A Thermodynamic Consistent Rate-Dependent Elastoplastic-Damage Model

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Article

International Journal of Damage


Mechanics
0(0) 1–24
A thermodynamic ! The Author(s) 2016
Reprints and permissions:
consistent rate-dependent sagepub.co.uk/journalsPermissions.nav
DOI: 10.1177/1056789516676882
elastoplastic-damage model ijd.sagepub.com

Mehdi Ganjiani

Abstract
In this article, a rate-dependent elastoplastic-damage constitutive model considering the effect of strain
rate has been developed. The derivation of this model has been established based on the irreversible
thermodynamics with internal variables within the fundamentals of Continuum damage mechanics (CDM).
For investigating the rate effect, an additional power function dependent on the effective strain rate has
been involved in the plastic dissipation function (dynamic plastic yield surface). The damage has been
assumed as a tensor-type variable and based on the energy equivalence hypothesis, the damage evolution
has been developed. The proposed constitutive model has been implemented into user-defined subrou-
tines UMAT and VUMAT in the finite-element program ABAQUS/(Standard and Explicit). For this pur-
pose, the implicit and explicit stress integration algorithms of the model have been explained. The model
has been validated by comparing the predicted results with experimental data conducted on Al2024-T3.
These experiments are including the tensile and double-notched tests. Furthermore, the numerical results
have been compared with some data available in the literature. The numerical examples show the excel-
lent correlation between experiments and simulations for stress (or force) and damage results.

Keywords
Rate-dependent constitutive relation, continuum damage mechanics, strain-rate sensitivity, dynamic yield
surface, Consistency model

Introduction
For the analysis of many rate-dependent processes such as metal forming, impact, crash, and others,
various viscoplastic models have been developed. Two kinds of models can be formulated to account
the viscoplastic behavior of materials: the overstress (Perzyna)-type model (Duvaut and Lions, 1972;
Perzyna, 1966) and the Consistency-type model (Heeres et al., 2002; Ponthot, 1995; Ristinmaa and
Ottosen, 2000; Saksala et al., 2015; Wang, 1997; Wang et al., 1997). In the Perzyna-type model,

Department of Mechanical Engineering, College of Engineering, University of Tehran, Tehran, Iran


Corresponding author:
Mehdi Ganjiani, Department of Mechanical Engineering, College of Engineering, University of Tehran, P.O. Box 11155-4563,
Tehran, Iran.
Email: ganjiani@ut.ac.ir

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


2 International Journal of Damage Mechanics 0(0)

a rate-independent yield function is used to describe the viscoplastic behavior of materials. The
overstress effect in this model means that this yield function can become larger than zero. Using the
overstress model, the consistency conditions are not fulfilled and the stress states outside the yield
surface are allowed (see Zaera and Fernández-Sáez, 2006). The main problems of the overstress
model have been discussed by Hashiguchi et al. (2005). In the Consistency-type model, the rate of
state variables is considered as the independent state variable. The rate of state variables is included
in the yield function and the time derivative of rate-dependent yield function illustrates the visco-
plastic behavior of materials. Additionally, the classical Kuhn–Tucker relations, along with the
consistency conditions, are valid for loading and unloading conditions.
Heeres et al. (2002) compared the elastic–viscoplastic characteristics of the Perzyna- and
Consistency-type models regarding the viscoplastic multiplier. In the Perzyna-type model,
the rate of viscoplastic multiplier is explicitly defined via an overstress function, while in the
Consistency-type model, it is governed by a non-homogeneous differential equation. As they illu-
strated, the different responses during the stress reversals are the dissimilarities of these two
models. Voyiadjis and Abed (2006) proposed a coupled temperature and strain-rate microstructure
physically based yield function to derive a kinematical model for thermo-viscoplastic deform-
ations of BCC metals. In their work, the viscoplastic multiplier is obtained using both the
Consistency- and Perzyna-type viscoplasticity models and in the case of the Perzyna
viscoplasticity model, the athermal yield function is employed instead of the static yield
function.Yu et al. (2009a, 2009b) performed the experimental studies for thin plate specimens of
DP600 steel at strain rates range from 104 s1 to 103 s1 . They proposed a plastic constitutive
relationship based on Khan–Huang model to describe the rate-dependent plastic behavior of
DP600 steel at various strain rates.
Continuum damage mechanics (CDM) is a theory which phenomenologically studies the deteri-
oration of material properties caused by growing of the damage. In this theory, the damage is
defined as an internal variable, which its evaluation is consistent with the thermodynamics. In
this theory, the elasticity and/or viscoplasticity (plasticity) theory is incorporated to illustrate the
behavior of materials (Al-Rub and Darabi, 2012; De Souza Neto and Peric, 1996; Ganjiani et al.,
2012a, 2012b; Gomez and Basaran, 2006; Grammenoudis et al., 2009a, 2009b; Lubarda et al., 1994;
Saanouni et al., 2003; Simo and Ju, 1987; Steinmann et al., 1994; Voyiadjis and Abed, 2006). It is
observed experimentally that the deformation response of material as well as damage evolution is
influenced by strain rate, temperature, history of loading, and stress (Børvik et al., 2001; Wang et al.,
2010; Zukas, 1990).
The study of the dynamic deformation and fracture of materials by taking into account the
damage evolution has become one of the research frontiers, receiving more and more attention
by the scientists until now (DiLellio and Olmstead, 2003; Johansson et al., 1999; Mahnken et al.,
1998).
There are two types of damage mechanics theories: one uses a damage potential surface and the
second theory does not use a potential surface and defines damage as a change in entropy produc-
tion rate (Basaran and Nie, 2004; Basaran and Yan, 1998; Yao and Basaran, 2013). The second
damage theory, which automatically defines damage metric directly as entropy generation rate
without intermediary, phenomenological parameters, and variables.
Coupling a damage model with Perzyna-type approach, as a theory which uses damage potential
surface, is widely used in the literature to describe the rate-dependent (viscous) damage behavior, but
this is not a viscous damage model to be consistent with irreversible thermodynamics.
In this tendency, Simo and Ju (1987) developed a rate-dependent (viscous) damage model that
produces retardation of micro-cracking at higher strain rates. Their model is a viscous regularization

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


Ganjiani 3
of rate-independent damage, with a structure analogous to that of viscoplasticity of the Perzyna-
type. Also, Johansson and Runesson (1997) presented a formulation of viscoplasticity theory,
with kinetic coupling to isotropic damage. They described the rate-dependent material behavior
and failure processes, including creep-rupture (for constant loading) and creep-fatigue (for cyc-
lic loading). They adopted the Duvaut–Lions formulation of viscoplasticity with general harden-
ing of the quasi-static yield surface. Chow et al. (2001) developed a viscoplastic constitutive
modeling of anisotropic damage for the prediction of forming limit curve (FLC). Their viscoplastic
constitutive formulation is based on the Perzyna-type model. Ren and Li (2013) represented the
nonlinear behavior and the rate-dependency of plain concrete by proposing a model in which its
rate-dependency is based on the Perzyna-type viscoplasticity. Carniel et al. (2015) presented a finite-
element formulation of generalized Kelvin–Voigt and Perzyna models, which are coupled with
Lemaitre’s model as a material degradation theory. They found the presence of damage expressively
affects the impact structural response. Al-Rub et al. (2015) used a viscodamage model, coupled with
Schapery viscoelasticity and Perzyna viscoplasticity constitutive models in order to simulate and
predict the inelastic and time-dependent damage behavior of polymeric materials and their
composites.
The lack of research on developing a viscous damage model which is consistent with irreversible
thermodynamics is obvious. In this analogy, Saksala et al. (2015) developed a constitutive model
combining continuum damage with embedded discontinuity for dynamic analyses of quasi-brittle
failure phenomena. Their model involves a rate-dependent continuum damage with isotropic
hardening formulated in the Consistency-type approach.
The main objective of this article is to propose a rate-dependent elastoplastic-damage constitutive
model in the framework of CDM based on the Consistency-type approach. So, a rate-dependent
elastoplastic-damage constitutive model in framework of CDM is introduced. The effect of strain
rate is introduced in the framework of a power function incorporated to the plastic dissipation
potential. This implies that the proposed model is based on the Consistency-type model. The appro-
priate internal variables are identified which are consistent in the context of irreversible thermo-
dynamics. For numerical purposes, the implicit and explicit algorithms of the model are presented.
Also, in order to show the capability of the model at high strain-rate deformation, the model is
implemented as a user-defined subroutine UMAT/VUMAT in the finite-element program
ABAQUS, and the simulation results are compared with the experiment data.

Theoretical background
Preliminaries
In the concept of damage mechanics, the effective stress tensor r is assumed to be

r ¼ MðDÞ : r ð1Þ

where MðDÞ is a fourth-order symmetric tensor which denotes the damage effect tensor (Chow and
Wang, 1988).
 
1 1 1 1 1 1
MðDÞ ¼ diagonal pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1  D1 1  D2 1  D3 ð1  D2 Þð1  D3 Þ ð1  D3 Þð1  D1 Þ ð1  D1 Þð1  D2 Þ
ð2Þ

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


4 International Journal of Damage Mechanics 0(0)

where D1 , D2 , and D3 are principal values of the damage tensor D. Based on the rules of irreversible
thermodynamics, the thermodynamic forces conjugated to the thermodynamic variables can be
defined as

ee ðr, DÞ ¼ M : C1 1
e0 : M :r ¼ Ce : r
@MðDÞ
Yðr, DÞ ¼ rT : M : C1
e0 : @D :r
ð3Þ
n
h ðrÞ ¼ Kr
Yh ðÞ ¼ Kd  expðmÞ

where Y is the damage energy release rate, h represents the isotropic hardening for plastic deform-
ation, and Yh is treated as isotropic hardening for damage evolution. r is the effective plastic strain
and  is the accumulated damage. The sound elastic stiffness tensor, Ce0 , is considered as:

Ce0 ¼ 2Is þ  I  I ð4Þ

with Is ¼ 12½ik jl þ il jk  ei ej ek el and I ¼ ij ei ej as the symmetric identity tensors of fourth- and
second order, and ,  as the Lame´ constants which are related to Young’s modulus E and
Poisson’s ratio . Using the hypothesis of normal dissipation, the evolution equations of the internal
variables can be obtained as

@Fp @Fp _ ¼ _d @F ,


d
@F d
e_ p ¼ _p , r_ ¼ _p , D _ ¼ _d ð5Þ
@r @h @Y @Yh

where Fp and Fd are limiting functions which denote respectively the plastic and damage dissipation
potentials (yield surfaces), and, _p and _d are two positive variables known as the plastic and damage
multipliers.

Damage yield function


The damage yield surface determines the possibility of damage propagation. The following relation-
ship is adopted for the damage yield criterion:
 1=2
1 T
Fd ðY, Yh ; D, , rÞ ¼ kYk  Y0  Yh ðÞ ¼ Y : LðrÞ: Y Y0  Yh ðÞ ð6Þ
2

where L is a fourth-order positive damage characteristic tensor describing the damage-induced


change in the damage surface. The parameter Y0 is the damage threshold used to specify the size
of the initial damage surface. The evolution law for damage is characterized as

d d
_ ¼ _d @F ¼ _d Nd ,
D _ ¼ _d
@F
¼ _d ð7Þ
@Y @Yh

with the following Kuhn–Tucker relations which control the damage evolution:

Fd  0, _d  0, _d Fd ¼ 0 ð8Þ

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


Ganjiani 5
The unit tensor Nd is normal to the damage surface, describing the direction of damage propa-
gation, defined as:

L:Y
Nd ¼ ð9Þ
2kYk

L ¼ 2 ð1  e r Þ Is þ 2e r I  I ð10Þ

where the characteristic tensor L is defined in Voigt notation as


2 3
2 2e r 2e r 0 0 0
6 2 2e r 0 0 0 7
6 7
6 2 0 0 0 7
L ¼ 6
6
7
7 ð11Þ
6 1  e r 0 0 7
4 1  e r 0 5
 r
syms 1e

with as a material constant.

Plastic yield function


The plastic yield criterion can be treated as a condition, which enables whether the plastic behavior
occurs or only elastic behavior happens. In this work, we developed the static/dynamic plastic yield
function for rate-independent/dependent deformations, respectively.

Static yield surface. The plastic dissipation potential, which distinguishes the plastic regime of deform-
ation from that elastic regime, is adopted as
 1=2
1 T 
Fp ðr, h ; D, rÞ ¼ krk  0  h ðrÞ ¼ r : H :r 0  h ðrÞ ð12Þ
2
 is the effective plastic characteristic tensor
where 0 is the initial strain hardening threshold and H
defined as

 ¼ MT ðDÞ: H : MðDÞ
H ð13Þ

The positive definite tensor H for orthotropic materials is represented by a 6  6 matrix in the
material principal coordinate system as
2 3
2 1 1 0 0 0
6 2 1 0 0 0 7
6 7
6 7
6 2 0 0 0 7
H ¼ 6
6
7 ð14Þ
6 6 0 0 7
7
6 7
4 6 0 5
syms 6

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


6 International Journal of Damage Mechanics 0(0)

It is assumed that the material is isotropic onset of deformation. Succeeding the loading, the
damage propagates and thus the material may behave in an anisotropic manner. In other words, the
anisotropic properties are induced in the behavior of material by damage propagation.

Dynamic yield surface. Assuming von-Mises plasticity with isotropic hardening, the yield function Fp is
written as:

Fp ðr, h ; D, r, r_Þ ¼ krk  h ðr, r_Þ ð15Þ

where h ðr, r_Þ is different for BCC and FCC materials (Liang and Khan, 1999; Rusinek et al., 2010;
Voyiadjis and Abed, 2005). We adopted h ðr, r_Þ for FCC materials as
 c
  r_
h FCC ðr, r_Þ ¼ 0 þ h ðrÞ ð16Þ
r_0

and for BCC materials as


 c
r_
h BCC ðr, r_Þ ¼ 0 þh ðrÞ ð17Þ
r_0

where r_ is the plastic strain rate, r_0 is a lower bound strain-rate and c is the strain-rate sensitivity
parameter. The yield function in equation (14) furnishes the constitutive equation for plastic strain
increment e_ p and the increment of isotropic hardening variable r_ as follows:

@Fp @Fp
e_ p ¼ _p ¼ _p Np , r_ ¼ _p ¼ _p ð18Þ
@r @ h

where Np which is a second-order tensor, represents the normal direction to the plastic yield surface
and is obtained as

H :r MT : H : M :r
Np ¼ ¼ ð19Þ
2 k rk 2krk

The plastic yield surface, equation (14), should also satisfy the following Kuhn–Tucker loading/
unloading conditions:

Fp  0, _p  0, _p Fp ¼ 0 ð20Þ

Stress update algorithm. The constitutive equations described in the previous sections are summarized
as

r ¼ Ce : e e
e_ p ¼ _p Np
_ ¼ _d Nd
D

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


Ganjiani 7

r_ ¼ _p
_ ¼ _d
F_p ðr, h ; D, r, r_Þ ¼ Fp,r : r_ þ Fp,D : D
_ þ Fp,r r_ þ Fp,_r r€ ¼ 0
F_d ðY, Yh ; D, , rÞ ¼ Fd,r : r_ þ Fd,D : D _ þ Fd, _ þ Fd,r r_ ¼ 0
Fp  0, _p  0, _p Fp ¼ 0
Fd  0, _d  0, _d Fd ¼ 0 ð21Þ

The aim of integration algorithm is that using the input data fen , epn , rn , Dn , rn , r_n , n g at time tn
and applying the strain increment e, the variables fenþ1 , epnþ1 , rnþ1 , Dnþ1 , rnþ1 , r_nþ1 , nþ1 g are
obtained by satisfying the loading–unloading conditions at time tnþ1 .

Implicit approach
This section presents the integration algorithm of the proposed constitutive equations using a back-
ward Euler implicit integration algorithm. At this algorithm, the plastic strain increment and the
increment of internal variables are calculated at the end time, i.e., tnþ1 . During this calculation, the
satisfaction of either/both plastic and damage surfaces is established. Therefore, the integration
algorithm is written as

enþ1 ¼ en þ e
epnþ1 ¼ epn þ nþ1
p
Npnþ1
d
Dnþ1 ¼ Dn þ nþ1 Ndnþ1
p
rnþ1 ¼ rn þ nþ1
p
r_nþ1 ¼ nþ1 =t ð22Þ
d
nþ1 ¼ n þ nþ1

rnþ1 ¼ Ce ðDnþ1 Þ : enþ1  epnþ1


Fpnþ1 ðrnþ1 , Dnþ1 , rnþ1 , r_nþ1 Þ ¼ 0
Fdnþ1 ðYnþ1 , Dnþ1 , nþ1 , rnþ1 Þ ¼ 0

The nonlinear equations within equation (21) are solved for the parameters with subscript n þ 1.
The equation (21)2 is rewritten in the following form:

epnþ1  epnþ1  epn ¼ nþ1


p
Npnþ1 ð23Þ

Substituting equations (21)1 and (22) into equation (21)6, we will get

rnþ1 ¼ Ce ðDnþ1 Þ : een þ e  epnþ1 ð24Þ

If we write Ce ðDnþ1 Þ as Ceðnþ1Þ , equation (23) is simplified to

p
C1 e
eðnþ1Þ : rnþ1 ¼ en þ e  enþ1 ð25Þ

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


8 International Journal of Damage Mechanics 0(0)

Multiplying both sides of equation (24) at CeðnÞ yields,


CeðnÞ : C1
eðnþ1Þ : rnþ1 ¼ CeðnÞ : een þ e  epnþ1
¼ CeðnÞ : een þ CeðnÞ : e  CeðnÞ : epnþ1

¼ rn þ CeðnÞ : e  CeðnÞ : epnþ1 ð26Þ


¼ rtrial
nþ1  CeðnÞ : epnþ1
p
¼ rtrial
nþ1  nþ1 CeðnÞ : Npnþ1

where rtrial
nþ1 ¼ rn þ CeðnÞ : e is the trial stress. In the integration of the model, a return-mapping
algorithm consisting the elastic predictor and plastic/damage corrector is used. During the plastic/
damage corrector, the total strain is constant and linearization is established based on the increment
of p and d .

Linearization. In the subsequent equations, we neglect the subscript n þ 1 for simplicity. Equation (21)
is written in the following form, which is consistent with the Newton–Raphson method:

Rr ¼ C1 1 trial p p
e : r  CeðnÞ : rnþ1 þ  N ¼ 0

RD ¼ D þ Dn þ d Nd ¼ 0
ð27Þ
Fp ¼ Fp ðr, D, r, r_Þ ¼ 0
Fd ¼ Fd ðY, D, , rÞ ¼ 0

Linearizing Equation (26) yields


       
@Rr ðkÞ @Rr ðkÞ @Rr ðkÞ pðkÞ @Rr ðkÞ dðkÞ
RðkÞ
r þ ðkÞ
: r þ ðkÞ
: D þ  þ  ¼0 ð28Þ
@r @D @p @d
       
@RD ðkÞ @RD ðkÞ @RD ðkÞ pðkÞ @RD ðkÞ dðkÞ
RðkÞ
D þ : rðkÞ þ : DðkÞ þ  þ  ¼0 ð29Þ
@r @D @p @d
 ðkÞ  p ðkÞ  p 
@Fp @F @F 1 @Fp ðkÞ pðkÞ
FpðkÞ þ : rðkÞ þ : DðkÞ þ þ  ¼0 ð30Þ
@r @D @p t @_p
 d ðkÞ  d ðkÞ  d ðkÞ  d ðkÞ
@F @F @F @F
FdðkÞ þ : rðkÞ þ : DðkÞ þ d
 dðkÞ
þ pðkÞ ¼ 0 ð31Þ
@r @D @ @p

If we focus on equations (27) and (28), the partial derivatives observed in these two equations are
obtained as

@Rr p @N
p
@Rr @C1 @Np
¼ C1
e þ  , ¼ e : r þ p
@r @r @D @D @D ð32Þ
@Rr @Rr
¼ Np , ¼0
@p @d

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


Ganjiani 9
and

@RD @Nd @RD @Nd


¼ d ¼ d Nd,r , ¼ I þ d ¼ I þ d Nd,D
@r @r @D @D
ð33Þ
@RD @Nd @RD
p
¼ d ¼ d Nd,p , ¼ Nd
@ @p @d
Henceforth, we write all equations in the matrix–vector format. Substituting equations (31)
and (32) in equations (27) and (28), the resultant equations can be written in the following matrix
format:

1 rðkÞ    
AðkÞ ðkÞ ¼  RðkÞ  NðkÞ ðkÞ ð34Þ
D

where
" # " # ðkÞ
p p
ðkÞ 1 C1 p p
e þ  N, r C1
e, D : r þ  N, D ðkÞ
Np 0
A ¼ , N ¼
d Nd,r I þ d Nd,D d Nd,p Nd
( ) p  ð35Þ
 ðkÞ  RðkÞ
r  ðkÞ   ðkÞ
R ¼ ,  ¼
RðkÞ
D
d ðkÞ

Solving equation (33) for the increment of stress and damage gives
( )
rðkÞ    
¼  AðkÞ RðkÞ  AðkÞ NðkÞ ðkÞ ð36Þ
DðkÞ

Also, the matrix form of equations (29) and (30) is


( )
 ðkÞ  h ðkÞ i rðkÞ h i
ðkÞ  ðkÞ 
F þ FrD þ F   ¼0 ð37Þ
DðkÞ

where
2 3
( ) " # pðkÞ 1
  FpðkÞ h i pðkÞ
F,r pðkÞ
F,D h i
6 F,p þ FpðkÞ
_
pðkÞ
F,d 7
FðkÞ ¼ , ðkÞ
FrD , FðkÞ t ,p
¼  ¼ 4 5 ð38Þ
FdðkÞ dðkÞ
F,r dðkÞ
F,D dðkÞ dðkÞ
F,p F,d

 
So, by substituting equation (35) into equation (36), and solving that for ðkÞ , we will get

 ðkÞ  h ðkÞ ðkÞ ðkÞ i1 n o


 ¼ FrD A N  FðkÞ

ðkÞ ðkÞ ðkÞ
FðkÞ  FrD A R ð39Þ

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


10 International Journal of Damage Mechanics 0(0)
 
Now, as fðkÞ g was calculated, the values of rðkÞ DðkÞ can be obtained from equation (35).
Therefore, other internal variables are updated at the end of time step as follows:

rðkþ1Þ ¼ rðkÞ þ rðkÞ , dðkþ1Þ ¼ dðkÞ þ dðkÞ


epðkþ1Þ ¼ epðkÞ þ epðkÞ , rðkþ1Þ ¼ rn þ pðkþ1Þ
ð40Þ
Dðkþ1Þ ¼ DðkÞ þ DðkÞ , 1
r_ðkþ1Þ ¼ r_n þ t pðkþ1Þ
pðkþ1Þ ¼ pðkÞ þ pðkÞ , ðkþ1Þ ¼ n þ dðkþ1Þ

The iterations are repeated until the criterion fFðkþ1Þ g 5 tol is established.

Algorithmic tangent modulus. In the implicit numerical methods, a suitable tangent modulus of the
proposed algorithm is required for the fast convergent. This modulus is defined as

 
dr
Calg ¼ ð41Þ
de nþ1

For obtaining an expression for the tangent modulus, the increment form of equation (21) at time
tnþ1 will be

dr ¼ Ce : de  dep  dC1
e : r
dep ¼ dp Np þ p dNp
dD ¼ dd Nd þ d dNd ð42Þ
 
p p p 1 p
p
dF ¼ F,r : dr þ F,D : dD þ F,p þ tF,p
_ dp þ Fp,d dd ¼ 0
dFd ¼ Fd,r : dr þ Fd,D : dD þ Fd,p dp þ Fd,d dd ¼ 0

By substituting equation (41)2 into equation (41) 1 and employing equation (41) 3, we arrive at the
following equation:
 
dr de
¼ ½A  ½A½Nfdg ð43Þ
dD 0

Also, by replacing equation (42) into equations (41)4 and (41)5, and solving them for dp and dd ,
we get
 
dp 1 de
¼ F  FrD AN fFrD Ag ð44Þ
dd 0

By substituting this result into equation (42), we have


 h i 
dr 1 de
¼ A  AN  F  FrD AN ½FrD A ð45Þ
dD 0

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


Ganjiani 11
Equation (44), indeed, represents a relationship for algorithm tangent modulus of stress and
damage tensors. The algorithm tangent modulus defined in equation (40) can be simply extracted
from equation (44).

Explicit approach
The elastic stress–strain relationship, equation (3) 2, can be rewritten in the following form:

e  ep ¼ M : C1
e0 : M :r ð46Þ

Therefore, the incremental form of equation (45) will give

@M
e  ep ¼ 2 M : C1
e0 : : r : D þ M : C1
e0 : M :r
ð47Þ
@D

where the first part in right-hand side of equation (46) can be interpreted as ed , the increment of
damage strain. Substituting equations (7)2 and (17)2 into equation (46), after some simple manipu-
lations, we obtain
 
@M
r ¼ M1 : Ce0 : M1 : e  p Np  2 d M : C1
e : : r: N d
ð48Þ
@D

The plastic and damage multipliers, p and d , are calculated by establishing the consistency
conditions for plastic and damage surfaces as follows:

@Fd @Fd
Fd ¼ : Y þ  ¼ 0 ð49Þ
@Y @

@Fp @Fp @Fp @Fp


Fp ¼ : r þ : D þ r þ _r ¼ 0 ð50Þ
@r @D @r @_r

The increment of plastic strain rate _r can be written as

p  np
_r ¼ ð51Þ
t

where np is the plastic multiplier which was calculated at previous step. Using equations (7)2 and
(17)2, equations (48) and (49) yield the following relations for plastic and damage multipliers:

dYh p @Fd
 ¼ : Y ð52Þ
d @Y
 
@ h 1 @ h 1 @ h p @Fp @Fp
þ p  n ¼ : r þ : D ð53Þ
@r t @_r t @_r @r @D

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


12 International Journal of Damage Mechanics 0(0)

The increment of damage energy release rate Y is derived as

@Y @Y
Y ¼ : r þ : D ¼ Y, r : r þ Y, D : D ð54Þ
@r @D

Substituting equation (53) into equation (51), we obtain the following relations for d and p :

d ¼ ?d : r
1 @ h
ð55Þ
p ¼ ?p : rþ @ h t @_1r @ h np
@r þ t @_r

with

Nd : Y,r
?d ¼ dY d d
d  N : Y, D : N
h

ð56Þ
p Np þ ?d : Fp,D : Nd
? ¼ @ h 1 @ h
@r þ t @_r

where ?d and ?p are second-order tensors. Substituting equation (54) into equation (47) and after
some mathematical manipulations, we obtain the following equation for :
  1 @h
1 p p d 1 @M
M : Ce0 : M þ ?  N þ2?  M : Ce0 : : r: N : r ¼ e @h t @_1r @h np Np
d
ð57Þ
@D @r þ t @_r

Explicit integration. In this section, the integration algorithm of the proposed constitutive equations
using a forward Euler explicit integration algorithm is presented. The first step in the algorithm is the
evaluation of the elastic trial state in which the increment is assumed to be purely elastic with no
evolution of internal variables. If we denote all quantities at time tn with subscript n, and those at the
next time increment with subscript n þ 1, then with the knowledge of the total strain increment, e,
together with the stress and damage at time tn , rn , and Dn , the trial variables can be obtained as
follows:
1
1
rtrial
nþ1 ¼ rn þ MðDn Þ: Ce0 : MðDn Þ : e ð58Þ
 
1 @MðDÞ
Ytrial trialT
nþ1 ¼ rnþ1 : MðDn Þ: Ce0 : : rtrial ð59Þ
@D n nþ1

With these trial variables, the damage and plastic yield conditions are checked:
1=2
1 trial T
Fp ¼ r : MT ðDn Þ: H : MðDn Þ : rtrial h ðrn , r_n Þ ð60Þ
2 nþ1 nþ1

1=2
1 trial T
Fd ¼ Y : Lðrn Þ: Ytrial Y0  Yh ðn Þ ð61Þ
2 nþ1 nþ1

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


Ganjiani 13
The process of deformation can be split into three stages: plastic deformation only, damage
propagation only, and plastic–damage deformation simultaneously. If Fp  0 and Fd  0, the elastic
step is accepted and all trial variables are the solutions, i.e., ð Þnþ1 ¼ ð Þtrial
nþ1 . When we have only
plastic deformation, Fp 4 0 and Fd  0, ?d ¼ 0, d ¼ 0, and
" #
1 1 @h
p
r ¼ M : C1
e0 : Mþ?  Np n : e @h t @_1r @h np Npn ð62Þ
@r þ t @_r

So, the plastic multiplier p will be calculated via equation (54)2. When we have only damage
propagation, Fp  0 and Fd 4 0, ?p ¼ 0, p ¼ 0, and
 1
d 1 @M
r ¼ M : C1
e0 : Mþ2?  M : C e0 : : r: N d
: e ð63Þ
@D n

Therefore, the damage multiplier d ¼ ?d : r will be obtained. In case of plastic and damage
deformations occurring simultaneously, Fp 4 0 and Fd 4 0, r is calculated by equation (56). The
integration to obtain all quantities at the end of the time step, t, may then be written as

epnþ1 ¼ epn þ p Npn , Dnþ1 ¼ Dn þ d Ndn


ð64Þ
rnþ1 ¼ rn þ p , nþ1 ¼ n þ d

and
h i
r_nþ1 ¼ p =t, rnþ1 ¼ Ceðnþ1Þ : C1
eðnÞ : r n þ e   p p
Nn ð65Þ

where Ceðnþ1Þ defined in equation (3)1 is the elastic stiffness tensor updated with Dnþ1 and CeðnÞ is that
calculated by Dn .

Results and discussions


In order to calibrate the parameters involved in the model, two materials, aluminum Al2024-T3 and
steel 1045, are employed. The author has conducted some experiments on Al2024-T3 to extract its
stress and damage behaviors and for steel 1045, data reported in the literature have been used. The
experiments were tensile and double-notched tests. In addition, these tests were verified by imple-
menting the proposed constitutive model in the finite-element program ABAQUS via a user’s mater-
ial subroutine code in which the details are explained subsequently.

Aluminum Al2024-T3
Tensile specimen on Al2024-T3. The mechanical behavior of materials has been characterized using the
tensile test. The dimensions of the tensile specimen are shown in Figure 1. This test was conducted at
different strain rates and their stress–strain diagrams are depicted in Figure 2(a). Based on this data,
we can conclude that the Al2024-T3 alloy has low sensitivity to the strain rate. The damage–strain
data extracted by the micro hardness measurement (Ganjiani, 2013) are shown in Figure 2(b).

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


14 International Journal of Damage Mechanics 0(0)

Figure 1. Dimensions of the tensile specimen (mm) used for characterization of the material at different strain rates.

Figure 2. The data of tensile test conducted on Al2024-T3 at different strain rates: (a) stress–strain and
(b) damage–strain.

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


Ganjiani 15
Identification of parameters for Al2024-T3. A trial-and-error procedure is adopted here to identify the
parameters involved in the model. The initial estimation of the parameters 0 , K, and n can be
obtained from the stress–strain curve. This estimation is achieved by adjusting the parameters K and
n to get the best fitting to the stress–strain curve. During this fitting, it is assumed the damage has no
effect on the stress curve. The damage threshold parameter Y0 is calculated using the relation Y2 0 =E
(Ganjiani et al., 2012b) where Y0 is the stress at which damage is propagated. It is noted that this
relation is valid only when the damage propagates in the elastic deformation. Using the data in
Figure 2, Y0 gets the value of 315 MPa so that damage threshold will be Y0 ¼ 1:42 MPa. By fixing
the identified parameters, other parameters Kd , m, and are determined exploiting the damage–
strain curve using a fitting procedure. Finally, all parameters are changed slightly in order to obtain
the best fitting to the stress and damage curves. The rate-sensitivity parameter c is determined using
the stress–strain data at different strain rates. In case of uniaxial state of loading at a specific strain,
the relative stress at two strain rates for FCC materials is given by
   
2 r_2
ln ¼ c ln ð66Þ
1 r_1

In case of Al2024-T3 as a FCC material, equation (65) at a specific strain r ¼ 0:1 has been plotted
in Figure 3 which leads to take c as 0:0052. This figure was obtained based on the data used in
Figure 2(a). This identified value of c implies that the material Al2024-T3 has low sensitivity to the
strain rate, which has also been observed by other researchers (Johnson and Cook, 1983; Johnson
et al., 1983). The identified material parameters for Al2024-T3 are listed in Table 1.
For validating the model and their calculated parameters, the proposed model has been imple-
mented in the finite-element program ABAQUS/Standard via a user’s material subroutine coded as

Figure 3. Correlation between stress and strain rate at a specific strain for Al2024-T3.

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


16 International Journal of Damage Mechanics 0(0)

Table 1. The determined parameters involved in the model for Al2024-T3.

 E ðGPaÞ
ðkg=m3 Þ 0 ðMPaÞ K ðMPaÞ n Y0 ðMPaÞ Kd ðMPaÞ m c

0.353 70 2770 325 950 0.68 1.42 14.5 2.5 3.13 0.0052

Figure 4. The simulation and experiment curves of tensile test for Al2024-T3: (a) stress and (b) damage.

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


Ganjiani 17
UMAT. The stress (damage)–strain curves are obtained by simulating a single-element mesh. The
predicted stress and damage are compared with experiments in Figure 4. The good agreement
between the stress–strain and damage–strain curves shows the capability of the model and validates
the identified parameters.

Figure 5. Geometry and dimension of double-notched specimen (mm) with 3.2 mm thickness.

Figure 6. Two mesh refinement types of the double-notched specimen: (a) coarse and (b) fine.

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


18 International Journal of Damage Mechanics 0(0)

Double-notched specimen. The double-notched test on the material Al2024-T3 was conducted to probe
the propagation of crack path. The double-notched specimen with its dimensions are shown in
Figure 5 with the thickness of 3:2 mm. The specimen was stretched with two different stain rates
(0.15 s1 and 1.5 s1), and subsequently the load and displacement history for subsequent analysis
were recorded.

Figure 7. Crack path in deformed double-notched specimen of Al2024-T3: (a) coarse mesh, (b) fine mesh, and
(c) experiment.

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


Ganjiani 19
In order to observe the failure phenomena in the simulation, we need to enable the
element deletion flag, which can only be used in the explicit solution. In this regard, the maximum
value of damage is employed as a criterion for the element deletion. When the maximum damage in
the elements reaches its critical value, the stresses and strain increments within the critical elem-
ents will be released. Based on Figure 2(b), the maximum damage used as the critical value is 0.24.
The simulation is performed using eight-node linear brick elements with reduced integration
(C3D8R; ABAQUS notation) including hourglass control. Time increments of the order 108 s
are used to satisfy the stability criteria. The double-notched test is simulated by implementing the
proposed constitutive model in the finite-element program ABAQUS/Explicit via a user’s material
subroutine coded as VUMAT. Two types of mesh refinement, coarse and fine, shown in Figure 6 are
used.
The crack propagation path for these two different meshes are displayed in Figure 7(a) and (b)
besides the fractured specimen which is shown in Figure 7(c). According to the results of Figure 7,
the path of crack growth is suitably predicted by the two mesh refinements. Furthermore, the load–
displacement curve of the specimen is shown in Figure 8. The location for calculating the load is a
cross section far from the fractured region. The force at any time is calculated by summing the loads
of nodes at this cross section. Although there is a slight deviation between the load–displacement
curves with respect to two levels of mesh refinement, the predicted results numerically are in good
agreement with those obtained experimentally.

Steel 1045
Identification of parameters for 1045. A similar procedure can be exploited to determine the model
parameters for steel 1045 as a BCC material. This task has been performed on the experimental data
after the works of Le Roy et al. (1981) and Jaspers and Dautzenberg (2002) (Figures 9 and 10). In
case of uniaxial loading at a prescribed strain, the following relationship can be derived from

Figure 8. Load-displacement curve at different strain-rates in double-notched test.

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


20 International Journal of Damage Mechanics 0(0)

equations (14) and (16) for BCC materials:


   
2  h ðrÞ r_2
ln ¼ c ln ð67Þ
1  h ðrÞ r_1

At a fixed strain r ¼ 0:15 and different strain rates, the corresponding stresses versus their strain
rates substituted into equation (66) are plotted in Figure 9(b). Exploiting this figure, the rate-
sensitivity parameter, c, is found to take the value 0:06.

Figure 9. Characterisitic curves of steel 1045: (a) stress–strain at different strain rates (q-static (Le Roy et al., 1981),
7500 s1 (Jaspers and Dautzenberg, 2002)) and (b) correlation between stress and strain rate at a specific strain.

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


Ganjiani 21
Table 2. The identified material parameters involved in the model for steel 1045.

 E ðGPaÞ
ðkg=m3 Þ 0 ðMPaÞ K ðMPaÞ n Y0 ðMPaÞ Kd ðMPaÞ m c

0.3 200 7872 302 750 0.55 1.81 138 –12 – 0.06

Figure 10. Behavior curves of steel 1045 at low (Le Roy et al., 1981) and high (Jaspers and Dautzenberg, 2002)
strain rates: (a) stress–strain and (b) damage–strain.

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


22 International Journal of Damage Mechanics 0(0)

It is noted that the author neglects the differences of damage growth in tension and compres-
sion states. So, in plotting equation (66), the compression data for the mode of high strain rate
(Jaspers and Dautzenberg, 2002) are adopted. Table 2 represents the identified parameters for
steel 1045.
The results of simulation on steel 1045 are compared with experiments in uniaxial tensile/com-
pressive tests. Figure 10(a) shows the results of stress–strain curve in two different strain rates. At
low strain rate (quasi-static), the experiments of tensile test after Le Roy et al. (1981) are adopted
and at high strain rate, the results of compressive test following Jaspers and Dautzenberg (2002) are
chosen. Furthermore, the corresponding predicted damage data are compared with those of the
experiments in Figure 10(b). Inasmuch as there are no damage data for steel 1045 at high strain rate,
we can only compare the damage results of simulation and experiment at low strain rate. These
comparisons show a satisfactory agreement between simulations and experiments.

Conclusions
In this article, a rate-dependent damage model has been presented in framework of CDM and
Consistency approach. The model has been presented based on the irreversible thermodynamics
with internal variables. In this regard, two plastic and damage surfaces have been defined to distin-
guish the plastic deformation and the damage growth. The plastic yield surface is presented in the
category of Consistency-type model in which the rate of state variables is considered as independent
state variables. The proposed constitutive equations have been integrated using two algorithms,
backward Euler implicit and forward Euler explicit.
The implicit algorithm has been employed in developing the subroutine UMAT whereas the
explicit one has been used in implementing the subroutine VUMAT. In order to validate the
model, the tensile and double-notched tests are investigated experimentally and numerically for
aluminum Al2024-T3. In addition, the numerical results of steel 1045, including stress and
damage, are compared to the published experimental data. These comparisons show the excellent
correlation between experiment and numerical data for stress (or force) and damage results.

Declaration of Conflicting Interests


The author(s) declared no potential conflicts of interest with respect to the research, authorship, and/or pub-
lication of this article.

Funding
The author(s) disclosed receipt of the following financial support for the research, authorship, and/or publi-
cation of this article: The author gratefully acknowledges the Iran National Science Foundation (INSF) for
providing the financial support.

References
Al-Rub RKA and Darabi MK (2012) A thermodynamic framework for constitutive modeling of time-and rate-
dependent materials. Part I: Theory. International Journal of Plasticity 34: 61–92.
Al-Rub RKA, Tehrani AH and Darabi MK (2015) Application of a large deformation nonlinear-viscoelastic
viscoplastic viscodamage constitutive model to polymers and their composites. International Journal of
Damage Mechanics 24: 198–244.
Basaran C and Nie S (2004) An irreversible thermodynamics theory for damage mechanics of solids.
International Journal of Damage Mechanics 13: 205–223.

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


Ganjiani 23
Basaran C and Yan C-Y (1998) A thermodynamic framework for damage mechanics of solder joints. Journal of
Electronic Packaging 120: 379–384.
Børvik T, Hopperstad O, Berstad T, et al. (2001) A computational model of viscoplasticity and ductile damage
for impact and penetration. European Journal of Mechanics-A/Solids 20: 685–712.
Carniel T, Muñoz-Rojas P and Vaz M (2015) A viscoelastic viscoplastic constitutive model including mech-
anical degradation: Uniaxial transient finite element formulation at finite strains and application to space
truss structures. Applied Mathematical Modelling 39: 1725–1739.
Chow CL and Wang J (1988) Ductile fracture characterization with an anisotropic continuum damage theory.
Engineering Fracture Mechanics 30: 547–563.
Chow CL, Yang XJ and Chu E (2001) Viscoplastic constitutive modeling of anisotropic damage under non-
proportional loading. Journal of Engineering Materials and Technology 123: 403–408.
De Souza Neto EA and Peric D (1996) A computational framework for a class of fully coupled models for
elastoplastic damage at finite strains with reference to the linearization aspects. Computer Methods in
Applied Mechanics and Engineering 130: 179–193.
DiLellio J and Olmstead W (2003) Numerical solution of shear localization in Johnson-Cook materials.
Mechanics of Materials 35: 571–580.
Duvaut G and Lions JL (1972) Les ine´quations en me´canique et en physique. Paris: Dunod.
Ganjiani M (2013) Identification of damage parameters and plastic properties of an anisotropic damage model
by micro-hardness measurements. International Journal of Damage Mechanics. DOI: 1056789513482598.
Ganjiani M, Naghdabadi R and Asghari M (2012a) Analysis of concrete pressure vessels in the framework of
continuum damage mechanics. International Journal of Damage Mechanics 21(6): 843–870.
Ganjiani M, Naghdabadi R and Asghari M (2012b) An elastoplastic damage-induced anisotropic constitutive
model at finite strains. International Journal of Damage Mechanics. DOI: 1056789512455937.
Gomez J and Basaran C (2006) Damage mechanics constitutive model for Pb/Sn solder joints incorporating
nonlinear kinematic hardening and rate dependent effects using a return mapping integration algorithm.
Mechanics of Materials 38: 585–598.
Grammenoudis P, Reckwerth D and Tsakmakis C (2009a) Continuum damage models based on energy equiva-
lence: Part I – Isotropic material response. International Journal of Damage Mechanics 18(1): 31–63.
Grammenoudis P, Reckwerth D and Tsakmakis C (2009b) Continuum damage models based on energy equiva-
lence: Part II – Anisotropic material response. International Journal of Damage Mechanics 18(1): 65–91.
Hashiguchi K, Okayasu T and Saitoh K (2005) Rate-dependent inelastic constitutive equation: The extension of
elastoplasticity. International Journal of Plasticity 21: 463–491.
Heeres OM, Suiker ASJ and de Borst R (2002) A comparison between the Perzyna viscoplastic model and the
Consistency viscoplastic model. European Journal of Mechanics-A/Solids 21: 1–12.
Jaspers SPFC and Dautzenberg JH (2002) Material behaviour in conditions similar to metal cutting: Flow
stress in the primary shear zone. Journal of Materials Processing Technology 122: 322–330.
Johansson M, Mahnken R and Runesson K (1999) Efficient integration technique for generalized viscoplas-
ticity coupled to damage. International Journal for Numerical Methods in Engineering 44: 1727–1747.
Johansson M and Runesson K (1997) Viscoplasticity with dynamic yield surface coupled to damage.
Computational Mechanics 20: 53–59.
Johnson G, Hoegfeldt J, Lindholm U, et al. (1983) Response of various metals to large torsional strains over a
large range of strain rates—Part 2: Less ductile metals. Journal of Engineering Materials and Technology 105:
48.
Johnson GR and Cook WH (1983) A constitutive model and data for metals subjected to large strains, high
strain rates and high temperatures. In: Proceedings of the 7th international conference on ballistics, The
Hague, Netherlands, 19–21 April, pp. 541–547.
Le Roy G, Embury JD, Edward G, et al. (1981) A model of ductile fracture based on the nucleation and growth
of voids. Acta Metallurgica 29: 1509–1522.
Liang R and Khan AS (1999) A critical review of experimental results and constitutive models for BCC and
FCC metals over a wide range of strain rates and temperatures. International Journal of Plasticity 15:
963–980.

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016


24 International Journal of Damage Mechanics 0(0)

Lubarda VA, Krajcinovic D and Mastilovic S (1994) Damage model for brittle elastic solids with unequal
tensile and compressive strengths. Engineering Fracture Mechanics 49(5): 681–697.
Mahnken R, Johansson M and Runesson K (1998) Parameter estimation for a viscoplastic damage model using
a gradient-based optimization algorithm. Engineering Computations 15: 925–955.
Perzyna P (1966) Fundamental problems in viscoplasticity. Advances in Applied Mechanics 9: 243–377.
Ponthot JP (1995) Radial return extensions for viscous-plasticity and lubricated friction. In: Transactions of the
13. International Conference on Structural Mechanics in Reactor Technology. v. 2.
Ren X and Li J (2013) A unified dynamic model for concrete considering viscoplasticity and rate-dependent
damage. International Journal of Damage Mechanics 22: 530–555.
Ristinmaa M and Ottosen NS (2000) Consequences of dynamic yield surface in viscoplasticity. International
Journal of Solids and Structures 37: 4601–4622.
Rusinek A, Rodrı́guez-Martı́nez JA and Arias A (2010) A thermo-viscoplastic constitutive model for FCC
metals with application to OFHC copper. International Journal of Mechanical Sciences 52: 120–135.
Saanouni K, Chaboche JL, Milne I, et al. (2003) Computational damage mechanics: Application to metal
forming simulation. In: Milne I, Ritchie RO and Karihaloo BL (eds) Comprehensive structural integrity:
Numerical and computational method (Vol 3). Elsevier: Pergamon, pp.321–376.
Saksala T, Brancherie D, Harari I, et al. (2015) Combined continuum damage-embedded discontinuity model
for explicit dynamic fracture analyses of quasi-brittle materials. International Journal for Numerical Methods
in Engineering 101: 230–250.
Simo JC and Ju JW (1987) Strain- and stress-based continuum damage models – I. Formulation. International
Journal of Solids and Structures 23: 821–840.
Steinmann P, Miehe C and Stein E (1994) Comparison of different finite deformation inelastic damage models
within multiplicative elastoplasticity for ductile materials. Computational Mechanics 13: 458–474.
Voyiadjis GZ and Abed FH (2005) Microstructural based models for BCC and FCC metals with temperature
and strain rate dependency. Mechanics of Materials 37: 355–378.
Voyiadjis GZ and Abed FH (2006) A coupled temperature and strain rate dependent yield function for dynamic
deformations of BCC metals. International Journal of Plasticity 22: 1398–1431.
Wang LL, Zhou FH, Sun ZJ, et al. (2010) Studies on rate-dependent macro-damage evolution of materials at
high strain rates. International Journal of Damage Mechanics.
Wang WM (1997) Stationary and Propagative Instabilities in Metals: A Computational Point of View. Delft, The
Netherlands: Delft University Press.
Wang WM, Sluys LJ and De Borst R (1997) Viscoplasticity for instabilities due to strain softening and strain-
rate softening. International Journal for Numerical Methods in Engineering 40: 3839–3864.
Yao W and Basaran C (2013) Computational damage mechanics of electromigration and thermomigration.
Journal of Applied Physics 114: 103708.
Yu H, Guo Y and Lai X (2009a) Rate-dependent behavior and constitutive model of DP600 steel at strain rate
from 104 to 103 s1. Materials & Design 30: 2501–2505.
Yu H, Guo Y, Zhang K, et al. (2009b) Constitutive model on the description of plastic behavior of DP600 steel
at strain rate from 104 to 103 s1. Computational Materials Science 46: 36–41.
Zaera R and Fernández-Sáez J (2006) An implicit consistent algorithm for the integration of thermoviscoplastic
constitutive equations in adiabatic conditions and finite deformations. International Journal of Solids and
Structures 43: 1594–1612.
Zukas JA (1990) High Velocity Impact Dynamics. New York: John Wiley.

Downloaded from ijd.sagepub.com at ATHABASCA UNIV LIBRARY on November 15, 2016

You might also like