS 1026
xa xa = 32.0 ksi xa xa = 32.0 ksi
20 xm xm = 6.52 ksi 20 xm xm = 6.52 ksi
t t
10 10

a b
0 0

-10 -10
0 0.5 1 1.5 2 0 0.5 1 1.5 2 2.5 3

(b) xp
(%) (b) xp (%)


30 30
2 2
(ksi) (ksi)

20 20
a b

10 10

0 0

-10 -10
0 0.5 1 1.5 2 0 0.5 1 1.5 2 2.5 3

(c) xp (%) (c) xp (%)


30 30
3 3
(ksi) 3 = 0 (ksi) 3 = 9

20 20

10 10

a b
0 0

-10 -10
0 0.5 1 1.5 2 0 0.5 1 1.5 2 2.5 3

(d) xp
(%) (d) xp (%)

Fig 6. Uniaxial backstresses by Chaboche Fig 7. Uniaxial backstresses by Chaboche model with
model with three decomposed hardening three decomposed hardening and 3=9; (a) total
and 3=0; (a) total backstress, x, (b) 1, backstress, x, (b) 1, (c) 2, (d) 3 against axial
(c) 2, (d) 3 against axial plastic strain. plastic strain.

It is clear from Eq. (11) that the difference in incremental hardening between points a and b

comes from the second rule. As 3 increases linearly from point a to b making 3b > 3a (Fig. 6d) and

22
Paper I-International Journal of Plasticity 16 (2000) 381-409

as 1a = 1b, so 2 has to decrease linearly at a rate defined by C3 between points a and b in order to

hold the equality of Eq. (10), which is also validated in Fig. 6c. Since 2a > 2b, Eq. (11) dictates that

dxb> dxa. This essentially results in the gradual stiffening of the loading curves with cycles. Similarly,

the unloading curves relax with increasing cycles, resulting in progressive decrease of ratcheting strain

with cycles. The point at which both the loading and unloading curves assume the same shape, the

ratcheting ceases completely (shakedown) as observed in Fig. 5c (for 3 = 0) and Fig. 6a.

If a slight nonlinearity is introduced in the third rule by assigning a relatively small value to 3 (= 9),

keeping other parameters the same, the ratcheting simulation improves as shown in Figs. 5c and 7a. This

small value does not introduce any noticeable change in the strain-controlled stable hysteresis loop

simulation. Note also that a non-zero 3 does not have any effect on 1 (Fig. 7b), but it changes the course

of 3 (Fig. 7d) and thereby of 2 (Fig. 7c), which improve the uniaxial ratcheting simulation and prevent

shakedown (Fig. 7a). When the third backstress (3) reaches its limiting value (near xp = 1.5% in Fig. 7d),

this model starts predicting a constant ratcheting rate which is also the trend in experimental responses.

The higher the value of 3, the quicker the third rule would reach its limiting state and, consequently, the

earlier the steady rate of ratcheting would start. The inception point of constant ratcheting rate also

depends on the stress limits of a stress cycle. One may consider 3 as a ratcheting parameter whose value

can be calibrated by matching a uniaxial ratcheting rate data.

II.4.1.1 Parameter Determination (Three Decomposed Rules): In this study, all parameters, except

3, are determined from a uniaxial strain-controlled stable hysteresis curve, not a monotonic curve as sug-

p
gested in the original model. This requires a hysteresis loop of reasonable strain limit ( L = 0.85% in

Fig. 5a) which ensures that all, except the third slightly nonlinear kinematic hardening variable get stabi-

lized within the strain limit. Also, a stabilized decomposed backstress (1 or 2) should have the same

23
Paper I-International Journal of Plasticity 16 (2000) 381-409

tensile and compressive levels within the strain range of the stable loop. In other words, for the loading

portion of the hardening curves, they should start from -Ci /i at the starting plastic strain Lp and reach

the value Ci/i at or prior to the final plastic strain, Lp as shown in Fig. 5a. In addition, the third linear

backstress 3 should pass through the origin (Fig. 5a).

The equations used for the loading part of the hardening curves are:

i + 0 = x (12)
i=1
Ci p p
i = ----- 1 2 exp i ( x ( L ) ) , for i = 1 and 2 (13)
i

p
where L is the strain limit of the stable hysteresis loop (Fig. 5a).

C1 should be a very large value to match the plastic modulus at the yielding and corresponding 1

also should be large enough to stabilize the hardening of 1 immediately. C3 is determined from the slope

of the linear segment of hysteresis curve at a high strain range. C2 and 2 are evaluated by trials to pro-

duce a good representation of the experimental stable hysteresis curve which also satisfy the relationship

C1 C2 C3 p p
------- + ------- + 0 = x ------- { x ( L ) } (14)
1 2 2

at or close to plastic strain Lp (fig. 5a).

Finally, 3 is determined from a uniaxial ratcheting experiment (xp vs. N plot) to produce the best

possible fit. It should be noted here that choosing a value of 3 relatively smaller in magnitude than 1 or

2 would produce no tangible effect on the stable hysteresis loop simulation, but would produce a signifi-

cant effect on the uniaxial ratcheting rate. The parameters determined following this method and used for

simulation by the Chaboche model (three decomposed rule) are:

24
Paper I-International Journal of Plasticity 16 (2000) 381-409

0=18.8 ksi, E=26300 ksi, =0.302 (common to all models)

C1-3 = 60000, 12856, 455 (ksi); 1-3 = 20000, 800, 9.

II.4.1.2. Ratcheting Simulations (Three Decomposed Rule): The above set of parameters is used

to simulate all uniaxial and biaxial ratcheting responses of stabilized carbon steels considered in this

study. Figure 8 shows the predictions by the Chaboche model (three decomposed rules) for all the

experiments. In these figures, the maximum peak strain in each cycle is plotted as a function of the

number of cycles, N. As observed in Figs. 8a and 8b, the Chaboche model has a tendency to over-predict

uniaxial ratcheting rates during initial cycles. As a result, the overall simulation for a low rate ratcheting

experiment (e.g. xa= 28.29 ksi in Fig. 8b) deviates from the experiment. Otherwise, the ratcheting rate

predictions for other uniaxial cases are quite reasonable (Figs. 8a,b). However, this model overpredicts

the ratcheting for all the biaxial loading cases (Figs. 8c,d,e,f).

One might argue here that the incorporation of more than three decomposed rules might improve

the ratcheting simulations. In order to verify this argument, simulations are obtained for all the

experiments by adding a fourth Armstrong-Frederick type decomposed hardening rule. It is observed that

the incorporation of a fourth rule improves the stable hysteresis loop simulation, especially at the knee

(not shown). But the ratcheting simulations for both uniaxial and biaxial experiments do not improve.

Hence, these results are not presented in this paper.

II.4.2. Chaboche Model - Fourth Rule with a Threshold: Realizing the above deficiencies of the

model, Chaboche [1991] added a fourth hardening rule with a concept of threshold in his model. This

kinematic hardening rule grows linearly to a certain threshold stress level and subsequently hardens

according to the Armstrong-Frederick rule, as follows:

ai
da i = --- C i d i a i 1 ------------- dp ,
2 p
for i = 4 (15)
3 f ( i )

25
Paper I-International Journal of Plasticity 16 (2000) 381-409

xp xp
4 4
CS 1026 CS 1026
xa = 32.0 ksi xm = 9.14 ksi xm = 6.5 ksi
(%) (%) a = 33.28 ksi
x
xa
3 3 xm xa= 32.12 ksi
t

xm = 6.52 ksi xa = 30.35 ksi


2 2

xm = 4.18 ksi
1 1 xa = 28.29 ksi
Experiment
C-H3 model

0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(a) (b)
p p
4 4
CS 1026 CS 1026
(2) (3)
(%) m = 9.65 ksi (%) xc = 0.50 %
(3) (4)

3 (1) xc = 0.40 % 3 (1) m = 4.88 ksi


(1) (2) xc = 0.50 % (2) (2) m = 7.32 ksi
(3) xc = 0.65 % (3) m = 9.65 ksi
(4) m =14.54 ksi

xc (1)
2 2
(2) m (4) (3)
(3) x
(2)
1 (1) 1
Experiment (1)
C-H3 model

0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(c) (d)
p p
4 4
(2) CS 1026 (2) CS 1018
(%) a = 2.36 ksi (%) a = 2.36 ksi
(1)
xc = 0.50 % xc = 0.50 %
3 3
(1)
(1) m = 9.65 ksi (1) m = 9.65 ksi
(2) m = 14.54 ksi (2) m = 14.54 ksi

xc
2 (2) 2 (2)
a xc
(1)
m a
(1) m
1 x 1
Experiment
x
C-H3 model

0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(e) (f)

Fig. 8. Ratcheting experiments (From Hassan and Kyriakides [1992], Hassan et al. [1992] and
Corona et al. [1996]) and predictions by Chaboche model with three decomposed rule (C-H3).
(a,b) axial strain at positive stress peaks of uniaxial cycles, (c,d) Circumferential strain peaks of
axial strain cycles with constant pressure, (e,f) Circumferential strain peaks under biaxial bow-tie
and reverse bow-tie cycles.

26
Paper I-International Journal of Plasticity 16 (2000) 381-409

This fourth rule simulates a constant linear hardening with a slope of C4 within the threshold a 4

and becomes nonlinear outside the threshold (see trace of 4 in Fig. 9a). When this fourth rule is added to

the previous three decomposed rules and the parameters are determined using the methodology described

in section II.4.2.1., the model predicts a reduced rate of ratcheting for initial cycles without jeopardizing

the simulation of the stable hysteresis loop. The improvement becomes evident if simulations in Figs. 5a

and 5b are compared to Figs. 9a and 9b, respectively. The simulation in Fig. 9b conforms better with the

experimental uniaxial ratcheting hysteresis loop compared to that in Fig. 5b. The reason for improvement

in uniaxial ratcheting simulation by the fourth rule with a threshold is the fact that within the threshold

level, this rule does not use its recall term and assumes linear hardening in a similar fashion as shown in

x Experiment x CS 1026
(ksi) 40 (ksi) 40 xa = 32.0 ksi
xm = 6.52 ksi
x
20 20
4 1 3
2

0 0

-20 -20

C-H4T model
-40 -40 Experiment

-1 Lp
p
-0.5 0 0.5 1 -0.5 0 0.5 1
L

(a) x (%)
p
(b) x (%)
p

xp
4
CS 1026 x
(%) xa = 32.0 ksi xa
xm = 6.52 ksi xm

3 t

1 Experiment
C-H4T model

0
0 10 20 30 40 50

(c) N

Fig 9. Predictions from Chaboche model with threshold (C-H4T) for (a) strain-controlled
stable hysteresis loop, (b) stress-controlled hysteresis loop, (c) axial strain at positive stress
peaks of uniaxial cycles.

27
Paper I-International Journal of Plasticity 16 (2000) 381-409

Fig. 6d. As stated before, a linear hardening is instrumental in stiffening the loading curves and relaxing

the unloading curves to reduce overall ratcheting with cycles. When the threshold level is reached, the

hardening becomes nonlinear again and the reduction of ratcheting is attenuated to avoid potential

shakedown. The threshold level, a 4 can also be considered a ratcheting parameter which is to be

determined from uniaxial ratcheting experiments.

II.4.2.1. Parameter Determination (Chaboche Model with Threshold): The parameter C1,1 and

C3 are determined using the same method as discussed in section II.4.1.1 using Eqs. (12) and (13). C2,

C4, 2 and 4 are evaluated by trials to produce a good representation of the experimental stable hysteresis

curve which also satisfy the relationship

C1 C2 C4 C p p
------- + ------- + ------- + a 4 + 0 = x ------3- { x ( L ) } (16)
1 2 4 2

at or close to the plastic strain Lp (Fig. 9a). The first trial value of a 4 can be taken close to the value of the

mean-stress in the uniaxial ratcheting experiment used for the parameter determination (Chaboche

[1991]). The 3 value is determined using the ratcheting rate from a uniaxial ratcheting experiment (xp

vs. N) as discussed in section II.4.1.1. Finally, the value of a 4 is varied a little to improve the simulation

for the uniaxial ratcheting experiment (xp vs. N) further. Care should be taken so that the value of a 4

selected does not deteriorate the simulation for the strain-controlled stable hysteresis loop. All parameters

determined accordingly and used in this study for simulation by the Chaboche model with threshold

are:

0=18.8 ksi, E=26300 ksi, =0.302 (common to all the models)

C1-4 = 60000, 3228, 455, 15000 (ksi); 1-4 = 20000, 400, 11, 5000; a4 = 5 ksi

28
Paper I-International Journal of Plasticity 16 (2000) 381-409

xp xp
4 4
CS 1026 CS 1026 xa = 33.28 ksi
xa = 32.0 ksi xm = 9.14 ksi xm = 6.5 ksi
(%) (%)
x
3 3 xa xa= 32.12 ksi
xm
t

xm = 6.52 ksi
2 2
xa = 30.35 ksi
xm = 4.18 ksi
1 1 xa = 28.29 ksi
Experiment
C-H4T model

0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(a) (b)

4 4
CS 1026 CS 1026
p (2)
m = 9.65 ksi p (3)
xc = 0.50 %
(%) (%)
(3) (4)

3 (1) xc = 0.40 % 3 (1) m = 4.88 ksi


(1) (2) xc = 0.50 % (2) (2) m = 7.32 ksi
(3) xc = 0.65 % (3) m = 9.65 ksi
(4) m =14.54 ksi

2 xc 2 (1)

(2) m (4) (3)


(3)
x (2)
1 (1) 1
Experiment (1)
C-H4T model

0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(c) (d)
p p
4 4
(2) CS 1026 (2) CS 1018
(%) a = 2.36 ksi (%) a = 2.36 ksi
(1)
xc = 0.50 % xc = 0.50 %
3 3
(1)
(1) m = 9.65 ksi (1) m = 9.65 ksi
(2) m = 14.54 ksi (2) m = 14.54 ksi

2 xc 2 (2)
(2) xc
a
(1)
a
m
(1) m
1 x 1
Experiment x
C-H4T model

0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(e) (f)

Fig. 10. Ratcheting experiments (From Hassan and Kyriakides [1992], Hassan et al. [1992] and
Corona et al. [1996]) and predictions by Chaboche model with threshold (C-H4T). (a,b) axial
strain at positive stress peaks of uniaxial cycles, (c,d) Circumferential strain peaks of axial
strain cycles with constant pressure, (e,f) Circumferential strain peaks under biaxial bow-tie
and reverse bow-tie cycles.

29
Paper I-International Journal of Plasticity 16 (2000) 381-409

II.4.2.2. Ratcheting Simulations (Chaboche model with Threshold): Figure 10 shows the

ratcheting simulations by the Chaboche model with threshold along with the experimental responses.

The simulations by this model for most of the uniaxial loading cases are similar to those by the Chaboche

model with three decomposed rules (compare Figs. 10a,b and 8a,b). The effect of the forth rule with a

threshold becomes pronounced for the experiments with lower rates of ratcheting (see Figs. 8b and 10b

for xa= 28.29 ksi). But overpredictions for all biaxial loading cases persist (shown in Figs. 10b,c,d,e)

even after introducing the fourth rule with a threshold level.

II.5. Ohno and Wang Model

II.5.1. Model Formulation and Ratcheting Responses: The Ohno-Wang model [1993] is also a

superposition of several kinematic hardening rules. They initially proposed the form:

M ai
da i = --- C i d i ai d -------------
H { a2 ( Ci i )2 }
2
d a i ,
p p
da = (17)
3 f ( i ) i
i=1

where, H stands for the Heaviside step function.

Here, each decomposed hardening rule simulates a linear hardening with a slope Ci until it reaches

the critical value Ci/i. After that it does not evolve at all, as shown qualitatively by solid lines in Fig. 11.

Consequently, the model becomes a multilinear model in a uniaxial case. It is demonstrated in section II.2

that the multilinear models produce no uniaxial ratcheting due to the prediction of closed hysteresis

loops. To avoid this limitation, Ohno and Wang [1993] introduces a slight nonlinearity for each rule at the

transition from linear hardening to the stabilized critical state as follows:

mi
M a i f ( i )
d a i , da i = --3- C i d i a i d -------------
-------------
2 p p
-
da = (18)
f ( i ) C i i
i=1

The slight nonlinearity is introduced by replacing the Heaviside step function with the multiplier

with power of mi in Eq. (18). The slight nonlinearities produced are shown qualitatively by dashed lines

30
Paper I-International Journal of Plasticity 16 (2000) 381-409

x,x
(4p,x4)
(3p,x3) x
(2p,x2)

2
(1p,x1)
3
L p 1
4
Lp Sp
(0p,x0) xp

Fig 11. Qualitative demonstration of decomposed hardening for a hysteresis curve in


Ohno-Wang model.

in Fig. 11. These nonlinearities prevent the stress-controlled hysteresis loops from closing and allow

uniaxial ratcheting to occur (see Fig. 5 in Ohno and Wang [1993] for details).

As several essentially linear hardening rules are employed to simulate a nonlinear hysteresis curve,

this model should use a large number of decomposed rules in order to produce a good representation of

the stable uniaxial hysteresis curve. In this study, ten hardening rules are found to be sufficient to obtain a

good stable loop simulation as demonstrated in Fig. 12a. In addition, this model uses the term

a
d p -------------
i in place of dp in the Armstrong-Frederick rule (Eq. 5). Although both these terms in a given
f ( )

i

model yield the same result in uniaxial cases, they produce different directions of kinematic hardening in

multiaxial loading cases. Inclusion of the former term improves the ratcheting simulation in multiaxial

cases as demonstrated by Ohno and Wang [1993] (Figs. 6 and 8 in the reference).

The Ohno-Wang model simulates the stress-controlled ratcheting hysteresis loop more closely than

the Chaboche model (compare Fig. 9b and 12b). It is evident from Fig. 12c that the overall ratcheting

simulation by the Ohno-Wang model is very good. The effect of nonlinearity induced by the parameter mi

31
Paper I-International Journal of Plasticity 16 (2000) 381-409

x Experiment x CS 1026
(ksi) 40 (ksi) 40 xa = 32.0 ksi
xm = 6.52 ksi
x
20 20

0 0

-20 -20

OW model
-40 -40 Experiment

-1 Lp
p
-0.5 0 0.5 1 -0.5 0 0.5 1
L
x (%)
p
x (%)
p

(a) (b)
xp
4
CS 1026 x
(%) xa = 32.0 ksi xa
xm = 6.52 ksi xm
t
3

1 Experiment
m i = 0.20
m i = 0.45
m i = 0.70
0
0 10 20 30 40 50

(c) N

Fig 12. Predictions from Ohno and Wang (OW) model (a) strain-controlled stable hysteresis
loop, (b) stress-controlled hysteresis loop, (c) axial strain at positive stress peaks of uniaxial
cycles.

on the ratcheting simulations is also shown in Fig. 12c. Note that the smaller the value of mi the more

nonlinear the decomposed rules are and, the higher the rate of ratcheting.

II.5.2. Parameter Determination: Parameters for this model are also determined from a uniaxial

stable hysteresis curve and a uniaxial ratcheting experiment. The stable hysteresis loading curve are

divided into several segments as shown in Fig. 11 and the corresponding parameters Ci and i for each

segment can be determined from the following equations:

32
Paper I-International Journal of Plasticity 16 (2000) 381-409

i i1 i+1 i
x x x x 2
C i = ---------------------- ---------------------- for i 1 ,
p p p p
i = ----------------
p
-;
p
(19a)
i i 1 i + 1 i i + 0

C
----- ii + 0
0
and finally C1 is determined using = x (19b)

where ix and pis are as indicated in Fig. 11. The power mi is assumed to be same for all segments and

should be determined from a uniaxial ratcheting experiment.

In the Chaboche model, the hardening rule (3) which is responsible for simulating the linear

segment of the hysteresis curve at higher strain ranges has a significant effect on the simulation of the rate

of uniaxial ratcheting. Figure 5c shows that an increase in the value of 3 increases the uniaxial ratcheting

rate and hastens the inception of the steady rate of ratcheting. In the Ohno model, the last hardening rule

(e.g. 4 in Fig. 11) also has a similar effect as discussed below.

If the strain-controlled stable hysteresis loop in Fig. 12a is represented by ten segments, the value of

10 is determined by taking 10p = Lp in Eq. (19). This essentially means that the backstress 10 reaches

its plateau at Lp and, 10 calculated from Eq. (19) becomes relatively large. Consequently, the Ohno-

Wang model overpredicts ratcheting in some cases as shown in Fig. 13a. This over-prediction is more

prominent for higher stress levels (xm= 9.14 ksi).

To overcome this problem, it is stipulated that the last linear segment (10th in this study) of the

stable hysteresis curve should be extended up to a high strain limit sp as shown by the long-dashed lines

in Fig. 11. This increase in the strain limit from Lp to sp reduces the value of 10 as determined by Eq.

(19). Figures 13b and 13c show the simulations for the same set of experiments when values of sp used

are 2.5% and 5%, respectively. Uniaxial ratcheting simulations are reasonable with sp= 2.5% (Fig. 13b),

33
Paper I-International Journal of Plasticity 16 (2000) 381-409

xp xp
4 4
CS 1026 ps = 0.85 % CS 1026 ps = 2.5 %
(%) xa = 32.0 ksi 10 segments (%) xa = 32.0 ksi 10 segments

3 3
xm = 6.52 ksi xm = 6.52 ksi
xm = 9.14 ksi
xm = 9.14 ksi
2 2

xm = 4.18 ksi
1 1
xm = 4.18 ksi
Experiment Experiment
OW model OW model
0 0
0 10 20 30 40 50 0 10 20 30 40 50

(a) N (b) N

xp xp
4 4
CS 1026 ps = 5.0 % CS 1026 ps = 5.0 %
(%) xa = 32.0 ksi 10 segments (%) xa = 32.0 ksi 12 segments

3 3
xm = 9.14 ksi xm = 6.52 ksi
xm = 9.14 ksi
xm = 6.52 ksi
2 2

xm = 4.18 ksi xm = 4.18 ksi


1 1

Experiment Experiment
OW model OW model
0 0
0 10 20 30 40 50 0 10 20 30 40 50

(c) N (d) N

Fig 13. Axial (peak) strain ratcheting simulations by Ohno and Wang (OW) model for
uniaxial loading with (a) sp=0.85%, 10 segments; (b) sp=2.5%, 10 segments; (c) sp=5%, 10
segments and (d) sp=5%, 12 segments.

but overprediction still persists for the higher stress level (xm= 9.14 ksi) as the simulated rate reaches a

steady state quickly. If sp is increased to 5% for the tenth segment, 10 decreases further. Consequently,

ratcheting simulation show prolonged nonlinearity and reduced strain (Fig. 13c).

In reality, the strain-controlled stable curve is not expected to maintain the same slope all the way

from Lp (0.85%) to sp = 5%. It is expected to show slight nonlinearity or a decrease in modulus with

increasing strain. This behavior can be modeled by adding a few more segments beyond Lp (0.85%),

34
Paper I-International Journal of Plasticity 16 (2000) 381-409

with gradually smaller slopes than previous segments. When two more segments are added to the

hysteresis curve beyond Lp (0.85%) and extended up to sp=5%, the simulations improve greatly as

shown in Fig. 13d. Simulations are now tracing the nonlinear ratcheting rate trend of the experiments

more closely.

Hence, the best and recommended way to determine the parameters (Ci and is) is from a hysteresis

curve with reasonably larger strain range. If such a hysteresis curve is not available, one should follow the

method discussed above and use more than one uniaxial ratcheting experiment to validate the parameters.

The parameters used in this study for simulations by the Ohno-Wang model are (with sp=5% and 12

segments):

0=18.8 ksi, E=26300 ksi, =0.302 (common to all the models)

C1-12 = 31940, 36214, 2520, 376, 11021, 4551, 3475, 2196, 857, 247, 98, 200 (ksi);

1-12 = 45203, 13944, 7728, 4955, 3692, 2135, 1230, 585, 295, 119, 50, 20;
mi = 0.45

II.5.3. Ratcheting Simulations: Figure 14 shows the simulations of the Ohno-Wang model using

the above set of parameters for all the experiments. As the parameters in this model are determined using

uniaxial experiments, the uniaxial ratcheting predictions by this model are quite good and comparable to

those obtained by the Chaboche model (compare Figs. 10a,b and 14a,b). Although this model performs

better in all biaxial cases compared to the Chaboche model, the trend of overprediction persists (compare

a
Figs. 10c,d,e,f and 14c,d,e,f). Note, however, that the incorporation of d p -------------
i in Eq. (18) yields
f ( i )

nonlinear biaxial ratcheting rates and the trend is similar to the experimental results.

35
Paper I-International Journal of Plasticity 16 (2000) 381-409

xp xp
4 4
CS 1026 CS 1026
(%) xa = 32.0 ksi (%) xm = 6.5 ksi
x xa = 33.28 ksi
3 3 xa
xm
xm = 9.14 ksi xm = 6.52 ksi xa= 32.12 ksi
t

2 2
xa = 30.35 ksi
xm = 4.18 ksi
1 1
xa = 28.29 ksi
Experiment
OW model
0 0
0 10 20 30 40 50 0 10 20 30 40 50

(a) N (b) N

p p
4 4
CS 1026 CS 1026
(%) m = 9.65 ksi (%) xc = 0.50 %

3 (1) xc = 0.40 % 3 (4) (1) m = 4.88 ksi


(2) xc = 0.50 % (2) m = 7.32 ksi
(3) (2) (3) xc = 0.65 % (3) (3) m = 9.65 ksi
(4) m =14.54 ksi
(2)
2 (1) xc 2
(3) (4) (3)
(2) m (1)
x (2)
1 (1) 1
Experiment (1)
OW model

0 0
0 10 20 30 40 50 0 10 20 30 40 50

(c) N (d) N

p p
4 4
CS 1026 CS 1018
(%) a = 2.36 ksi (%) a = 2.36 ksi
(2) xc = 0.50 % xc = 0.50 %
3 3
(2)
(1) (1) m = 9.65 ksi (1) m = 9.65 ksi
(2) m = 14.54 ksi (2) m = 14.54 ksi

2 xc 2 (1)
(2) (2)
a xc
(1)
a
m
1 1 (1) m
x
Experiment
x
OW model

0 0
0 10 20 30 40 50 0 10 20 30 40 50

(e) N (f) N

Fig. 14. Ratcheting experiments (From Hassan and Kyriakides [1992], Hassan et al. [1992]
and Corona et al. [1996]) and predictions by Ohno and Wang (OW) model. (a,b) axial strain at
positive stress peaks of uniaxial cycles, (c,d) Circumferential strain peaks of axial strain cycles
with constant pressure, (e,f) Circumferential strain peaks under biaxial bow-tie and reverse
bow-tie cycles.

36
Paper I-International Journal of Plasticity 16 (2000) 381-409

II.6. Guionnet model

The Chaboche and Ohno-Wang models, discussed above, include parameters which are determined

from uniaxial experiments only. These models have been shown to predict the uniaxial ratcheting

responses well, but the predictions are not as good for biaxial ratcheting responses. Guionnet [1992]

proposes a model which uses some parameters that are determined from biaxial ratcheting experiments.

This model is expected to simulate the biaxial ratcheting responses better and therefore, is studied in this

paper.

The Guionnet model basically modifies the original Armstrong-Frederick hardening rule by

incorporating the effect of accumulated plastic strain in it. For cyclically stabilized material, the

kinematic hardening rule in this model is reduced to the form:

p
, n = 2--- --------
-
2 p d
m1 (19)
da = m p1 --3- C 1 ( a n ) d ( 2 )a dp 3 dp

n n
where = p 1 = p 1M , for p 1 = p 1M ,

n p 1M
= p 1M ----------------------- , for p 1 p 1M ,
p 1M + p 1

Q IK
p1 = dp , and p 1M = dp
IK IK 1

Here, p1 is the accumulated plastic strain between the last reversal (Ik) and the current loading point

(Q), and p1M is the accumulated plastic strain between the last two reversals (Ik-1 and Ik). The parameters

C and 1 are similar to those in Armstrong-Frederick model and are determined from a uniaxial stable

hysteresis curve. Two ratcheting parameters, 2 and are determined using a biaxial ratcheting response

(experiment (2) in Fig. 16c). No clear guidelines are provided by Guionnet [1992] to determine m and n.

These parameters do not affect the ratcheting simulations greatly. Therefore, the values of m and n as

used in Guionnet [1992] are also used in this study.

37
Paper I-International Journal of Plasticity 16 (2000) 381-409

x
Experiment x
CS 1026
(ksi) 40 (ksi) 40 xa = 32.0 ksi
xm = 6.52 ksi
x
20 20

0 0

-20 -20

Guionnet model
-40 -40 Experiment

-1 -0.5 0 0.5 1 -0.5 0 0.5 1

x (%)
p
x (%)
p

(a) (b)
Fig 15. Predictions from Guionnet model (a) strain-controlled stable hysteresis loop, (b) stress-
controlled hysteresis loop

The stable, uniaxial hysteresis loop simulation by this model is similar to that from Armstrong-
Frederick model (compare Fig. 15a to Fig. 4a). This model predicts the uniaxial stress-controlled
hysteresis loop poorly and also, overpredicts the amount of ratcheting strain in the first cycle as shown in
Fig. 15b. Figure 16 shows the simulations by the Guionnet model for all the experiments considered in
this study using the following parameters:

0=18.8 ksi, E=26300 ksi, =0.302 (common to all the models)

C = 3450 ksi, 1=150, 2=20, m=0.8, n=0.075, =0.4

The ratcheting strains for the uniaxial experiments are overpredicted by this model (Fig. 16a,b).

This model also has numerical divergence problems for higher uniaxial ratcheting rate responses.

However, its biaxial ratcheting predictions for the loading history shown in Fig. 1b are very good (Figs.

16c,d) despite the constant rate of ratcheting prediction. For the bow-tie loading case (Fig. 16e), this

model overpredicts with constant rate of ratcheting, whereas for the reverse bow-tie (Fig. 16f), it

underpredicts the ratcheting. Similar erratic behavior is also observed in the simulations for bow-tie and

reverse bow-tie loading cycles obtained by the Dafalias-Popov uncoupled model with Armstrong-

Frederick kinematic hardening rule (Corona et al. [1996]).

38
Paper I-International Journal of Plasticity 16 (2000) 381-409

xp xp
4 4
CS 1026 CS 1026
(%) xa = 32.0 ksi (%) xm = 6.5 ksi
x
3 xa 3
xm xm = 6.52 ksi xa= 32.12 ksi
t

2 2
xa = 30.35 ksi
xm = 4.18 ksi

1 1
Experiment xa = 28.29 ksi
Guionnet model

0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(a) (b)
p p
4 4
CS 1026 CS 1026
m = 9.65 ksi xc xc = 0.50 %
(%) (%)
m
3 (1) xc = 0.40 % 3 (1) m = 4.88 ksi
(2) xc = 0.50 % x (2) m = 7.32 ksi
(3) xc = 0.65 % (3) m = 9.65 ksi
(4) m =14.54 ksi
2 2
(3)
(2) (4) (3)
(2)

1 (1) 1
Experiment (1)
Guionnet model

0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(c) (d)
p p
4 4
CS 1026 CS 1018
(%) a = 2.36 ksi (%) a = 2.36 ksi
xc = 0.50 % xc = 0.50 %
3 3

(2) (1) m = 9.65 ksi (1) m = 9.65 ksi


(1) (2) m = 14.54 ksi (2) m = 14.54 ksi

(2) xc
2 2 (2)
a xc
(1)
m a
(1) m
1 x 1
Experiment (2)
x
Guionnet model
(1)

0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(e) (f)

Fig. 16. Ratcheting experiments (From Hassan and Kyriakides [1992], Hassan et al. [1992]
and Corona et al. [1996]) and predictions by Guionnet model. (a,b) axial strain at positive
stress peaks of uniaxial cycles, (c,d) Circumferential strain peaks of axial strain cycles with
constant pressure, (e,f) Circumferential strain peaks under biaxial bow-tie and reverse bow-tie
cycles.

39
Paper I-International Journal of Plasticity 16 (2000) 381-409

III. DISCUSSION AND CONCLUSIONS

Five well-known constitutive models are evaluated in this paper in terms of simulations for

ratcheting responses from a series of uniaxial and biaxial experiments. Six uniaxial and ten biaxial

ratcheting responses on stabilized carbon steels are collected from literature (Hassan and Kyriakides

[1992], Hassan et al. [1992] and Corona et al. 1996]) for the study. The rate-independent and cyclically

stable ratcheting response simulations by Prager [1956]), Mroz [1967], Armstrong and Frederick [1966],

Chaboche [1986, 1991], Guionnet [1992], Ohno and Wang [1993] are evaluated.

All these models are categorized as coupled models due to their plastic modulus calculation being

coupled with the kinematic hardening rule of the models. Sequential development of different modeling

features, such as, linear kinematic hardening (Prager [1956]), multilinear model (Mroz [1967]), nonlinear

kinematic hardening (Armstrong and Frederick [1966], Guionnet [1992]), decomposed nonlinear

kinematic hardening (Chaboche and his coworkers [1979,1986]), and finally decomposed nonlinear

kinematic hardening with threshold (Chaboche [1991], Ohno and Wang [1993]) are presented. The

bearings of each of these features on the simulation of ratcheting responses are elaborated.

For uniaxial stress-controlled history (Fig. 1a), the linear kinematic hardening and multilinear

models produce closed hysteresis loops and hence, cannot simulate a ratcheting response. On the other

hand, the nonlinear kinematic hardening models (Armstrong and Frederick [1966], Guionnet [1992])

grossly overpredict the uniaxial ratcheting strains. This overprediction problem is remedied by Chaboche

[1986] through decomposing the back stress into a number Armstrong and Frederick type nonlinear

kinematic hardening rules. The original Chaboche [1986] model with three or four decomposed rules,

with at least one rule as linear hardening, improves the ratcheting simulation for the initial cycles, but

always stabilizes to shakedown with persistent cycling. Incorporation of a slight nonlinearity into the

linear rule, through a small 3 value instead of 3 = 0, improves the models capability of simulating

40
Paper I-International Journal of Plasticity 16 (2000) 381-409

steady rate of ratcheting and prevents shakedown. This model simulates all the uniaxial ratcheting

experiments, except the one with small rate of ratcheting, reasonably well (Figs. 8a,b). In this case, the

ratcheting strains are overpredicted due to overpredictions during the initial cycles. Subsequently,

Chaboche [1991] introduces the concept of threshold on backstress into the fourth nonlinear kinematic

hardening rule. This model incorporates a linear segment in the uniaxial hysteresis curve within the

threshold level (see Figs. 9a and 9b), thus makes the hysteresis curve stiffer and reduces the rate of

ratcheting compared to the models without any threshold. This visible linear segment in the hysteresis

loop seems unnatural, but works well in improving the uniaxial ratcheting simulation.

The Ohno-Wang model [Ohno and Wang, 1993] is also composed of a number of the Armstrong

and Frederick type kinematic hardening rules. This model reproduces uniaxial hysteresis loops by using a

number of almost linear segments and thus needs a large number of kinematic hardening rules to

simulate the hysteresis loop. The hysteresis loop simulations however are very good as shown in Figs.

12a and 12b. The simulations of all the uniaxial ratcheting experiments by the Ohno-Wang model are also

very good.

None of the coupled models studied performed satisfactorily in simulating ratcheting responses of

biaxial loading histories in Figs. 1b, 1c and 1d. The linear kinematic hardening model [Prager, 1956]

overpredicts ratcheting strains during the initial cycles which are followed by shakedown after a few

more cycles for all biaxial cases. The nonlinear kinematic hardening rule [Armstrong and Frederick,

1966] grossly overpredicts the strains in all biaxial ratcheting experiments. Multisurface Mroz model (not

studied here) also overpredicts the ratcheting responses under multiaxial stresses (Mroz and Rodzik

[1996], Jiang and Sehitoglu [1996b]). The multilinear Besseling [1958] model is found to overpredict the

ratcheting strains in most biaxial cases (Yuan, et al., [1999]). The Chaboche [1991] and the Ohno-Wang

[1993] models, which use decomposed nonlinear kinematic hardening rules with threshold, also

41
Paper I-International Journal of Plasticity 16 (2000) 381-409

overpredict the biaxial ratcheting responses.

The models proposed by Prager, Mroz, Besseling, Chaboche and Ohno-Wang determine model

parameters using uniaxial experiments only. Therefore, the kinematic hardening rule in each of these

models with the assumption of stable yield surface fails to simulate the biaxial ratcheting strains

satisfactorily. Hassan et al. [1992] and Corona et al. [1996] explained graphically the reason behind this

weakness of models. Each of these kinematic hardening rules stabilizes the backstress loading paths

differently after a few initial cycles and thereby, each rule dictates different normal directions to the yield

surface. These normal directions determine the amount of ratcheting strain during an increment. In

reality, the yield surface changes shape during plastic deformations [Phillips and Lee, 1979]. Hence, the

normals to the actual yield surface are different from those predicted by models [Corona et al., 1996].

This difference can be compensated, if some model parameters are calibrated using biaxial ratcheting

experiments. This methodology is demonstrated by the Guionnet [1992] model which simulates the

circumferential ratcheting strain in the constant pressure biaxial experiment impressively. Unfortunately,

when the loading changes to the biaxial bow-tie histories, which are different from the biaxial experiment

used for parameter calibration, this model fails to simulate the ratcheting responses.

The above observations indicate that the incorporation of parameters, which improve multiaxial

ratcheting simulations while being dormant for uniaxial ratcheting responses, in a models kinematic

hardening rule can lead to a better constitutive modeling scheme (Bari and Hassan [1999b]). Uncoupling

of the plastic modulus calculation from its kinematic hardening rule (uncoupled model) and determining

the parameters in the kinematic rule using multiaxial ratcheting responses can be another viable

alternative in achieving a constitutive model for ratcheting simulation (Bari and Hassan [1999a]). Finally,

towards achieving a robust model for ratcheting simulation, the incorporation of yield surface shape

change (formative hardening rule) into the modeling scheme seems to be a more rewarding approach.

42
Paper I-International Journal of Plasticity 16 (2000) 381-409

Acknowledgment - The financial supports from the Center for Nuclear Power Plant Structures, Equipment
and Piping and the Department of Civil Engineering at North Carolina State University are gratefully
acknowledged.

NOMENCLATURE

a = deviatoric backstress tensor



s = deviatoric stress tensor

n = unit normal to yield surface at the current stress point

p
dp = d = magnitude of plastic strain increment tensor

t = history or time
E = Youngs modulus
H = generalized plastic modulus
N = number of loading cycles
= total backstress tensor

d = incremental total backstress tensor

x = total axial backstress
i = axial decomposed backstress
= strain tensor

p
d = plastic strain increment tensor

x = axial strain
xc = amplitude of axial strain cycle
xp = maximum axial strain in a cycle
xp = plastic axial strain
Lp = plastic axial strain amplitude of a strain-controlled hysteresis loop
= circumferential strain
p = maximum circumferential strain in a cycle
= stress tensor

o = size of yield surface
x = total axial stress
xa = amplitude of axial stress cycle
xm = mean of axial stress cycle
= circumferential stress
a = amplitude of circumferential stress cycle
m = mean of circumferential stress cycle

43
Paper I-International Journal of Plasticity 16 (2000) 381-409

REFERENCES

ANSYS (1998) ANSYS users Manual, Revision 5.4. ANSYS Inc. Providence, Houston, PA 15342-

0065, U.S.A.

Armstrong, P.J. and Frederick, C.O. (1966) A Mathematical Representation of the Multiaxial Bauscinger

Effect. CEGB Report No. RD/B/N 731.

Bari, S. and Hassan. T. (1999a) Uncoupled Constitutive Models for Simulating Ratcheting Responses. To

be submitted to the International Journal of Solids and Structures.

Bari, S. and Hassan. T. (1999b) An Improved Constitutive Model for Multiaxial Ratcheting. To be sub-

mitted to the Acta Mechanica.

Besseling, J.F. (1958) A Theory of Elastic, Plastic and Creep Deformations of an Initially Isotropic Mate-

rial. Journal of Applied Mechanics, Vol 25, pp. 529-536.

Chaboche, J.L., Dang-Van, K. and Cordier, G. (1979) Modelization of the Strain Memory Effect on the

Cyclic Hardening of 316 Stainless Steel. Proceedings of the 5th International Conference on SMiRT,

Div. L, Berlin, Germany.

Chaboche, J.L. (1986) Time-Independent Constitutive Theories For Cyclic Plasticity. International Jour-

nal of Plasticity, Vol 2, pp. 149-188.

Chaboche, J.L. (1991) On Some Modifications of Kinematic Hardening to Improve the Description of

Ratcheting Effects. International Journal of Plasticity, Vol 7, pp. 661-678.

Chaboche, J.L. (1994) Modeling of ratchetting: evaluation of various approaches. European Journal of

Mechanics, A/Solids, Vol 13, pp. 501-518.

Corona, E., Hassan, T. and Kyriakides, S. (1996) On the Performance of Kinematic Hardening Rules in

Predicting a Class of Biaxial Ratcheting Histories. International Journal of Plasticity, Vol 12, pp. 117-

145.

Dafalias, Y.F. and Popov, E.P. (1976) Plastic Internal Variables Formalism of Cyclic Plasticity. Journal of

44
Paper I-International Journal of Plasticity 16 (2000) 381-409

Applied Mechanics, Vol 43, pp. 645-650.

Delobelle, P., Robinet, P. and Bocher, L. (1995) Experimental Study and Phenomenological Modelization

of Ratchet Under Uniaxial and Biaxial Loading on an Austenitic Stainless Steel. International Journal

of Plasticity, Vol 11, pp. 295-330.

Drucker, D.C. and Palgen, L. (1981) On stress-Strain Relations Suitable for Cyclic and Other Loadings.

Journal of Applied Mechanics, Vol 48, pp. 479-485.

Guionnet, C. (1992) Modeling of Ratcheting in Biaxial Experiments. Journal of Engineering Materials

and Technology, Vol 114, pp. 56-62.

Hassan, T. and Kyriakides, S. (1992) Ratcheting in Cyclic Plasticity, Part I: Uniaxial Behavior. Interna-

tional Journal of Plasticity, Vol 8, p. 91-116.

Hassan, T., Corona, E. and Kyriakides, S. (1992) Ratcheting in Cyclic Plasticity, Part II: Multiaxial

Behavior. International Journal of Plasticity, Vol 8, p. 117-146.

Hassan, T. and Kyriakides, S. (1994a) Ratcheting of Cyclically Hardening and Softening Materials, Part

I: Uniaxial Behavior. International Journal of Plasticity, Vol 10, pp. 149-184.

Hassan, T. and Kyriakides, S. (1994b) Ratcheting of Cyclically Hardening and Softening Materials, Part

II: Multiaxial Behavior. International Journal of Plasticity, Vol 10, pp. 185-212.

Ishikawa, H. and Sasaki, K. (1988) Yield Surfaces of SUS 304 Under Cyclic Loading. Journal of Engi-

neering Materials and Technology, Vol 110, pp. 365-371.

Jhansale, H.R. (1975) A New Parameter for the Hysteretic Stress-Strain Behaviour of Metals. Journal of

Engineering Materials and Technology, pp. 33-38.

Jiang, Y. and Sehitoglu, H. (1996a) Modeling of Cyclic Ratchetting Plasticity, Part I: Development of

Constitutive Relations. Journal of Applied Mechanics, Vol 63, pp. 720-725.

Jiang, Y. and Sehitoglu, H. (1996b) Comments on the Mroz Multiple Surface Type Plasticity Models.

International Journal of Solids and Structures, Vol 33, pp. 1053-1068.

45
Paper I-International Journal of Plasticity 16 (2000) 381-409

McDowell, D.L. (1995) Stress State Dependence of Cyclic Ratcheting Behavior of Two Rail Steels.

International Journal of Plasticity, Vol 11, pp. 397-421.

Morrow, J. (1965) Cyclic Plastic Strain Energy in Fatigue of Metals. Proceedings of Symp. on Internal

Friction Damping in Cyclic Plasticity, ASTM STP 378, pp. 45-87.

Mroz, Z. (1967) On the Description of Anisotropic Work Hardening. Journal of the Mechanics and Phys-

ics of Solids, Vol 15, pp. 163-175.

Mroz, Z. and Rodzik, P. (1996) On Multisurface and Integral Description of Anisotropic Hardening Evo-

lution of Metals. European Journal of Mechanics, A/Solids, Vol 15, pp. 1-28.

Ohno, N. and Wang, J.-D. (1993) Kinematic Hardening Rules with Critical State of Dynamic Recovery,

Part I: Formulations and Basic Features for Ratcheting Behavior. International Journal of Plasticity,

Vol 9, pp. 375-390.

Ohno, N. (1997) Current State of the Art in Constitutive Modeling for Ratcheting. Proceedings of the

14th International Conference on SMiRT, Lyon, France, pp. 201-212.

Phillips, A. and Lee, C.W. (1979) Yield Surfaces and Loading Surfaces. Experiments and Recommenda-

tions. International Journal of Solids and Structures, Vol. 15, pp. 715-729.

Prager, W. (1956) A New Method of Analyzing Stresses and Strains in Work Hardening Plastic Solids.

Journal of Applied Mechanics, Vol 23, pp. 493-496.

Tseng, N.T. and Lee, G.C. (1983) Simple Plasticity Model of the Two-Surface Type. ASCE Journal of

Engineering Mechanics, Vol 109, pp. 795-810.

Tuegel, E.J. (1987) Measurements of Kinematic and Isotropic Hardening in 304 Stainless Steel During

Cyclic Deformations. Res Mechanica, Vol 22, pp. 65-78.

Voyiadjis, G.Z. and Sivakumar, S.M. (1991) A Robust Kinematic, A Robust Kinematic Hardening Rule

for Cyclic Plasticity with Ratcheting Effects, Part I: Theoretical Formulation. Acta Mechanica, Vol 90,

pp. 105-123.

46
Paper I-International Journal of Plasticity 16 (2000) 381-409

Voyiadjis, G.Z. and Basuroychowdhury, I.N. (1998) A Plasticity Model for Multiaxial Cyclic Loading

and Ratchetting. Acta Mechanica, Vol 126, pp. 19-35.

Xia, Z. and Ellyin, F. (1997) A Constitutive Model with Capability to Simulate Complex Multiaxial

Ratcheting Behaviour of Materials. International Journal of Plasticity, Vol 13, pp. 127-142.

Yuan, X., Hassan, T. and Bari, S. (1999) Improved Ratcheting Simulations by ANSYS and ABAQUS.

(Under preparation).

47
CHAPTER 3

UNCOUPLED CONSTITUTIVE MODELS FOR RATCHETING


SIMULATIONS

It has been realized from the study in chapter 2 that the distinction of the roles
played by the kinematic hardening rule and the plastic modulus is central to the
understanding of the performance of a cyclic plasticity model. In coupled models, when
kinematic hardening rule parameters are calibrated from uniaxial loading condition, the
ratcheting simulations under multiaxial loading condition suffer and vice versa. In an
effort to avoid this problem in cyclic plasticity modeling, the uncoupled models are
introduced where the plastic modulus may be indirectly influenced by the kinematic
hardening rule but its calculation does not depend on it. Thus, the simulation of uniaxial
ratcheting is dictated only by the plastic modulus calculation scheme, whereas the
simulation of multiaxial ratcheting is mainly influenced by the kinematic hardening rule
adopted in the model. In this study, the modified Dafalias-Popov [1976] model is used as
the basic plastic modulus calculation scheme and eight different kinematic rules are
evaluated in terms of their ratcheting simulations for the same set of uniaxial and
multiaxial loading histories considered in chapter 2. A journal paper "Kinematic
Hardening Rules in Uncoupled Modeling for Multiaxial Ratcheting Simulation" written
from this part of the study is accepted for publication in the International Journal of
Plasticity. This paper is enclosed next as paper II. Readers are referred to this paper for
details of the study conducted in this chapter.

48
Paper II-To be published in the International Journal of Plasticity (in press)

Kinematic Hardening Rules in Uncoupled Modeling for


Multiaxial Ratcheting Simulation
Shafiqul Bari and Tasnim Hassan*
Center for Nuclear Power Plant Structures, Equipment and Piping
Department of Civil Engineering, North Carolina State University,
Raleigh, NC 27695-7908; e-mail: thassan@eos.ncsu.edu

Keywords: Cyclic plasticity, multiaxial ratcheting and constitutive modeling


Shortened Title: Kinematic Hardening Rules for Ratcheting

ABSTRACT

An earlier paper by the authors evaluated the performance of several coupled models in simulating a

series of uniaxial and biaxial ratcheting responses. This paper evaluates the performance of various

kinematic hardening rules in an uncoupled model for the same set of ratcheting responses. A modified

version of the Dafalias-Popov uncoupled model has been demonstrated to perform well for uniaxial

ratcheting simulation. However, its performance in multiaxial ratcheting simulation is significantly

influenced by the kinematic hardening rules employed in the model. Performances of eight different

kinematic hardening rules, when engaged with the modified Dafalias-Popov model, are evaluated against

a series of rate-independent multiaxial ratcheting responses of cyclically stabilized carbon steels. The

kinematic hardening rules proposed by Armstrong-Frederick, Voyiadjis-Sivakumar, Phillips, Tseng-Lee,

Kaneko, Xia-Ellyin, Chaboche and Ohno-Wang are examined. The Armstrong-Frederick rule performs

reasonably for one type of the biaxial ratcheting response, but fails in others. The Voyiadjis-Sivakumar

rule and its constituents, the Phillips and the Tseng-Lee rules, can not simulate the biaxial ratcheting

responses. The Kaneko rule, composed of the Ziegler and the prestress directions, and the Xia-Ellyin

rule, composed of the Ziegler and Mroz directions, also fail to simulate the biaxial ratcheting responses.

The Chaboche rule, with three decomposed Armstrong-Frederick rules, performs the best for the whole

set of ratcheting responses. The Ohno-Wang rule performs well for the data set, except for one biaxial

* The corresponding author.

49
Paper II-To be published in the International Journal of Plasticity (in press)

response where it predicts shakedown with subsequent reversal of ratcheting.

I. INTRODUCTION

Pioneering works of Dafalias and Popov [1976], Yoshida et al. [1978], and Chaboche et al. [1979]

developed new concepts in the modeling of ratcheting responses. Motivated by these works, many

researchers have proposed improved models for simulation of ratcheting (Voyiadjis and Sivakumar,

[1991], Voyiadjis and Basuroychowdhury [1998], Hassan and Kyriakides [1992], Hassan et al. [1992],

Ohno and Wang [1993], Chaboche [1994], Delobelle et al. [1995], McDowell [1995], Jiang and

Shehitoglu [1996], Xia and Ellyin [1997] and others). Despite the considerable efforts made so far,

existing constitutive models fail to simulate ratcheting from a broad class of loading histories (Corona et

al. [1996], Bari and Hassan [2000a]).

Performance studies of many existing models demonstrate that the simulation for uniaxial

ratcheting primarily depends on the plastic modulus calculation scheme, whereas the simulation for

ratcheting in multiaxial loading depends essentially on the kinematic hardening rule employed in the

model (Hassan and Kyriakides [1992], Hassan et al. [1992], and Corona et al. [1996]). In this paper,

ratcheting from uniaxial loading is referred to as uniaxial ratcheting and that from multiaxial loading is

referred to as multiaxial ratcheting. Note that in case of multiaxial loading, ratcheting may occur along

one or more directions depending on the control mode of the prescribed loading.

An anatomical study of several coupled constitutive models for simulating ratcheting responses in

both uniaxial and biaxial loading has recently been performed by Bari and Hassan [2000a]. In coupled

models, the plastic modulus calculations are directly tied to the kinematic hardening rule of the model

through the consistency condition. The models proposed by Chaboche [1991,1994] and Ohno and Wang

[1993] perform well to simulate uniaxial ratcheting responses, but consistently overpredict the biaxial

ratcheting responses. This limitation of the models can be attributed to the determination of the model

50
Paper II-To be published in the International Journal of Plasticity (in press)

parameters from uniaxial responses and the assumption of the von-Mises yield surface. The

Guionnet [1992] model, on the other hand, determines two of its parameters using biaxial

responses and thus, improves the simulation for a set of biaxial ratcheting responses. This model,

however, fails to simulate the uniaxial ratcheting and other biaxial ratcheting responses. The

authors have improved the biaxial ratcheting simulations of the coupled models by introducing a

parameter in the hardening rule. This new parameter is determined from a biaxial ratcheting

experiment and thus produces representative plastic strain rate directional properties for

multiaxial ratcheting, but does not influence the uniaxial ratcheting simulation. This improved

coupled model and its simulations will be presented in a forthcoming paper by the authors (Bari

and Hassan [2000b]).

Another viable alternative in achieving an improved constitutive model for ratcheting

simulation is uncoupling of the plastic modulus calculation from the kinematic hardening rule

(Hassan et al. [1992]). In the uncoupled models, the plastic modulus may be indirectly influenced

by the kinematic hardening rule but its calculation does not depend on it. Thus, the simulation of

uniaxial ratcheting is dictated only by the plastic modulus calculation scheme, whereas the

simulation of multiaxial ratcheting is mainly influenced by the kinematic hardening rule adopted

in the model. Following the pioneering work of Mroz [1967], new concepts of uncoupled models

have been introduced by Dafalias and Popov [1976], Drucker and Palgen [1981], Tseng and Lee

[1983] and others. Hassan and Kyriakides [1992, 1994a, 1994b], Hassan et al. [1992] and Corona

et al. [1996] evaluated several of these uncoupled models and various kinematic hardening rules

for ratcheting simulation. They have developed an improved version of the Dafalias-Popov model

that can simulate uniaxial ratcheting responses quite well. This model, however, fails to simulate a

broad class of biaxial ratcheting responses with kinematic hardening rules of Ziegler [1959],

51
Paper II-To be published in the International Journal of Plasticity (in press)

Armstrong-Frederick [1966], Mroz [1967], Phillips and Lee [1979] and Tseng and Lee [1983].

In order to identify and develop a robust constitutive model for ratcheting simulation, this paper

evaluates the performances of several kinematic hardening rules with the improved Dafalias and Popov

model. The set of uniaxial and biaxial ratcheting responses that is adopted by the authors to evaluate the

coupled models (Bari and Hassan [2000a]) is also used in this study. The evaluation in this paper starts

with a well-known kinematic hardening rulethe Armstrong and Frederick [1966] rule. Following this,

the kinematic hardening rules by Voyiadjis and Sivakumar [1991], Phillips and his coworkers

[1972,1979,1984], Tseng and Lee [1983], Kaneko [1981,1984], and Xia and Ellyin [1994,1997] are

evaluated. Finally, an improved model, composed of the modified Dafalias-Popov model (Hassan and

Kyriakides [1992, 1994a]) and the kinematic hardening rule by Chaboche [1991, 1994], is presented

along with its ratcheting simulations. The simulations from the Ohno-Wang [1993] kinematic rule, when

incorporated in the modified Dafalias-Popov model, is also presented.

II. EXPERIMENTAL RESPONSES STUDIED

Model performances are evaluated with respect to a series of uniaxial and biaxial ratcheting

responses of cyclically stable carbon steels acquired from Hassan and Kyriakides [1992], Hassan et al.

[1992] and Corona et al. [1996]. The philosophy behind studying the ratcheting responses of cyclically

stable materials is discussed in Bari and Hassan [2000a]. A brief review of the ratcheting responses

studied is given below (see Hassan and Kyriakides [1992], Hassan et al. [1992] and Corona et al. [1996]

for details on the experiments).

The loading histories of the experiment set are shown in Fig. 1. The history in Fig. 1a, which results

in uniaxial ratcheting, involves axial stress cycles with a mean stress. The readers are referred to Hassan

and Kyriakides [1992] for demonstration of a typical uniaxial ratcheting response (Fig. 7 in the

reference). Data from these experiments are presented in Figs. 3a and 3b, where the peak axial strains

52
Paper II-To be published in the International Journal of Plasticity (in press)


xc xc
xc
x a a
xa
xm m
m m
t
x x x
(a) (b) (c) (d)

Fig. 1. Loading histories; (a) Uniaxial stress cycle, (b) Axial strain cycle with constant internal
pressure, (c) biaxial bow-tie cycle, (d) biaxial reverse bow-tie cycle.

(xp) of each cycle are plotted against the number of cycles, N. The loading history in Fig. 1b involves

axial strain cycles in the presence of a constant internal pressure. This history results in circumferential

ratcheting as demonstrated in Fig.1 of Hassan et al. [1992]. Two sets of data obtained from this biaxial

loading history are studied (see Figs. 3c and 3d). In Figs. 3c and 3d, the peak circumferential strains (p)

in each cycle are plotted against the number of cycles, N. The bow-tie and reverse bow-tie loading

histories, shown in Figs. 1c and 1d, also result in circumferential ratcheting as demonstrated in Figs. 10

and 11 of Corona et al. [1996]. The data from the bow-tie histories are shown in Figs. 3e and 3f.

III. THEORY AND RESULTS

III. 1. Dafalias and Popov Model

Implementation of the Dafalias-Popov model [1976] involves the following equations:

p 1 f f
i. the flow rule: d = ---- ------ d ------ , (1)
H

3 12
ii. the von-Mises yield surface: f ( ) = --- ( s a ) ( s a ) = 0 , (2)
2

3 12
iii. the bounding surface: F ( ) = --- ( s b ) ( s b ) = b , (3)
2

p p
iv. the kinematic hardening rule: da = g ( , , a, d, d , etc ) ; (4)

53
Paper II-To be published in the International Journal of Plasticity (in press)

p
where, is the stress tensor, is the plastic strain tensor, s is the deviatoric stress tensor, is the

current center of the yield surface, a is the current center of the yield surface in the deviatoric space, is

the current center of the bounding surface, b is the current center of the bounding surface in the

deviatoric space, H is the plastic modulus, 0 and b are the sizes of the yield and bounding surfaces,

respectively. indicates the MacCauley bracket and the inner product a b = a ij b ij . For a cyclically

stable material, the values of 0 and b remain unchanged during plastic deformation. Unlike a coupled

model, the Dafalias and Popov model has the flexibility to adopt any kinematic hardening rule.

In the Dafalias-Popov model [1976], the plastic modulus H is evaluated using

12
H = E o + h ---------------- , , = 3--- ( s s ) ( s s )
P a
in
h = ------------------------------- . (5)
in m 2
1 + b ---------
2 b

The parameters of the model are shown in Fig. 2, where and in Eq. (5), Eop is the plastic modulus at the

bound or bounding surface, ( s ) is the image of ( s ) on the bounding surface, is the distance in stress

space of the current stress state from the bounding surface image point, in is the distance in stress space

2 Yield
Bound P Surface
Eo Bounding
b Surface

o
o


p 1


in n



n

(a) (b)

Fig. 2. Definition of parameters of the Dafalias-Popov model; (a) in uniaxial -p space,


(b) in biaxial 1-2 stress space.

54
Paper II-To be published in the International Journal of Plasticity (in press)

of the recent yielding state from the corresponding bounding surface image point, and a, b, and m are

model constants (see Dafalias and Popov [1976] for details of the model).

The original Dafalias-Popov model [1976] simulates an eventual shakedown for the uniaxial

ratcheting response. Hassan and Kyriakides [1992,1994a] developed a modified version of this model to

improve its uniaxial ratcheting simulation. The improvement incorporates relaxation of the bounding

surface through its kinematic rule

p
E ds n
db = da m dM and d M = 1 -----b- -------------
- (6)
H m n

P P 12 3 (s a) 3 (s s)
where, Eb = Eo + Cr[(b b) b n] , n = --- ----------
- , m = --- ----------
- . (7)
2 2
0

Ebp defines the modulus for the translation of the bounding surface center, not the bound modulus, which

is always Eop; m is the unit vector along the direction ( s s ) , n is the unit vector along the normal of the

yield surface and Cr is a model parameter.

Note that in the Dafalias and Popov model [1976] the plastic modulus calculation (Eq. 5) is not

directly dependent on the yield surface kinematic hardening rule (Eq. 4). In fact, the uniaxial stable

hysteresis loop and ratcheting simulations by the Dafalias-Popov model are entirely controlled by the

parameters involved in the plastic modulus calculations (Eqs. 5 and 6). In this model, the kinematic

hardening rule specifies the direction of movement of the yield surface center. During a uniaxial stress

cycle, the yield surface moves along the stress direction only. Hence, the Dafalias-Popov model always

produces the same uniaxial stress-strain response irrespective of the kinematic hardening rule employed

with the model. It is the multiaxial ratcheting simulation that gets affected by different kinematic

hardening rules. For a given kinematic hardening rule, the amount of movement along the specified

direction of the rule is determined using the consistency condition

f ( + d d ) = 0 . (8)

55
Paper II-To be published in the International Journal of Plasticity (in press)

The parameter determination scheme in the Dafalias-Popov model for the cyclically stabilized

materials is described in Hassan and Kyriakides (1992,1994a). In this scheme, o, b, Eop, a, b, m, and E

are determined by using a stable hysteresis curve and Cr is determined from the ratcheting rate in an

uniaxial experiment. The parameters determined and used for simulations are:

E = 26300 ksi, = 0.302, 0 = 18.8 ksi, b = 38.66 ksi, E0p = 274 ksi,

a = 71100, b = 27, m = 2, Cr = 8.

III. 2. Kinematic Hardening Rules and Ratcheting Simulations

III.2.1. Armstrong and Frederick Rule

The Armstrong and Frederick [1966] kinematic hardening rule is used with the modified Dafalias and

Popov model (Hassan and Kyriakides [1992, 1994a]) in the form

12
da = ( --- Cd a dp )
2 p p 2 p p
where dp = d = --- d d . (9)
3 3

The parameter C (= 140 ksi) is determined from the ratcheting rate of a constant pressure biaxial

experiment. In this study, the experiment (2) in Fig. 3c is used for this purpose. The value of the scalar

is determined using the consistency condition, Eq. (8), for each loading increment. Figure 3 shows the

simulations of the ratcheting responses by the Armstrong-Frederick hardening rule. The good simulations

for the uniaxial ratcheting responses, shown in Figs. 3a and 3b, are not surprising, as these are the

outcome of the improved Dafalias-Popov model (Hassan and Kyrikides [1992,1994a]). As stated earlier,

any other kinematic hardening rule would also produce the same uniaxial simulation with this uncoupled

model.

The ratcheting simulations for the biaxial loading history of Fig. 1b with the Armstrong-Frederick

kinematic hardening rule are shown in Figs. 3c and 3d. The simulations exhibit constant rate of ratcheting

and conform reasonably well with the experiments. It should be pointed out here that the Armstrong-

56
Paper II-To be published in the International Journal of Plasticity (in press)

xp xp
4 4
CS 1026 CS 1026
xa = 32.0 ksi xm = 9.14 Ksi xm = 6.5 ksi
(%) (%) a = 33.28 Ksi
x
xa
3 3 xm
t

xm = 6.52 Ksi xa= 32.12 Ksi


2 2
xa = 30.35 Ksi

xm = 4.18 Ksi
1 1
Experiment
DP model xa = 28.29 Ksi

0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(a) (b)
p p
4 4
CS 1026 CS 1026
(%) m = 9.65 Ksi (%) xc = 0.50 %
xc

3 (1) xc = 0.40 % 3 m (1) m = 4.88 Ksi


(2) xc = 0.50 % (2) m = 7.32 Ksi
(3) xc = 0.65 % x (3) m = 9.65 Ksi
(4) m =14.54 Ksi

2 2
(3) (2) (4) (3)
(2)

1 (1) 1 (1)

Experiment
Armstrong-Frederick
0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(c) (d)
p p
4 4
CS 1026 CS 1018
(%) a = 2.36 Ksi (%) a = 2.36 Ksi
xc = 0.50 % xc = 0.50 %
3 3

(2) (1) m = 9.65 Ksi (1) m = 9.65 Ksi


(1)
(2) m = 14.54 Ksi (2) m = 14.54 Ksi

2 (2) xc 2 (2)

(1) a xc
m a
1 1 (1)
x m
(2)
Experiment x
(1)
Armstrong-Frederick
0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(e) (f)

Fig. 3. Ratcheting Predictions by the Armstrong-Frederick rule with the modified Dafalias-Popov
model (DP). (a,b) Axial strain at positive stress peaks from uniaxial cycles, (c,d) circumferential
strain peaks from axial strain cycles with constant pressure, (e,f) circumferential strain peaks from
bow-tie and reverse bow-tie cycles. Experiments are from Hassan and Kyriakides [1992], Hassan et
al. [1992] and Corona et al. [1996].

57
Paper II-To be published in the International Journal of Plasticity (in press)

Frederick kinematic hardening rule in a coupled model overpredicts the ratcheting for this set of biaxial

experiments (Bari and Hassan, 2000a). In an uncoupled model, the parameter C for this rule is

determined from a biaxial ratcheting response and hence, the simulations for similar biaxial responses are

improved. However, with the same set of parameters, this rule overpredicts the ratcheting rate in the bow-

tie experiments (see Fig. 3e) and underpredicts in the reverse bow-tie experiments (see Fig. 3f).

III.2.2. Voyiadjis and Sivakumar Rule

The kinematic hardening rule proposed by Voyiadjis and Sivakumar [1991,1994] is a combination

of the Phillips (Phillips and Tang [1972], Phillips and Lee [1979]) and Tseng-Lee [1983] rules. They

reason that the Phillips rule, which stipulates the evolution of the backstress along the stress rate

direction, follow the experimental trend better, but it does not ensure the tangential nesting of the yield

and other surfaces in a multi-surface plasticity model. In contrast, the Tseng-Lee rule invokes the desired

nesting feature for both the proportional and non-proportional loading. Hence, Voyiadjis and Sivakumar

[1991] propose a rule to appropriately blend the deviatoric stress rate and the Tseng-Lee rules, to include

both the requirement of surface nesting and the experimental observation of the yield surface movement.

Note that, in the improved Dafalias-Popov model, the surface nesting is inherently ensured with any

kinematic hardening rule. In order to provide a broader perspective of this kinematic hardening rule, the

Phillips [1972,1979] and Tseng-Lee [1983] rules are also evaluated here along with the rule proposed by

Voyiadjis and Sivakumar [1991].

Phillips Rule: Phillips and his co-workers [1972,1979,1984] recommend a kinematic hardening

rule which prescribes the yield surface translation in the stress rate direction as follows:

da = d ( v D ) (10)

12 12
- , x for any tensor x means --- x x = --- x ij x ij
ds 3 3
where, v D = l = --------- 2
, and d is a scalar variable to
ds 2

be determined from the consistency condition.

58
Paper II-To be published in the International Journal of Plasticity (in press)

Tseng and Lee Rule: The Tseng-Lee [1983] rule can be described with the help of Fig. 4 as

follows:

da = d ( v T ) , (11)

a' a ds
- , a' = ( s + 'l ) ----- ( s + 'l b ) , l = ---------
- , ' is calculated using
0
where, v T = ----------------
a' a b ds

2 2 2
' = ( s b ) l + { ( s b ) l } ( s b ) ( s b ) + --- b , (12)
3

which is obtained from the condition (see Fig. 4)

2 2
( s + 'l b ) ( s + 'l b ) = --- b , (13)
3

and d is a scalar variable.

When the above two rules are used with the modified Dafalias-Popov model, the value of the scalar

d, in Eqs. (10) and (11), is determined using the consistency condition, Eq. (8) during each load

increment. Note that, unlike the Armstrong-Frederick kinematic hardening rule discussed in section

III.2.1, these hardening rules do not include any parameters which can be calculated using ratcheting

responses.

n

'
S2 ds

s a
a'

b
Bounding
Surface

S1
S3
Fig. 4. Yield surface movement in the Tseng-Lee kinematic hardening rule

59
Paper II-To be published in the International Journal of Plasticity (in press)

p p
4 4
CS 1026 CS 1026
(%) m = 9.65 Ksi (%) xc = 0.50 %
(1) m = 4.88 Ksi
3 (1) xc = 0.40 % 3 (2) m = 9.65 Ksi
(2) xc = 0.50 % (3) m =14.54 Ksi


xc xc
2 2
(2) m (3) (2) m
x x
(1)
1 1 (1)
Experiment
Phillips
Tseng-Lee
0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(a) (b)

2 4
CS 1026 CS 1018
p (2)
a = 2.36 Ksi p a = 2.36 Ksi
(%) (%)
xc = 0.50 % xc = 0.50 %
1 (1) (1) m = 9.65 Ksi 3
(2) m = 14.54 Ksi
(1) m = 9.65 Ksi
(2) m = 14.54 Ksi
0 2 (2)

xc
xc
a
a
-1 1 (1)
m
Experiment m
x Phillips x
Tseng-Lee
-2 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(c) (d)

Fig. 5. Ratcheting Predictions by the Phillips and Tseng-Lee rules with the modified Dafalias-
Popov model. (a,b) Circumferential strain peaks from axial strain cycles with constant pressure,
(c,d) circumferential strain peaks from bow-tie and reverse bow-tie cycles. Experiments are from
Hassan et al. [1992] and Corona et al. [1996].

Figure 5 shows the biaxial ratcheting simulations by the Phillips and Tseng-Lee rules along with the

experimental data. Since the uniaxial ratcheting simulations are not influenced by the kinematic

hardening rule employed in an uncoupled model, they are not presented again in Fig 5. This figure

demonstrates that both the Phillips and Tseng-Lee rules fail to simulate biaxial ratcheting for the

experiments studied. The Phillips rule can not even reproduce the effect of different levels of internal

pressure for experiments shown in Figs. 5b and 5d.

60
Paper II-To be published in the International Journal of Plasticity (in press)

The Voyiadjis-Sivakumar rule blends the stress rate and Tseng-Lee directions by introducing two

parameters and as follows

ba ba
= ---------------------------
, = l ---------------
- . (14)
2 ba
--- ( b 0 )
3

Depending on the value of , the v D (Eq. 10) and v T (Eq. 11) directions are combined using

0 vD + 1 vT
= ----------------------------------
1
- , 0 + 1 = 1 (15)
0 vD + 1 vT

Where, 0() and 1() are blending functions satisfying the boundary conditions
0 = 1 , 1 = 0 for = 0; thus, = v D
1


0 = 0 , 1 = 1 , for = 1; thus, = v T
1


The direction is further combined with v D using
1


' 0 + ' 1 v D
1

2
- , ' 0 + ' 1 = 1
= ------------------------------------
1 . (16)
' 0 + ' 1 v D

Where, ' 0 ( ) and ' 1 ( ) are blending functions satisfying the boundary conditions

' 0 = 1 , ' 1 = 0 , for = 0; thus, =


2 1


' 0 = 0 , ' 1 = 1 , for = 1; thus, = v D
2

Voyiadjis and Sivakumar [1991] propose the kinematic rule in the form

da = d ( ).
2
(17)

Where, the scalar d is determined using the consistency condition, Eq. (8) during each loading

increment. They propose to determine the blending functions 0(), 1(), ' 0 ( ) and ' 1 ( ) from

experimental responses, but have not provided any guideline to obtain these functions. To evaluate the

performance of their hardening rule the following functions are assumed in this study:

q
0 = 1 , thus, 1 = q (18)

1q
' 0 = 1 , thus, ' 1 = 1 q . (19)

61
Paper II-To be published in the International Journal of Plasticity (in press)

When, q = 1, both these functions vary linearly with and , respectively. If the value of q is increased

further, the dominance of the Phillips rule increases over the Tseng-Lee rule.

Figure 6 shows the simulations for the biaxial ratcheting responses by the Voyiadjis-Sivakumar

kinematic hardening rule for q = 1 and q = 10. Since, both the Phillips and Tseng-Lee rules fail to

simulate the biaxial ratcheting responses (Fig. 5), the Voyiadjis-Sivakumar rule that is a combination of

these two rules, also fails to simulate these responses as shown in Fig 6. Note that, for q = 1, the

simulations fall in between the simulations by the Phillips and Tseng-Lee rules, as can be observed by

p p
4 4
CS 1026 CS 1026
(%) m = 9.65 Ksi (%) xc = 0.50 %
(1) m = 4.88 Ksi
3 (1) xc = 0.40 % 3 (2) m = 9.65 Ksi
(2) xc = 0.50 % (3) m =14.54 Ksi


xc
2 xc 2
(2) (3) (2) m
m
x
(1) x
1 1 (1)
Experiment
Voyiadjis-Sivakumar (q=1)
Voyiadjis-Sivakumar (q=10)
0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(a) (b)
p p
2 4
CS 1026 CS 1018
(2)
(%) a = 2.36 Ksi (%) a = 2.36 Ksi
xc = 0.50 % xc = 0.50 %
1 (1) (1) m = 9.65 Ksi 3
(2) m = 14.54 Ksi
(1) m = 9.65 Ksi
(2) m = 14.54 Ksi
xc
0 a 2 (2)

xc
m
a
x (1)
-1 1
m
Experiment
Voyiadjis-Sivakumar (q=1) x
Voyiadjis-Sivakumar (q=10)
-2 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(c) (d)

Fig. 6. Ratcheting Predictions by the Voyiadjis-Sivakumar rule with the modified Dafalias-Popov
model. (a,b) Circumferential strain peaks from axial strain cycles with constant pressure, (c,d)
circumferential strain peaks from bow-tie and reverse bow-tie cycles. Experiments are from
Hassan et al. [1992] and Corona et al. [1996].

62
Paper II-To be published in the International Journal of Plasticity (in press)

comparing Figs. 5 and Fig. 6. It is also observed from these figures that, for q = 10, the simulations reside

very close to those obtained from the Phillips rule. The authors would like to point out here that, in the

experiments by Phillips and his coworkers [1972,1979,1984]), the yield surfaces change shape when they

translate in the stress-rate direction. This might be a reason why the von-Mises yield surface with

kinematic hardening rules along the stress-rate direction can not represent multiaxial ratcheting

responses.

III.2.3. Kaneko Rule

Kaneko [1981,1984] proposes a kinematic hardening rule that combines the prestress path

direction with the Ziegler [1959] direction as follows:

(k)
da = d C ( s a ) + ( s s ) (20)
k

where the s ( k ) are the prestress points in stress space when the stress history changes directions (see

Kaneko [1981] for descriptions and examples). For the biaxial loading history of Fig. 1b, this rule with

Dafalias-Popov model produces negative ratcheting as shown in Fig. 7 (for C= 50). If the value of C is

increased, the Ziegler direction becomes dominant and simulates no ratcheting at all. Thus, the inclusion

of the prestress path direction with the Ziegler rule produces adverse effect on the ratcheting simulation

for the experiments considered.

III.2.4. Xia and Ellyin Rule

Xia and Ellyin [1994,1997] propose a kinematic hardening rule in the following form:

da = d ( s a ) , for monotonic loading (ML) (21a)


da = d ( s s ) , for reloading (RL). (21b)


When the yield and bounding surfaces are in contact at the current stress point, the subsequent

plastic loading is defined as monotonic loading. In this case, both the surfaces translate together along

63
Paper II-To be published in the International Journal of Plasticity (in press)

p
2
CS 1026
(%) (2) m = 9.65 Ksi

1 (1) (1) xc = 0.40 %


(2) xc = 0.50 %


xc
-1
m Experiment
Kaneko
x Xia-Ellyin
-2
0 10 20 30 40 50

Fig. 7. Ratcheting Predictions by the Kaneko and Xia-Ellyin rules with the modified
Dafalias-Popov model for circumferential strain peaks from axial strain cycles with
constant pressure. Experiments are from Hassan et al. [1992].

the Ziegler [1959] direction. On the other hand, when the yield surface is not in contact with the bounding

surface, the subsequent plastic loading increment is defined as reloading and in such a case, the yield

surface moves along the Mroz [1967] direction and the bounding surface moves according to Eq. (6).

This kinematic hardening rule with the modified Dafalias-Popov model overpredicts the ratcheting in

constant pressure biaxial experiments as shown in Fig. 7. This rule also overpredicts the ratcheting rate

for all other biaxial cases considered and hence, are not presented here. The overprediction by this rule is

mainly caused by the Mroz rule (Hassan et al. [1992], Corona et. al [1996]).

III.3. An Improved Uncoupled Model

The above results clearly indicate that many existing kinematic hardening rules when implemented

with an uncoupled model cannot simulate a broad set of ratcheting responses. Since the Armstrong-

Frederick rule performs reasonably well with the uncoupled models in simulating many ratcheting

responses (discussed in section III.2.1.), the hardening rule proposed by Chaboche and his co-workers

[1979,1991], which is the superposition of several Armstrong-Frederick rules, becomes a natural choice

as a viable hardening rule to be used with the uncoupled models. The Ohno-Wangs [1993] modification

64
Paper II-To be published in the International Journal of Plasticity (in press)

to the Chaboche rule is also studied and presented in this context.

After evaluating the Chaboche kinematic hardening rule with two, three and four decomposed

terms, it has been found that with the modified Dafalias-Popov model three decomposed terms

3
i ( --- C i d a i dp ) ,
2
d a i ,
p
da = d ai = 3
(22)

i=1

yields the optimum performance in simulation of the biaxial ratcheting responses. In Eq. (22), the scalar

is determined using the consistency condition, Eq. (8), for each loading increment and Ci, i are model

parameters to be determined from multiaxial responses. It is demonstrated by Bari and Hassan [2000a]

that in the Chaboche [1991] coupled model, each decomposed rule has its specific purpose in reproducing

different features of the uniaxial, nonlinear stable hysteresis loop. However, with the Dafalias-Popov

uncoupled model, this hardening rule loses the physical bearing for uniaxial loading responses as the

plastic modulus is calculated by the uncoupled model itself. Therefore, with an uncoupled model the

parameters of the Chaboche rule have significance only in multiaxial loading conditions and are

recommended to be determined from multiaxial responses.

In order to determine the parameters of the Chaboche rule, the qualitative nature of the parameters

observed in uniaxial loading cases (Bari and Hassan [2000a]) is also adopted here. Since C = 140 ksi in

the original Armstrong-Frederick rule yields good results for most of the ratcheting responses (shown in

Fig. 3), the same value is used here for the third rule (C3 = 140 and 3 = 1). The rest of the parameters are

determined by trials to conform with the nonlinear ratcheting rate of a constant pressure biaxial

experiment (test (2) of Fig. 8c). The general guideline followed is that the first decomposed rule (i = 1)

should have a low C1 value with a large 1 to represent the initial high rate of ratcheting region, while C2

and 2 should have relatively moderate values to represent the transitional nonlinear region. The values of

the parameters determined are: C1-3 = 4, 20, 140; 1-3 = 1200, 20, 1.

65
Paper II-To be published in the International Journal of Plasticity (in press)

xp xp
4 4
CS 1026 CS 1026
xa = 32.0 ksi xm = 9.14 Ksi xm = 6.5 ksi
(%) (%) a = 33.28 Ksi
x
3 3 xm xa
t

xm = 6.52 Ksi xa= 32.12 Ksi


2 2
xa = 30.35 Ksi

xm = 4.18 Ksi
1 1
Experiment
DP model xa = 28.29 Ksi

0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(a) (b)
p p
4 4
CS 1026 CS 1026

(%) xc m = 9.65 Ksi (%) xc = 0.50 %

3 m (1) xc = 0.40 % 3 (1) m = 4.88 Ksi


x (2) xc = 0.50 % (2) m = 7.32 Ksi
(3) xc = 0.65 % (3) m = 9.65 Ksi
(4) m =14.54 Ksi

2 2
(3) (2) (4) (3)
(2)
(1)
1 1 (1)
Experiment
Chaboche
Ohno-Wang
0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(c) (d)
p p
4 4
CS 1026 CS 1018
(%) a = 2.36 Ksi (%) a = 2.36 Ksi
xc = 0.50 % xc = 0.50 %
3 3

(2) (1) m = 9.65 Ksi (1) m = 9.65 Ksi


(1) (2) m = 14.54 Ksi (2) m = 14.54 Ksi

(2) xc
2 2 (2)
a
(1) xc
m
a
(1)
1 x 1
Experiment m
Chaboche x
Ohno-Wang
0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(e) (f)

Fig. 8. Ratcheting Predictions by the Chaboche and Ohno-Wang rules with the modified Dafalias-
Popov model (DP). (a,b) Axial strain at positive stress peaks from uniaxial cycles, (c,d)
circumferential strain peaks from axial strain cycles with constant pressure, (e,f) Circumferential
strain peaks from bow-tie and reverse bow-tie cycles. Experiments are from Hassan and
Kyriakides [1992], Hassan et al. [1992] and Corona et al. [1996].

66
Paper II-To be published in the International Journal of Plasticity (in press)

Figure 8 shows the simulations by this model for all the experiments considered. Note again that,

the kinematic hardening rule with an uncoupled model do not have any effect on the uniaxial ratcheting

simulations (compare Figs. 3a,b to 8a,b). The kinematic hardening rule influences only the biaxial

ratcheting simulations, which is evident from Figs. 8c and 8d. The nonlinear nature of the ratcheting rates

in constant pressure biaxial experiments are reproduced well with the improved model. Moreover, the

ratcheting simulations for both the bow-tie loading histories are also improved (see Figs. 8e,f) by the

Chaboche hardening rule.

a
It has been shown by Bari and Hassan [2000a] that introduction of the term d p -------------
i replacing
f ( i )

dp in the Armstrong-Frederick kinematic rule, that is proposed by Ohno-Wang [1993], is instrumental in

reducing the overprediction of the rate of ratcheting. This modification by Ohno-Wang [1993] is also

evaluated in this paper with the modified Dafalias-Popov model in the following form:

i --- C i d a i d -------------
3 ai
da = d a i , = 2 p p
(23)
3 f ( i )
d ai

i=1

where f ( i ) is the von-Mises function defined in Eq. 2.


The parameter determination scheme followed in this rule is the same as in case of the Chaboche rule.

The parameters obtained are: C1-3 = 4, 20, 30; 1-3 = 400, 12, 1.

Figure 8 also shows the simulations by the Ohno-Wang kinematic hardening rule (Eq. 23). It is

observed in the figure that this rule simulates all the biaxial ratcheting responses more-or-less similar to

those as obtained from the Chaboche rule (Eq. 22). Only, the nonlinear nature of ratcheting rates is more

pronounced in simulations by this rule.

The Chaboche (Eq. 22) and the Ohno-Wang (Eq. 23) kinematic rules are further evaluated for a

reverse bow-tie response with larger number of cycles as shown in Fig. 9. It is observed in this figure that

the Ohno-Wang rule simulates shakedown followed by reversal of ratcheting, which seemingly defies the

67
Paper II-To be published in the International Journal of Plasticity (in press)

p
4
CS 1018
xc
(%) a = 2.36 Ksi
xc = 0.50 % a
3
m = 14.54 Ksi
m

1
Experiment
Chaboche
Ohno-Wang
0
0 10 20 30 40 50

Fig. 9. Ratcheting Predictions by the Chaboche and Ohno-Wang rules with the
modified Dafalias-Popov model for circumferential strain peaks from reverse
bow-tie cycles. Experiments are from Corona et al. [1996].

experimental trend. On the other hand, the Chaboche rule predicts a constant rate of ratcheting at higher

cycles, which does not match the actual material responses exactly, but is clearly more representative.

IV. DISCUSSION AND CONCLUSIONS

In a coupled model, the plastic modulus calculation is directly tied to the kinematic hardening rule.

As a result, if all the parameters in the hardening rule are calibrated from uniaxial responses, the

multiaxial simulations by the coupled model suffer and vice versa (Bari and Hassan [2000a]). The

uncoupling of plastic modulus calculation from the kinematic hardening rule allows the plastic modulus

parameters to be determined from uniaxial responses and the kinematic rule parameters from multiaxial

responses. These two parameters sets are uncoupled, i.e, the uniaxial simulations always remain

unaffected by the selection of the kinematic hardening rule in the model.

This study evaluates eight different kinematic hardening rules in order to identify or develop a

kinematic hardening rule that works well with an uncoupled model for both uniaxial and multiaxial

ratcheting simulation. The modified Dafalias-Popov model that has been demonstrated to be one of the

68
Paper II-To be published in the International Journal of Plasticity (in press)

best uncoupled model for uniaxial ratcheting simulation (Hassan and Kyriakides [1992,1994a]) is used in

this study. Six uniaxial and ten biaxial ratcheting responses on stabilized carbon steels, acquired from

Hassan and Kyriakides [1992], Hassan et al. [1992] and Corona et al. 1996], are considered for evaluation

of the kinematic rules. The rate-independent and cyclically stable material ratcheting simulations from

the kinematic hardening rules by Armstrong and Frederick [1966], Phillips [1972,1979], Tseng and Lee

[1983], Voyiadjis and Sivakumar [1991,1994], Kaneko [1981,1984], Xia and Ellyin [1994,1997],

Chaboche [1991] and Ohno-Wang [1993] are evaluated. Many of these rules has demonstrated success in

simulating some specific type of ratcheting experiment. In this paper, however, their performances are

evaluated against a broad set of ratcheting experiments.

The original Armstrong and Frederick [1966] rule performs very well for all the constant pressure

biaxial ratcheting responses when the parameter C is determined from one of these experiments. But this

rule fails to simulate the ratcheting responses for bow-tie histories.

The Voyiadjis and Sivakumar [1991] rule is a combination of the Phillips (Phillips and Tang [1972],

Phillips and Lee [1979]) and Tseng-Lee [1983] rules. None of these three rules perform satisfactorily for

the set of biaxial experiments considered. In reality, when the center of the yield surface translates in the

stress-rate direction, the yield surface changes its shape too (Phillips and his co-workers

[1972,1979,1984]). Thus any effort to use a hardening rule that dictates the yield surface translation in the

stress-rate direction is expected to be futile for ratcheting simulations, unless, the change of the shape of

yield surface is also incorporated in the modeling.

The Kaneko [1981,1984] rule is a modification over the Ziegler [1959] rule. It has been shown by

Hassan et al. [1992] that the Ziegler rule ceases to simulate biaxial ratcheting strains after the first cycle

for the loading history in Fig. 1b. The Kaneko rule predicts negative ratcheting for this history due to the

addition of the prestress path dependence. The Xia and Ellyin [1994,1997] rule, on the other hand,

69
Paper II-To be published in the International Journal of Plasticity (in press)

overpredicts all the biaxial ratcheting cases due to the influence of the Mroz [1967] rule.

The Chaboche [1991] kinematic hardening rule, with three decomposed terms, when employed in

the modified Dafalias-Popov model (Hassan and Kyriakides [1992,1994a]) performs the best for the

whole set of ratcheting responses considered. This rule also captures the transient ratcheting rates at

initial cycles and the constant ratcheting rates at higher cycles as observed in material ratcheting

responses of stabilized carbon steels.

The ratcheting simulations by the Ohno-Wang rule, which is a variation of the Chaboche rule, are

reasonable for small number of cycles, but predicts shakedown and subsequent reversal of ratcheting for

higher number of cycles in the reverse bow-tie loading case (Fig. 9).

In an effort to obtain a constitutive model to simulate both uniaxial and multiaxial ratcheting, the

modified Dafalias-Popov uncoupled model recognizes the essential difference in the modeling of uniaxial

and multiaxial ratcheting and separates the two cases by uncoupling the plastic modulus calculation from

the kinematic hardening rule. As a result, with a suitable kinematic hardening rule (the Chaboche rule in

this study) whose parameters are determined from a biaxial ratcheting experiment, this uncoupled model

manifest success in simulating all the ratcheting responses considered. The improvement in simulation by

this model is found to be significant when the simulations are compared to those from many well known

coupled models (see Bari and Hassan [2000a]).

Since the yield surface undergoes shape changes during plastic deformation (Phillips and his co-

workers [1972,1979,1984], and others), the normals to the actual yield surfaces are different from the

ones predicted by any models that use the von-Mises surface. This would result in differences between

the actual and the predicted ratcheting strains, unless it is compensated for by an artificial movement of

the idealized yield surface through a kinematic hardening rule (Corona et al. [1996]). Most of the rules

70
Paper II-To be published in the International Journal of Plasticity (in press)

which are subsets of the Ziegler, Mroz and Phillips directions can not perform this compensation

effectively. The Armstrong-Frederick rule with one parameter, calibrated from a biaxial ratcheting

experiment, performs the compensation reasonably for similar biaxial experiments. However, this rule

fails when the loading history is changed significantly (for instance, the bow-tie case) from the calibrated

experiment. On the other hand, the Chaboche rule that combines three Armstrong-Frederick rules

performs the required compensation more effectively in disparate loading cases, when its six parameters

are calibrated from a biaxial ratcheting experiment. However, the lack of physical meaning of the

parameters of the Chaboche rule makes the parameter determination scheme less definitive. Further

validation of this modeling scheme using different nonproportional ratcheting experiments is needed in

order to verify its generality. The authors strongly feel that, in order to achieve a more general and robust

ratcheting simulation model, the incorporation of yield surface shape change into the modeling scheme is

likely to be a more rewarding approach.

Acknowledgment - The financial supports from the Center for Nuclear Power Plant Structures, Equipment
and Piping and the Department of Civil Engineering at North Carolina State University are gratefully
acknowledged.

NOMENCLATURE

a = deviatoric backstress tensor



da = incremental deviatoric backstress tensor

b = deviatoric bounding surface center tensor

db = incremental deviatoric bounding surface center tensor

s = deviatoric stress tensor

ds = incremental deviatoric stress tensor

s = deviatoric stress tensor at bounding surface image point

n = unit normal to yield surface at current stress point

m = unit normal along ( s s ) direction

dp = d p = magnitude of plastic strain increment tensor

t = history or time
E = Youngs modulus
H = generalized plastic modulus

71
Paper II-To be published in the International Journal of Plasticity (in press)

Eop = plastic modulus of bounding surface


Ebp = modulus for translation of bounding surface center
N = number of loading cycles
= backstress tensor

d = incremental backstress tensor

= bounding surface center tensor

= strain tensor

= distance of current stress state from bounding surface image point
in = distance of recent yield state from corresponding bounding surface image point
p
d = plastic strain increment tensor

x = axial strain
xc = amplitude of axial strain cycle
xp = maximum axial strain in a cycle
= circumferential strain
p = maximum circumferential strain in a cycle
= stress tensor

d = incremental stress tensor

= stress tensor at the bounding surface image point

o = size of yield surface
b = size of bounding surface
x = total axial stress
xa = amplitude of axial stress cycle
xm = mean of axial stress cycle
= circumferential stress
a = amplitude of circumferential stress cycle
m = mean of circumferential stress cycle

REFERENCES

Armstrong, P.J. and Frederick, C.O. (1966) A Mathematical Representation of the Multiaxial Bausch-

inger Effect. CEGB Report No. RD/B/N 731.

Bari, S. and Hassan. T. (2000a) Anatomy of Coupled Constitutive Models for Ratcheting Simulation.

International Journal of Plasticity, Vol 16, pp. 381-409.

Bari, S. and Hassan. T. (2000b) An Advancement in Cyclic Plasticity Modeling for Multiaxial Ratcheting

Simulation. Accepted for publication in the International Journal of Plasticity, Sept. 2000.

72
Paper II-To be published in the International Journal of Plasticity (in press)

Chaboche, J.L., Dang-Van, K. and Cordier, G. (1979) Modelization of the Strain Memory Effect on the

Cyclic Hardening of 316 Stainless Steel. Proceedings of the 5th International Conference on SMiRT,

Div. L, Berlin, Germany.

Chaboche, J.L. (1991) On Some Modifications of Kinematic Hardening to Improve the Description of

Ratcheting Effects. International Journal of Plasticity, Vol 7, pp. 661-678.

Chaboche, J.L. (1994) Modeling of Ratchetting: Evaluation of Various Approaches. European Journal of

Mechanics, A/Solids, Vol 13, pp. 501-518.

Corona, E., Hassan, T. and Kyriakides, S. (1996) On the Performance of Kinematic Hardening Rules in

Predicting a Class of Biaxial Ratcheting Histories. International Journal of Plasticity, Vol 12, pp. 117-

145.

Dafalias, Y.F. and Popov, E.P. (1976) Plastic Internal Variables Formalism of Cyclic Plasticity. Journal of

Applied Mechanics, Vol 43, pp. 645-650.

Delobelle, P., Robinet, P. and Bocher, L. (1995) Experimental Study and Phenomenological Modelization

of Ratcheting Under Uniaxial and Biaxial Loading On an Austenitic Stainless Steel. Internatinal Jour-

nal of Plasticity, Vol 11, pp. 295-330.

Drucker, D.C. and Palgen, L. (1981) On stress-Strain Relations Suitable for Cyclic and Other Loadings.

Journal of Applied Mechanics, Vol 48, pp. 479-485.

Guionnet, C. (1992) Modeling of Ratcheting in Biaxial Experiments. Journal of Engineering Materials

and Technology, Vol 114, pp. 56-62.

Hassan, T. and Kyriakides, S. (1992) Ratcheting in Cyclic Plasticity, Part I: Uniaxial Behavior. Interna-

tional Journal of Plasticity, Vol 8, p. 91-116.

Hassan, T., Corona, E. and Kyriakides, S. (1992) Ratcheting in Cyclic Plasticity, Part II: Multiaxial

Behavior. International Journal of Plasticity, Vol 8, p. 117-146.

73
Paper II-To be published in the International Journal of Plasticity (in press)

Hassan, T. and Kyriakides, S. (1994a) Ratcheting of Cyclically Hardening and Softening Materials, Part

I: Uniaxial Behavior. International Journal of Plasticity, Vol 10, pp. 149-184.

Hassan, T. and Kyriakides, S. (1994b) Ratcheting of Cyclically Hardening and Softening Materials, Part

II: Multiaxial Behavior. International Journal of Plasticity, Vol 10, pp. 185-212.

Jiang, Y. and Sehitoglu, H. (1996) Modeling of Cyclic Ratchetting Plasticity, Part I: Development of Con-

stitutive Relations. Journal of Applied Mechanics, Vol 63, pp. 720-725.

Kaneko, K. (1981) Proposition of New Translation Rule in Kinematic Hardening. Bulletin of the JSME,

Vol 24, pp. 9-14.

Kaneko, K. (1984) Development of a New Practical Plastic Constitutive Model for Anisotropic Metals

After Various Preloading. Bulletin of the JSME, Vol 27, pp. 2687-2693.

McDowell, D.L. (1995) Stress State Dependence of Cyclic Ratcheting Behavior of Two Rail Steels.

International Journal of Plasticity, Vol 11, pp. 397-421.

Mroz, Z. (1967) On the Description of Anisotropic Work Hardening. Journal of the Mechanics and Phys-

ics of Solids, Vol 15, pp. 163-175.

Ohno, N. and Wang, J.-D. (1993) Kinematic Hardening Rules with Critical State of Dynamic Recovery,

Part I: Formulations and Basic Features for Ratcheting Behavior. International Journal of Plasticity,

Vol 9, pp. 375-390.

Phillips, A. and Tang, J.-L. (1972) The Effect of Loading Path on the Yield Surface at Elevated Tempera-

tures. International Journal of Solids and Structures, Vol. 8, pp. 463-474.

Phillips, A. and Lee, C.W. (1979) Yield Surfaces and Loading Surfaces. Experiments and Recommenda-

tions. International Journal of Solids and Structures, Vol. 15, pp. 715-729.

Phillips, A. and Lu, W.-Y. (1984) An Experimental Investigation of Yield Surfaces And Loading Surfaces

of Pure Aluminum With Stress-Controlled And Strain-Controlled Paths of Loading. Journal of Engi-

neering Materials and Technology, Trans. ASME, Vol. 106, pp. 349-354.

74
Paper II-To be published in the International Journal of Plasticity (in press)

Tseng, N.T. and Lee, G.C. (1983) Simple Plasticity Model of the Two-Surface Type. ASCE Journal of

Engineering Mechanics, Vol 109, pp. 795-810.

Voyiadjis, G.Z. and Sivakumar, S.M. (1991) A Robust Kinematic Hardening Rule for Cyclic Plasticity

with Ratcheting Effects, Part I: Theoretical Formulation. Acta Mechanica, Vol 90, pp. 105-123.

Voyiadjis, G.Z. and Sivakumar, S.M. (1994) A Robust Kinematic Hardening Rule for Cyclic Plasticity

with Ratcheting Effects, Part II: Application to Nonproportional Loading Cases. Acta Mechanica, Vol

107, pp. 117-136.

Voyiadjis, G.Z. and Basuroychowdhury, I.N. (1998) A Plasticity Model for Multiaxial Cyclic Loading

and Ratchetting. Acta Mechanica, Vol 126, pp. 19-35.

Xia, Z. and Ellyin, F. (1994) Biaxial Ratcheting Under Strain or Stress-Controlled Axial Cycling With

Constant Hoop Stress. Journal of Applied Mechanics, Vol 61, pp. 422-428.

Xia, Z. and Ellyin, F. (1997) A Constitutive Model with Capability to Simulate Complex Multiaxial

Ratcheting Behaviour of Materials. International Journal of Plasticity, Vol 13, pp. 127-142.

Yoshida, F., Tajima, N., Ikegami, K. and Shiratori, E. (1978) Plastic Theory of Mechanical Ratcheting.

Bulletin of the JSME, Vol 21, pp. 389-397.

Ziegler, H. (1959) A Modification of Paragers Hardening Rule. Quarterly of Applied Mechanics, Vol 17,

pp. 55-65.

75
CHAPTER 4

IMPROVEMENT IN COUPLED CONSTITUTIVE MODELS FOR


RATCHETING SIMULATIONS

Many researchers have conducted studies in order to improve the


multiaxial ratcheting simulations from the coupled models. Most of these studies have
incorporated new parameters or directional properties in the hardening rules, which are
effective only in multiaxial loading conditions. This chapter investigates four such
relatively recent coupled models in terms of their ratcheting simulations against the same
loading set considered in the previous chapters. This study also proposes an kinematic
hardening rule in the framework of the Chaboche model for improved ratcheting
simulations. A journal paper titled "An Advancement in Cyclic Plasticity Modeling for
Multiaxial Ratcheting Simulation" is written from this part of the study, which is
accepted for publication in the International Journal of Plasticity. This paper is enclosed
as paper III in this chapter. Readers are referred to this paper for details of the study
conducted in this chapter.

76
Paper III-To be published in the International Journal of Plasticity (in press)

An advancement in cyclic plasticity modeling for multiaxial


ratcheting simulation
Shafiqul Bari and Tasnim Hassan*
Center for Nuclear Power Plant Structures, Equipment and Piping
Department of Civil Engineering, North Carolina State University,
Raleigh, NC 27695-7908; e-mail: thassan@eos.ncsu.edu

Keywords: Cyclic plasticity, multiaxial ratcheting, constitutive modeling


Shortened Title: An advancement in modeling for multiaxial ratcheting

ABSTRACT
In a search for a constitutive model for ratcheting simulations, the models by Chaboche, Ohno-

Wang, McDowell, Jiang-Sehitoglu, Voyiadjis-Basuroychowdhury and AbdelKarim-Ohno are evaluated

against a set of uniaxial and biaxial ratcheting responses. With the assumption of invariant shape of the

yield surface during plastic loading, the ratcheting simulations for uniaxial loading are primarily a

function of the plastic modulus calculation, whereas the simulations for multiaxial loading are sensitive

to the kinematic hardening rule of a model. This characteristic of the above mentioned models is

elaborated in this paper. It is demonstrated that if all parameters of the kinematic hardening rule are

determined from uniaxial responses only, these parameters primarily enable a better plastic modulus

calculation. However, in this case the role of the kinematic hardening rule in representing the ratcheting

responses for multiaxial loading is under-appreciated. This realization motivated many researchers to

incorporate multiaxial load dependent terms or parameters into the kinematic hardening rule. This paper

evaluates some of these modified rules and finds that none is general enough to simulate the ratcheting

responses consistently for the experiments considered. A modified kinematic hardening rule is proposed

using the idea of Delobelle and his coworkers in the framework of the Chaboche model. This new rule

introduces only one multiaxial load dependent parameter to the Chaboche model, but performs the best in

simulating all the ratcheting responses considered.

* The corresponding author.

77
Paper III-To be published in the International Journal of Plasticity (in press)

I. INTRODUCTION

An earlier paper by the authors evaluates several coupled cyclic plasticity models in terms of their

ability to simulate a set of uniaxial and multiaxial ratcheting responses (Bari and Hassan, 2000a). In these

models, the plastic modulus calculation is coupled with the kinematic hardening rule through the

consistency condition. Among the coupled models evaluated, the Chaboche (1991) and the Ohno-Wang

(1993) models show promise. Both of these models demonstrate success in simulating ratcheting for

uniaxial loading cases, but consistently overpredict ratcheting for multiaxial loading cases.

Such performance by the models is related to an inherent feature of modelingsimulations of

ratcheting responses in uniaxial loading primarily depend on the plastic modulus, whereas the

simulations in multiaxial loading depend on both the plastic modulus and the kinematic hardening rule,

with significant influence from the latter. This feature is elaborated through Fig. 1. For a uniaxial load

increment AB, in Fig. 1a, the normal direction, n , at the stress point B remains parallel to that at A

irrespective of the kinematic hardening rule adopted in a model. This is also true for an increment CD

during reverse loading. As a result, during uniaxial loading cycles OBDBD...., direction and magnitude

f
of the term ------

in the flow rule (Eq. 2 in section III) remain unchanged and the plastic strain increments

become a function of the plastic modulus (H) only. Therefore, simulations of uniaxial ratcheting

responses depend entirely on the accuracy of the plastic modulus calculation of a model.

This feature in multiaxial loading is demonstrated for a biaxial loading history OACDCD.....,

obtained by superimposing a constant 2 stress to a stress cycle along 1 (Fig. 1b), which is a small

deviation from the uniaxial loading cycle in Fig. 1a. During a biaxial loading increment BC (Fig. 1b),

normal direction at C changes from that at B. Similar changes occur continuously throughout the loading

history. It is clearly observed in Fig. 1 that normal directions in biaxial loading are significantly different

from those in uniaxial loading. This change in normal direction during a biaxial loading results from the

78
Paper III-To be published in the International Journal of Plasticity (in press)

1 1
n
n

n C n
B
d d

A B

d d
E n

O 2 O A 2
C
D
n
D
n

(a) (b)

Fig. 1. Yield surface translation and normal directions during (a) uniaxial and (b) biaxial loading
cycles.

shift of the stress point along the yield surface when it translates. Hence, the normal direction in

multiaxial loading is a function of the direction and magnitude of the yield surface translation, which, in

turn, is dictated by the kinematic hardening rule of a model. Different kinematic hardening rule produces

different translation direction and, thereby, greatly varied normal directions for a given stress history. As

a result, the simulation of ratcheting responses under multiaxial loading by a plasticity model depends

significantly on its kinematic hardening rule.

In coupled models, the plastic modulus (H) is calculated using the kinematic hardening rule and the

consistency condition. Usually, the parameters of a kinematic hardening rule are determined from

uniaxial loading responses. These parameters are, in effect, calibrated to produce a better representation

of the plastic modulus only, and consequently, fall short in predicting the representative yield surface

translation and subsequent normal directions. As a result, ratcheting simulations by many coupled models

are rarely good in multiaxial loading cases (Bari and Hassan, 2000a).

79
Paper III-To be published in the International Journal of Plasticity (in press)

One solution to this problem can be achieved by uncoupling the plastic modulus calculation from

the kinematic hardening rule and, thereby, employing the kinematic hardening rule only to dictate the

yield surface translation. In this methodology, parameters for the plastic modulus calculation are

determined from uniaxial responses and parameters for the kinematic hardening rule are determined

using multiaxial responses (Hassan et al., 1992). The authors have demonstrated reasonable success from

such a modeling scheme in simulating ratcheting responses for a set of uniaxial and biaxial experiments

(see Bari and Hassan, 2000b).

A solution to the problem for the coupled models has been attempted by incorporating multiaxial

terms and parameters into the Chaboche or the Ohno-Wang model (McDowell, 1995, Jiang and

Sehitoglu, 1996a, Voyiadjis and Basuroychowdhury, 1998, and AbdelKarim and Ohno, 2000). Most of

these modified models have simulated the ratcheting responses from tension-shear biaxial loading cycles

and rolling/sliding contact problems (see also McDowell, 1997, Jiang and Sehitoglu, 1996b, 1999, Ekh et

al., 2000, Ringsberg, 2000). In this study these models are evaluated using a different set of ratcheting

responses. This study also proposes an improved kinematic hardening rule by incorporating only one

multiaxial parameter to the Chaboche model.

II. EXPERIMENTAL RESPONSES STUDIED

Model performance in this study is evaluated with respect to the same set of ratcheting responses

used in the earlier works by the authors (Bari and Hassan, 2000a,b). This data set on cyclically stabilized

carbon steels is acquired from Hassan and Kyriakides (1992), Hassan et al. (1992) and Corona et al.

(1996), and is briefly discussed below.

The loading histories of the experiment set are shown in Fig. 2. The history in Fig. 2a, which results

in axial strain ratcheting, involves axial stress cycles with a mean stress. Readers are referred to Fig. 7 in

Hassan and Kyriakides (1992) for demonstration of a typical uniaxial ratcheting response. Data from

80
Paper III-To be published in the International Journal of Plasticity (in press)


xc xc
xc
x a a
xa
xm m
m m
t
x x x
(a) (b) (c) (d)
Fig. 2. Loading histories studied; (a) Uniaxial stress cycle, (b) Axial strain cycle with constant
internal pressure, (c) biaxial bow-tie cycle, (d) biaxial reverse bow-tie cycle.

uniaxial loading experiments are presented in Figs. 3a and 3b, where the peak axial strain (xp) in each

cycle is plotted against the number of cycles, N. The biaxial loading history in Fig. 2b involves axial

strain cycles in the presence of a constant internal pressure. This loading history results in circumferential

strain ratcheting as demonstrated in Fig. 1 of Hassan et al. (1992). Ratcheting responses with this biaxial

loading history are shown in Figs. 3c and 3d, where the peak circumferential strain (p) in each cycle is

plotted against the number of cycles, N. The bow-tie and reverse bow-tie loading histories, shown in Figs.

2c and 2d, also result in circumferential strain ratcheting as demonstrated in Figs. 10 and 11 of Corona et

al. (1996). Response data from the bow-tie histories are shown in Figs. 3e and 3f. An important feature of

the bow-tie histories is that they have resemblance to the loading histories observed in piping components

(Corona et al., 1996). Hence, a success in simulating these responses by a model would imply its

suitability for analysis and design of ratcheting in piping components. In this study, the ratcheting

responses from the uniaxial loading experiments are referred to as uniaxial ratcheting, and those from the

biaxial loading experiments are referred to as biaxial ratcheting.

81
Paper III-To be published in the International Journal of Plasticity (in press)

III. CYCLIC PLASTICITY MODELS

Basic components of the rate-independent plasticity models evaluated are:

12
von-Mises yield criterion: f ( ) = 3--- ( s a ) ( s a ) = 0 , (1)
2

1 f f
flow rule: d
p
= ---- ------ d ------ , (2)
H

p p
kinematic hardening rule: da = g ( , , a, d, d , etc ) , (3)

p
where, is the stress tensor, is the plastic strain tensor, s is the deviatoric stress tensor, is the

current center of the yield surface, a is the current center of the yield surface in the deviatoric space, 0 is

the size of the yield surface (constant for a cyclically stable material), and H is the plastic modulus. Also,

indicates the MacCauley bracket ( a = 0 if a < 0 ; a = a if a 0 ) and the inner product

a b = a ij b ij . The study is limited to the cyclically stable material responses.


In coupled models, the plastic modulus H is evaluated using the consistency condition, f = 0 , the

kinematic hardening rule (Eq. 3), the flow rule (Eq. 2), and the yield criterion (Eq. 1) (see Chaboche,

1986). From the point of view of multiaxial ratcheting simulation, the primary distinction between

different models lies in the kinematic hardening rule. All of the models studied are modifications over

either the Chaboche (1991) or the Ohno-Wang (1993) rules, which are briefly reviewed below. The

kinematic hardening rules proposed by McDowell (1995), Jiang and Sehitoglu (1996a), Voyiadjis and

Basuroychowdhury (1998), and AbdelKarim and Ohno (2000) are then evaluated. Finally, a new

kinematic hardening rule is proposed in the framework of the Chaboche (1991, 1994) model and

improvement in simulations using this rule is demonstrated.

82
Paper III-To be published in the International Journal of Plasticity (in press)

III.1. Chaboche Model

Chaboche (1991, 1994) proposes a decomposed nonlinear kinematic hardening rule in the form:

4
2
d a i ,
p
da = da i = --- C d i a i dp , for i = 1,2,3 and
3 i
i=1

ai
da i = --- C i d i a i 1 ------------- dp ,
2 p
for i = 4 (4)
3 f ( i )

p 2 p p
12
3 12
where dp = d = --- d d and f ( i ) = --- a i a i .
3 2

As can be observed in Eq. 4, the first three decomposed rules are the same as the Armstrong-Frederick

(1966) hardening rule. The fourth rule contains a threshold level of backstress, a4, that makes the

dynamic recovery term inactive within the threshold (Chaboche, 1991). Outside the threshold, the fourth

rule evolves according to the Armstrong-Frederick rule. The role of each decomposed rule is discussed

elaborately in Bari and Hassan (2000a), where it is also demonstrated that the Chaboche model performs

well in simulating the uniaxial ratcheting responses (see Figs. 3a and 3b). However, this model

overpredicts ratcheting strains for all biaxial loading responses considered in the study (see Figs. 3c to

3f). The parameters used in the simulations are:

0 = 18.8 ksi, E = 26300 ksi, = 0.302 (common to all the models)

C1-4 = 60000, 3228, 455, 15000 (ksi); 1-4 = 20000, 400, 11, 5000; a4 = 5 ksi.

All the parameters in this model are determined using uniaxial responses (or, in effect, are chosen for

better plastic modulus calculation). These parameters can not produce representative yield surface normal

directions (or plastic strain rate directions) and hence, fail to simulate the biaxial ratcheting responses

considered.

83
Paper III-To be published in the International Journal of Plasticity (in press)

xp xp
4 4
CS 1026 CS 1026 xa = 33.28 ksi
xa = 32.0 ksi xm = 9.14 ksi xm = 6.5 ksi
(%) (%)
x
xa
3 3 xm xa= 32.12 ksi
t

xm = 6.52 ksi
2 2
xa = 30.35 ksi
xm = 4.18 ksi
1 1 xa = 28.29 ksi
Experiment
Chaboche
Ohno-Wang
0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(a) (b)
p p
4 4
CS 1026 CS 1026
(%) m = 9.65 ksi (%) xc = 0.50 %

3 (1) xc = 0.40 % 3 (1) m = 4.88 ksi


(2) xc = 0.50 % (2) m = 7.32 ksi
(3) xc = 0.65 % (3) m = 9.65 ksi
(4) m =14.54 ksi

2 xc 2
(3)
(2) m (4) (3)
x (2)
1 (1) 1
Experiment (1)
Chaboche
Ohno-Wang
0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(c) (d)
p p
4 4
CS 1026 CS 1018
(%) a = 2.36 ksi (%) a = 2.36 ksi
xc = 0.50 % xc = 0.50 %
3 3
(1) m = 9.65 ksi (1) m = 9.65 ksi
(2) m = 14.54 ksi (2) m = 14.54 ksi

2 xc 2
(2)
a (2) xc
(1)
m a
(1)
1 x 1 m
Experiment
Chaboche x
Ohno-Wang
0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(e) (f)

Fig. 3. Ratcheting predictions by the Chaboche and the Ohno-Wang models. (a,b) Axial strain at
positive stress peaks from uniaxial cycles, (c,d) circumferential strain peaks from axial strain cycles
with constant pressure, (e,f) circumferential strain peaks from bow-tie and reverse bow-tie cycles.
Experiments are from Hassan and Kyriakides (1992), Hassan et al. (1992) and Corona et al. (1996).

84
Paper III-To be published in the International Journal of Plasticity (in press)

III.2. Ohno-Wang Model

Ohno and Wang (1993) also introduce thresholds to control evolution of decomposed kinematic

hardening rules. Each decomposed rule stops evolving outside of its threshold, Ci/i. The Ohno-Wang

kinematic hardening rule has the form:

mi
M a i f ( i )
da i = --- C i d i a i d -------------
-------------
2
d a i ,
p p
da = - . (5)
3 f ( i ) C i i
i=1

where, f ( i ) is the von-Mises function as defined in Eq. 1. All the parameters in this model are

determined using responses under uniaxial loading. Hence, the uniaxial ratcheting simulations from the

model are quite good and comparable to those from the Chaboche model (Figs. 3a,b). The parameters

used in the simulations with the Ohno-Wang model (M = 12) are:

0 = 18.8 ksi, E = 26300 ksi, = 0.302 (common to all the models)

C1-12 = 31940, 36214, 2520, 376, 11021, 4551, 3475, 2196, 857, 247, 98, 200 (ksi);

1-12 = 45203, 13944, 7728, 4955, 3692, 2135, 1230, 585, 295, 119, 50, 20;

mi = 0.45, i = 1, 12

A detailed discussion of the Ohno-Wang model and its parameter determination scheme can be found in

Bari and Hassan (2000a).

The simulations by the Ohno-Wang model for the biaxial loading cases are shown in Figs. 3c to 3f.

Note in these figures that the amount by which ratcheting is overpredicted by this model is generally

reduced compared to the simulations by the Chaboche model. This is mainly caused by the term

a
d p -------------
i that replaces dp (plastic strain increment magnitude) of the Armstrong-Frederick (1966)
f ( )

i

rule. This new term is instrumental in introducing some effects of multiaxial loading and thus improves

85
Paper III-To be published in the International Journal of Plasticity (in press)

the models multiaxial ratcheting simulation. Moreover, it does not affect uniaxial ratcheting simulation,

as it reduces to dp for uniaxial loading. Although the nonlinear nature of the biaxial ratcheting responses

is captured well by the Ohno-Wang model (Figs. 3c to 3f), overpredictions still persist for all biaxial

loading cases presented. Similar to the Chaboche model, as all of the parameters of this model are

determined using only uniaxial responses, these parameters fall short in simulating the multiaxial

ratcheting responses considered.

III.3. McDowell Model

In an effort to improve the capability of the Ohno-Wang (1993) model for multiaxial ratcheting

simulations, McDowell (1995) modifies the exponents mi in Eq. 5 as follows:

ai Bi p
m i = A i n' -------------
, ---------- ( s a ) .
d 3
where n' = --------
= (6)
f ( i ) dp 2 0

The exponents mi, a constant in the Ohno-Wang rule, is proposed to be dependent on the noncoaxiality of

the plastic strain rate and the backstress in this model. Also, note that the expression for the exponents mi

(Eq. 6) include the constants Bi, which can be calibrated using multiaxial ratcheting responses to

influence the multiaxial ratcheting simulations without affecting the uniaxial simulations.

a
For uniaxial loading, the exponents mi reduce to the constants Ai as the terms n' -------------
i become
f ( i )

unity. As a result, for uniaxial loading, this rule reduces to the Ohno-Wang rule, which is also validated

by the ratcheting simulations in Fig. 4a. As the parameters of the Ohno-Wang model are determined from

uniaxial responses, these parameters can also be used with the McDowell model (with Ai = mi = 0.45 for

i = 1, 12). When the multiaxial parameters Bi are zeros, for all 12 decomposed terms, the biaxial

ratcheting simulations by the McDowell model exactly match the simulations by the Ohno-Wang model.

With increasing Bi, the McDowell model predicts higher ratcheting strains than the Ohno-Wang model

86
Paper III-To be published in the International Journal of Plasticity (in press)

xp p
4 4
CS 1026 CS 1026
(%) xa = 32.0 ksi (%) m = 9.65 ksi
xc = 0.50 %
3 3
xm = 9.14 ksi xm = 6.52 ksi
xc

m
2 2
x

xm = 4.18 ksi
1 Experiment 1 Experiment
Ohno-Wang Ohno-Wang
McDowell, Bi = 2 McDowell, Bi = 2
Jiang-Sehitoglu Jiang-Sehitoglu
0 0
0 10 20 30 40 50 0 10 20 30 40 50

(a) N (b) N

p p
4 4
CS 1026 CS 1018
(%) a = 2.36 ksi (%) a = 2.36 ksi
xc = 0.50 % xc = 0.50 %
3 m = 9.65 ksi 3 m = 9.65 ksi

xc xc

a a
2 2
m m

x x
1 Experiment 1 Experiment
Ohno-Wang Ohno-Wang
McDowell, Bi = 2 McDowell, Bi = 2
Jiang-Sehitoglu Jiang-Sehitoglu
0 0
0 10 20 30 40 50 0 10 20 30 40 50

(c) N (d) N

Fig. 4. Ratcheting predictions by the McDowell and the Jiang-Sehitoglu models. (a) Axial strain at
positive stress peaks from uniaxial cycles, (b) circumferential strain peaks from axial strain cycles
with constant pressure, (c,d) circumferential strain peaks from bow-tie and reverse bow-tie cycles.
Experiments are from Hassan and Kyriakides (1992), Hassan et al. (1992) and Corona et al. (1996).

for biaxial loading cases, as shown in Figs. 4b to 4d for Bi = 2. These results demonstrate that the

incorporation of a multiaxiality effect into the exponents mi according to the McDowell (1995) model

(Eq. 6) is not conducive to better biaxial ratcheting simulations for the experiments considered in this

study.

87
Paper III-To be published in the International Journal of Plasticity (in press)

III.4. Jiang-Sehitoglu Model

Jiang and Sehitoglu (1996a) proposed another variation of the Ohno-Wang (1993) model, with

exponents mi dependent on the noncoaxiality of plastic strain rate and the backstress, as follows:

mi
M
2 f ( i )
d a i ,
p
da = da i = --- C i d i -------------
- a i dp . (7)

i=1
3 C i i

a p
m i = A 0i 2 i , = ---------- ( s a ) .
d 3
where, n' -------------
f ( )
n' = --------
dp 2 0
i

Similar to the McDowell (1995) model, the exponents mi in this model assume constant values (mi = A0i)

and the model reduces to the Ohno-Wang model for uniaxial loading. Hence, the Ohno-Wang model

parameters can also be used with this model and consequently, its uniaxial ratcheting simulations remain

the same as the Ohno-Wang model (Fig. 4a). The Jiang-Sehitoglu (1996a) model, however, overpredicts

the ratcheting responses for the biaxial loading cases as shown in Figs. 4b to 4d. As this model replaces

p a
i - in the Ohno-Wang model by dp (compare Eq. 5 with Eq. 7), it overpredicts the
the term d -------------
f ( )
i

ratcheting responses in a manner similar to the Chaboche (1991) model for the biaxial loading histories

considered.

III.5. Voyiadjis-Basuroychowdhury Model

Voyiadjis-Basuroychowdhury (1998) and Basuroychowdhury-Voyiadjis (1998) attempt to improve

the Chaboche (1991) kinematic hardening rule by adding the direction of the stress rate ds . Their rule is

used in a slightly modified form in this study as follows:

M
2
d a i ,
p
da = da i = --- C d i a i dp + i ---- ds , for i = 1,2,3 and
3 i m
i=1

ai
da i = --- C i d i a i 1 ------------- dp + i ---- ds ,
2 p
for i = 4. (8)
3 f ( i ) m

88
Paper III-To be published in the International Journal of Plasticity (in press)

where i is a material parameter, is the distance of the current stress point from the bounding surface

along the direction of ds , m is the chord length of the bounding surface along the same direction. Due to

the dependence of and m on the stress rate, the numerical solution for Eq. (8) becomes computationally

intensive because of the iterations involved when strains are prescribed as loading. Therefore, and m

along the ( s s ) direction (Mroz, 1967) are used in this study as follows:

3 12 2 3
= --- ( s s ) ( s s ) and m = --- --- ( s b ) ( s s ) . (9)
2 2

The deviatoric stress point s ( ) is the image of s ( ) on the bounding surface and b is the current center

of the bounding surface in the deviatoric space (see Fig. 5 for different parameters in the stress space

where is the current center of the bounding surface). Intersections of the yield and the bounding

2 Yield
Bound Surface
Eo P Bounding
b Surface

o
o



p 1

m
n


n

(a) (b)
Fig. 5. Definition of parameters and m in the modified Voyiadjis-Basuroychowdhury
model; (a) in uniaxial -p space, (b) in biaxial 1-2 stress space.

surfaces are avoided by translating the bounding surface according to Dafalias and Popov (1976):

p
E ds n 3 (s s) 3 (s a)
db = da dM , d M = 1 ------0- --------------
- , = --- ----------
- , and n = --- ----------
- , (10)
H n 2 2 0

89
Paper III-To be published in the International Journal of Plasticity (in press)

where E0p defines the plastic modulus at the bounding surface, is the unit vector along ( s s ) and n is

the unit normal to the yield surface at the current stress point. It is important to note here that the purpose

of incorporation of the bounding surface in this model is not to calculate the plastic modulus using the

concept of Dafalias and Popov (1976), but to improve the multiaxiality effect in the modeling.

All parameters of this model which are to be determined using the uniaxial responses are kept the

same as in the Chaboche model (see Section III.1). As stated by Voyiadjis and Basuroychowdhury

(1998), and also observed in this study, parameters 1 and 2 do not have much influence on ratcheting

simulations. Therefore, the same values (1, 2 = 0.15) as used by Voyiadjis and Basuroychowdhury

(1998) are also used in this study. Ratcheting simulations from Eq. 8, for two sets of 3, 4 values (=

0.025, 0.05), are compared to the simulations from the Chaboche model in Fig. 6. As seen in this figure,

the modified Voyiadjis and Basuroychowdhury rule (Eq. 8) is performing inadequately for one uniaxial

experiment (Fig. 6a). For the biaxial loading cases, the simulations by this rule are not improving

compared to the simulations by the Chaboche model (Figs. 6b to 6d).

III.6 AbdelKarim-Ohno Model

AbdelKarim and Ohno (2000) propose a new kinematic hardening rule combining the initial

(multilinear) version of the Ohno-Wang (1993) and the Armstrong-Frederick (1966) rules as follows:

M
da = d a i ,
2
da i = --- C i d
3
p
i i ai dp i ai H { 3--2- ai ai ( C i i ) 2 } d i (11)

i=1

p ai
where, d i = d -------------
- i dp ,
Ci i

and H stands for the Heaviside step function ( H ( a ) = 0 if a < 0 ; H ( a ) = 1 if a 0 ). For i = 0 (i = 1,

M), Eq. 11 reduces to the multilinear Ohno-Wang (1993) rule,

90
Paper III-To be published in the International Journal of Plasticity (in press)

xp p
4 4
CS 1026 x CS 1026
(%) xa = 32.0 ksi xa (%) xc = 0.50 %
xm
xm = 6.52 ksi m = 9.65 Ksi
t
3 3

xc

m
2 2
x

1 Experiment 1 Experiment
VB (3,4 =.025) VB (3,4 =.025)
VB (3,4 =.05) VB (3,4 =.05)
Chaboche Chaboche
0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(a) (b)
p p
4 4
CS 1026 CS 1018
(%) a = 2.36 Ksi (%) a = 2.36 Ksi
xc = 0.50 % xc = 0.50 %
3 m = 9.65 Ksi 3 m = 9.65 Ksi

xc xc

a a
2 2
m m

x x
1 Experiment 1 Experiment
VB (3,4 =.025) VB (3,4 =.025)
VB (3,4 =.05) VB (3,4 =.05)
Chaboche Chaboche
0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(c) (d)

Fig. 6. Ratcheting predictions by the modified Voyiadjis-Basuroychowdhury (VB) and the


Chaboche models. (a) Axial strain at positive stress peaks from uniaxial cycles, (b) circumferential
strain peaks from axial strain cycles with constant pressure, (c,d) circumferential strain peaks from
bow-tie and reverse bow-tie cycles. Experiments are from Hassan and Kyriakides (1992), Hassan et
al. (1992) and Corona et al. (1996).
a
2
da i = --- C i d
3
p
i ai H { 3--2- ai ai ( C i i ) 2 } d p -------------
C
i - (12)
i i

and, for i = 1 (i = 1, M), Eq. 11 reduces to the Armstrong-Frederick rule,

2 p
da i = --- C i d
3
i ai dp , (13)

as d i in the third term in Eq. 11 never becomes active. The multilinear version of the Ohno-Wang

model (Eq. 12) simulates shakedown (cessation of ratcheting) for uniaxial loading and some ratcheting

91
Paper III-To be published in the International Journal of Plasticity (in press)

for biaxial loading conditions (see Figs. 7a and 7b for i = 0). On the other hand, superposition of several

Armstrong-Frederick rules (Eq. 13) overpredicts ratcheting for both uniaxial and biaxial loading (see

Figs. 7a and 7b for i = 1). The same set of Ohno-Wang model parameters (except mi), presented in

Section III.2, are used in the above two simulations.

AbdelKarim and Ohno (2000) propose to blend the Armstrong-Frederick and multilinear Ohno-

Wang rules through a small value of i (between 0 and 1). For i = 0.07, this model produces a better

biaxial ratcheting simulation, especially in reproducing the recorded steady rate of ratcheting (see Fig.

7b). With the same value of i (= 0.07) this model, however, underpredicts the ratcheting in uniaxial

loading cases (Fig. 7a). The simulations by this model for other experiments, with i = 0.07, are shown in

Figs. 7c to 7f. It is observed from these figures that the simulations for all of the biaxial loading cases are

improved compared to the simulations by the Ohno-Wang model (compare Figs. 3c, 3e and 3f with Figs.

7d, 7e and 7f, respectively). However, the uniaxial ratcheting simulations by this model have deteriorated

(underpredicting) compared to the simulations by the Ohno-Wang model (compare Figs. 7c and 3a). It

should be noted that AbdelKarim and Ohno (2000) have verified their model using uniaxial experiments

with low rates of ratcheting only (see Fig. 6 in the reference).

If the AbdelKarim-Ohno (Eq. 11) and the Ohno-Wang (Eq. 5) kinematic hardening rules are

scrutinized, it is realized that both these models are quite comparable. For instance, i = 0 in Eq. 11 and

m i = in Eq. 5 yield multilinear version of the Ohno-Wang model, and thus simulate shakedown for

uniaxial loading and some ratcheting for biaxial loading (see Figs. 7a and 7b). Whereas i = 1 in Eq. 11

and m i = 0 in Eq. 5 yield overpredictions of ratcheting for both uniaxial and multiaxial loading. For the

latter case, the kinematic rules in both Eqs. 5 and 11 behave like the Armstrong-Frederick rule. The

parameters i and mi influence both the uniaxial and biaxial ratcheting responses. If the parameter i is

determined from a biaxial ratcheting response (as is done in this study), Eq. 11 yields a good

92
Paper III-To be published in the International Journal of Plasticity (in press)

xp p
4 4
CS 1026 CS 1026
(%)
Experiment xa = 32.0 ksi (%) m = 9.65 ksi
Abdel-Ohno xm = 6.52 ksi xc = 0.50 %
i = 1
3 3

i = 1
i = 0.07
2 x 2
i = 0
xa xc
xm
t m
1 i = 0.07 1
x
Experiment
i = 0 Abdel-Ohno
0 0
0 10 20 30 40 50 0 10 20 30 40 50

(a) N (b) N

xp p
4 4
CS 1026 CS 1026
xa = 32.0 ksi xm = 9.14 ksi Experiment m = 9.65 Ksi
(%) (%)
i = 0.07 Abdel-Ohno i = 0.07
x
3 3 (1) xc = 0.40 %
xa xm = 6.52 ksi (2) xc = 0.50 %
xm
(3) xc = 0.65 %
t
xm = 4.18 ksi
2 2 (3)
(2)

xc
1 1 (1)
m
x

0 0
0 10 20 30 40 50 0 10 20 30 40 50

(c) N (d) N

p p
4 4
CS 1026 CS 1018
(%) Experiment a = 2.36 Ksi (%) Experiment a = 2.36 Ksi
Abdel-Ohno xc = 0.50 % Abdel-Ohno xc = 0.50 %
i = 0.07 i = 0.07
3 3
(1) m = 9.65 Ksi (1) m = 9.65 Ksi
(2) m = 14.54 Ksi (2) m = 14.54 Ksi
(2) (2)
2 2
(1)
xc xc
a (1)
a
1 1
m m
x
x
0 0
0 10 20 30 40 50 0 10 20 30 40 50

(e) N (f) N

Fig. 7. Ratcheting predictions by the AbdelKarim-Ohno model. (a,c) Axial strain at positive stress
peaks from uniaxial cycles, (b,d) circumferential strain peaks from axial strain cycles with constant
pressure, (e,f) circumferential strain peaks from bow-tie and reverse bow-tie cycles. Experiments
are from Hassan and Kyriakides (1992), Hassan et al. (1992) and Corona et al. (1996).

93
Paper III-To be published in the International Journal of Plasticity (in press)

representation of the yield surface normal directions, but a somewhat poor estimation of the plastic

modulus. As a result, the biaxial ratcheting simulations for all cases are improved, whereas the uniaxial

ratcheting simulations are deteriorated (Fig. 7). On the other hand, if the parameter mi is determined from

a uniaxial ratcheting response, Eq. 5 produces good uniaxial ratcheting simulations but a consistent

overprediction for all the biaxial ratcheting responses (Fig. 3).

IV. AN IMPROVED COUPLED MODEL

The basis of the improved model is derived from the model by Burlet and Cailletaud (1986), who

modifies the radial evanescence term in the Armstrong-Frederick (1966) hardening rule as follows:

2 p 2 f 3 (s a)
da = --- Cd ( a n )n dp , where n = --- ------ = --- ----------
- . (14)
3 3 2 0

The plastic modulus expression obtained with this hardening rule by satisfying the consistency condition

( f = 0 ) is the same as that obtained from the Armstrong-Frederick rule. Therefore, for uniaxial loading,

these two rules produce the same simulation, which is the overprediction of ratcheting under uniaxial

loading (Bari and Hassan, 2000a). For multiaxial loading, the radial evanescence term ( ( a n )n dp ) of the

Burlet and Cailletaud rule essentially yields a tensor along the plastic strain-rate direction. As a result,

this rule behaves like the Prager (1956) rule and predicts shakedown for biaxial loading cases (see Fig. 8

for ' = 0). On the other hand, Bari and Hassan (2000a) demonstrated that the Armstrong-Frederick

(1966) rule overpredicts ratcheting for the same biaxial loading cases. To obtain ratcheting simulations

for multiaxial loading in between these two extremes (shakedown and overprediction), the idea of

Delobelle et al. (1995) is borrowed to introduce a parameter ' in Eq. 14 as follows:

da = --- Cd { 'a + ( 1 ' ) ( a n )n } dp ,


2 p
(15)
3

which yields the Burlet-Cailletaud rule for ' = 0, and the Armstrong-Frederick rule for ' = 1.

94
Paper III-To be published in the International Journal of Plasticity (in press)

The plastic modulus expression (H) obtained for this rule (Eq. 15) does not include '. This means

that ' can be calibrated using a multiaxial ratcheting response and thus can influence multiaxial

ratcheting simulations without having any bearing on the plastic modulus calculation or the uniaxial

responses. However, the plastic modulus expression obtained from Eq. 15 is also the same as that

obtained from the Armstrong-Frederick rule, which overpredicts the uniaxial ratcheting. This drawback

of the rule (Eq. 15) can be fixed by superimposing several rules in the framework of the Chaboche (1991)

model as follows:

4
da i = --- C i d i { 'a i + ( 1 ' ) ( a i n )n } dp
2
d a i ,
p
da = for i = 1,2,3 and
3
i=1

ai
da i = --- C i d i { 'a i + ( 1 ' ) ( a i n )n } 1 ------------- dp ,
2 p
for i = 4. (16)
3 f ( i )

Equation 16 becomes the Chaboche (1991) rule for ' = 1 and the superimposed version of the Burlet and

Cailletaud (1986) rule for ' = 0. The plastic modulus expression (H) obtained from this rule is

independent of ' and the same as that obtained from the Chaboche rule. Therefore, all of the parameters

of the Chaboche model (which are determined using uniaxial responses) can be used with this modified

rule (Eq. 16). The new parameter ' is completely uncoupled from the plastic modulus calculation, and its

value can be calibrated using a multiaxial ratcheting response. Figure 8 shows the ratcheting simulations

for a constant pressure biaxial experiment by this improved rule for different values of '. It is observed

from this figure that, for ' = 0, this model predicts shakedown like the Prager model and, for ' = 1, this

model overpredicts ratcheting like the Chaboche model. However, for ' = 0.18, the ratcheting simulation

by this model improves compared to the simulations by the Chaboche and the Ohno-Wang models (Fig.

8).

A similar approach is also adopted by Geyer (1995) using only two decomposed rules following the

methodology of Burlet and Cailletaud (1987). This model, however, fails to reproduce several of the

95
Paper III-To be published in the International Journal of Plasticity (in press)

p
4
CS 1026
(%) = 9.65 Ksi
xc = 0.50 %
3
xc

m
2 x

Experiment
1 = 0.18
= 0
= 1 (Chaboche)
Ohno-Wang

0
0 10 20 30 40 50

Fig 8. Ratcheting predictions by the improved rule (Eq. 16) for various ' and the Ohno-
Wang model. The experimental data is obtained from Hassan et al. (1992)

ratcheting responses considered by Portier et al. (2000). It has been demonstrated by the authors (Bari

and Hassan, 2000a) that, while superposition of at least three (Armstrong-Frederick) decomposed rules is

necessary, four decomposed rules with a threshold term in the fourth rule proposed by Chaboche (1991)

performs the best in simulating the uniaxial hysteresis loops and ratcheting responses. This framework of

the Chaboche (1991) model is adopted here with the proposed kinematic hardening rule (Eq. 16). The

improved ratcheting simulations by the new rule, with ' = 0.18 and other parameters of the Chaboche

model (see section III.1), are presented in Fig. 9. As expected, the uniaxial ratcheting simulations are the

same as those from the Chaboche model (compare Figs. 3a, 3b to Figs. 9a, 9b, respectively). The

ratcheting simulations by this new rule for all biaxial loading cases conform reasonably well with the

experimental responses (see Figs. 9c to 9f). The simulations exhibit desired nonlinear ratcheting for lower

cycles and constant ratcheting rate for higher cycles. The rates of ratcheting at higher cycles match well

with the rates observed in the experimental responses.

96
Paper III-To be published in the International Journal of Plasticity (in press)

xp x
4 4
CS 1026 CS 1026 xa = 33.28 Ksi
xa = 32.0 ksi xm = 9.14 Ksi xm = 6.5 Ksi
(%) (%)
x
3 3 xa xa= 32.12 Ksi
xm
t
xm = 6.52 Ksi
2 2
xa = 30.35 Ksi
xm = 4.18 Ksi
1 1 xa = 28.29 Ksi
Experiment
Improved model
= 0.18
0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(a) (b)
p p
4 4
CS 1026 CS 1026
(%) xc m = 9.65 Ksi (%) xc = 0.50 %

m
3 (1) xc = 0.40 % 3 (1) m = 4.88 Ksi
x (2) xc = 0.50 % (2) m = 7.32 Ksi
(3) xc = 0.65 % (3) m = 9.65 Ksi
(4) (4) m =14.54 Ksi
(3)
2 (2) 2 (3)
(3) (4) (3)
(1) (2)
(2) (2)
(1) (1)
1 1
Experiment (1)
Improved model
= 0.18
0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(c) (d)
p p
4 4
CS 1026 CS 1018
(%) a = 2.36 Ksi (%) a = 2.36 Ksi
xc = 0.50 % xc = 0.50 %
3 3
(2) (1) m = 9.65 Ksi (1) m = 9.65 Ksi
(2) m = 14.54 Ksi (2) m = 14.54 Ksi
(2) (1) xc (2)
2 2
a (2)
(1) (1) xc
m
a
1 x 1 (1)
m
Experiment
Improved model x
= 0.18
0 0
0 10 20 30 40 50 0 10 20 30 40 50

N N
(e) (f)

Fig. 9. Ratcheting predictions by the Improved model (Eq. 16). (a,b) Axial strain at positive stress
peaks from uniaxial cycles, (c,d) circumferential strain peaks from axial strain cycles with constant
pressure, (e,f) circumferential strain peaks from bow-tie and reverse bow-tie cycles. Experiments
are from Hassan and Kyriakides (1992), Hassan et al. (1992) and Corona et al. (1996).

97
Paper III-To be published in the International Journal of Plasticity (in press)

V. DISCUSSION AND CONCLUSIONS

Evaluation of several coupled models demonstrates that most of the models are not general enough

to simulate the set of uniaxial and biaxial ratcheting responses considered in this study. This drawback of

the models is believed to be related to the assumption of the yield surface shape remain invariant during

plastic loading. With this simplifying assumption, the evolution of the yield surface normal directions

(plastic strain rate directions) is influenced primarily by the kinematic hardening rule of a model. Since

different kinematic rules produce varied estimates for the yield surface normals, the multiaxial ratcheting

simulations are sensitive to the kinematic hardening rule employed in a model.

Moreover, the simulation of ratcheting in uniaxial loading depends only on the plastic modulus

calculation. In coupled models, the kinematic hardening rule defines both the plastic modulus and the

yield surface translation in stress space. If all parameters of a kinematic hardening rule are calibrated

using only uniaxial responses, this rule can accurately estimate the plastic modulus, but its role in

dictating the yield surface translation is underestimated. Consequently, multiaxial ratcheting simulations

by the rule become less representative of the experimental responses.

For example, in the Chaboche (1991) and the Ohno-Wang (1993) models, all parameters are

determined using uniaxial responses. These models perform reasonably well in simulating uniaxial

ratcheting responses (Figs. 3a and 3b), but overpredict consistently for all biaxial ratcheting responses

considered (Figs. 3c to 3f). The Ohno-Wang model, however, predicts a reduced rate of ratcheting

compared to the Chaboche model for all biaxial responses. This improvement can be attributed to the

a
term d p -------------
i , which introduces a multiaxiality effect in the model without affecting its uniaxial
f ( i )

ratcheting simulation.

Therefore, in order to achieve better simulations in both uniaxial and multiaxial cases, it is essential

for a coupled model to introduce special multiaxiality terms or parameters in the kinematic hardening

98
Paper III-To be published in the International Journal of Plasticity (in press)

rule. These terms and parameters should remain dormant under uniaxial loading (or do not take part in

the plastic modulus calculation), but play an important role in defining representative yield surface

translation and thus plastic strain rate directions. As the actual yield surface shape deforms during plastic

loading (Phillips and Tang, 1972, Phillips and Lee, 1979), the lack of exactness introduced in the

modeling through invariant yield surface shape assumption is compensated to some extent by calibrating

these parameters using multiaxial ratcheting responses.

This paper evaluates several coupled models that vie to improve their multiaxial ratcheting

simulations by introducing new parameter(s) or term(s) in the kinematic hardening rule. The models

considered are proposed by McDowell (1995), Jiang and Sehitoglu (1996a), Voyiadjis and

Basuroychowdhury (1998), and AbdelKarim and Ohno (2000). All of these models are modifications of

either the Chaboche (1991) or the Ohno-Wang (1993) model. In this paper, the performances of these

models are evaluated against a set of ten biaxial ratcheting responses.

McDowell (1995) modifies the Ohno-Wang (1993) model by allowing the exponents mi (compare

Eqs. 5 and 6) to depend on the noncoaxiality of plastic strain rate and backstress. This model also

introduces parameters Bi that influence only the multiaxial responses and have no bearing on the uniaxial

responses. McDowells modification, however, does not manifest improvement over the Ohno-Wang

model in simulating the biaxial ratcheting responses considered in this study (see Figs. 4b to 4d).

Jiang and Sehitoglu (1996a) modify the exponents mi of the Ohno-Wang model to depend on the

noncoaxiality of plastic strain rate and backstress in a slightly different manner than the McDowell

a
(1995) model (compare Eqs. 5, 6 and 7). In addition, this model repudiates the term d p -------------
i of the
f ( i )

Ohno-Wang (1993) model in favor of dp. As a result, the model overpredicts the ratcheting responses in

biaxial loading (see Figs. 4b to 4d) similar to the Chaboche (1991) model.

Voyiadjis and Basuroychowdhury (1998) attempt to improve the Chaboche (1991) model by

99
Paper III-To be published in the International Journal of Plasticity (in press)

incorporating the stress-rate direction in the kinematic hardening rule. A slightly modified form of the

Voyiadjis and Basuroychowdhury (1998) rule is evaluated in this study. The uniaxial ratcheting

simulations by this rule (Eq. 8) are inadequate, whereas the biaxial ratcheting simulations do not

demonstrate any improvement over the Chaboche (1991, 1994) rule (see Fig. 6).

AbdelKarim and Ohno (2000) proposed a new kinematic hardening rule (Eq. 11) by blending the

Armstrong-Frederick (1966) and the multilinear Ohno-Wang (1993) rules through a parameter i. As the

parameter i is determined using a biaxial ratcheting response in this study, this rule simulates all the

biaxial responses considered quite well (Figs. 7f to 7f). The new rule, however, underpredicts the uniaxial

ratcheting responses (Fig. 7c). Conversely, if i were determined from a uniaxial ratcheting response,

good simulations would have been obtained for the uniaxial responses (Ohno-AbdelKarim, 2000), but

multiaxial ratcheting responses would have been overpredicted. This behavior has been observed with the

Ohno-Wang (1993) model, where the parameters mi are determined to represent a uniaxial ratcheting

response (Fig. 3; see Bari and Hassan, 2000a for details). In case of the AbdelKarim-Ohno and the

Voyiadjis-Basuroychowdhury models, as the new terms or parameters affect both the uniaxial and

multiaxial responses it is not possible to improve the ratcheting simulations for the whole set of responses

studied.

Using the concept introduced by Delobelle et al. (1995) to appropriately blend the Armstrong-

Frederick (1966) and the Burlet and Cailletaud (1986) rules, an improved kinematic hardening rule is

proposed in the framework of the Chaboche (1991) model. This hardening rule includes a parameter '

that does not take part in the plastic modulus calculation, but produces meaningful influence on the yield

surface translation under multiaxial loading. As a result, uniaxial ratcheting simulations by this model are

exactly the same as those by the Chaboche (1991) model, regardless of the value of ' (compare Figs. 9a

and 9b to Figs. 3a and 3b, respectively). When ' is calibrated using a biaxial ratcheting experiment (see

Fig. 8), simulations by this rule manifest improvement for the biaxial ratcheting responses considered

100
Paper III-To be published in the International Journal of Plasticity (in press)

(see Figs. 9c to 9f). The authors have critically evaluated a number of coupled and uncoupled models

against the same set of uniaxial and biaxial ratcheting responses (see Bari and Hassan, 2000a,b). Among

these models, the proposed kinematic hardening rule performs the best in simulating the whole set of

responses considered. Moreover, the parameter determination scheme of this model is simple and

systematic.

The authors would like to emphasize that the calibration of the new parameter in the modified rule

using a biaxial ratcheting experiment merely tries to compensate for the lack of exactness of the model

introduced by various assumptions. As a result, applicability of the model to multiaxial loading histories

depends strongly on the ratcheting response used for the parameter calibration. Realizing this

characteristic of modeling, Corona et al. (1996) have demonstrated the challenge in simulating multiaxial

ratcheting responses. This study proposes a new kinematic hardening rule to successfully meet this

challenge. Such a success does not, by any means, imply that the new rule is general enough for

simulation of ratcheting in other multiaxial histories or structures. Further validation is needed to explore

its general purpose application. The authors, however, believe that a more generalized ratcheting

simulation model can be achieved through the incorporation of yield surface shape change (formative

hardening rule) into the modeling schemes.

Acknowledgment - The financial supports from the Center for Nuclear Power Plant Structures, Equipment
and Piping and the Department of the Army, Army Research Office, Grant No. DAAD19-99-1-0312 are
gratefully acknowledged.

101
Paper III-To be published in the International Journal of Plasticity (in press)

NOMENCLATURE

a = deviatoric backstress tensor



da = incremental deviatoric backstress tensor

b = deviatoric bounding surface center tensor

db = incremental deviatoric bounding surface center tensor

p
dp = d = magnitude of plastic strain increment tensor

m = chord length of the bounding surface
n = unit normal to yield surface at current stress point

t = history or time
s = deviatoric stress tensor

s = deviatoric stress tensor at bounding surface image point

E = Youngs modulus
Eop = plastic modulus of bounding surface
H = generalized plastic modulus
N = number of loading cycles
= backstress tensor

= bounding surface center tensor
= distance of current stress state from bounding surface image point
= strain tensor

p
d = plastic strain increment tensor

x = axial strain
xc = amplitude of axial strain cycle
xp = maximum axial strain in a cycle
= circumferential strain
p = maximum circumferential strain in a cycle
= unit normal along ( s s ) direction

= stress tensor

d = incremental stress tensor



= stress tensor at the bounding surface image point

o = size of yield surface


b = size of bounding surface
x = total axial stress
xa = amplitude of axial stress cycle
xm = mean of axial stress cycle
= circumferential stress
a = amplitude of circumferential stress cycle
m = mean of circumferential stress cycle

102
Paper III-To be published in the International Journal of Plasticity (in press)

REFERENCES

AbdelKarim, M., Ohno, N., 2000. Kinematic hardening model suitable for ratcheting with steady-state.

Int. J. Plasticity, 16, 225-240.

Armstrong, P.J., Frederick, C.O., 1966. A mathematical representation of the multiaxial bauschinger

effect. CEGB Report No. RD/B/N 731.

Bari, S., Hassan. T., 2000a. Anatomy of coupled constitutive models for ratcheting simulation. Int. J.

Plasticity, 16, 381-409.

Bari, S., Hassan. T., 2000b. Kinematic hardening rules in uncoupled modeling for multiaxial ratcheting

simulation. Int. J. Plasticity. (In press)

Basuroychowdhury, I.N., Voyiadjis, G.Z., 1998. A multiaxial cyclic plasticity model for non-proportional

loading cases. Int. J. Plasticity, 14, 855-870.

Burlet, H., Cailletaud, G., 1986. Numerical techniques for cyclic plasticity at variable temperature.

Engng. Compt., 3, 143-153

Burlet, H., Cailletaud, G., 1987. Modeling of cyclic plasticity in finite element codes. Proc. of 2nd Int.

Conf. on Constitutive Laws for Engng. Mater.: Theory and Applications, 1157-1164.

Chaboche, J.L., 1986. Time-independent constitutive theories for cyclic plasticity. Int. J. Plasticity, 2,

149-188.

Chaboche, J.L., 1991. On some modifications of kinematic hardening to improve the description of

ratcheting effects. Int. J. Plasticity, 7, 661-678.

Chaboche, J.L., 1994. Modeling of ratchetting: evaluation of various approaches. Eur. J. Mech., A/Solids,

13, 501-518.

Corona, E., Hassan, T., Kyriakides, S., 1996. On the performance of kinematic hardening rules in

predicting a class of biaxial ratcheting histories. Int. J. Plasticity, 12, 117-145.

Dafalias, Y.F., Popov, E.P., 1976. Plastic internal variables formalism of cyclic plasticity. ASME J. Appl.

103
Paper III-To be published in the International Journal of Plasticity (in press)

Mech., 43, 645-650.

Delobelle, P., Robinet, P., Bocher, L,. 1995. Experimental study and phenomenological modelization of

ratcheting under uniaxial and biaxial loading on an austenitic stainless steel. Int. J. Plasticity, 11, 295-

330.

Drucker, D.C., 1951. A more fundamental approach to stress-strain relations. Proc. of the First U.S.

National Congress of App. Mech., ASME, 487-491.

Ekh, M., Johansson, A., Thorberntsson, H., Josefson, B.L., 2000. Models for cyclic ratcheting

plasticityintegration and calibration, ASME J. Engng. Mater. Techn., 122, 49-55.

Geyer, P., 1995. On use of radial evanescence remain term in kinematic hardening. Trans. of 13th Int.

Conf. on SMiRT, II, 699-704.

Guionnet, C., 1992. Modeling of ratcheting in biaxial experiments. ASME J. Engng. Mater. Techn., 114,

56-62.

Hancell, P.J., Harvey, S.J., 1979. The use of kinematic hardening models in multi-axial cyclic plasticity.

Fatigue of Engng. Mater. Structures, 1, 271-279.

Hassan, T., Kyriakides, S., 1992. Ratcheting in cyclic plasticity, part I: uniaxial behavior. Int. J. Plasticity,

8, 91-116.

Hassan, T., Corona, E., Kyriakides, S., 1992. Ratcheting in cyclic plasticity, part II: multiaxial behavior.

Int. J. Plasticity, 8, 117-146.

Hassan, T., Kyriakides, S., 1994a. Ratcheting of cyclically hardening and softening materials, part I:

uniaxial behavior. Int. J. Plasticity, 10, 149-184.

Hassan, T., Kyriakides, S., 1994b. Ratcheting of cyclically hardening and softening materials, part II:

multiaxial behavior. Int. J. Plasticity, 10, 185-212.

Jiang, Y., Sehitoglu, H., 1996a. Modeling of cyclic ratchetting plasticity, part I: development of

constitutive relations. ASME J. App. Mech., 63, 720-725.

104
Paper III-To be published in the International Journal of Plasticity (in press)

Jiang, Y., Sehitoglu, H., 1996b. Modeling of cyclic ratchetting plasticity, part II: comparison of model

simulations with experiments. ASME J. App. Mech., 63, 726-733.

Jiang, Y., Sehitoglu, H., 1999. A model for rolling contact failure. Wear, 224, 38-49.

McDowell, D.L., 1995. Stress state dependence of cyclic ratcheting behavior of two rail steels. Int. J.

Plasticity, 11, 397-421.

McDowell, D.L., 1997. An approximate algorithm for elastic-plastic two-dimensional rolling/sliding

contact. Wear, 211, 237-246.

Mroz, Z., 1967. On the description of anisotropic work hardening. J. Mech. Phys. Solids, 15, 163-175.

Ohno, N., Abdel-Karim, M., 2000. Uniaxial ratcheting of 316FR steel at room temperaturepart II:

constitutive modeling and simulation. ASME J. App. Mech., 122, 35-41.

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. Plasticity, 9, 375-390.

Phillips, A., Tang, J.-L., 1972. The effect of loading path on the yield surface at elevated temperatures.

Int. J. Solids Structures, 8, 463-474.

Phillips, A., Lee, C.W., 1979. Yield surfaces and loading surfaces: experiments and recommendations.

Int. J. Solids Structures, 15, 715-729.

Portier, L., Calloch, S., Marquis, D., Geyer, P., 2000. Ratcheting under tension-torsion loadings:

experiments and modelling. Int. J. Plasticity, 16, 303-335.

Prager, W., 1956. A new method of analyzing stresses and strains in work hardening plastic solids.

ASME J. App. Mech., 23, 493-496.

Ringsberg, J.W., 2000. Cyclic ratcheting and failure of a pearlitic rail steel, Fatigue Fract. Engng. Mater.

Struct., in press.

Voyiadjis, G.Z., Basuroychowdhury, I.N., 1998. A plasticity model for multiaxial cyclic loading and

ratchetting. Acta Mech., 126, 19-35.

105
CHAPTER 5

RATCHETING SIMULATIONS USING FORMATIVE HARDENING


RULE

5.1. INTRODUCTION

It has been argued in the previous chapter that in order to have a generalized
ratcheting simulation model that performs well for various multiaxial loading histories,
the yield surface shape change (formative hardening rule) should be considered in cyclic
plasticity modeling. Phillips and his co-workers [1972,1979,1984] along with many other
researchers (Naghdi et. al [1959], Ivey [1961], Ikegami [1979]) have demonstrated that
the yield surface changes shape with progressive plastic loading. This change in shapes

312 C
III

I II
O
B A 11

IV
D

Experiment Yield Surface


(Phillips and Lu, 1984)
von-Mises Surface

Fig. 5.1. Comparison of experimental and von-Mises yield surfaces at different


stress points

106
results in yield surfaces that are quite different from the simple von-Mises yield surface
as shown in Fig. 5.1. Consequently, the normal directions calculated from the von-Mises
yield surface are not always representative of the experimental normal directions
(incremental plastic strain directions) observed in multiaxial loading conditions.
Although it has been demonstrated in this study that an improved representation of
normals to the von-Mises yield surface can be achieved by introducing a multiaxial
parameter into the Chaboche [1991] kinematic hardening rule, the general applicability of
this model for various multiaxial loading histories is yet to be validated. It is our
postulation that incorporating the shape change of the yield surface into cyclic plasticity

312 PR
B'

B
A' D'

E'
A D

11

Fig. 5.2. Yield surface deformation law from Phillips and Tang [1972]

modeling would result in a more robust and general ratcheting simulation model.
Development of such a model is attempted in this chapter.

Many researchers like Phillips and Tang [1972], Shiratori et. al [1979], Ortiz and
Popov [1983], Eisenberg and Yen [1981,1984], Ishikawa and Sasaki [1988], Wegener
and Schlegel [1996], Kurtyka and Zyczkowski [1996] have tried to model the deformed

107
shape of the yield surface with progressive plastic loading. Most of these models are
quite complex and numerically extensive and hence are impractical for implementation
with a cyclic plasticity model.

Phillips and Tang [1972] present experimental results of subsequent yield surfaces
of pure aluminum under tension-torsion loading histories. From these experiments, they
outline a qualitative model for deformed shapes of the yield surface in stress space. They
propose that the subsequent yield surface is the superposition of a rigid body translation
in the prestressing direction and of a deformation in the same direction as illustrated in
Fig. 5.2. The deformation reduces the width of the yield surface in the prestressing
direction in such a way that the motion of the forward part ABD of the yield surface is
less than the rear part AED (see Fig. 5.2). They have not put forward any modeling
scheme to implement this idea.

Eisenberg and Yen [1981] present a theory of multiaxial anisotropic plasticity that
is general enough in its structure to accommodate the hardening mechanism of distorted
yield surface. In their 1984 paper (Eisenberg and Yen [1984]), they present a modeling
scheme to implement the qualitative hardening model proposed by Phillips and Tang
[1972] in an effort to simulate the experimentally observed yield surface deformations.
To describe both deformation and translation of the yield surface in deviatoric stress
space, the subsequent yield state sij (see Fig. 5.3) is described by

s ij a ij + Rij = S ij (5.1)

where Sij is the initial stress state, Rij is a measure of the deformation of the yield surface
and aij is the center of the yield surface. The center of the yield surface is defined as the
intersection point of the loading direction and the hyperplane that divides the yield
surface into its forward and rear subdomains. The initial von-Mises surface is

S ij S ij = 0
3 2
2 (5.2)

where 0 is the initial yield surface size. Thus, the yielding condition becomes:

108
( s ij a ij + Rij )( s ij a ij + Rij ) = 0 .
3 2
2 (5.3)

To implement the hardening rule proposed by Phillips and Tang [1972],


Eisenberg and Yen [1984] develop a numerical scheme to determine Rij in such a way
that the subsequent yield state sij takes the form

n
a kl ) = S ij , where Tijkl = Tijkl
(n) (n) (m )
Tijkl ( s kl . (5.4)
m =1

It can be observed from Eq. 5.4 that after n loading increments Tijkl assumes 2n different
values for 2n subdomains (forward and rear parts) of the yield surface and the description

s(1)

(0)
0'(aij) s(0)
s
s(1)
b(sij) 0,0'
sij-aij ij
0 ij
Rij
a(Sij) Sij

(b)
(a)
Fig. 5.3 A schematic diagram of deformation and translation of yield
surface from Eisenberg and Yen [1984]

of the yield surface becomes quite complex and extensive numerically. Therefore, for
simulation of multiaxial cyclic loading responses, where large numbers of loading
increments are used, this model appears to be less attractive.

Ortiz and Popov [1983] put forward a distorted yield function rule that contains a
state tensor of second order. They propose the diameter J 2 of the elastic region to be a
Fourier-cosine series of the angle to the backstress direction. Because it is not possible to
construct an invariant sine of the angle to the backstress direction, the yield function is

109
only applicable to yield conditions that are symmetric to the back stress direction. This
effectively means that this yield function is not suitable for non-proportional loading.

Ishikawa and Sasaki [1988] also makes effort to describe the deformed yield
surface during plastic loading with the use of a fourth order state tensor. For the tension-
torsion space, after simplification, their model includes four anisotropic coefficients
(parameters) of the fourth order tensor whose values are to be determined from non-
proportional cyclic loading. However, they only demonstrate the suitability of their
model for proportional loading and a very special type of non-proportional loading. Also,
the model lacks guidelines for the selection of effective parameters and their
determination for general multiaxial loading condition.

Wegener and Schlegel [1996] put forward and implement a very extensive
numerical model to describe the distorted yield surface shapes observed by Phillips and
Tang [1972]. After evaluating several yielding functions with second to fourth order state
tensors, they demonstrate success in reproducing the experimentally observed yield
surfaces by using a yielding function that contains a sixth order state tensor. The Ortiz
and Popov [1983] model is a special case of this generalized yield function. Because a
sixth order tensor is used, the number of coefficients to be determined for this rule is
large and their determination requires several yield surface experiments. This model
defines a quality function whose values are to be minimized using the Monte-Carlo
simulation or other similar artificial intelligence programs. The variation of different
parameters with plastic loading does not seem to follow any general trend that can be
modeled to use this method in general multiaxial loading condition. Therefore, even
though this model demonstrates success in simulating the experimentally observed yield
surfaces, the numerical scheme is very complex, extensive, and requires many
experimental data for parameter determination. As a result, the model becomes
impractical to implement in a constitutive modeling scheme for simulation of cyclic
responses of materials and structures.

Yoshida et. al [1977] and Shiratori et. al [1979] propose a so-called equi-plastic-
strain surface as the plastic potential surface in place of the subsequent yield surfaces.

110
The locus of equal plastic strains measured from the unloading point in the stress space is
defined as the equi-plastic-strain surface. Figure 5.4 shows a qualitative diagram of the
equi-plastic-strain surface. For any loading path OSQiPi, after unloading at S, the loading
2

d p (1)
d p ( 2 )
P1
P2
d p ( n )
Q1 S
Pn Q2
Qn

F0
Fn O 1

Fig. 5.4 Qualitative equi-plastic-strain surface in Ilyushin [1954] vector stress


space

path SQi (for i = 1,2,....n) is the elastic loading stage within the subsequent yield surface
where no plastic strains accumulate. The point Qi (for i = 1,2,....n) falls on the initial yield
surface F0. If the accumulated plastic strains

( Pi )

p= d p , i = 1,2,.....n (5.5)
( Qi )

for paths QiPi are equal, then the locus of these equal plastic strain points Pi (for i =
1,2,....n) approximately forms a closed surface f called the equi-plastic-strain surface. The
subsequent yield surface Fn touches the equi-plastic-strain surface f at loading point Pn
where both the surfaces have the same normal. Therefore, from the normality rule of the
plastic strain increment, the equi-plastic-strain surface can be considered as a plastic

111
potential surface and the plastic strain increment can be derived from the principle of
maximum plastic work as follows:

d p = d ( grad f ) . (5.6)

The assumption of the equi-plastic-strain surface f as the plastic potential surface


is supported to some degree by the facts that (i) for p = 0, the equi-plastic-strain surface f0
is same as the yield surface F0 and (ii) for large p, it becomes the isotropic hardening
surface. Shiratori et. al [1979] demonstrate from a set of experiments (see Fig. 3 and 4 in
the reference) that the normal directions to the equi-plastic-strain surface at stress points
do match with the experimentally observed directions of plastic strain increments.

The main advantages of this type of potential surface over its yield surface
counterpart are that (i) the equi-plastic-strain surfaces assume much simpler shapes in the
Ilyushin [1954] stress space, and (ii) the numerical modeling of this plastic potential
surface is simpler. It also contains fewer parameters that are to be determined from
experimental responses. Yoshida et. al [1977] and Shiratori et. al [1979] have conducted
several equi-plastic-strain surface experiments on Brass, Aluminum alloy, Ni-Cr Mo steel
and Mild steels and show that the equi-plastic-strain surfaces match experimentally
observed shapes of the yield surfaces for p 0 ,i.e., a rounded corner in the loading side
and a flattening corner at the rear side. The shape of the subsequent equi-plastic-strain
surfaces gradually approach a concentric isotropic hardening hypersphere with increase
of the plastic strains as shown qualitatively in Fig. 5.5. Also, the equi-plastic-strain
surfaces generated for loading history OSBC are found to be symmetric with respect to
the preloading direction 1 .

Shiratori et. al [1979] propose an expression for the equi-plastic-strain surfaces in


the moving stress space i of the fixed Ilyushin space i (Ilyushin[1954]). This model
has only four parameters that are determined from equi-plastic-strain surface experiments
and several others which are determined from a uniaxial ratcheting experiment. This
model is demonstrated to perform well in simulating the stress and plastic strain
responses from a number of non-proportional loading histories. It also simulates

112
ratcheting responses successfully for two types of cyclic loading histories; axial stress
cycles on steady internal pressure similar to Fig. 1.1 and axial stress (or shear stress)
cycles on steady torsion (or tension). However, this model has some limitations,
especially in simulations of the ratcheting responses under general multiaxial loading
conditions as follows.

2
2

d p
1
C B S

O 1
f0

f3 f1

f2

Fig. 5.5. Subsequent equi-plastic-strain surfaces and plastic strain


increment directions

(i) It uses the equivalent stress-strain idealization to represent the general


multiaxial stress-strain states. This idealization undermines the effect of the loading
history in ratcheting responses under general multiaxial loading.

(ii) This model ignores the Bauchinger effect in uniaxial cyclic loading that is
critical for uniaxial ratcheting responses.

(iii) It assumes that line about which the equi-plastic-strain surfaces are
symmetric always passes through the origin in Ilyushin stress space. In all the
experiments presented by Shiratori et. al [1979], the unloading occurs after sufficient
113
hardening that takes the unloading point very close to the isotropic surface (see Fig. 4 and
Fig. 18 in the reference). As a result, the symmetry lines appear to pass through the origin
in all these experiments. However, in case of an unloading point where the equi-plastic-
strain surface is much away from the isotropic surface, this symmetry line may not pass
through the origin. For example, if the unloading occurs at points S or C in Fig. 5.5, the
symmetry line is expected to pass through or in the vicinity of the origin, as the normals
at those points also pass through or in the vicinity of the origin. However, if unloading
starts at some point like B, the assumed symmetry line is much different from the normal
or the incremental plastic strain direction at B. Therefore, at unloading point B, the
normal directions to the yield surface and to the equi-plastic-strain surface would be
different and the basic premise of using the equi-plastic-strain surface as a plastic
potential surface would be violated.

(iv) The methodology for the determination of plastic modulus and other
parameters presented in Shiratori et. al [1979] can not be applied in combined stress
strain controlled loading conditions.

Regardless of these limitations, the Shiratori et. al [1979] model shows promise
and seems to be simpler in taking into account the effect of the distorted yield surfaces in
cyclic plasticity modeling. Therefore, in this study, efforts have been made to present a
constitutive model that includes the equi-plastic-strain surface and try to address the
above mentioned limitations in order to improve the simulations of ratcheting responses
in general multiaxial loading. Two types of equi-plastic-strain surface models are studied
within the uncoupled modeling scheme of Dafalias and Popov [1976]. Performances of
these models are evaluated against a set of constant pressure biaxial ratcheting
experiments (Fig 1.1b). These modeling schemes and numerical results are presented
below.

114
5.2. PLASTIC POTENTIAL THEORY BASED ON THE EQUI-
PLASTIC-STRAIN SURFACE

5.2.1. Coordinate system

The equi-plastic-strain surfaces are represented in the five-dimensional Cartesian


deviatoric stress and plastic strain vector spaces as introduced by Ilyushin [1954]. Stress
and plastic strain vector components in the spaces i and i (i = 1,2,....5) are given as
follows:

1 = 32 S11 , 2 = 2
3
( S11 + 2 S 22 ), 3 = 3S12 , 4 = 3S 23 , 5 = 3S 32 ; (5.7)

1 = 11 , 2 = ( 11 + 2 22 ) , 3 = 12 , 4 = 23 , 5 = 31
p p p 1 p p p 2 p p 2 p p 2 p
3 3 3 3
(5.8)

where S ij and ij are deviatoric stress and plastic strain tensor components in the three-
p

dimensional Cartesian coordinate system respectively. In the Ilyushin space, the stress
and the plastic strain vectors are given as

5 5
= i e i and = i ei
p p
(5.9)
i =1 i =1

where e i are the unit vectors along the Ilyushin axes.

5.2.2. Flow rule

Upon considering the equi-plastic-strain surface as the plastic potential surface,


the plastic strain increment can be derived from the normality rule and the principal of
maximum plastic work as follows:

1
d p
= d n G = ( nG d )nG (5.10)
H

f
where nG = , f is the equi-plastic-strain surface and H is the plastic modulus.
f

115
5.2.3. Model I

(a) Equi-plastic-strain surface

Although both Yoshida et. al [1977] and Shiratori et. al [1979] propose two
similar type of expressions of the equi-plastic-strain surface in the moving stress space
i of the fixed space i , they do not provide any basis for their selection. On the other
hand, Kurtyka and Zyczkowski [1996] present a general form for defining any
anisotropically deformed surface in the Ilyushin vector space. In this study, a special case
of this general form is adopted to describe the shape of the equi-plastic-strain surface.
With the assumption that the equi-plastic-strain surface is symmetric with respect to a
direction PR , the expression for the equi-plastic-strain surface is proposed in this study
as follows:

( 1 0 *) 2 5
( k vY ) 2
+ =1 (5.11)
R1 + 2d 1 ( 1 0 *) d1
2 2 2
k =2 Rk

where R1 = ( N M ) / 2 , Rk = B C * , d 1 = N 0 * R1 .

All the parameters and variables in Eq. 5.11 are shown in Fig. 5.6. As adopted by
Shiratori et. al [1979], the expressions for 0 * and C * are assumed to depend on the

accumulation of plastic strain from unloading and ( p + q ) as follows:

m1( + dp) m2 ( + dp)


0 * = C1( p + q ) e * and C * = C 2 ( p + q )e * (5.12)

where p = v b + B N , q = M - v b + B and is * the accumulated plastic strain

after unloading. It should be noted that the equi-plastic-strain surface in this model
gradually approaches a bounding surface for very large * .

(b) Plastic modulus H

It has been demonstrated in this study (chapter 3) that the plastic modulus H
calculation scheme by Dafalias and Popov [1976] is successful in simulating the
116
k (k = 2,...,5)
k
C*

vb
nG

0*
b Rk

R1 PR
d1
1
vy b
1

q
M
p
N
Bounding surface

Fig. 5.6. Equi-plastic-strain surface I in the moving stress space i of the


Ilyushin space

ratcheting responses when used with appropriate hardening rule. Therefore, this scheme
is also adopted in this study to calculate the plastic modulus as follows:

a
H = E0 + h( ) , h=
p
(5.13)
in
1 + b( in ) m
2 B

where = and B = nG + b . The in is the initial after unloading, E0


p

is the plastic modulus of the bounding surface, B is the image point of on the

bounding surface and b is the bounding surface center. The parameters a, b and m are
determined from a uniaxial ratcheting test as described in the paper II.

117
(c) Hardening rule

It has been observed in experiments from Shiratori et. al [1979] that p and q

decrease with the increase in plastic strain value. With them, 0 * and C * also
decrease according to Eq. 5.12 and the equi-plastic-strain surface gradually approaches
the bounding surface. When p and q become zero, the equi-plastic-strain surface

converges with the bounding surface. This hardening mechanism of the equi-plastic-
strain surface is modeled in this study as follows:

d M = K q + dvb , d N = K p + dvb , dvb = 0 dp .


p
(5.14)

It should be observed in Eq. 5.14 that the bounding surface does not grow in size,
but translates along the negative PR direction. The magnitude of the scalar K is
determined from the consistency condition at each loading increment.

(d) Symmetry direction

The experimental data from Shiratori et. al [1979] shows that the equi-plastic-
strain surfaces appear to be symmetric with respect to a line that passes through the origin
when the unloading points are very close to the bounding surface. However, there is no
experimental data available showing the shapes of the equi-plastic-strain surface when
unloading occurs far away from the bounding surface. In order to hold the basic premise
that the equi-plastic-strain surface can be used as a plastic potential surface, the line of
symmetry for the equi-plastic-strain surface has to be along the plastic strain increment
(or normal to the equi-plastic-strain surface) at the unloading stress point. Therefore, in
this study, it is assumed that

PR = nG . (5.15)

(e) Ratcheting simulations

The performance of the above mentioned modeling scheme is evaluated against


some biaxial ratcheting responses from Hassan and Kyriakides [1992] on stabilized

118
carbon steels where cyclic axial strains are applied on straight pipe with steady internal
pressure (see loading history (b) in Fig. 1.1). The material constants C1, C2, m1 and m2
that define the characteristic values for 0 * and C * are supposed to be determined
from equi-plastic-surface experiments in multiaxial loading condition according to
Shiratori et. al [1979]. As there is no experimental data available on the shape of the equi-
plastic-strain surface in stabilized carbon steels, these parameters are selected from a
constant pressure biaxial ratcheting experiment in this study. Fig. 5.7 shows the
simulation of ratcheting responses by this model along with the experimental responses
under biaxial loading condition. The parameters used are:

4
p

CS 1026
=====m = 9.65 ksi

3.5 x (1) xc = 0.40 %


(2) xc = 0.50 %
3 m
2.5 x Experiment
Model I
2

(2)
1.5

(1)
1
(2)

0.5
(1)

0
0 10 20 30 40
N

Fig. 5.7. Biaxial ratcheting predictions by model I and experiments from Hassan
and Kyriakides [1992].

C1 = 0.35, C2 = 0.40, m1 = 310 and m2 = 210.

The simulated accumulation of the hoop strains from this model tends to stabilize after
some ratcheting during a couple of initial cycles. The predicted ratcheting responses for

119
both axial strain amplitudes with the same internal pressure behave in similar manner.
This deviation of ratcheting responses from experimental responses can be attributed to
the assumption that the line of symmetry is along the plastic strain increment at the
unloading point (Eq. 5.15). With this assumption, this symmetry direction gradually
flattens until it becomes parallel to the 1 direction when the hoop strain component of
plastic strain increment becomes zero. As a result, shakedown of ratcheting responses
occurs.

5.2.4. Model II

In order to gain a better perspective on the use of equi-plastic-strain surface with


the uncoupled modeling scheme for ratcheting simulations, another variation of the equi-
plastic-strain surface and corresponding hardening rule is examined in this study. This
modeling scheme (model II) differs from the model I in three key aspects: the symmetry
direction, the plastic modulus H calculation, and the bounding surface evolution.

(a) Symmetry direction

Although it is argued in model I that the symmetry direction may not necessarily
pass through the origin of the Ilyushin vector stress space in all loading histories, it has
been observed in the experiments by Shiratori et. al [1979] that the equi-plastic-strain
surfaces tend be symmetric with respect to a line passing through the vicinity of the
origin. It should be noted that, for the ratcheting experiments considered in this study to
evaluate models I and II, enough plastic strains are accumulated in each cycle to bring the
unloading points sufficiently close to the bounding surface. For these experiments, the
direction of the plastic strain increments (normals to the equi-plastic-strain surface) at the
unloading points also tend to pass through the origin. Therefore, in model II, the
symmetry direction is considered to be the direction of the line passing through the origin
and the unloading point (OS direction in Fig. 5.8).

(b) Equi-plastic-strain surface

In model II, the equi-plastic-strain surface is proposed to be symmetric with

120
k (k= 2,...,5)
Dafalias-Popov k
bounding surface C *

nG
Rk
*
0
a PR
R1
F = B + b b d1 1
S
0
1

q
M
p
N

Isotropic bounding
surface

Fig. 5.8. Equi-plastic-strain surface II in the moving stress space i of the


Ilyushin space

respect to the direction PR , as follows:

( 1 0 *) 2 5
( k ) 2
+ =1 (5.16)
R1 + 2d1 ( 1 0 *) d 1
2 2 2
k =2 Rk

where R1 = ( N M ) / 2 , Rk = F C * , d 1 = N 0 * R1 , F = B + b .

All the parameters and variables in Eq. 5.16 are illustrated in Fig. 5.8. The new parameter
b is the increase in size of the isotropic bounding surface during plastic loading from its
original size b . The expressions for 0 * and C * are assumed to depend on the

accumulation of plastic strain after preloading and ( p + q ) as follows:

m1( + dp) m2 ( + dp)


0 * = C1( p + q ) e * and C * = C2 ( p + q )e * (5.17)

121
where p = F N , q = M + F and * is the accumulated plastic strain after

unloading.

(c) Plastic modulus H

The Dafalias and Popov [1976] plastic modulus calculation scheme (Eq. 5.13) is
also adopted in this model. However, an additional Dafalias-Popov type kinematic
bounding surface is introduced in this model as shown in Fig. 5.8 where, b is the center
of the kinematic bounding surface. The image point of on the kinematic bounding
surface is used as a measure of in Eq. 5.13. The same parameter determination scheme

as used in Eq. 5.13 is adopted in model II as well.

(d) Hardening rule

In this model, the equi-plastic-strain surface gradually approaches the concentric


isotropic bounding surface, which also grows in all direction with plastic loading. When
p and q become zero, the equi-plastic-strain surface converges with the isotropic
bounding surface. This hardening mechanism of the equi-plastic-strain surface is
modeled as follows:

d M = K q + dvb , d N = K p + dvb , d b = 0 dp .
p
(5.18)

The kinematic bounding surface center b grows according to the evolution model
proposed by Seyed-Ranjbari [1986] (also Hassan and Kyriakides [1994]) where
a = 0 nG in this model. The readers are referred to Eqs. 6 and 7 of the enclosed
paper II for expressions of this evolution. The magnitude of the scalar K in Eq. 5.18 is
determined from the consistency condition at each loading increment.

(e) Ratcheting simulations

The performance of the modeling scheme II is also evaluated against the same set
of biaxial ratcheting responses as used in evaluating model I. The same guidelines to
determine material constants C1, C2, m1 and m2 as adopted in model I are applied in this

122
4

p
CS 1026
=====m = 9.65 ksi


3.5 x (1) xc = 0.40 %
(2) xc = 0.50 %
3 m (3) xc = 0.65 %

2.5 x
Experiment
Model II
2
(2)
1.5 (3)
(1)
1

0.5

0
0 10 20 30 40
N

Fig. 5.9. Biaxial ratcheting predictions by model II and experiments from Hassan
and Kyriakides [1992].

model. Fig. 5.9 shows the simulation of ratcheting responses by model II with
parameters:

C1 = 0.36, C2 = 0.42, m1 = 310 and m2 = 210.

The simulated ratcheting responses for all axial strain amplitudes show an almost
constant rate of ratcheting. For initial cycles, ratcheting simulations match the
experimental responses reasonably, but fail to represent the ratcheting rates observed in
the experiments for higher cycles (see Fig. 5.9). In addition, the changes in the ratcheting
rates observed with different axial strain amplitudes in the experiments of Fig. 5.9 are not
reproduced by model II.

123
5.3. CONCLUSION AND DISCUSSIONS

Pioneering experimental studies on the yielding criterion of metals by Phillips and


his co-workers [1972,1979,1984] have demonstrated that the yield surfaces undergo both
translation and distortion during plastic loading. These ensuing shapes of the yield
surfaces are quite complex and different from the shapes of the von-Mises yield surfaces
(see Fig. 5.1). As a result, cyclic plasticity models with the von-Mises yield surface
would rarely perform well for general multiaxial loading conditions. The limitation
introduced by the assumption of the von-Mises yielding in ratcheting simulation models
can be compensated to some degree by calibrating some of the kinematic hardening
parameters using multiaxial ratcheting responses as has been demonstrated in paper III.
However, the generality of this methodology is yet to be validated. Therefore, in the final
stage of this study, efforts have been made to evaluate the prospect and potentials of
incorporating the anisotropic deformation of the yield surface in cyclic plasticity
modeling in order improve the ratcheting simulations in multiaxial, nonproportional
loading conditions.

From the experimental trends observed for the shapes of the yield surface, Phillips
and Tang [1972] put forward a qualitative model of the translation and distortion of the
yield surfaces during plastic loading. Ortiz and Popov [1983], Eisenberg and Yen
[1981,1984], Ishikawa and Sasaki [1988], and Wegener and Schlegel [1996] propose
analytical and numerical modeling schemes to implement the yield surface shape change
observations of Phillips and Tang [1972]. After careful scrutiny of these modeling
schemes, it has been observed that most of these models use complex and numerically
extensive schemes with higher order state tensors. As a result, these models become less
attractive (in some cases, impractical) for implementation with a cyclic plasticity model.

Shiratori and his co-workers (Yoshida et. al [1977], Shiratori et. al [1979]), on the
other hand, propose a so-called equi-plastic-strain surface that can be used as the plastic
potential surface instead of the yield surface generally used in cyclic plasticity modeling.
This equi-plastic-strain surface can effectively incorporate the effect of distorted yield
surface in the flow rule as the yield surface and the equi-plastic-strain surface have the

124
same normal direction at the current stress point. Experimental results of Yoshida et. al
[1977] and Shiratori et. al [1979] also demonstrate the basis behind the applicability of
using the equi-plastic-strain surface as the plastic potential surface for incremental plastic
strains. Moreover, the formulation for the equi-plastic-strain surface is simpler compared
to other deformed yield surface formulations.

Towards developing a generalized ratcheting simulation model, the use of the


equi-plastic-strain surface as the plastic potential surface is examined in this study. Some
key limitations of the Shiratori et. al [1979] model are identified. Two modeling schemes
(models I and II) are developed to eliminate these limitations of the Shiratori et. al [1979]
model. These new models are evaluated against a set of biaxial ratcheting experiments on
stabilized carbon steels from Hassan and Kyriakides [1992]. Model I predicts
nonlinearity in the ratcheting rates for initial cycles, but shakes down at higher cycles
(see Fig. 5.7). The model II, on the other hand, simulates almost constant rates of
ratcheting and fails to represent the change in the ratcheting rate with variation in the
axial strain amplitudes (see Fig. 5.9).

There are no experimental results available on the equi-plastic-strain surface


evolution for stabilized carbon steels in multiaxial cyclic loading conditions. Some very
critical aspects of modeling like the symmetry direction and the evolution of the isotropic
bound need to be verified from experimental observations. This study demonstrates the
basic methodology and promise in incorporating the equi-plastic-strain surface as the
plastic potential surface into the cyclic plasticity modeling. However, some experimental
studies on the shapes of the equi-plastic-strain surface, especially when unloading occurs
far away from the bounding surface, are essential in order to have a more representative
modeling scheme.

125
CHAPTER 6

CONCLUSIONS AND RECOMMENDATIONS

6.1. CONCLUSIONS

All of the enclosed journal papers (papers I, II and III) written from this study
contain conclusions from the corresponding part of the study. Conclusions drawn from
study of constitutive modeling with anisotropic deformation of yield surface are
presented in chapter 5. A summary of all these conclusions is presented in the following.

Coupled models ( paper I)

For uniaxial stress-controlled history (Fig. 1.1a), the linear kinematic hardening
[Prager, 1956] and multilinear models produce closed hysteresis loops and, hence, cannot
simulate the ratcheting response. The Prager model overpredicts ratcheting strains during
the initial cycles, which are followed by shakedown after a few more cycles for all biaxial
loading cases.

The nonlinear kinematic hardening model (Armstrong and Frederick [1966])


overpredicts the ratcheting responses for both uniaxial and biaxial loading histories
considered in this study.

The original Chaboche [1986] model with three or four decomposed rules, with at
least one rule as linear hardening, improves the ratcheting simulation for the initial
cycles, but always stabilizes to shakedown with persistent cycling. Incorporation of a
slight nonlinearity into the linear rule improves the model's capability to simulate steady
rate of ratcheting and prevents shakedown.

Chaboche [1991] model introduces the concept of a threshold backstress into the
fourth nonlinear kinematic hardening rule. This model incorporates a linear segment in
the uniaxial hysteresis curve within the threshold level and, thus, makes the hysteresis

126
curve stiffer and reduces the rate of uniaxial ratcheting compared to the models without
any threshold. However, this model overpredicts all of the biaxial ratcheting responses
considered.

The Ohno-Wang [1993] model is also composed of a number of the Armstrong


and Frederick type kinematic hardening rules. This model reproduces uniaxial hysteresis
loops by using a number of almost linear segments and, thereby, needs a large number of
kinematic hardening rules to simulate the hysteresis loop. The hysteresis loop simulations
however are very impressive. In addition, simulations by the Ohno-Wang model for all of
the uniaxial ratcheting experiments are reasonably good. However, the Ohno-Wang
[1993] model overpredicts all of the biaxial ratcheting responses considered in this study.

The Guionnet [1992] model grossly overpredicts the ratcheting responses for
uniaxial loading cases. However, this model performs impressively in simulating the
circumferential strain ratcheting for the biaxial experiments with constant pressure. When
the loading changes to the biaxial bow-tie histories, this model fails to simulate the
ratcheting responses.

The limitations introduced by the idealized yield surface shape and other
simplified assumptions in the coupled cyclic plasticity modeling can be compensated if
some model parameters are calibrated using multiaxial ratcheting experiments or if the
hardening rule is uncoupled from the plastic modulus.

Uncoupled models (paper II)

This study evaluates eight different kinematic hardening rules in order to identify
or develop a kinematic hardening rule that works well with the Dafalias-Popov uncoupled
model for both uniaxial and multiaxial ratcheting simulation.

The original Armstrong and Frederick [1966] rule performs very well for all the
constant pressure biaxial ratcheting responses, but fails for the ratcheting responses in
bow-tie loading histories.

The Voyiadjis and Sivakumar [1991] rule is a combination of the Phillips

127
[1972,1979] and Tseng-Lee [1983] rules. None of these three rules perform satisfactorily
for the set of biaxial ratcheting experiments considered.

The Kaneko [1981,1984] rule is a modification over the Ziegler [1959] rule. This
rule predicts negative ratcheting for the biaxial history of Fig. 1.1b due to the addition of
the prestress path dependence.

The Xia and Ellyin [1994,1997] rule overpredicts all of the biaxial ratcheting
responses due to the influence of the Mroz [1967] rule.

The Chaboche [1991] kinematic hardening rule, with three decomposed terms,
performs the best for the whole set of ratcheting responses considered. This rule also
captures the transient ratcheting rates at initial cycles and the constant ratcheting rates at
higher cycles as observed in material ratcheting responses of stabilized carbon steels.

The ratcheting simulations by the Ohno-Wang [1993] rule, which is a variation of


the Chaboche rule, are reasonable for small number of cycles. But this model predicts
shakedown and subsequent reversal of ratcheting for higher number of cycles in the
reverse bow-tie loading case.

Improved coupled models (paper III)

In cyclic plasticity modeling, the simulations of ratcheting responses in uniaxial


loading primarily depend on the plastic modulus, whereas the simulations in multiaxial
loading depend on both the plastic modulus and the kinematic hardening rule, with
significant influence from the latter.

McDowell [1995] modifies the Ohno-Wang [1993] model by introducing some


parameters that depend on the noncoaxiality of plastic strain rate and backstress.
McDowell's modification, however, does not manifest improvement over the Ohno-Wang
model in simulating the biaxial ratcheting responses considered in this study.

Jiang and Sehitoglu [1996] modify the exponents of the Ohno-Wang model to
depend on the noncoaxiality of plastic strain rate and backstress in a slightly different
manner than the McDowell model. In addition, this model uses the equivalent plastic
128
strain increments in the hardening rules. As a result, the model overpredicts the ratcheting
responses in biaxial loading similar to the Chaboche [1991] model.

Voyiadjis and Basuroychowdhury [1998] attempts to improve the Chaboche


model by incorporating the stress-rate direction in the kinematic hardening rule. The
uniaxial ratcheting simulations by this rule are inadequate, whereas the biaxial ratcheting
simulations do not demonstrate any improvement over the Chaboche rule.

AbdelKarim and Ohno [2000] proposed a new kinematic hardening rule by


blending the Armstrong-Frederick [1966] and the multilinear Ohno-Wang [1993] rules.
As the multiaxiality parameter is determined using a biaxial ratcheting response in this
study, this rule simulates all the biaxial responses considered quite well. This rule,
however, underpredicts the uniaxial ratcheting responses.

The improved coupled model proposed in this study by appropriately blending the
Armstrong-Frederick [1966] and the Burlet and Cailletaud [1986] rules in the framework
of the Chaboche [1991] model manifest significant improvement in simulating ratcheting
responses for all of the biaxial loading cases considered. The study has critically
evaluated a number of coupled and uncoupled models against the same set of uniaxial
and biaxial ratcheting responses. Among these models, the proposed kinematic hardening
rule performs the best in simulating the whole set of responses considered. Moreover, the
parameter determination scheme of this model is simple and systematic.

Formative Hardening rules with anisotropic deformation of yield surface

Most of the models that try to formulate the anisotropic deformation of the yield
surface use complex and numerically extensive schemes with higher order state tensors.
As a result, these models become less attractive (in some cases, impractical) for
implementation in a cyclic plasticity model.

Shiratori and his co-workers (Yoshida et. al [1977], Shiratori et. al [1979]), on the
other hand, propose a so-called equi-plastic-strain surface that can be used as the plastic
potential surface. The formulation for the equi-plastic-strain surface is simpler compared
to other deformed yield surface formulations.
129
This study demonstrates the basic methodology and promise in incorporating the
equi-plastic-strain surface as the plastic potential surface into the cyclic plasticity
modeling. More experimental results on the shapes of the equi-plastic-strain surface,
especially when unloading occurs far away from the bounding surfaces are essential in
order to have a representative modeling scheme.

6.2. RECOMMENDATIONS FOR FUTURE RESEARCH

From the knowledge gained in this study by investigating a number of cyclic


plasticity models in terms of their ratcheting simulation performance, the following
recommendations are made for future studies:

(i) The single multiaxiality parameter introduced in the proposed coupled model
of this study is assumed to be constant for all loading conditions. However, this
parameter should include the effect of plastic loading history in order to predict
ratcheting response in general multiaxial loading cases. Studies can be conducted to
evaluate the performance of the proposed model in general multiaxial loading conditions
by identifying the effect of loading histories on this parameter.

(ii) In an effort to incorporate the yield surface shape changes in cyclic plasticity
modeling, this study demonstrates the prospect of employing the equi-plastic-strain
surface as the plastic potential surface. Extensive experimental studies should be
conducted to observe the shapes of the equi-plastic-strain surfaces in general multiaxial
loading conditions.

(iii) Studies should be conducted to represent the plastic potential surfaces,


including the yield surfaces, in a simple, theoretically sound and numerically inexpensive
manner.

(iv) This study concentrates on the rate-independent ratcheting responses of


stabilized carbon steels. Future studies should be conducted to extend these models to
cyclically hardening and softening materials with loading rate and temperature variation.

130
REFERENCE

AbdelKarim, M., Ohno, N., (2000). Kinematic Hardening Model Suitable for Ratcheting
with Steady-state. International Journal of Plasticity, 16, 225-240.

Armstrong, P.J. and Frederick, C.O. (1966) A Mathematical Representation of the


Multiaxial Bauscinger Effect. CEGB Report No. RD/B/N 731.

Besseling, J.F. (1958) A Theory of Elastic, Plastic and Creep Deformations of an Initially
Isotropic Material. Journal of Applied Mechanics, Vol 25, pp. 529-536.

Burlet, H., Cailletaud, G., (1987). Modeling of Cyclic Plasticity in Finite Element Codes.
Proc. of 2nd Int. Conf. on Constitutive Laws for Engineering Materials: Theory and
Applications, pp. 1157-1164.

Chaboche, J.L. (1986) Time-Independent Constitutive Theories For Cyclic Plasticity.


International Journal of Plasticity, Vol 2, pp. 149-188.

Chaboche, J.L. (1991) On Some Modifications of Kinematic Hardening to Improve the


Description of Ratcheting Effects. International Journal of Plasticity, Vol 7, pp. 661-
678.

Corona, E., Hassan, T. and Kyriakides, S. (1996) On the Performance of Kinematic


Hardening Rules in Predicting a Class of Biaxial Ratcheting Histories. International
Journal of Plasticity, Vol 12, pp. 117-145.

Dafalias, Y.F. and Popov, E.P. (1976) Plastic Internal Variables Formalism of Cyclic
Plasticity. Journal of Applied Mechanics, Vol 43, pp. 645-650.

Delobelle, P., Robinet, P., Bocher, L,. (1995). Experimental study and phenomenological
modelization of ratcheting under uniaxial and biaxial loading on an austenitic
stainless steel. International Journal of Plasticity, 11, 295-330.

Eisenberg, M.A. and Yen, C.F. (1981) A Theory of Multiaxial Anisotropic


Viscoplasticity. Journal of Applied Mechanics, Vol 48, pp. 276-284.
131
Eisenberg, M.A. and Yen, C.F. (1984) The Anisotropic Deformation of Yield Surface.
Journal of Engineering Materials and Technology, Vol 105, pp. 355-360.

Guionnet, C. (1992) Modeling of Ratcheting in Biaxial Experiments. Journal of


Engineering Materials and Technology, Vol 114, pp. 56-62.

Hassan, T. and Kyriakides, S. (1992) Ratcheting in Cyclic Plasticity, Part I: Uniaxial


Behavior. International Journal of Plasticity, Vol 8, pp. 91-116.

Hassan, T. and Kyriakides, S. (1994) Ratcheting of Cyclically Hardening and Softening


Materials, Part II: Multiaxial Behavior. International Journal of Plasticity, Vol 10,
pp. 185-212.

Hassan, T., Corona, E. and Kyriakides, S. (1992) Ratcheting in Cyclic Plasticity, Part II:
Multiaxial Behavior. International Journal of Plasticity, Vol 8, p. 117-146.

Ikegami, K (1979) Experimental Plasticity on the Anisotropy of Metals. Colloques


Internationaux du CNRS, No. 295 - Comportment Mechanique Des Solids
Anisotropes, Coll. 115, Paris, pp. 201-242.

Ilyushin, A.A. (1954) On the Relation Between Stresses and Small Strains in the
Mechanics of Continua (in Russian). Prikl. Mat. Mech., Vol 18, pp. 641-666.

Ishikawa, H. and Sasaki, K. (1988) Yield Surfaces of SUS 304 Under Cyclic Loading.
Journal of Engineering Materials and Technology, Vol 110, pp. 365-371.

Ivey, H.J. (1961) Plastic Stress Strain Relations and Yield Surfaces for Aluminum
Alloys. Journal of the Mechanical Engineering Science, Vol 3, pp. 15-31.

Jiang, Y., Sehitoglu, H., (1996) Modeling of cyclic ratchetting plasticity, part I:
development of constitutive relations. ASME J. App. Mech., 63, 720-725.

Kaneko, K. (1981) Proposition of New Translation Rule in Kinematic Hardening.


Bulletin of the JSME, Vol 24, pp. 9-14.

Kaneko, K. (1984) Development of a New Practical Plastic Constitutive Model for

132
Anisotropic Metals After Various Preloading. Bulletin of the JSME, Vol 27, pp.
2687-2693.

Kurtyka, T. and Zyczkowski, M. (1996) Evolution Equations For Distortional Plastic


Hardening. International Journal of Plasticity, Vol 12, pp. 191-213.

McDowell, D.L. (1995) Stress State Dependence of Cyclic Ratcheting Behavior of Two
Rail Steels. International Journal of Plasticity, Vol 11, pp. 397 421.

Mroz, Z. (1967) On the Description of Anisotropic Work Hardening. Journal of the


Mechanics and Physics of Solids, Vol 15, pp. 163-175.

Naghdi, P.M., Essenburg, F. and Koff, W. (1958) An Experimental Study of Initial and
Subsequent Yield Surfaces in Plasticity. Journal of Applied Mechanics, Vol 25, pp.
201-209.

Ohno, N. and Wang, J.-D. (1993) Kinematic Hardening Rules with Critical State of
Dynamic Recovery, Part I: Formulations and Basic Features for Ratcheting
Behavior. International Journal of Plasticity, Vol 9, pp. 375-390.

Ortiz, M. and Popov, E.P. (1983) Distortional Hardening Rules for Metal Plasticity.
ASCE Journal of Engineering Mechanics, Vol 109, pp. 1042-1057.

Phillips, A. and Lu, W.-Y. (1984) An Experimental Investigation of Yield Surfaces And
Loading Surfaces of Pure Aluminum With Stress-Controlled And Strain-Controlled
Paths of Loading. Journal of Engineering Materials and Technology, Trans. ASME,
Vol 106, pp. 349-354.

Phillips, A., Lee, C.W., (1979) Yield surfaces and loading surfaces: experiments and
recommendations. International Journal of Solids and Structures, Vol 15, pp. 715-
729.

Phillips, A., Tang, J.-L., (1972) The effect of loading path on the yield surface at elevated
temperatures. International Journal of Solids and Structures, Vol 8, pp. 463-474.

133
Prager, W. (1956) A New Method of Analyzing Stresses and Strains in Work Hardening
Plastic Solids. Journal of Applied Mechanics, Vol 23, pp. 493-496.

Seyed-Ranjbari, M. (1986) Further Development, Multiaxial Formulation, and


Implementation of the Bounding Surface Plasticity Model for Metals. Ph.D.
dissertation, University of California, Davis.

Shiratori, E., Ikegami, K. and Yoshida, F. (1979) Analysis of Stress-Strain Relations by


Use of An Anisotropic Hardening Plastic Potential. Journal of the Mechanics and
Physics of Solids, Vol 27, pp. 213-229.

Tseng, N.T. and Lee, G.C. (1983) Simple Plasticity Model of the Two Surface Type.
ASCE Journal of Engineering Mechanics, Vol 109, pp. 795-810.

Voyiadjis, G.Z. and Sivakumar, S.M. (1991) A Robust Kinematic Hardening Rule for
Cyclic Plasticity with Ratcheting Effects, Part I: Theoretical Formulation. Acta
Mechanica, Vol 90, pp. 105-123.

Voyiadjis, G.Z. and Sivakumar, S.M. (1994) A Robust Kinematic Hardening Rule for
Cyclic Plasticity with Ratcheting Effects, Part II: Application to Nonproportional
Loading Cases. Acta Mechanica, Vol 107, pp. 117-136.

Voyiadjis, G.Z., Basuroychowdhury, I.N., (1998) A plasticity model for multiaxial cyclic
loading and ratchetting. Acta Mechanica., 126, 19-35.

Wegener, K. and Schlegel, M. (1996) Suitability of Yield Functions for the


Approximation of Subsequent Yield Surfaces. International Journal of Plasticity,
Vol 12, pp. 1151-1177.

Xia, Z. and Ellyin, F. (1994) Biaxial Ratcheting Under Strain or Stress-Controlled Axial
Cycling With Constant Hoop Stress. Journal of Applied Mechanics, Vol 61, pp. 422-
428.

Xia, Z. and Ellyin, F. (1997) A Constitutive Model with Capability to Simulate Complex
Multiaxial Ratcheting Behaviour of Materials. International Journal of Plasticity,

134
Vol 13, pp. 127-142.

Yoshida, F., Murakami, T., Kaneko, K., Ikegami, K. and Shiratori, E. (1977) Plastic
Behaviour in the Combined Loading along Straight Stress Paths With a Bend.
Bulletin of the JSME, Vol 20, pp. 1549-1556.

Ziegler, H. (1959) A Modification of Parager's Hardening Rule. Quarterly of Applied


Mechanics, Vol 17, pp. 55-65.

135

You might also like