Sie sind auf Seite 1von 24

International Journal of Plasticity 20 (2004) 1851–1874

www.elsevier.com/locate/ijplas

Associative coupled thermoplasticity at


finite strain with temperature-dependent
material parameters

Marko Cana  *, Josip Brnic
dija
Department of Engineering Mechanics, Faculty of Engineering, University of Rijeka, Vukovarska 58,
HR-51000 Rijeka, Croatia
Received in final revised form 15 November 2003
Available online 17 January 2004

Abstract

The paper deals with associative coupled thermoplasticity at finite strains. J2 plasticity
model is based on hyperelastic formulation with multiplicative decomposition of deformation
gradient into elastic and plastic parts. As a novel aspect, temperature dependence of all ma-
terial parameters is introduced. For such model, variational procedure is carried out and
discretization by the mixed finite element method is performed. Consistent linearization is
carried out and algorithmic elasto-plastic tangent moduli are given. Verification of algorithm
is provided by two examples.
Ó 2003 Elsevier Ltd. All rights reserved.

Keywords: Finite strain; Thermoplasticity; Coupled systems; Finite element method

1. Introduction

It is well known that the plastic deformation of metals is accompanied by the heat
generation. This means that the energy balance equation, as an equation that gov-
erns temperature evolution, should involve several terms arising from thermome-
chanical coupling. One term relates heat production to recoverable deformations.
Second term defines heating arising from dissipation of mechanical work during

*
Corresponding author. Tel.: +385-51-651-496; fax: +385-51-651-490.

E-mail address: markoc@rijeka.riteh.hr (M. Cana dija).


0749-6419/$ - see front matter Ó 2003 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijplas.2003.11.016
1852 
M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874

breaking of internal bonds in crystal lattice. Third term describes stored energy of
cold work, which is motivated by the rearrangement of various defects in structure
during plastic deformation. These effects are important in some cases in small strain
regime and cannot be avoided in the most cases in finite strain regime. In addition,
heating leads toward softening of metals through lowering of yield strength of metals
and variation of other material properties with temperature. Since such variation of
material properties can also trigger generation of plastic strain, we can generally
state that thermomechanical processes undergoing finite plastic strains can be either
mechanically or thermally driven. In both cases, frequently large variations in
temperature field are involved and therefore calculations should use variable me-
chanical properties.
Temperature dependency of yield stress curve was previously addressed in Simo
and Miehe (1992) and Armero and Simo (1993), for example. Introduction of
temperature dependency of all material parameters in question is relatively rare in
literature. It can be found specialized for the solidification problem in de Saracibar
et al. (1999), Cervera et al. (1999), Celentano (2001) and Celentano (2002). Tem-
perature dependency of the part of mechanical work that is stored as energy of cold
work in crystal lattice can also play significant role in a proper simulation of de-
formation process. This problem is far from solved (see Kamlah and Haupt, 1998
and Rosakis et al., 2000) due to large number of variables involved. Besides tem-
perature, the most influential variables are strain and strain rate.
Numerical investigations in the thermoplasticity are frequently oriented toward
simulation of necking process (see Lehmann and Blix, 1985; Simo and Miehe,
1992; Wriggers et al., 1992; Armero and Simo, 1993; Ibrahimbegovic and Chorfi,
2002). In addition to thermomechanical nature, issues regarding stability are also
addressed within the context of necking. Therefore, necking is a fairly demanding
test problem for a numerical procedure. Generalizations that include rate de-
pendent (thermoviscoplastic) at finite strains material behavior are presented in
Srikanth and Zabaras (1999), Lion (2000), Batra and Chen (2001), Ibrahimbe-
govic and Chorfi (2002), Haslach (2002) and Adam and Ponthot (2003). Most of
the above-cited works are based on hyperelastic formulation and multiplicative
decomposition of deformation gradient into elastic and plastic part. Since mul-
tiplicative decomposition introduced in this way involves so-called intermediate
(stress-free) configuration, this can lead to certain problems. One of the most
serious difficulties is inability to describe anisotropy. This can be avoided with
the concept of material isomorphism. Consequently, isomorphism is recently in-
troduced into thermoplasticity, Bertram (2003) and Dachkovski and B€ ohm
(2004).
This research is directed toward formulating a more general framework for
thermoplasticity in the finite strain regime. In particular, a temperature dependency
of all material properties is introduced. In that way, we believe that a more variety of
problems can be analyzed.
The core of this research is today already classical paper by Simo and Miehe
(1992). Consequently, proposed model is based on the hyperelastic formulation and
multiplicative decomposition of deformation gradient.

M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874 1853

Paper is structured as follows. Section 2 describes local governing equations and


proposed thermoplastic model. It starts with a brief description of kinematics and
introduces basic notation. Summary of balance equations follow and local dissipation
inequalities arising from the second law of thermodynamics are presented. This gives
insight into evolution equations for stress and entropy. A J2 model of multiplicative
thermoplasticity proposed by Simo and Miehe (1992) is augmented with temperature-
dependent material properties giving more general model. A thermoplastic model is
completed by specifying a flow rule and an adequate yield criterion. Section ends with
description of temperature evolution and suitable definition of mechanical dissipation
and structural elasto-plastic heating in this context. Additional terms arise in the
structural elasto-plastic heating as a result of the temperature dependency.
Next section deals with the numerical implementation of the proposed continuum
model. It is based on isothermal staggered split into mechanical and thermal phase.
Spatial discretization is carried out by mixed finite element method. Corresponding
algorithmically consistent tangent operators are presented for both phases. Due to
temperature dependency, thermal tangent operator takes new form.
Proposed model is applied on two examples previously considered by Simo and
Miehe (1992) and others. Introduction of temperature-dependent properties gives
some new insight into these examples.
Some conclusions are drawn at the end of the paper.

2. Local governing equations and thermoplastic model

2.1. Basic kinematics

Let open set X 2 Rndim with smooth boundary oX, X ¼ X [ oX, describes refer-
ence placement of a continuum body B, ndim being the space dimension. Then con-
figuration is one-to-one smooth mapping uð; tÞ : X ! Rndim and defines
deformation of such body, where t 2 ½0; T  is the time period of interest. We assume
that inverse u1 exists and that both u and u1 are differentiable as many times as
necessary. Particles of body are labeled with their initial coordinates X or in current
configuration with x. Deformation gradient is derivative of the configuration
ouðXÞ
FðXÞ ¼ : ð1Þ
oX
Definition of the deformation gradient implies that two configurations can be
compared without knowledge of the intermediate configurations. Only initial and
current configurations are accounted for. Jacobian J ¼ detF gives volume change
and relates initial and current densities as q0 ¼ J q.
We assume that there exists a unique intermediate local configuration which
characterize elastic response. Neighborhood of each material point X 2 B is firstly
plastically stressed to the intermediate configuration and then elastically into final
configuration. This gives rise to the well-known multiplicative decomposition of the
deformation gradient
1854 
M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874

FðX; tÞ ¼ Fe ðX; tÞFp ðX; tÞ; ð2Þ


where superscripts ‘‘e’’ and ‘‘p’’ denote elastic and plastic part of deformation gra-
1
dient, respectively. In that way, unloading ðFe Þ would lead toward stress-free in-
termediate ‘‘configuration’’. Generally, unloading of a plastically deformed body
results in a residual stress field in a continuum body B. Therefore, intermediate
‘‘configuration’’ introduced in this way violates basic assumptions of a continuity
and is valid only locally. Plastic part of deformation gradient Fp represents plastic
flow of material by dislocation movement through the crystal lattice.
We also introduce mapping hð; tÞ : X ! R that defines absolute temperature
distribution in X.
In the text that follows, two strain measures are used: the plastic part of the right
Cauchy–Green tensor
T
Cp ¼ ðFp Þ Fp ; ð3Þ
and elastic left Cauchy–Green tensor
1
be ¼ FðCp Þ FT : ð4Þ

2.2. Balance equations

The coupled thermomechanical initial boundary value problem is governed by


momentum and energy balance equations, given in local form

q0 V_ ¼ DIV P þ f
: ð5Þ
q0 E_ þ DIV Q ¼ S  D þ q0 R
In above equations q0 is the reference density, f are prescribed body forces per unit
reference volume, DIV is divergence operator in reference configuration, PðX; tÞ and
SðX; tÞ are first and second Piola–Kirchhoff stress tensors, respectively; superim-
posed dots denote time derivatives. V is the velocity field, EðX; tÞ is the internal
energy per unit mass, QðX; tÞ is the heat flux, RðX; tÞ is the heat input per unit mass,
DðX; tÞ is the material rate of the right Cauchy–Green strain tensor CðX; tÞ, related as
2DðX ; tÞ ¼ oCðX ; tÞ=ot (Marsden and Hughes, 1994).
In addition to (5)1 boundary conditions in the time period of interest ½0; T  are
also supplied; deformation is prescribed at the part of the boundary ou X  oX
u ¼ u; ð6Þ
and nominal traction vector is prescribed at part of the boundary os X  oX with unit
normal N
t ¼ PN ¼ t; ð7Þ
alongside with the standard assumption ou X \ ot X ¼ ; and ou X [ ot X ¼ oX. For the
quasi-static case of interest, only initial conditions for configuration are specified
u ¼ u0 : ð8Þ
We also introduce free energy as a part of the internal energy E available for work
at constant temperature

M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874 1855

W ¼ E  HN ; ð9Þ
or in spatial configuration with N ð X ; tÞ ¼ gðx; tÞ and HðX ; tÞ ¼ hðx; tÞ
w ¼ e  hg: ð10Þ
N and g being material and spatial entropy, respectively. We will use the same
functional form as in Simo and Miehe (1992), namely e ¼ eðbe ; n; ge Þ. Here n rep-
resents internal strain hardening variable and ge elastic part of entropy.
A thermoplastic material is described in full by specifying a free energy function, a
constitutive equation for the heat flux vector, evolution equations for internal
variables and a yield criterion. For the analysis of the hyperelastic material behavior
considered in this work, the form of free energy function is of prime interest, since it
represents a basis for the stress calculation.
It should be noted, however, that principle of objectivity restricts present hy-
perelastic formulation to isotropic materials, Simo and Miehe (1992) or Marsden
and Hughes (1994). In that way, anisotropy cannot be treated with this theory. Free
energy function and yield criterion must be also consistent with such observations.

2.3. J2 model of multiplicative plasticity

The basis for calculation of stress tensor in hyperelastic materials is the Helmholtz
free energy function. Definition of this function follows Simo and Miehe (1992) at
this point but with more general form of free energy function, more specifically with
temperature dependency of all material properties.
Free energy function is defined in the terms of variables given in spatial config-
uration
 
^ ¼ T^ ðhÞ þ MðJ
w ^ ; hÞ þ U ^ be ; h þ Kðn;
^ ðJ ; hÞ þ W ^ hÞ: ð11Þ

We define terms in the free energy function as follows. The purely thermal part T^ ðhÞ
represents potential for the purely thermal entropy. Following form will be con-
sidered:
Z h Z h
d^
h
^
T ðhÞ ¼  dh q0 c0 ð^
hÞ ; ð12Þ
h0 h0 ^
h
where c0 ðhÞ is temperature-dependent heat capacity of the material at constant de-
formation. With this definition at hand, specific heat capacity in the case of material-
dependent properties takes form
 
^ ¼ h o2 T^  h o2 M
cðhÞ ¼ h o2hh w ^ þU ^ þW
^ þK^
hh hh
 
¼ c0 ðhÞ  h o2hh M^ þU^ þW^ þK ^ : ð13Þ
^ ; hÞ describes thermoelastic coupling arising from thermal dilatations.
Potential MðJ
Although several options are available at this point, linear theory is widely accepted
choice for metals
1856 
M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874

h i
^ ; hÞ ¼ ðh  h0 Þ  3aðhÞoJ U
MðJ ^ ðJ ; hÞ ; ð14Þ

where aðhÞ is linear coefficient of thermal dilatation and h0 is reference tempera-


^ ðJ ; hÞ corresponds to volume changes and is taken to
ture. Volumetric potential U
be
 
^ 1 1 2
U ðJ ; hÞ ¼ jðhÞ ðJ  1Þ  ln J ; ð15Þ
2 2
where jðhÞ stands for bulk modulus. This is a basis for calculation of pressure part of
stress tensor. For a recent review of various alternatives for volumetric potential, see
Doll and Schweizerhoff (2000). Potential for deviatoric part of stress tensor is
  h   i
W^ be ; h ¼ 1 lðhÞ tr be  3 ; ð16Þ
2
e
where lðhÞ stands for temperature-dependent shear modulus. b represents isochoric
e 2=3 e
part of be , related through b ¼ ðJ e Þ b and J e ¼ detFe . This potential defines
change in free energy caused by isochoric deformations. Hardening potential rep-
resents an extension of the infinitesimal case, Armero and Simo (1993)
9
^ hÞ ¼ 1 hðhÞn2 þ ½y0 ðhÞ  y1 ðhÞH ðnÞ >
Kðn; (2 =
ð1edn Þ ; ð17Þ
H ðnÞ ¼ n  for d ¼
6 0 >
d ;
0 for d ¼ 0
where d is saturation exponent and functions y0 ðhÞ; y1 ðhÞ; hðhÞ describe linear soft-
ening
9
y0 ðhÞ ¼ y0 ðh0 Þ½1  x0 ðh  h0 Þ =
hðhÞ ¼ hðh0 Þ½1  xh ðh  h0 Þ ; ð18Þ
;
y1 ðhÞ ¼ y1 ðh0 Þ½1  xh ðh  h0 Þ
while x0 and xh are flow stress softening and hardening softening parameter, re-
spectively. Eqs. (17) and (18) imply that solely isotropic hardening is considered in
this model. For extensions to kinematic hardening in the purely mechanical case
reader is refereed to Simo (1988a,b, Parts I and II) and Ibrahimbegovic and Chorfi
(2000) and for the thermoviscoplastic case to Ibrahimbegovic and Chorfi (2002)
among others.

2.4. Local dissipation inequalities

System of partial differential Eq. (5) is constrained by the second law of ther-
modynamics, Truesdell and Noll (1965)
)
1
cloc ¼ g_  hr þ qh div q P 0
; ð19Þ
ccon ¼  qhq 2  grad h P 0

where cloc denotes local entropy production, ccon entropy due to heat conduction, q
current density, r heat supply per unit mass and q is heat flux vector. First of the

M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874 1857

above equations ensures that thermal energy cannot be converted into mechanical
work in a system of uniform temperature and without heat sources; opposite process
are allowable. Second equation does not allow that heat spontaneously flows from
colder to warmer bodies.
Energy balance Eq. (5)2 and Legendre transformation (10) can be rearranged to
give following equation, Simo and Miehe (1992)
1
hc ¼ _e þ gh
_ þ s  d  q  grad h P 0; ð20Þ
h
where the product D ¼ hc ¼ hðcloc þ ccon Þ is called internal dissipation, s is the
Kirchhoff stress tensor and d is the spatial rate of deformation tensor.
In the line with (19), dissipation production can be separated into two parts –
local dissipation and dissipation due to heat conduction, Armero and Simo (1993)
D ¼ Dloc þ Dcon ; ð21Þ
where are

Dloc ¼ hcloc ¼ _e þ gh
_ þ s  dP0
: ð22Þ
Dcon ¼ hccon ¼  1h q  grad h P 0
With the general definition of free energy function (10) and specific form for the
problem at hand (11) and time derivative of some kinematic quantities, local dissi-
pation can be transformed into the following form:

Dloc ¼ ½oh w  ðg  gp Þh_ þ ½s  2obe wbe   d


 
1 e 1
e
þ ð2ob wb Þ   ðLm b Þðb Þ
e
e
þ g_ p h  on w  n_ P 0; ð23Þ
2
where Lm ðÞ denotes Lie derivative and gp is the plastic part of entropy caused by
dissipative plastic structural changes. Following arguments in Coleman and Gurtin
(1967), constitutive equations for the Kirchhoff stress tensor and entropy are ob-
tained:

s ¼ 2obe wbe
: ð24Þ
g ¼ gp  oh w
e
The free energy function is dependent on isochoric part b of left elastic Cauchy–
e
Green tensor b through deviatoric potential, Eq. (16). Therefore, using chain rule
with (24) we obtain following suitable form of evolution equation for Kirchhoff
stress tensor:
 e

s ¼ 2 obe w  obe b be : ð25Þ

At this moment it is convenient to introduce elastic entropy corresponding to the


non-plastic dissipative processes
ge ¼ oh w: ð26Þ
1858 
M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874

Now, Eq. (23) can be rewritten into reduced form of dissipation inequality, Simo and
Miehe (1992)
 
1 e 1
e
Dloc ¼ s   ðLm b Þðb Þ þ bn_ þ g_ p h ¼ Dmech þ Dther P 0; ð27Þ
2
where b ¼ on w and
 
1
Dmech ¼ s   ðLm be Þðbe Þ1 þ bn_ Dther ¼ g_ p h: ð28Þ
2
This equation should be augmented by adequate constitutive equations.

2.5. Maximal plastic dissipation and flow rule

Thermoelastic domain is defined as


E :¼ fðs; b; hÞ : /ðs; b; hÞ 6 0g: ð29Þ
Boundary oE defines yield surface in stress space and therefore plastic state of ma-
terial, while interior defines elastic state of material. States outside domain are non-
admissible in classical plasticity. Function / in (29) is known as yield criteria. There
are several possible choices for this function, however, for most metals von Mises
yield criteria is adequate (Khan and Huang, 1995). In the isotropic hardening con-
text it takes the following form:
rffiffiffi
2

/ðs; b; hÞ ¼ kdev sk þ b  ry ðhÞ 6 0; ð30Þ


3
where ry ðhÞ is yield stress function.
Now the principle of maximum plastic dissipation can be applied (Simo and
Hughes, 1998 and Simo and Miehe, 1992): from all admissible states ðs ; b ; h Þ
plastic dissipation attains its maximum for the actual state ðs; b; hÞ. Standard opti-
mization procedure (Luenberger, 1984) yields following evolution equations:
1
 Lm be ¼ kos /  be ; n_ ¼ kob /; g_ p ¼ koh /; ð31Þ
2
with
kP0 /ðs; b; hÞ 6 0 k  /ðs; b; hÞ ¼ 0; ð32Þ
and consistency condition
_ b; hÞ ¼ 0:
k  /ðs; ð33Þ

2.6. Temperature evolution

To obtain temperature evolution equation, we start from the balance of energy


Eq. (5)2 . Using Piola transformation (Marsden and Hughes, 1994) and Kirchhoff
stress tensor as stress measure, we obtain

M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874 1859

Q
J DIV þ r ¼ e_  s  d: ð34Þ
J
This equation is transformed using (22)1
Q
J DIV þ r ¼ hg_  Dloc ¼ hðg_  g_ p Þ  Dmech : ð35Þ
J
Now, from (24)2 with (10) and (28) after some elementary transformations we obtain
_
hg_ ¼ hg_ p  oh ½s  d  Dmech   h o2hh wh: ð36Þ
The last term in above equation is the specific heat capacity, defined by (13), while
the term in brackets is known as elasto-plastic structural heating
H ¼ oh ½s  d  Dmech : ð37Þ
Therefore, Eq. (36) can be written

hðg_  g_ p Þ ¼ ch_ þ H: ð38Þ


Eq. (35) with (38) now gives equation that governs temperature evolution
Q
ch_ ¼ Dmech  H  J DIV þ r: ð39Þ
J
We also need to specify boundary conditions for the absolute temperature at
boundary oh X  oX

h ¼ h; ð40Þ
and prescribe heat flux at the boundary oQ X  oX with normal N
Q ¼ Q  N; ð41Þ
where oh X \ oQ X ¼ ; and oh X [ oQ X ¼ oX. Initial condition is
h ¼ h0 : ð42Þ
Finally, constitutive equation for the heat flux needs to be specified. For most
metals Fourier law of heat conduction is usual choice
Q ¼ K  GRAD H: ð43Þ
Since we are dealing with isotropic materials, tensor function that describes heat
conduction coefficients KðHÞ collapses to the scalar valued function kðhÞ. Positive
value of heat conduction coefficient fulfills constraint imposed by inequality (22)2 .

2.7. Mechanical dissipation and structural elasto-plastic heating

Mechanical dissipation enters into the temperature evolution Eq. (39). Therefore,
this term needs to be defined. During plastic deformation a large amount of me-
chanical plastic power is dissipated as heat. However, plastic power is not completely
transformed into thermal power. Instead, a part of expended work is left in the
1860 
M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874

crystal lattice as a result of dislocations interaction, among other phenomena.


Amount of stored energy is dependent on several variables where strain rate, strain
level and temperature level are the most significant. To be consistent with such ex-
perimental observation, mechanical dissipation is frequently calculated as fraction of
used plastic power Ppmech (Simo and Miehe, 1992)
rffiffiffi
p 2
Dmech ¼ vPmech ¼ v k b  ry ðhÞ ; ð44Þ
3
where v 2 ½0; 1 is dissipation factor. As described in Cervera et al. (1999), Zdebel
and Lehmann (1987), Kamlah and Haupt (1998) and Rosakis et al. (2000), this value
is usually treated as a constant with values 0.85–0.95. However, as noted in Rosakis
et al. (2000) this factor can reach even value 0.30 in certain cases. Moreover, same
work presents results when amount of stored energy varies from about 0.60 to 0.30
over plastic strain. This is a strong argument that amount of stored energy should be
treated as function. This is particularly true for the aluminum and titan. Never-
theless, due to lack of information on specific materials, to be more precise which
variables influence on the process and in what amount, this work will treat dissi-
pation factor as constant. We emphasize that this approach can lead toward errors
that cannot be neglected in calculations of temperature and other fields. From the
numerical aspect, introduction of functional dependence of dissipation factor would
lead toward further complication of tangent stiffness matrices but should not rep-
resent a significant problem.
Eq. (36) implies existence of additional coupling between mechanical and thermal
field through elasto-plastic structural heating H. In the simpler case of thermo-
elasticity externally applied mechanical energy results in variation of strain field. Eq.
(36) implies that such variations also must induce variations in temperature field.
Since this is accompanied by the heat flow, increase of entropy also takes place. This
process is known as thermoelastic dissipation. In the specific case when external
loading is accompanied by the small temperature variations, one can conclude that
deformations can be calculated without considering thermal expansion of the body
in question. On the other side, if non-homogenous temperature field is the source of
strain in the body, it is reasonable to conclude that these strains have a small in-
fluence on the temperature field. Therefore, frequent practice in linearized thermo-
elasticity is to neglect this coupling term what makes analysis significantly simpler.
However, in cases when larger variations in strain and temperature field occur,
question on justification of disregarding of this term becomes more complicated. In
the line with such considerations, we do not neglect this term.
Elasto-plastic structural heating (38) can be recast into the following form:
 
H ¼ h o2hbe w  b_ e þ o2hn wn_ ; ð45Þ

or with (26)
 
H ¼ h obe ge  b_ e þ on ge n_ : ð46Þ

M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874 1861

If structure of free energy function w is now considered, we get


 
H ¼ h o2hbe W  b_ e þ o2hbe U  b_ e þ o2hbe M  b_ e þ o2hn K  n_
 e

¼ h o2hbe W obe b  b_ e þ o2hJ U obe J  b_ e þ o2hJ Mobe J  b_ e þ o2hn K  n_ : ð47Þ

For potentials defined in Section 2.3, we get following partial derivatives:


9
o2hbe W ¼ 12 oh lI >
>
>
>
2 1 1
ohJ U ¼ 2 oh j J  J =
; ð48Þ
o2hJ M ¼ 3 oh aðh  h0 Þo2JJ U þ aðh  h0 Þo3JJ h U þ ao2JJ U >>
>
>
o2 K ¼ no h þ ðo y  o y Þð1  edn Þ ;
hn h h 0 h 1

with
 
1 1
o3JJ h U ¼ oh j 1 þ 2 : ð49Þ
2 J

Following methodology in Rosakis et al. (2000), when mechanical dissipation is


calculated with (44) structural elasto-plastic heating should exclude any hardening
effects. In this case, we arrive at expression that is used in further elaboration
 e

H ¼ h o2hbe W obe b  b_ e þ o2hJ U obe J  b_ e þ o2hJ Mobe J  b_ e ; ð50Þ

with partial derivatives defined with (48) and (49).

3. Numerical implementation

Numerical solution of the coupled thermomechanical initial boundary value


problem can be approached by the simultaneous time-stepping algorithms or
staggered time-stepping algorithms. Simultaneous approach leads toward larger
systems of equations with unsymmetrical stiffness matrices. Staggered schemes
uncouple problem into two smaller problems that typically involve symmetric
systems of equations. Such approach also opens possibility of employment of
different time scales for thermal and mechanical phase. However, uncoupling often
leads toward algorithms that are not unconditionally stable. For example, widely
used isothermal split into thermal and mechanical phase is only conditionally
stable (Simo and Miehe, 1992). Nevertheless, unconditional stability can be
achieved by the so-called adiabatic split (Armero and Simo, 1992; Armero and
Simo, 1993). In the adiabatic split, isothermal mechanical phase is substituted with
adiabatic mechanical phase in which entropy is fixed. Change of entropy occurs in
second phase. Although instability of the isothermal split is not significant for
metal materials (Simo and Miehe, 1992) it is obviously an undesirable phenome-
non. Since instability problems can be expected only at strong coupling effects, we
will use more classical isothermal split.
1862 
M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874

3.1. Isothermal staggered split

Standard isothermal split uncouples problem into isothermal (mechanical) and


heat conduction (thermal) phase. In the thermal phase displacements are held con-
stant, while in the mechanical phase temperature field is fixed. For the time inte-
gration backward Euler difference scheme is used.
Evolution of the primary variables ðu; m; hÞ in the mechanical phase is governed by
the following system of differential equations:
9
u_ ¼ m =
q0 m_ ¼ DIV½sðu; h; kÞFt  þ f : ð51Þ
_ ;
ch ¼ 0
Quasi-static problem is obtained by letting reference density q0 tends toward zero.
Eq. (51)3 enforces constancy of temperature during this phase. If Euler backward
integration scheme is applied to (51) for the time interval of interest X  ½tn ; tnþ1  we
arrive at
~ nþ1 un
u
9
¼ ~m nþ1 >
=
Dt

0 ¼ DIV ðdev snþ1 þ pnþ1 Jnþ1 IÞFt


nþ1 þ f nþ1 ; ð52Þ
~
hnþ1 hn
>
;
Dt
¼ 0
where we also employed definition of first Piola–Kirchhoff stress tensor and sepa-
ration of stress tensor into deviatoric and pressure part. Boundary conditions are
prescribed at C ¼ oX for configuration u : Cu  ½tn ; tnþ1  and for surface tractions
t : Ct  ½tn ; tnþ1  with Cu \ Ct ¼ f0g and Cu [ Ct ¼ oX. Body forces f nþ1 ðX ; tÞ are
prescribed functions in X  ½tn ; tnþ1 . Initial values are converged values from pre-
vious increment fben ; nn ; hn g. Solution of this system gives intermediate values
f~benþ1 ; ~nnþ1 ; ~
hnþ1 g that are initial values for subsequent thermal phase.
Thermal phase follows mechanical phase. In this phase following differential
equations need to be solved:
9
u_ ¼ 0 =
m_ ¼ 0 : ð53Þ
;
ch_ ¼ DIV½Ft qðu; hÞ þ Dmech ðh; kÞ  Hðu; m; hÞ þ R
Eq. (53)1 does not allow any alternations of configuration in this phase. Application
of Euler backward scheme yields
9
unþ1  u ~ nþ1 ¼ 0 >
=
mnþ1  ~m nþ1 ¼ 0 : ð54Þ
hnþ1 ~
t
>
hnþ1
cnþ1 Dt ¼ DIV Fnþ1 qnþ1 þ ðDmech Þnþ1  Hnþ1 þ Rnþ1 ;

Boundary conditions are prescribed temperatures h : Ch  ½tn ; tnþ1  and heat flux at
the boundary with normal n defined as q ¼ q  n, q : Cq  ½tn ; tnþ1  with
Ch \ Cq ¼ f0g and Ch [ Cq ¼ oX. Heat flux at the boundary can be of convective
nature, for example
ðqc Þnþ1 ¼ hnþ1 ðhe  hnþ1 Þ; ð55Þ

M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874 1863

where hnþ1 is convective coefficient, or radiation boundary condition


ðqr Þnþ1 ¼ ðjr Þnþ1 ðhr  hnþ1 Þ; ð56Þ
where ðjr Þnþ1 is a coefficient
h i
2 2
ðjr Þnþ1 ¼ ðhr Þnþ1 ðhr Þ þ ðhnþ1 Þ ðhr þ hnþ1 Þ; ð57Þ

where ðhr Þnþ1 represents coefficient determined from emissivity of interacting mate-
rials and geometric factors. hr is temperature of the radiative source. It is to be
emphasized that except for large differences in temperature, radiative heat exchange
is usually neglected in analysis of metals. Heat source Rnþ1 ðX ; tÞ is treated as given in
X  ½tn ; tnþ1 .
Initial values for this phase are converged values from previous mechanical phase
f~benþ1 ; ~
nnþ1 ; ~
hnþ1 g. Solution gives converged values at the end of current increment
fbenþ1 ; nnþ1 ; hnþ1 g.

3.2. Discrete variational formulation for mixed finite element method implementation

In a standard finite element method displacements are used to calculate strain,


stress and other relevant fields. However, in the analysis of almost incompressible
media, displacement based finite elements lose its effectiveness. This means that al-
though such elements can be used in almost incompressible analyses, significantly
greater number of elements compared to similar compressible analysis is required.
The problem lies in calculation of pressure part of stress tensor due to its connection
to volume change. Unwanted effects in this regime include ill-conditioning of stiffness
matrix, spurious pressure modes and locking. In the case of totally incompressible
media, such as plastic deformation, purely displacement based finite element can not
be used.
Therefore, in order to avoid difficulties associated with displacement based finite
element method, an alternative approach should be employed. We follow procedure
proposed in Simo and Miehe (1992) and use mixed finite element method.
In the mechanical part of the algorithm besides displacement field, additional field
for pressure p and Jacobian J are introduced. In this way, pressure part of stress
tensor is completely decoupled from deviatoric part of stress tensor and displacement
field. Starting from equilibrium Eq. (52)2 , for all admissible test functions u fol-
lowing weak form of linearized equation at time tnþ1 is obtained:
Z Z Z
 

ru  dev snþ1 þ div u pnþ1 Jnþ1 dX þ tnþ1  u dC þ f nþ1  u dX ¼ 0;




X oX X
ð58Þ
with
ru ¼ GRAD u Ft 
nþ1 ¼ grad u : ð59Þ
In addition to (58), due to mixed formulation following equations also needs to be
established:
1864 
M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874

R  
p  detFnþ1  J nþ1 dX ¼ 0
R 
X
; ð60Þ
X
J  pnþ1 þ oJ U J nþ1 ; hnþ1 þ oJ M J nþ1 ; hnþ1 dX ¼ 0
Eqs. (26) and (46) and the structure of the free energy function (11) implies that
elastic entropy is also a function of volume change and temperature. Therefore,
additional mixed fields of elastic entropy ge and temperature h are introduced in
thermal phase. Now, from (54)3 with convective (55) and radiative boundary con-
ditions (56) and (57) for all admissible test functions # we arrive at the following
weak formulation at time instant tnþ1 :
R n  h hnþ1 ~hnþ1 i

o )
X
# c Dt
 ðD mech Þ nþ1 þ Hnþ1 þ ½ k nþ1 rh nþ1  r#  dX
R R R ; ð61Þ
 X # Rnþ1 dX  oX # ½ðqc Þnþ1 þ ðqr Þnþ1  dC  oX # ½qnþ1  dC ¼ 0
and due to mixed interpolations
R  )
g  ðhnþ1 þ hnþ1 Þ dX ¼ 0
X
R  h  h
  eh h
  h
 i :
h h
X
h   ðge Þnþ1  oh M J ; h  oh W b ; h  oh U J ; h dX ¼ 0
ð62Þ
Configuration and temperature fields in above equations are approximated via
standard isoparametric formulation, see Hughes (2000) for example. Mixed fields
are approximated an order lower than configuration and temperature field. Fre-
quent choices are Q1/P0 and Q2/P1 elements (Sussman and Bathe, 1987; Brezzi
and Fortin, 1991). Although mixed variables can be continuous at the mesh level,
continuity only at the element level is usually preferred. In this approach mixed
variables can be statically condensed out at the element level prior to the global
assemblage. In the post-processing phase these are recovered again at the element
level.

3.3. Algorithmically consistent tangent matrices

Solution of non-linear problem described above can be obtained via Newton–


Raphson method. In order to achieve quadratic convergence exhibited by this
procedure, consistent linearization of the problem must be carried out (Simo and
Taylor, 1985). Tangent moduli obtained in that way must be consistent with inte-
gration procedure used. Any deviation from the algorithmically consistent tangent
moduli results in the loss of quadratic convergence of Newton–Raphson procedure.
Use of tangent moduli consistent with continuum rates and not with numerical al-
gorithm also results in deterioration of convergence.
Linearization of the mechanical weak form (58) in the direction of the displace-
ment u leads toward tangent matrices that are algorithmically consistent. For these
types of analyses, stiffness matrix for the mechanical phase KM is decomposed into
geometric part KMgeo , material part KMmat and part that corresponds to follower
forces effect KML

KM ¼ KMgeo þ KMmat þ KML : ð63Þ



M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874 1865

Geometric and follower forces part takes usual form, see Wriggers (2001) or Zie-
nkiewicz and Taylor (2000). Material part is obtained after laborious procedure and
therefore only final results are summarized below
Z
KMmat ¼ ru : ½ pJ ðI  IÞ  2I þ cep
M
X
Z

: ru dX þ divu o2JJ U þ o2JJ M divu dX; ð64Þ


X

with div is operator introduced into Simo et al. (1985) and where are
cep e;trial
M ¼ cM þ cpM ; ð65Þ
and
 
1  e;trial  1 2
ce;trial
M ¼ 2l tr b I  I  I  dev strial  I
3 3 3
trial

þ I  dev s ; ð66Þ


cpM ¼ ap b1 ce;trial  b2 n  n  b3 n  dev ½n2   b4 n  dev strial ; ð67Þ
with coefficients
h e;trial i 9
o2 K
b0 ¼ 1 þ nne;trial
; b1 ¼ 23 lDk tr b 1
; >
>
kdev s k
trial >
>
l tr b
h e;trial in o   n o =
2
b2 ¼ 3 tr b l b0 þ b1 l ; b3 ¼ 2dev s  b0 þ b1 ; >
1 trial 1 ð68Þ
n o >
>
>
;
b4 ¼ 23 Dk lb1 0  23

and plastic flag



1 for /trial P 0;
ap ¼ ð69Þ
0 for /trial < 0:
Linearization of the thermal weak form in the direction of temperature change #
yields tangent operator for the thermal phase
Z   Z Z
h  hn 1 
KT ¼ # oh c þc # dX þ # oh H # dX þ r# foh krh þ kr#g dX
X Dt Dt X X
2 0 o h ry
13
Z rffiffiffi e;trial
 oh l
6 vDk B 2 2 Dk tr b C7
 # 4 @ ohn K  oh ry þ ðb  ry Þ on K A5# dX
X Dt 3

e;trial þ l
tr b
Z Z
þ # ½oh hðhe  hÞ  h# dC þ # ½oh jr ðhr  hÞ  jr # dC: ð70Þ
oX oX

Thermal tangent operator involves derivatives of the material parameters oh c, oh k,


oh l, oh jr , oh h. These are calculated from the chosen functional dependency on the
temperature.
1866 
M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874

4. Numerical simulations

4.1. Expansion of a thickwalled sphere

As an illustration of proposed numerical model an expansion of a thickwalled


sphere is considered. This example was previously considered in Simo and Miehe
(1992) in the context of constant material parameters, with the exception of tem-
perature dependency of yield curve.
Initial geometry of the sphere is given by inner radius A ¼ 10 mm and outer
B ¼ 20 mm. Temperature is constant throughout the sphere and equals h0 ¼ 293 K.
The sphere is loaded by internal pressure of pA ¼ 187:5 MPa. This pressure causes
initial straining of the sphere that can be characterized as purely elastic. To initiate
plastic deformation a thermal shock by temperature increase #B ¼ 333:333 K on the
outer boundary is introduced. Therefore, this is a thermally driven problem. The
process is simulated in time period t ¼ 7 s.
Material properties are given in Table 1. Material properties are taken to be linear
functions of current temperature.
Due to symmetry, only one quarter of the sphere cross-section was discretized by
a total of 100 axisymmetric mixed isoparametric quadrilateral finite elements. Cal-
culation was performed in 100 equal time steps. The convergence tolerance is 107
times the maximum value attained by length of variables vector during the incre-
ment. To deal with non-linearities, Newton–Raphson algorithm was used. Tem-
perature integration was performed by backward Euler integration scheme.
As already mentioned, pressure initially causes only elastic deformation. How-
ever, instant application of temperature increase leads toward occurrence of two
plastic zones: at the outer layer, where thermal shock is applied and consequently
flow stress is lower; and at the inner layer, where stresses are the greatest. These two
initially narrow zones propagate until the sphere is completely plastified, Fig. 3. In
the time instant of full plastification a fast increase of inner radius occurs, leading
toward so-called ‘‘blow-up’’. Temperature dependency of material properties leads
toward slower propagation of plastic strain and consequently later initiation of
‘‘blow-up’’ behavior, Fig. 1. However, at the instant of ‘‘blow-up’’, increase of

Table 1
Material properties of a thickwalled sphere
Shear modulus l 30:7h þ 8:99  104 MPa
Bulk modulus j 66:6h þ 1:95  105 MPa
Flow stress y0 300 MPa
Hardening modulus h 700 MPa
Density q 7:8  109 N s2 /mm4
Heat expansion coefficient a 2:0  108 h þ 4:04  106 K1
Conductivity k 1:66  102 h þ 5:02  101 N/s K
Heat capacity at constant deformation c0 3:1  103 h þ 2:8135 N/mm2 K
Dissipation factor v 0.9
Flow stress softening x0 0.003 K1
Hardening softening xh 0

M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874 1867

4.5

4.0

3.5
Displacement, mm
3.0

2.5

2.0

1.5 Const. mat. prop.

1.0 Temperature dep.


mat. prop.
0.5

0.0
0 1 2 3 4 5 6 7
Time, s

Fig. 1. Evolution of displacements at the inner side of sphere.

temperature at inner side is equal ( 120 °C) in constant and variable properties
problem, Fig. 2. The principal reason of ‘‘blow-up’’ expansion is sufficient lowering
of the yield stress through softening of the sphere. From Fig. 2 it is also evident that
variable properties algorithm yields slightly lower values of temperatures than
constant one. The sudden expansion of the sphere is clearly visible in Fig. 1. After
some time, expansion velocity starts to decrease and continue to do so toward the

350

300
Temperature increase, K

250

200

150 A - Const. mat. prop.

B - Temperature dep. mat.


100 prop.
Expansion init. - case A
50
Expansion init. - case B

0
0 1 2 3 4 5 6 7
Time, s

Fig. 2. Evolution of temperature increase at the inner side of sphere.


1868 
M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874

end of process. At the end of the process, sphere thickness is decreased from 10 to
7.46 mm. Furthermore, due to expansion, inner surface at which pressure acts is
significantly greater at the end. Since pressure is held constant in the sphere interior,
nodal forces should increase during time. In order to accomplish this, a simplified
version of follower forces algorithm (Wriggers, 2001) is also introduced.

Fig. 3. Equivalent plastic strain at t ¼ 0:84 s. Time step before full plastification (t ¼ 0:98 s).

Fig. 4. Temperature increase at t ¼ 7 s.



M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874 1869

Table 2
Material properties of the cylindrical specimen
Shear modulus l 3:07  101 h þ 8:99  104 MPa
Bulk modulus j 6:66  101 h þ 1:95  105 MPa
Flow stress y0 450 MPa
Linear hardening h 129.24 MPa
Saturation hardening y0;1 715 MPa
Hardening exponent d 16.93
Density q 7:8  109 N s2 /mm4
Expansion coefficient a 2:0  108 h þ 4:04  106 K1
Heat conductivity k 1:66  102 h þ 5:02  101 N/s K
Heat capacity at const. def. c0 3:10  103 h þ 2:8135 N/mm2 K
Dissipation factor v 0.9
Flow stress softening x0 0.002 K1
Hardening softening xh 0.002 K1

Fig. 3 presents two zones of plastic deformation an increment prior to full plas-
tification of the sphere. White zone represents elastic part of sphere. Increase in
temperature at the end of deformation is presented at Fig. 4.

4.2. Non-isothermal necking of a cylindrical specimen

This example is concerned with the simulation of experimental analysis of a cy-


lindrical specimen under axial load. Specimen is 53.334 mm long and 6.413 mm in
radius. The total axial elongation of the specimen is prescribed to be 16 mm.
Stretching is performed through time period of 8 s with a constant rate of stretch.
The same example is presented in Simo and Miehe (1992) with constant material

160
Const. mat. prop.
140
Temperature dep.
mat. prop.
120
Temperature, K

100

80

60

40

20

0
0 2 4 6 8 10 12 14 16
Elongation, mm

Fig. 5. Evolution of temperature increase at the middle of specimen (at surface). Constant and variable
material properties.
1870 
M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874

90000

80000

70000

60000
Force, N

50000

40000

30000

20000 Const. mat. prop.

10000 Temperature dep. mat.


prop.

0
0 2 4 6 8 10 12 14 16
Elongation, mm

Fig. 6. F –DL curves for constant and variable material properties.

parameters and used as a verification model for this work. Similar example, but with
the viscoplastic material behavior and also constant material parameters was treated
in Ibrahimbegovic and Chorfi (2002).
Due to axial symmetry, only one-quarter of the longitudinal cross-section needs
to be discretized. Furthermore, specimen ends are constrained in order to prevent
contraction. Finite element mesh consisted of 200 mixed quadrilateral isoparametric
finite elements, with a total of 242 nodes. Temperature field of the specimen is ini-
tially homogenous with reference temperature 293 K.
Heat exchange with the environment is considered through heat convection.
Coefficient of convection is taken to be h ¼ 17:5  105 N/mm s K. Environment
temperature does not change through the whole process and equals 293 K.
Material properties are linear functions of temperatures, Table 2.
Numerical simulation is performed in 200 equal time steps. Convergence toler-
ance was: variables (107 ), residuum (107 ) and energy (107 ). Newton–Raphson
method was used. Temperature integration was again performed by backward Euler
integration scheme.
This problem within the context of isothermal material behavior is of bifurcation
nature and frequent solution of this problem is introduction of linear imperfection in
the specimenÕs radius. However, we consider non-isothermal problem and in this
case such geometrical imperfection is no longer needed. Non-homogenous temper-
ature field is trigger for necking behavior. In this case, shortly before necking takes
place maximal temperature difference is 0.062 K. Although this is a relatively small
value, it is sufficient to initiate necking.
Fig. 5 presents evolution of temperature increase at the middle of specimen (at
surface). Constant properties lead toward reduction of released heat at the end of the

M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874 1871

Fig. 7. Temperature increase field (a) and von Mises stresses (b) at the end of the process.

process (Simo and Miehe, 1992; Ibrahimbegovic and Chorfi, 2002), while variable
properties yield increase of released heat. Difference in temperatures at the end be-
tween these two cases is about 40 K. Interestingly, the trend of increase of released
heat was also experimentally and numerically noticed by Lehmann and Blix (1985).
Therefore, this result at least qualitatively corresponds to these experimental
investigations.
Difference in F –DL curves for both cases are barely noticeable, Fig. 6.
Some other results are presented on Figs. 7 and 8. As expected, temperature field
is nearly constant towards the end of specimen, while in the middle strong self-
heating induced by the plastic deformation takes place, Fig. 7(a). In accordance to
this behavior, heat flux vector is directed toward colder parts of specimen, Fig. 8(a).
Inspection of radial component of heat flux vectors, Fig. 8(b), reveals heat exchange
with environment through convection.
1872 
M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874

Fig. 8. Heat flux vector (a) and radial component of heat flux vector (b) at the end of the process.

5. Concluding remarks

A formulation of coupled thermoplasticity suitable for finite strain analysis with


temperature-dependent properties has been presented. The constitutive model was
introduced within the context of continuum mechanics. Detailed J2 plasticity model
suitable for such formulation is described. A consistent tangent operator for both
phases of isothermal algorithmic split is given. Spatial discretization was carried out
by mixed finite element method. Such approach successfully circumvents problems
arising from isochoric behavior of plastic deformation. Numerical assessment of
proposed formulation was carried out by the means of two examples. Obtained
numerical results shows good correlation with similar existing numerical and ex-
perimental results. These investigations have given some new insight into standard
test examples and justified proposed more general approach.

M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874 1873

In authorsÕ opinion, further research should be directed toward coupling between


effects at micro and macro scale. Inclusion of rate-dependent plasticity, damage
evolution and solid–solid phase change should be also addressed.

References

Adam, L., Ponthot, J.P., 2003. A coupled thermo-viscoplastic formulation at finite strains for the
numerical simulation of superplastic forming. J. Mater. Process. Technol. 139, 514–520.
Armero, F., Simo, J.C., 1992. A new unconditionally stable fractional step method for non-linear coupled
thermomechanical problems. Int. J. Numer. Meth. Eng. 35, 737–766.
Armero, F., Simo, J.C., 1993. A priori stability estimates and unconditonally stable product formula
algorithms for nonlinear coupled thermoplasticity. Int. J. Plasticity 9, 749–782.
Batra, R.C., Chen, L., 2001. Effect of viscoplastic relations on the instability strain, shear band initiation
strain, the strain corresponding to the minimum shear band spacing, and the band width in
thermoviscoplastic material. Int. J. Plasticity 17, 1465–1489.
Bertram, A., 2003. Finite thermoplasticity based on isomorphisms. Int. J. Plasticity 19, 2027–2050.
Brezzi, F., Fortin, M., 1991. Mixed and Hybrid Finite Element Methods. Springer, New York.
Celentano, D.J., 2001. A large strain thermoviscoplastic formulation for the solidifcation of S.G. cast iron
in a green sand mould. Int. J. Plasticity 17, 1623–1658.
Celentano, D.J., 2002. A thermomechanical model with microstructure evolution for aluminium alloy
casting processes. Int. J. Plasticity 18, 1291–1335.
Cervera, M., de Saracibar, C.A., Chiumenti, M., 1999. Thermo-mechanical analysis of industrial
solidification processes. Int. J. Numer. Meth. Eng. 46, 1575–1591.
Coleman, B.D., Gurtin, M.E., 1967. Thermodynamics with internal state variables. J. Chem. Phys. 47,
597–613.
Dachkovski, S., B€ ohm, M., 2004. Finite thermoplasticity with phase changes based on isomorphisms. Int.
J. Plasticity 20, 323–334.
de Saracibar, C.A., Cervera, M., Chiumenti, M., 1999. On the formulation of coupled thermoplastic
problems with phase change. Int. J. Plasticity 15, 1–34.
Doll, S., Schweizerhoff, K., 2000. On the development of volumetric strain energy functions. ASME J.
Appl. Mech. 67, 17–21.
Haslach, H.W., 2002. A non-equilibrium thermodynamic geometric structure for thermoviscoplasticity
with maximum dissipation. Int. J. Plasticity 18, 127–153.
Hughes, T.J.R., 2000. The Finite Element Method – Linear Static and Dynamic Finite Element Analysis.
Dover, Mineola, NY.
Ibrahimbegovic, A., Chorfi, L., 2000. Viscoplasticity model at finite deformations with combined isotropic
and kinematic hardening. Comput. Struct. 77, 509–525.
Ibrahimbegovic, A., Chorfi, L., 2002. Covariant principal axis formulation of associated coupled
thermoplasticity at finite strains and its numerical implementation. Int. J. Solids Struct. 39, 499–528.
Kamlah, M., Haupt, P., 1998. On the macroscopic description of stored energy and self heating during
plastic deformation. Int. J. Plasticity 13, 893–911.
Khan, A.S., Huang, S., 1995. Continuum Theory of Plasticity. Wiley, New York.
Lehmann, Th., Blix, U., 1985. On the coupled thermo-mechanical process in the necking problem. Int. J.
Plasticity 1, 175–188.
Lion, A., 2000. Constitutive modelling in finite thermoviscoplasticity: a physical approach based on
nonlinear rheological models. Int. J. Plasticity 16, 464–494.
Luenberger, D.G., 1984. Linear and Nonlinear Programming. Addison-Wesley, Reading, MA.
Marsden, J.E., Hughes, T.J.R., 1994. Mathematical Theory of Elasticity. Dover, Mineola, NY.
Rosakis, P., Rosakis, A.J., Ravichandran, G., Hodowany, J., 2000. A thermodynamic internal variable
model for the partition of plastic work into heat and stored energy in metals. J. Mech. Phys. Solids 48,
581–607.
1874 
M. Canad
ija, J. Brnic / International Journal of Plasticity 20 (2004) 1851–1874

Simo, J.C., 1988a. A framework for finite strain elastoplasticity based on maximum plastic dissipation and
the multiplicaitve decomposition. Part I. Continuum formulation. Comput. Methods Appl. Mech.
Eng. 66, 199–219.
Simo, J.C., 1988b. A framework for finite strain elastoplasticity based on maximum plastic dissipation and
the multiplicative decomposition. Part II. Computational aspects. Comput. Methods Appl. Mech. Eng.
68, 1–31.
Simo, J.C., Hughes, T.J.R., 1998. Computational Inelasticity. Springer, New York.
Simo, J.C., Miehe, C., 1992. Associative coupled thermoplasticity at finite strains: formulation, numerical
analysis and implementation. Comput. Methods Appl. Mech. Eng. 98, 41–104.
Simo, J.C., Taylor, R.L., 1985. Consistent tangent operators for rate-independent elastoplasticity.
Comput. Methods Appl. Mech. Eng. 48, 101–118.
Simo, J.C., Taylor, R.L., Pister, K.S., 1985. Variational and projection methods for the volume constraint
in finite deformation elasto-plasticity. Comput. Methods Appl. Mech. Eng. 51, 177–208.
Srikanth, A., Zabaras, N., 1999. A computational model for the finite element analysis of thermoplasticity
coupled with ductile damage at finite strains. Int. J. Numer. Meth. Eng. 45, 1569–1609.
Sussman, T., Bathe, K.J., 1987. A finite element formulation for nonlinear incompressible elastic and
inelastic analysis. Comput. Struct. 26, 357–409.
Truesdell, C., Noll, W., 1965. The non-linear field theories. In: Handbuch der Physik, vol. III/3. Springer,
Berlin.
Wriggers, P., 2001. Nichtlineare Finite-Elemente Methoden. Springer, Berlin.
Wriggers, P., Miehe, C., Kleiber, M., Simo, J.C., 1992. On the coupled thermomechanical treatment of
necking problems via finite element method. Int. J. Numer. Meth. Eng. 33, 869–883.
Zdebel, U., Lehmann, Th., 1987. Some theoretical considerations and experimental investigations on a
constitutive law in thermoplasticity. Int. J. Plasticity 3, 369–389.
Zienkiewicz, O.C., Taylor, R.L., 2000. The finite element method. In: Solid Mechanics, vol. 2.
Butterworth–Heinemann, Oxford.

Das könnte Ihnen auch gefallen