Sie sind auf Seite 1von 46

International Journal of Plasticity 25 (2009) 1833–1878

Contents lists available at ScienceDirect

International Journal of Plasticity


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

A large deformation theory for rate-dependent elastic–plastic


materials with combined isotropic and kinematic hardening
David L. Henann, Lallit Anand *
Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139, USA

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

Article history: We have developed a large deformation viscoplasticity theory with


Received 24 August 2008 combined isotropic and kinematic hardening based on the dual
Received in final revised form 25 November decompositions F ¼ Fe Fp [Kröner, E., 1960. Allgemeine kontinu-
2008
umstheorie der versetzungen und eigenspannungen. Archive for
Available online 7 December 2008
Rational Mechanics and Analysis 4, 273–334] and Fp ¼ Fpen Fpdis
[Lion, A., 2000. Constitutive modelling in finite thermoviscoplastic-
ity: a physical approach based on nonlinear rheological models.
Keywords:
Large deformation viscoplasticity
International Journal of Plasticity 16, 469–494]. The elastic distor-
Isotropic hardening tion Fe contributes to a standard elastic free-energy wðeÞ , while Fpen ,
Kinematic hardening the energetic part of Fp , contributes to a defect energy wðpÞ – these
Time-integration procedure two additive contributions to the total free energy in turn lead to
Finite elements the standard Cauchy stress and a back-stress. Since Fe ¼ FFp1
and Fpen ¼ Fp Fp1
dis , the evolution of the Cauchy stress and the back-
stress in a deformation-driven problem is governed by evolution
equations for Fp and Fpdis – the two flow rules of the theory.
We have also developed a simple, stable, semi-implicit time-
integration procedure for the constitutive theory for implementa-
tion in displacement-based finite element programs. The proce-
dure that we develop is ‘‘simple” in the sense that it only
involves the solution of one non-linear equation, rather than a sys-
tem of non-linear equations. We show that our time-integration
procedure is stable for relatively large time steps, is first-order
accurate, and is objective.
Ó 2008 Elsevier Ltd. All rights reserved.

1. Introduction

In classical small deformation theories of metal plasticity, when attempting to model the Bausch-
inger phenomenon (Bauschinger, 1886) and other phenomena associated with cyclic loading, kine-

* Corresponding author. Tel.: +1 617 253 1635.


E-mail address: anand@mit.edu (L. Anand).

0749-6419/$ - see front matter Ó 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijplas.2008.11.008
1834 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

matic strain hardening is often invoked by introducing a symmetric and deviatoric stress-like tensorial
internal variable Tback called the back-stress, which acts to oppose the Cauchy stress T in the formu-
lation of an appropriate yield criterion and a corresponding flow rule for the material. The simplest
model for the evolution of such a back-stress is Prager’s linear kinematic-hardening rule (Prager,
1949):

T_ back ¼ BE_ p so that Tback ¼ BEp ; ð1:1Þ


p
where B > 0 is a back-stress modulus and E is the plastic stain tensor in the standard decomposition
E ¼ Ee þ Ep of the infinitesimal strain tensor E into elastic and plastic parts. A better description of
kinematic hardening is given by an evolution equation for the back-stress Tback proposed initially by
Armstrong and Frederick (1966):1
def
T_ back ¼ BE_ p  cTback e_ p ; with e_ p ¼ jE_ p j; ð1:2Þ
where the term ðcTback e_ p Þ with c > 0 represents a dynamic recovery term, which is collinear with the
back-stress Tback and is proportional to the equivalent plastic strain rate e_ p . For monotonic uniaxial
loading, the evolution of back-stress, instead of being linear as prescribed by (1.1), is then exponential
with a saturation value. The Armstrong–Frederick model for kinematic hardening is widely used as a
simple model to describe cyclic loading phenomena in metals. For a discussion of the Armstrong–Fred-
erick model and its various extensions to model cyclic plasticity in the small deformation regime see
Lemaitre and Chaboche (1990) and the recent review by Chaboche (2008).
While there have been numerous attempts to extend the Armstrong–Frederick rule to large deforma-
tions, a suitable extension is still not widely agreed upon. Dettmer and Reese (2004) have recently re-
viewed and numerically analyzed several previous models for large deformation kinematic hardening
of the Armstrong–Frederick type. One class of theories (cf., Dettmer and Reese, 2004, for a list of refer-
ences to the literature) follow Prager and Armstrong and Frederick, and introduce a stress-like internal
variable to model kinematic hardening; these theories require suitable frame-indifferent evolution
equations to generalize an evolution equation resembling (1.2) to large deformations. An alternative
to such theories are theories based on a proposal by Lion (2000), who in order to account for energy storage
mechanisms associated with plastic flow, introduced a kinematic constitutive assumption that the plastic
distortion Fp in the standard Kröner (1960) elastic–plastic decomposition F ¼ Fe Fp of the total deforma-
tion gradient, may be further multiplicatively decomposed into energetic and dissipative parts Fpen and
Fpdis , respectively, as Fp ¼ Fpen Fpdis (cf., Lion’s Fig. 5).2 The elastic distortion Fe contributes to a standard elas-
tic free-energy wðeÞ , while Fpen contributes a defect energy wðpÞ – these two additive contributions to the total
free energy in turn lead to the standard Cauchy stress and a back-stress. Since Fe ¼ FFp1 and Fpen ¼ Fp Fp1 dis ,
the evolution of the Cauchy stress and the back-stress in a deformation-driven problem is governed by evo-
lution equations for Fp and Fpdis – the two flow rules of the theory. An important characteristic of such a the-
ory is that since Fp is invariant under a change in frame, then so also are Fpen and Fpdis , and the resulting
evolution equations for the back-stress are automatically frame-indifferent, and do not require special con-
siderations such as those required when generalizing equations of the type (1.2) to proper frame-indiffer-
ent counterparts. Examples of kinematic-hardening constitutive theories based on Lion-type Fp ¼ Fpen Fpdis
decomposition may be found in Dettmer and Reese (2004), Menzel et al. (2005), Håkansson et al.
(2005), Wallin and Ristinmaa (2005), and Vladimirov et al. (2008).
While Dettmer and Reese (2004) in Section 4.3 of their paper, and more recently Vladimirov et al.
(2008) in Section 2.2 of their paper have presented a generalization of the Armstrong–Frederick type
of kinematic-hardening rate-independent plasticity theory using the Lion decomposition Fp ¼ Fpen Fpdis ,
we find their presentation rather brief, and (in our view) not sufficiently expository. The primary focus
of the papers by Dettmer and Reese and Vladimirov et al., was the development of suitable time-inte-
gration algorithms, and they have developed several fully-implicit time-integration procedures for

1
Also recently published as Frederick and Armstrong (2007).
2
Lion uses the notation FM  F, Fe  Fe , Fin  Fp , Fine  Fpen , Finp  Fpdis . Based on a statistical-volume-averaging argument for the
response of a polycrystalline aggregate, Clayton and McDowell (2003, Eq. (13)) have also introduced a Lion-type decomposition
Fp ¼ F ~i  Fp a ‘‘meso-incompatibility tensor.”
~p ; they call F
~i F
en
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1835

their model; one such integration procedure involves the solution of a system of 13 highly non-linear
equations – cf., Eq. (53) of Vladimirov et al. (2008). Motivated by the work of these authors, it is the
purpose of this paper to:

(1) Develop a thermodynamically consistent rate-dependent plasticity theory with combined isotro-
pic and kinematic hardening of the Armstrong–Frederick type, based on the dual decomposi-
tions F ¼ Fe Fp and Fp ¼ Fpen Fpdis . The limit m ! 0, where m is a material rate-sensitivity
parameter in our theory, corresponds to a rate-independent model. Importantly, we develop
the theory based on the principle of virtual power and carefully lay down all assumptions
and specializations that we have adopted so that it may in the future be possible to generalize
the theory with other, more flexible isotropic and kinematic-hardening models.
(2) Develop a simple, stable, semi-implicit time-integration procedure for our rate-dependent con-
stitutive theory for implementation in displacement-based finite element programs. The proce-
dure that we develop is ‘‘simple” in the sense that it involves the solution of only one stiff non-
linear equation,3 rather than a system of non-linear equations. We show that our time-integration
procedure is stable for relatively large time steps, is first-order accurate, and is objective.

2. Kinematics

We consider a homogeneous body B identified with the region of space it occupies in a fixed refer-
ence configuration, and denote by X an arbitrary material point of B. A motion of B is then a smooth one-
to-one mapping x ¼ vðX; tÞ with deformation gradient, velocity and velocity gradient given by4

F ¼ rv; v ¼ v;
_ _ 1 :
L ¼ grad v ¼ FF ð2:1Þ
To model the inelastic response of materials, we assume that the deformation gradient F may be
multiplicatively decomposed as (Kröner, 1960)
F ¼ Fe Fp : ð2:2Þ
As is standard, we assume that
J ¼ det F > 0;
and consistent with this we assume that
def def
J e ¼ det Fe > 0; J p ¼ det Fp > 0; ð2:3Þ
e p
so that F and F are invertible. Here, suppressing the argument t:

 Fp ðXÞ represents the local inelastic distortion of the material at X due to a ‘‘plastic mechanism.” This
local deformation carries the material into – and ultimately ‘‘pins” the material to – a coherent struc-
ture that resides in the structural space5 at X (as represented by the range of Fp ðXÞ);
 Fe ðXÞ represents the subsequent stretching and rotation of this coherent structure, and thereby rep-
resents the corresponding ‘‘elastic distortion.”

3
With reference to the power-law model (9.35) in Section 9.3, the stiffness of the equations depends on the strain-rate-
sensitivity parameter m, and the stiffness increases to infinity as m tends to zero, the rate-independent limit. For small values of m
special care is required to develop stable constitutive time-integration procedures; cf., e.g., Lush et al. (1989).
4
Notation: We use standard notation of modern continuum mechanics. Specifically: r and Div denote the gradient and
divergence with respect to the material point X in the reference configuration; grad and div denote these operators with respect to
the point x ¼ vðX; tÞ in the deformed body; a superposed dot denotes the material time-derivative. Throughout, we write
Fe1 ¼ ðFe Þ1 , Fp> ¼ ðFp Þ> , etc. We write trA, symA, skwA, A0 , and sym0 A, respectively, for the trace, symmetric, skew, deviatoric,
and symmetric–deviatoric
pffiffiffiffiffiffiffiffiffiffiffi parts of a tensor A. Also, the inner product of tensors A and B is denoted by A : B, and the magnitude of A
by jAj ¼ A : A.
5
Also sometimes referred to as the ‘‘intermediate” or ‘‘relaxed” local space at X.
1836 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

We refer to Fp and Fe as the inelastic and elastic distortions.

By (2.1)3 and (2.2),

L ¼ Le þ Fe Lp Fe1 ; ð2:4Þ

with

Le ¼ F_ e Fe1 ; Lp ¼ F_ p Fp1 : ð2:5Þ

As is standard, we define the total, elastic, and plastic stretching and spin tensors through
9
D ¼ symL; W ¼ skwL; >
=
e e e e
D ¼ symL ; W ¼ skwL ; ð2:6Þ
>
;
Dp ¼ symLp ; Wp ¼ skwLp ;

so that L ¼ D þ W, Le ¼ De þ We , and Lp ¼ Dp þ Wp .
The right and left polar decompositions of F are given by

F ¼ RU ¼ VR; ð2:7Þ

where R is a rotation (proper-orthogonal tensor), while U and V are symmetric, positive-definite ten-
sors with
pffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffi
U¼ F> F; V¼ FF> : ð2:8Þ
Also, the right and left Cauchy–Green tensors are given by

C ¼ U2 ¼ F> F; B ¼ V2 ¼ FF> : ð2:9Þ


e p
Similarly, the right and left polar decompositions of F and F are given by
Fe ¼ Re Ue ¼ Ve Re ; Fp ¼ Rp Up ¼ Vp Rp ; ð2:10Þ
e p e e p p
where R and R are rotations, while U , V , U , V are symmetric, positive-definite tensors with
pffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffi
Ue ¼ Fe> Fe ; Ve ¼ Fe Fe> ; Up ¼ Fp> Fp ; Vp ¼ Fp Fp> : ð2:11Þ
Also, the right and left elastic Cauchy–Green tensors are given by

Ce ¼ Ue2 ¼ Fe> Fe ; Be ¼ Ve2 ¼ Fe Fe> ; ð2:12Þ


and the right and left plastic Cauchy–Green tensors are given by

Cp ¼ Up2 ¼ Fp> Fp ; Bp ¼ Vp2 ¼ Fp Fp> : ð2:13Þ

2.1. Incompressible, irrotational plastic flow

We make two basic kinematical assumptions concerning plastic flow:

(i) First, we make the standard assumption that plastic flow is incompressible, so that

J p ¼ det Fp ¼ 1 and tr Lp ¼ 0: ð2:14Þ


Hence, using (2.2) and (2.14)1 ,

J e ¼ J: ð2:15Þ
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1837

(ii) Second, from the outset we constrain the theory by limiting our discussion to circumstances
under which the material may be idealized as isotropic. For isotropic elastic–viscoplastic theo-
ries utilizing the Kröner decomposition it is widely assumed that the plastic flow is irrotational
in the sense that6
Wp ¼ 0: ð2:16Þ
p p
Then, trivially, L  D and

F_ p ¼ Dp Fp ; with tr Dp ¼ 0: ð2:17Þ

Thus, using (2.1), (2.4), (2.5), and (2.17), we may write (2.4) for future use as

grad v ¼ Le þ Fe Dp Fe1 ; with trDp ¼ 0: ð2:18Þ

2.2. Decomposition of Fp into energetic and dissipative parts

Next following Lion (2000), in order to account for energy storage mechanisms associated with plastic
flow, we introduce an additional kinematic constitutive assumption that Fp may be multiplicatively
decomposed into an energetic Fpen and a dissipative part Fpdis as follows:
Fp ¼ Fpen Fpdis ; det Fpen ¼ det Fpdis ¼ 1: ð2:19Þ
We call the range of Fpdis ðXÞ
the local substructural space. Thus Fpen
maps material elements from the
p
substructural space to the structural space (which is the range of F ðXÞ). Lion (2000) calls the substruc-
tural space as the ‘‘intermediate configuration of kinematic hardening.” A ‘‘physical interpretation” of
such an ‘‘intermediate configuration” is not completely clear7; here we simply treat it as a mathematical
construct resulting from Lion’s presumed decomposition (2.19).
Then, from (2.5)

Lp ¼ Lpen þ Fpen Lpdis Fp1


en ð2:20Þ
with

Lpen ¼ F_ pen Fen


p1
with trLpen ¼ 0; Lpdis ¼ F_ pdis Fdis
p1
with trLpdis ¼ 0; ð2:21Þ
and corresponding stretching and spin tensors through

6
A note on role of the plastic spin in isotropic solids: There are numerous publications discussing the role of plastic spin in theories
based on the Kröner decomposition (cf., e.g., Dafalias, 1998). Based on the classical principle of frame-indifference (cf., Section 3),
Green and Naghdi (1971) (cf., also Casey and Naghdi, 1980), introduced the notion of a change in frame of the intermediate structural
space. This notion leads to transformation laws for Fe and Fp of the form

F
p
¼ QFp ; F
e
¼ Fe Q > ;
in which Q ðX; tÞ is an arbitrary time-dependent rotation of the structural space. Unfortunately, Green and Naghdi viewed struc-
tural frame-indifference as a general principle; that is, a principle that stands at a level equivalent to that of conventional frame-
indifference. This view has been refuted by many workers (cf., e.g., Dashner, 1986, and the references therein). While we agree
with the view that structural frame-indifference is not a general principle, this hypothesis may be used to represent an important
facet of the behavior of a large class of polycrystalline materials based on the Kröner decomposition. Indeed, for polycrystalline
materials without texture the structural space is associated with a collection of randomly oriented lattices, and hence the evo-
lution of dislocations through that space should be independent of the frame with respect to which this evolution is measured.
Using the notion of a change in frame of the structural space, Gurtin and Anand (2005, p. 1714) have shown that within a frame-
work for isotropic plasticity based on the Kröner decomposition (which we follow here), one may assume, without loss in gener-
ality, that the plastic spin vanishes Wp ¼ 0.
Here, we adopt the assumption Wp ¼ 0 from the outset; we do so based solely on pragmatic grounds: when discussing finite
deformations for isotropic materials a theory without plastic spin is far simpler than one with plastic spin.
7
However, see Clayton and McDowell (2003) for a ‘‘statistical-volume-averaging” argument for the response of a polycrystalline
aggregate.
1838 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

)
Dpen ¼ symLpen ; Wpen ¼ skwLpen ;
ð2:22Þ
Dpdis ¼ symLpdis ; Wpdis ¼ skwLpdis ;

so that Lpen ¼ Dpen þ Wpen and Lpdis ¼ Dpdis þ Wpdis .


Paralleling the assumption of plastic irrotationality for isotropic materials, Wp ¼ 0, we assume that
Wpdis ¼ 0: ð2:23Þ
Then, trivially Lpdis ¼ Dpdis , and

F_ pdis ¼ Dpdis Fpdis with trDpdis ¼ 0: ð2:24Þ


Since
 
Wp ¼ Wpen þ skw Fpen Dpdis Fp1
en ; ð2:25Þ

in order to enforce the constitutive constraint Wp ¼ 0, we require that both


Wpen ¼ 0; ð2:26Þ
and
 
skw Fpen Dpdis Fp1
en ¼ 0: ð2:27Þ

Also, (2.20) reduces to


 
Dp ¼ Dpen þ sym Fpen Dpdis Fp1
en ; with trDpen ¼ trDpdis ¼ 0: ð2:28Þ

The right and left polar decompositions of Fpen and Fpdis are given by
Fpen ¼ Rpen Upen ¼ Vpen Rpen ; Fpdis ¼ Rpdis Updis ¼ Vpdis Rpdis ; ð2:29Þ
where Rpen and Rpdis are rotations, while Upen , Vpen , Updis , Vpdis are symmetric, positive-definite tensors with
qffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Upen ¼ Fp> p
en Fen ; Vpen ¼ Fpen Fp> en ; Updis ¼ Fp> p
dis Fdis ; Vpdis ¼ Fpdis Fp> dis : ð2:30Þ

Also, the corresponding right and left Cauchy–Green tensors are given by

Cpen ¼ Up2 p> p


en ¼ Fen Fen ; Bpen ¼ Vp2 p p>
en ¼ Fen Fen ; ð2:31Þ
Cpdis ¼ Up2
dis ¼ Fp> p
dis Fdis ; Bpdis ¼ Vp2 p p>
dis ¼ Fdis Fdis : ð2:32Þ

3. Frame-indifference

Changes in frame (observer) are smooth time-dependent rigid transformations of the Euclidean
space through which the body moves. We require that the theory be invariant under such transforma-
tions, and hence under transformations of the form
v ðX; tÞ ¼ Q ðtÞðvðX; tÞ  oÞ þ yðtÞ; ð3:1Þ
with Q ðtÞ a rotation (proper-orthogonal tensor), yðtÞ a point at each t, and o a fixed origin. Then, under
a change in frame, the deformation gradient transforms according to
F ¼ QF; ð3:2Þ
and hence
C ¼ C that is; C is invariant; ð3:3Þ
also F_  ¼ Q F_ þ Q_ F, and by (2.1)3 ,

L ¼ QLQ > þ Q_ Q > : ð3:4Þ


D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1839

Thus,

D ¼ QDQ > ; W ¼ QWQ > þ Q_ Q > : ð3:5Þ


e p  e p
Moreover, ðF F Þ ¼ Q ðF F Þ, and therefore, since observers view only the deformed configuration,

Fe ¼ QFe ; Fp ¼ Fp ; and thus Fp is invariant: ð3:6Þ


Hence, by (2.5),

Le ¼ QLe Q > þ Q_ Q > ; ð3:7Þ


and
Lp ; Dp ; and Wp are invariant: ð3:8Þ
Further, by (2.10),

Fe ¼ Re Ue ! QFe ¼ QRe Ue ;
Fe ¼ Ve Re ! QFe ¼ QVe Q > QRe ;

and we may conclude from the uniqueness of the polar decomposition that

Re ¼ QRe ; Ve ¼ QVe Q > ; Ue ¼ Ue : ð3:9Þ


e e
Hence, from (2.12), B and C transform as

Be ¼ QBe Q > and Ce ¼ Ce ; ð3:10Þ


p
further, since F is invariant,

Bp and Cp are also invariant: ð3:11Þ


p
Finally, since F is invariant under a change in frame,

Fpen and Fpdis are invariant;


as are all of the following quantities based on Fpen and Fpdis :

Upen ; Cpen ; Vpen ; Bpen ; Rpen ; Updis ; Cpdis ; Vpdis ; Bpdis ; Rpdis :

4. Development of the theory based on the principle of virtual power

Following Gurtin and Anand (2005), the theory presented here is based on the belief that

 the power expended by each independent ‘‘rate-like” kinematical descriptor be expressible in terms of an
associated force system consistent with its own balance.

However, the basic ‘‘rate-like” descriptors, namely, v, Le , Dp , Dpen , and Dpdis are not independent, since
by (2.18) and (2.28) they are constrained by

grad v ¼ Le þ Fe Dp Fe1 ; trDp ¼ 0; ð4:1Þ


and
 
Dp ¼ Dpen þ sym Fpen Dpdis Fp1
en ; trDpen ¼ trDpdis ¼ 0; ð4:2Þ

and it is not apparent what forms the associated force balances should take. It is in such situations that
the strength of the principle of virtual power becomes apparent, since the principle of virtual power
automatically determines the underlying force balances.
1840 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

We denote by Pt an arbitrary part (subregion) of the deformed body with n the outward unit nor-
mal on the boundary @Pt of Pt . The power expended on Pt by material or bodies exterior to Pt results
from a macroscopic surface traction tðnÞ, measured per unit area in the deformed body, and a macro-
scopic body force b, measured per unit volume in the deformed body, each of whose working accom-
panies the macroscopic motion of the body; the body force b presumed to account for inertia; that is,
granted the underlying frame is inertial,
b ¼ b0  qv;
_ ð4:3Þ
with b0 the noninertial body force and q the mass density in the deformed body. We therefore write
the external power as
Z Z
Wext ðPt Þ ¼ tðnÞ  vda þ b  vdv: ð4:4Þ
@Pt Pt

We assume that power is expended internally by an elastic stress Se power-conjugate to Le , a plastic


stress Tp power-conjugate to Dp , an energetic stress Sen conjugate to Dpen , and a dissipative stress Tdis
conjugate to symðFpen Dpdis Fen
p1
Þ. Since Dp , Dpen , and symðFpen Dpdis Fp1
en Þ are symmetric and deviatoric, we as-
sume that Tp , Sen , and Tdis are symmetric and deviatoric. We write the internal power as
Z   
Wint ðPt Þ ¼ Se : Le þ J 1 Tp : Dp þ Sen : Dpen þ Tdis : symðFpen Dpdis Fp1
en Þ dv;
Pt
Z   
¼ Se : Le þ J 1 Tp : Dp þ Sen : Dpen þ sym0 ðFp> p> p
en Tdis Fen Þ : Ddis dv:
Pt
 
The term J arises because Tp : Dp þ Sen : Dpen þ Tdis : sym Fpen Dpdis Fp1
1
en is measured per unit volume in
the structural space, but the integration is carried out within the deformed body. Defining a new stress
measure
def
Sdis ¼ sym0 ðFp> p>
en Tdis Fen Þ; ð4:5Þ
the internal power may more compactly be written as
Z   
Wint ðPt Þ ¼ Se : Le þ J 1 Tp : Dp þ Sen : Dpen þ Sdis : Dpdis dv: ð4:6Þ
Pt

4.1. Principle of virtual power

Assume that, at some arbitrarily chosen but fixed time, the fields v, Fe , Fpen (and hence F, Fp , Fpdis ) are
known, and consider the fields v, Le , Dp , Dpen , and Dpdis as virtual velocities to be specified independently
in a manner consistent with the constraints (4.1) and (4.2). That is, denoting the virtual fields by v ~, L~e ,
~ pen , and D
~ p, D
D ~ p to differentiate them from fields associated with the actual evolution of the body, we
dis
require that
~ ¼ L~e þ Fe D
grad v ~ p Fe1 ; ~ p ¼ 0;
trD ð4:7Þ
and
 
D ~ p þ sym Fp D
~p ¼ D ~ p p1 ; ~ p ¼ 0:
~ p ¼ trD
en en dis Fen trD en dis ð4:8Þ

More specifically, we define a generalized virtual velocity to be a list


~; L~e ; D
V ¼ ðv ~ p; D
~p ; D
en
~ p Þ;
dis

consistent with (4.7) and (4.8). Then, writing


Z Z )
Wext ðPt ; VÞ ¼ ~ da þ
tðnÞ  v ~dv;
bv
@Pt Pt
Z    ð4:9Þ
Wint ðPt ; VÞ ¼ Se : L
~e þ J 1 Tp : D
~ p þ Sen : D
~ p þ Sdis : D
en
~p
dis dv;
Pt
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1841

respectively, for the external and internal expenditures of virtual power, the principle of virtual power is
the requirement that the external and internal powers be balanced. That is, given any part Pt ,
Wext ðPt ; VÞ ¼ Wint ðPt ; VÞ for all generalized virtual velocities V: ð4:10Þ

4.1.1. Consequences of frame-indifference


We assume that the internal power Wint ðPt ; VÞ is invariant under a change in frame and that the
virtual fields transform in a manner identical to their nonvirtual counterparts. Then given a change in
frame, invariance of the internal power requires that
Wint ðPt ; V Þ ¼ Wint ðPt ; VÞ; ð4:11Þ
where Pt and Wint ðPt ; V Þ
represent the region and the internal power in the new frame. In the new
frame Pt transforms rigidly to a region Pt , the stresses Se , Tp , Sen , and Sdis transform to Se , Tp , Sen , and
~e transforms to
Sdis , while L
~ e Q > þ Q_ Q > ;
L~e ¼ Q L
~ pen , and D
~ p, D
and D ~ p are invariant. Thus, under a change in frame Wint ðPt ; VÞ transforms to
dis

Wint ðPt ; V Þ ¼ Wint ðPt ; V Þ


Z     
¼ Se : Q L ~ e Q > þ Q_ Q > þ J 1 Tp : D ~ p þ S : D
~ p þ S : D~p dv
en en dis dis
P
Z t  
¼ ðQ > Se Q Þ : L~e þ Se : ðQ_ Q > Þ þ J 1 Tp : D
~ p þ S : D
en
~ p þ S : D
en dis
~p
dis dv:
Pt

Then (4.11) implies that


Z   
ðQ > Se Q Þ : L~e þ Se : ðQ_ Q > Þ þ J 1 Tp : D
~ p þ S : D
en en dis
~p
~ p þ S : D
dis dv
Pt
Z   
¼ Se : L~e þ J 1 Tp : D ~ p þ Sen : D
en
~p
~ p þ Sdis : D
dis dv; ð4:12Þ
Pt

or equivalently, since the part Pt is arbitrary,


 
ðQ > Se Q Þ : L~e þ Se : ðQ_ Q > Þ þ J 1 Tp : D
~ p þ S : D
en en dis
~p
~ p þ S : D
dis
 
¼ Se : L~e þ J 1 Tp : D ~ p þ Sen : D ~p :
~ p þ Sdis : D
en dis ð4:13Þ

Also, since the change in frame is arbitrary, if we choose it such that Q is an arbitrary time-independent
rotation, and that Q_ ¼ 0, we find from (4.13) that
  h     i
ðQ > Se Q Þ  Se : L~e þ J 1 ðTp  Tp Þ : D en en dis
~ p ¼ 0:
~ p þ S  Sdis : D
~ p þ S  Sen : D
dis

~e , D
Since this must hold for all L ~ p, D ~ p , we find that the stress Se transforms according to
~ pen , and D
dis

Se ¼ QSe Q > ; ð4:14Þ


p
while T , Sen , and Sdis are invariant
Tp ¼ Tp ; Sen ¼ Sen ; Sdis ¼ Sdis : ð4:15Þ
Next, if we assume that Q ¼ 1 at the time in question, and that Q_ is an arbitrary skew tensor, we find
from (4.13), using (4.14), (4.15) that

Se : Q_ ¼ 0;
or that the tensor Se is symmetric,
Se ¼ Se> : ð4:16Þ
e
Thus, the elastic stress S is frame-indifferent and symmetric.
1842 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

4.1.2. Macroscopic force balance


In applying the virtual balance (4.10) we are at liberty to choose any V consistent with the con-
~ is arbitrary, and
straints (4.7) and (4.8). Consider a generalized virtual velocity with V for which v
~p ¼ D
D ~ pen ¼ D
~ p  0, so that
dis

L~e ¼ grad v
~: ð4:17Þ
For this choice of V, (4.10) yields
Z Z Z
~da þ
tðnÞ  v ~dv ¼
bv Se : grad v
~dv: ð4:18Þ
@Pt Pt Pt

Further, using divergence theorem, (4.18) yields


Z Z
ðtðnÞ  Se nÞ  v
~da þ ðdivSe þ bÞ  v
~dv ¼ 0: ð4:19Þ
@Pt Pt

~, standard variational arguments yield the traction condition


Since (4.19) must hold for all Pt and all v
tðnÞ ¼ Se n; ð4:20Þ
and the local force balance
divSe þ b ¼ 0: ð4:21Þ
e
This traction condition and force balance and the symmetry and frame-indifference of S are classical
conditions satisfied by the Cauchy stress T, an observation that allows us to write
def
T ¼ Se ð4:22Þ
and to view

T ¼ T> ð4:23Þ
as the standard macroscopic Cauchy stress and (4.21) as the local macroscopic force balance. Since we are
working in an inertial frame, so that (4.3) is satisfied, (4.21) reduces to the local balance law for linear
momentum,
divT þ b0 ¼ qv;
_ ð4:24Þ
with b0 the noninertial body force.

4.1.3. Microscopic force balances


To discuss the microscopic counterparts of the macroscopic force balance, first consider a general-
ized virtual velocity with

v ~ p  0;
~  0 and D ð4:25Þ
dis

~ p arbitrarily, and let


also, choose the virtual field D

L~e ¼ Fe D
~ p Fe1 ~p ¼ D
and Den
~ p; ð4:26Þ

consistent with (4.7) and (4.8). Thus

T : L~e ¼ ðFe> TFe> Þ : D


~ p: ð4:27Þ
Next, define an elastic Mandel stress by

def
Me ¼ JFe> TFe> ; ð4:28Þ
which in general is not symmetric. Then, on account of our choice (4.25)1 the external power vanishes,
so that, by (4.10), the internal power must also vanish, and satisfy
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1843

Z
Wint ðPt ; VÞ ¼ J 1 ðTp þ Sen  Me Þ : D
~ p dv ¼ 0:
Pt

~ p , a standard argument
Since this must be satisfied for all Pt and all symmetric and deviatoric tensors D
yields the first microforce balance

ð4:29Þ

Next consider a generalized virtual velocity with

v ~ p  0;
~  0 and D ð4:30Þ
~ p arbitrarily, and let
also, choose the virtual field D dis
 
~ p p1
~ p ¼ sym Fp D
D en en dis Fen ð4:31Þ

consistent with (4.8). Thus


~ p ¼ ðFp> Sen Fp> Þ : D
Sen : D ~p : ð4:32Þ
en en en dis

Next, define a new stress measure by


def
Mpen ¼ Fp> p>
en Sen Fen ; ð4:33Þ
which in general is also not symmetric. For lack of better terminology, we call Mpen
a plastic Mandel
stress. Then, on account of our choice (4.30)1 the external power vanishes, so that, by (4.10), the inter-
nal power must also vanish, and satisfy
Z
 
Wint ðPt ; VÞ ¼ ~ p dv ¼ 0:
Sdis  Mpen : D dis
Pt

~ p , we obtain the sec-


Since this must be satisfied for all Pt and all symmetric and deviatoric tensors D dis
ond microforce balance

ð4:34Þ

5. Free-energy imbalance

We consider a purely mechanical theory based on a second law requiring that the temporal increase
in free energy of any part Pt be less than or equal to the power expended on Pt . The second law therefore
takes the form of a dissipation inequality
Z _
wJ 1 dv 6 Wext ðPt Þ ¼ Wint ðPt Þ; ð5:1Þ
Pt

with w the free energy, measured per unit volume in the structural space.
Since J 1 dv with J ¼ det F represents the volume measure in the reference configuration, and since
Pt deforms with the body,
Z _ Z
wJ 1 dv ¼ _ 1 dv:
wJ
Pt Pt

Thus, since Pt is arbitrary, we may use (4.6), (4.22), and (4.23) to localize (5.1); the result is the local
free-energy imbalance
1844 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

w_  JT : De  Tp : Dp  Sen : Dpen  Sdis : Dpdis 6 0: ð5:2Þ


We introduce a new stress measure
def
Te ¼ JFe1 TFe> ; ð5:3Þ
e
which is symmetric, since T is symmetric; T represents a second Piola stress with respect to the inter-
mediate structural space. Note that by using the definition (5.3), the Mandel stress defined in (4.28) is
related to the stress measure Te by
Me ¼ Ce Te : ð5:4Þ
e
Next, differentiating (2.12)1 results in the following expression for the rate of change of C ,

C_ e ¼ Fe> F_ e þ F_ e> Fe :
Hence, since Te is symmetric,

Te : C_ e ¼ 2Te : ðFe> F_ e Þ ¼ 2ðFe Te Fe> Þ : ðF_ e Fe1 Þ ¼ 2ðFe Te Fe> Þ : De ;


or, using (5.3) we obtain
1 e _e
JT : De ¼ T :C : ð5:5Þ
2
Use of (5.5) in (5.2) allows us to express the free-energy imbalance in an alternate convenient form as
1
w_  Te : C_ e  Tp : Dp  Sen : Dpen  Sdis : Dpdis 6 0: ð5:6Þ
2
Finally, we note that w is invariant under a change in frame since it is a scalar field, and on account
of the transformation rules discussed in Section 3, the transformation rules (4.14) and (4.15), and the
definitions (4.28), (5.3), (4.33), and (5.4), the fields
Ce ; Cpen ; Bpen ; Dp ; Dpen ; Dpdis ; Te ; Tp ; Sen ; Sdis ; Me ; and Mpen ð5:7Þ
are also invariant.

6. Constitutive theory

To account for the classical notion of isotropic strain hardening we introduce a positive-valued sca-
lar internal state-variable S, which has dimensions of stress. We refer to S as the isotropic deformation
resistance. Since S is a scalar field, it is invariant under a change in frame.
Next, guided by the free-energy imbalance (5.6), and by experience with previous constitutive the-
ories, we assume the following special set of constitutive equations:
9
w¼w  ðeÞ ðCe Þ þ w
 ðpÞ ðBp Þ; >
en >
>
>
 e ðCe Þ;
Te ¼ T >
>
>
>
 p >
=
Sen ¼ Sen ðB Þ; en
ð6:1Þ
Tp ¼ T p ðDp ; SÞ; >
>
>
>
dis ðDp ; dp ; SÞ;
Sdis ¼ S >
>
dis >
>
p
>
;
S_ ¼ hðd ; SÞ;
where
p def
d ¼ jDp j ð6:2Þ
p
is the scalar flow rate corresponding to D . Note that since w is the free energy per unit volume of the
 ðeÞ is chosen to depend on Ce and w
structural space, w  ðpÞ on Bp , because Ce and Bp are squared stretch-
en en
like measures associated with the intermediate structural space. Also note that we have introduced a
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1845

p
possible dependence of Sdis on d because for some materials we anticipate that Dpdis (and thereby Sdis )
p
may be constrained by the magnitude of d .
The energy w  ðeÞ represents the standard energy associated with intermolecular interactions, and the
energy w ðpÞ is a ‘‘defect-energy” associated with plastic deformation. For metallic materials w  ðpÞ may be
attributed to an energy stored in the complex local microstructural defect state, which typically in-
cludes dislocations, sub-grain boundaries, and local incompatibilities at second-phase particles and
grain boundaries. At the macroscopic continuum level, the ‘‘defect-energy” w  ðpÞ leads to the develop-
ment and evolution of the energetic stress Sen , and this allows one to phenomenologically account for
strain-hardening phenomena commonly called kinematic hardening. Isotropic hardening is modeled by
the evolution (6.1)6 of the isotropic deformation resistance S.
Finally, note that on account of the transformation rules listed in the paragraph containing (5.7)
and since S is also invariant, the constitutive equations (6.1) are frame-indifferent.

6.1. Thermodynamic restrictions

From (6.1)1
 ðeÞ ðCe Þ
@w  ðpÞ ðBp Þ
@w
w_ ¼ e : C_ e þ en
: B_ pen ;
@C @Bpen
and, using8

B_ pen ¼ Dpen Bpen þ Bpen Dpen ; ð6:3Þ


and the symmetry of Bpen 
and @ w=@B p
en ,
 
 ðpÞ Bp
@w  ðpÞ ðBp Þ  p p
@w 
p
en
: B_ pen ¼ p
en
: Den Ben þ Bpen Dpen
@Ben @Ben
  ðpÞ p 
@ w ðBen Þ p
¼2 B : Dpen :
@Bpen en
0

Hence, satisfaction of the free-energy imbalance (5.6) requires that the constitutive equations (6.1)
satisfy
  ðeÞ ðCe Þ    ðpÞ p  
1 e e @w en ðBp Þ  2 @ w ðBen Þ Bp
T ðC Þ  e : C_ e þ S en p en : Dpen
2 @C @Ben 0
þT dis ðDp ; dp ; SÞ : Dp P 0;
 p ðDp ; SÞ : Dp þ S ð6:4Þ
dis dis

and hold for all arguments in the domains of the constitutive functions, and in all motions of the body.
Thus, sufficient conditions that the constitutive equations satisfy the free-energy imbalance are that9

(i) the free-energy determine the stresses Te and Sen via the stress relations
 ðeÞ e
 e ðCe Þ ¼ 2 @ w ðC Þ ;
T ð6:5Þ
@Ce
  ðpÞ p 
en ðBp Þ ¼ 2 @ w ðBen Þ Bp
S : ð6:6Þ
en p en
@Ben 0

(ii) the plastic distortion-rates Dp and Dpdis satisfy the dissipation inequality
Tp : Dp þ Sdis : Dpdis P 0: ð6:7Þ
We assume henceforth that (6.5) and (6.6) hold in all motions of the body. We assume further that the
material is strictly dissipative in the sense that

8
Recall that Wpen ¼ 0, cf., (2.26).
9
We content ourselves with constitutive equations that are only sufficient, but generally not necessary for compatibility with
thermodynamics.
1846 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

Tp : Dp > 0 whenever Dp –0; ð6:8Þ


and
Sdis : Dpdis > 0 whenever Dpdis –0: ð6:9Þ
Some remarks: It is possible to enhance the theory by introducing another positive-valued dimen-
sionless strain-like scalar internal variable j which contributes an additional defect energy w  ðpÞ ðjÞ,
iso
p
which evolves as j evolves according to an evolution equation of the type j_ ¼ hðd ; jÞ. This will lead
def  ðpÞ
to a scalar internal stress KðjÞ ¼ @ w iso ðjÞ=@ j, and the dissipation inequality (6.7) will be modified as

Tp : Dp þ Sdis : Dpdis  K j_ P 0: ð6:10Þ


Although a term such as  ðpÞ ð
jÞ might be needed for a more complete characterization of the ‘‘stored
w iso
energy of cold work,” and a more precise determination of dissipation and thereby heat generation (cf.,
e.g., Rosakis et al., 2000; Ristinmaa et al., 2007), here, since our interest is in an isothermal theory, we
refrain from introducing such additional complexity. However, we do wish to emphasize that our
internal variable S, which represents an isotropic deformation resistance to plastic flow, is in no sense
introduced as being derived from a defect energy w  ðpÞ ðjÞ – it is introduced on a purely pragmatic phe-
iso
nomenological basis. Note that Ristinmaa et al. (2007) consider only ‘‘isotropic” hardening theories,
while the theory under consideration here is for combined isotropic and kinematic hardening, and
the term w ðpÞ ðBp Þ already accounts for some aspects of the ‘‘stored energy of cold work.”
en
Vladimirov et al. (2008) have considered a combined isotropic and kinematic-hardening theory
with an additional defect energy wiso ðjÞ, and write their corresponding stress as @wiso ðjÞ=@ j ¼
ðRðjÞÞ  KðjÞ. They associate the stress ðRðjÞÞ as the strain-hardening contribution to the isotropic
pffiffiffiffiffiffiffiffi p
deformation resistance, and take j_  ð 2=3Þd , so that j represents an accumulated plastic strain. It
is our opinion that for materials which show strong isotropic strain hardening, the theory of Reese and
co-workers will grossly underestimate the amount of dissipation. This is the major reason why we do not
follow the approach of Reese and co-workers, but directly introduce our stress-like internal variable S
to represent isotropic strain hardening.

6.2. Isotropy

The following definitions help to make precise our notion of an isotropic material (cf., Anand and
Gurtin, 2003):

þ
(i) Orth = the group of all rotations (the proper-orthogonal group);
(ii) the symmetry group GR , is the group of all rotations of the reference configuration that leaves the
response of the material unaltered;
(iii) the symmetry group GS at each time t, is the group of all rotations of the structural space that
leaves the response of the material unaltered;
(iv) the symmetry group GSS at each time t, is the group of all rotations of the substructural space that
leaves the response of the material unaltered.

Note that there is a slight difficulty in attaching a ‘‘physical meaning” to a rotation in GSS . Since the
substructural space itself is a mathematical construct defined as the range of the linear map Fpdis , a rota-
tion Q 2 GSS should be considered as a rotation in a ‘‘thought experiment” involving the substructural
space; such a rotation may never be ‘‘physically attainable.” We now discuss the manner in which the
basic fields transform under such transformations, granted the physically natural requirement of
invariance of the internal power or equivalently, the requirement that

Te : C_ e ; Tp : Dp ; Sen : Dpen ; and Sdis : Dpdis be invariant: ð6:11Þ

6.2.1. Isotropy of the reference space


Let Q be a time-independent rotation of the reference configuration. Then F ! FQ , and hence
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1847

Fp ! Fp Q ; Fpdis ! Fpdis Q and Fe and Fpen are invariant; and hence Ce and Cpen are invariant;
ð6:12Þ
so that, by (2.5), (2.12), and (2.21),

C_ e ; Dp ; Dpen ; and Dpdis are invariant:


We may therefore use (6.11) to conclude that
Te ; Tp ; Sen ; and Sdis are invariant: ð6:13Þ
Thus, with reference to the special constitutive functions (6.1), we note the these functions are not af-
fected by rotations Q 2 GR .

6.2.2. Isotropy of the structural space


Next, let Q , a time-independent rotation of the intermediate structural space, be a symmetry transfor-
mation. Then F is unaltered by such a rotation, and hence

Fe ! Fe Q ; Fp ! Q > Fp ; Fpen ! Q > Fpen ; and Fpdis is invariant; ð6:14Þ


and also

Ce ! Q > Ce Q ; C_ e ! Q > C_ e Q ; Dp ! Q > Dp Q ; Bpen ! Q > Bpen Q ;


Dpen ! Q > Dpen Q ; and Dpdis is invariant: ð6:15Þ
Then (6.15) and (6.11) yield the transformation laws

Te ! Q > Te Q ; Tp ! Q > Tp Q ; Sen ! Q > Sen Q ; and Sdis is invariant: ð6:16Þ


Thus, with reference to the constitutive equations (6.1) we conclude that
9
 ðeÞ ðCe Þ ¼ w
w  ðeÞ ðQ > Ce Q Þ; >
>
>
p
 ðB Þ ¼ w
ðpÞ  ðpÞ ðQ > Bp Q Þ; >
>
w en en >
=
> e e  e ðQ > Ce Q Þ;
Q T ðC ÞQ ¼ T ð6:17Þ
>
>
Q >S en ðBp ÞQ ¼ S en ðQ > Bp Q Þ; >
>
en en >
>
 p ðDp ; SÞQ ¼ T
p ðQ > Dp Q ; SÞ; ;
Q >T
must hold for all rotations Q in the symmetry group GS at each time t, while the constitutive functions
dis ðDp ; dp ; SÞ, and hðdp ; SÞ are invariant to such rotations.
S dis
We refer to the material as one which is structurally isotropic if
þ
GS ¼ Orth ; ð6:18Þ
so that the response of the material is invariant under arbitrary rotations of the intermediate struc-
tural space at each time t. Henceforth, we assume that (6.18) holds. In this case, the response functions
 ðeÞ , w
w  ðpÞ , T en , and T
e , S  p must each be isotropic.

6.2.3. Isotropy of the substructural space


Finally, let Q , a time-independent rotation of the substructural space, be a symmetry transformation.
Then F, Fe , and Fp are unaltered by such a rotation, and hence

Fpen ! Fpen Q ; Fpdis ! Q > Fpdis ; ð6:19Þ


and also

Dpdis ! Q > Dpdis Q ; and Dpen is invariant: ð6:20Þ


Then (6.20) and (6.11) yield the transformation laws

Sdis ! Q > Sdis Q ; and Sen is invariant: ð6:21Þ


1848 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

Thus, with reference to the constitutive equations (6.1) we conclude that


p dis ðQ > Dp Q ; d ; SÞ
dis ðDp ; d ; SÞQ ¼ S p
Q >S dis dis ð6:22Þ
must hold for all rotations Q in the symmetry group GSS at each time t, while the other constitutive
functions are invariant to such rotations.
We refer to the material as one which is substructurally isotropic if
þ
GSS ¼ Orth ; ð6:23Þ
so that the response of the material is also invariant under arbitrary rotations of the substructural
dis is an isotro-
space at each time. Henceforth, we assume that (6.23) also holds at each time, so that S
pic function.
Thus, by assumption, all the tensorial response functions in (6.1) are isotropic. We confine our
attention to materials which may be adequately defined by such isotropic functions, and refer to such
materials as isotropic.

6.3. Consequences of isotropy of the free energy

 ðeÞ ðCe Þ is an isotropic function of Ce , it has the representation


Since w
 ðeÞ ðCe Þ ¼ w
w ~ ðeÞ ðICe Þ; ð6:24Þ
where
 
ICe ¼ I1 ðCe Þ; I2 ðCe Þ; I3 ðCe Þ
is the list of principal invariants of Ce . Thus, from (6.5),
~ ðeÞ
 e ðCe Þ ¼ 2 @ w ðICe Þ ;
Te ¼ T ð6:25Þ
@Ce
 e ðCe Þ is an isotropic function of Ce . Then since the Mandel stress is defined by (cf., (5.4))
and T
Me ¼ Ce Te ;
 e ðCe Þ is isotropic, we find that Te and Ce commute,
and T
Ce Te ¼ Te Ce ; ð6:26Þ
e
and hence that the elastic Mandel stress M is symmetric.
Further the defect free energy has a representation
 
w ~ ðpÞ ðI p Þ where I p ¼ I1 ðBp Þ; I2 ðBp Þ; I3 ðBp Þ ;
 ðpÞ ¼ w ð6:27Þ
Ben Ben en en en

and this yields that


~ ðpÞ ðI p Þ
@w Ben
Bpen ð6:28Þ
@Bpen
is a symmetric tensor, and (6.6) reduces to
!
~ ðpÞ ðI p Þ
@w
en ðBp Þ ¼ 2 Ben
Sen ¼ S Bpen ; ð6:29Þ
en
@Bpen
0

a symmetric and deviatoric tensor.


Also the plastic Mandel stress defined in (4.33) is
!
~ ðpÞ ðI p Þ
@w Ben
Mpen ¼ 2Fp> Bpen Fp>
en ;
en
@Bpen
0
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1849

" ! #
~ ðpÞ ðI p Þ
@w Ben
¼ 2 Fp> Bpen Fp> ;
en
@Bpen en
0
" ! #
~ ðpÞ ðI p Þ
@w Ben
¼ 2 Fp> Fpen ; ð6:30Þ
en
@Bpen
0

~ ðpÞ ðI p Þ=@Bp is symmetric, this stress is symmetric and deviatoric.


and since @ w Ben en

7. Flow rules

Henceforth, in accord with common terminology we call the symmetric and deviatoric tensor Sen a
back-stress, and we denote an effective deviatoric Mandel stress by
  def
Meeff 0
¼ Me0  Sen : ð7:1Þ
Then, upon using the constitutive relation (6.1)4 and the first microforce balance (4.29), together with
symmetry of the Mandel stress discussed above, a central result of the theory is the first flow rule
 
Meeff  p ðDp ; SÞ:
¼T ð7:2Þ
0

Next, on account of the symmetry and deviatoric nature of the plastic stress Mpen (cf., (6.30)), the second
microforce balance (4.34) and the constitutive equation (6.4)5 yield the second flow rule of the theory:
dis ðDp ; d ; SÞ: p
Mpen ¼ S dis ð7:3Þ
We now make two major assumptions concerning the plastic flow for isotropic materials:

(1) Codirectionality hypotheses: Recall


p p
d ¼ jDp j; and let ddis ¼ jDpdis j;
also let
def Dp def Dpdis
Np ¼ p and Npdis ¼ p ð7:4Þ
d ddis
denote the directions of plastic flow whenever Dp –0 and Dpdis –0. Then the dissipation inequality
(6.7) may be written as
   
 p ðdp ; Np ; SÞ : Np dp þ S
T dis ðdp ; Np ; dp ; SÞ : Np dp P 0: ð7:5Þ
dis dis dis dis

Guided by (7.5), we assume henceforth that the dissipative flow stress Tp is parallel to and
points in the same direction as Np so that
 p ðdp ; Np ; SÞ ¼ Yðdp ; Np ; SÞNp ;
T ð7:6Þ
where
p  p ðdp ; Np ; SÞ : Np
Yðd ; Np ; SÞ ¼ T ð7:7Þ
represents a scalar flow strength of the material.
Similarly we assume that dissipative flow stress Sdis is parallel to and points in the same direc-
tion as Npdis so that
dis ðdp ; Np ; dp ; SÞ ¼ Y dis ðdp ; Np ; dp ; SÞNp ;
S ð7:8Þ
dis dis dis dis dis

where
p p dis ðdp ; Np ; dp ; SÞ : Np
Y dis ðddis ; Npdis ; d ; SÞ ¼ S dis dis dis ð7:9Þ
represents another scalar flow strength of the material.
We refer to the assumptions (7.6) and (7.8) as the codirectionality hypotheses.10

10
These assumptions corresponds to the classical notion of maximal dissipation in Mises-type theories of metal plasticity.
1850 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

(2) Strong isotropy hypotheses:


p
We also assume that the scalar flow strength Yðd ; Np ; SÞ is independent of the flow direction Np ,
so that
p
Yðd ; SÞ: ð7:10Þ
p p
Similarly, we also assume that the scalar flow strength Y dis ðddis ; Npdis ; d ; SÞ is independent of the
flow direction Npdis , so that
p p
Y dis ðddis ; d ; SÞ: ð7:11Þ
We refer to the assumptions (7.10) and (7.11) as the strong isotropy hypotheses.

Thus, using (7.10) and (7.6), the flow rule (7.2) reduces to,
p
ðMeeff Þ0 ¼ Yðd ; SÞNp ; ð7:12Þ
which immediately gives
ðMeeff Þ0
Np ¼ ; ð7:13Þ
jðMeeff Þ0 j
and
p
jðMeeff Þ0 j ¼ Yðd ; SÞ: ð7:14Þ
p
When jðMeeff Þ0 j
and S are known, (7.14) serves as an implicit equation for the scalar flow rate d .
Next, using (7.11) and (7.8), the flow rule (7.3) reduces to,
p p
Mpen ¼ Y dis ðddis ; d ; SÞNpdis ; ð7:15Þ
which gives
Mpen
Npdis ¼ ; ð7:16Þ
jMpen j
and
p p
jMpen j ¼ Y dis ðddis ; d ; SÞ: ð7:17Þ
p
When jðMeeff Þ0 j
and S are known, (7.17) serves as an implicit equation for the scalar flow rate ddis .
Finally, using (7.6), (7.10), and (7.14), as well as (7.8), (7.11), and (7.17) the dissipation inequality
(6.7) reduces to
p p
jðMeeff Þ0 jd þ jMpen jddis P 0: ð7:18Þ

8. Summary of the constitutive theory

In this section we summarize our constitutive theory.

(1) Free energy.


~ ðeÞ ðICe Þ þ w
w¼w ~ ðpÞ ðI p Þ; ð8:1Þ
Ben

e
where ICe and IBpen are the lists of the principal invariants of C and Bpen , respectively.
(2) Cauchy stress.

def   ~ ðeÞ ðICe Þ


@w
T ¼ J 1 Fe Te Fe> where Te ¼ 2 : ð8:2Þ
@Ce
(3) Mandel stresses. Back-stress. The elastic Mandel is given by
Me ¼ Ce Te ; ð8:3Þ
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1851

and is symmetric.
The back-stress is given by
!
~ ðpÞ ðI p Þ
@w Ben
Sen ¼ 2 Bpen ; ð8:4Þ
@Bpen
0

and is symmetric deviatoric. The stress difference


Meeff ¼ Me  Sen ð8:5Þ
is called an effective elastic Mandel stress.
The plastic Mandel stress is given by
" ! #
~ ðpÞ ðI p Þ
@w Ben
Mpen ¼ 2 Fp> Fpen ; ð8:6Þ
en
@Bpen
0

and is symmetric and deviatoric.


(4) Flow rules. The evolution equation for Fp is

F_ p ¼ Dp Fp ; ð8:7Þ
p
with D given by

p ðMeeff Þ0
Dp ¼ d Np ; Np ¼ ; ð8:8Þ
jðMeeff Þ0 j
p
where the scalar flow rate d is obtained by solving the scalar strength relation
p
jðMeeff Þ0 j ¼ Yðd ; SÞ; ð8:9Þ
p
for given ðMeeff Þ0 and S, where Yðd ; SÞ is a rate-dependent flow strength. The evolution equation
for Fpdis is

F_ pdis ¼ Dpdis Fpdis ; ð8:10Þ


with Dpdis given by

p Mpen
Dpdis ¼ ddis Npdis ; Npdis ¼ ; ð8:11Þ
jMpen j
p
where the scalar flow rate ddis is obtained by solving the scalar strength relation
p p
jMpen j ¼ Y dis ðddis ; d ; SÞ; ð8:12Þ
p p p
for given Mpen ,
d , and S, where Y dis ðddis ; d ; SÞ is another rate-dependent flow strength.
(5) Evolution equation for S.
p
S_ ¼ hðd ; SÞ: ð8:13Þ

The evolution equations for Fp , Fpdis , and S need to be accompanied by initial conditions. Typical ini-
tial conditions presume that the body is initially (at time t ¼ 0, say) in a virgin state in the sense that
FðX; 0Þ ¼ Fp ðX; 0Þ ¼ Fpdis ðX; 0Þ ¼ 1; SðX; 0Þ ¼ S0 ð¼ constantÞ; ð8:14Þ
e p p
so that by F ¼ F F and F ¼ Fpen Fpdis e
we also have F ðX; 0Þ ¼ 1 and Fpen ðX; 0Þ ¼ 1.

9. Specialization of the constitutive equations

In this section, based on experience with existing recent theories of viscoplasticity with isotropic
and kinematic hardening, we specialize our constitutive theory by imposing additional constitutive
assumptions.
1852 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

9.1. Free-energy wðeÞ

The spectral representation of Ce is


X
3
Ce ¼ xei rei  rei ; with xei ¼ ke2
i ; ð9:1Þ
i¼1

where ðre1 ; re2 ; re3 Þ are the orthonormal eigenvectors of Ce and Ue , and ðke1 ; ke2 ; ke3 Þ are the positive eigen-
values of Ue . Instead of using the invariants ICe , the free-energy wðeÞ for isotropic materials may be
alternatively expressed in terms of the principal stretches as

 ðeÞ ðke ; ke ; ke Þ:
wðeÞ ¼ w ð9:2Þ
1 2 3

Then, by the chain-rule and (8.2)2 , the stress Te is given by

 ðeÞ ðke ; ke ; ke Þ
@w X3  ðeÞ ðke ; ke ; ke Þ @ke X
@w 3  ðeÞ ðke ; ke ; ke Þ @ xi
1 @w
Te ¼ 2 1
e
2 3
¼2 1
e
2 3 i
e ¼ e
1 2 3
: ð9:3Þ
@C i¼1
@ki @C i¼1
ki @kei @Ce

Assume that the squared principal stretches xei are distinct, so that the xei and the principal directions
rei may be considered as functions of Ce ; then
@ xei
¼ rei  rei ; ð9:4Þ
@Ce
and, granted this, (9.4) and (9.3) imply that
X3
1 @w ðeÞ ðke ; ke ; ke Þ
Te ¼ e
1 2 3
rei  rei : ð9:5Þ
i¼1
k i @kei

Further, from (8.2)1 ,


!
1 e e e> e e e e e> e
X
3  ðeÞ ðke ; ke ; ke Þ
@w
T¼J F T F 1
¼J R U T U R 1
¼J R kei 1 2 3
ri  ri Re> :
e e
ð9:6Þ
i¼1
@kei

Next, since Me ¼ Ce Te (cf., (8.3)), use of (9.1) and (9.5) gives the Mandel stress as
X
3  ðeÞ ðke ; ke ; ke Þ
@w
Me ¼ kei 1 2 3
rei  rei : ð9:7Þ
i¼1
@kei

Let

def
X
3
Ee ¼ ln Ue ¼ Eei rei  rei ; ð9:8Þ
i¼1

denote the logarithmic elastic strain with principal values


def
Eei ¼ ln kei ; ð9:9Þ
and consider an elastic free-energy function of the form

w  ðeÞ ðEe ; Ee ; Ee Þ;
 ðeÞ ðke ; ke ; ke Þ ¼ w ð9:10Þ
1 2 3 1 2 3

so that, using (9.7),


X3  ðeÞ ðEe ; Ee ; Ee Þ
@w
Me ¼ 1 2 3
rei  rei : ð9:11Þ
i¼1
@Eei
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1853

We consider the following simple generalization of the classical strain-energy function of infinitesimal
isotropic elasticity which uses a logarithmic measure of finite strain11

 ðeÞ ðEe Þ ¼ GjEe j2 þ 1 KðtrEe Þ2 ;


w ð9:12Þ
0
2
where
G > 0 and K > 0; ð9:13Þ
are the shear modulus and bulk modulus, respectively. Then, (9.11) gives
Me ¼ 2GEe0 þ KðtrEe Þ1 ð9:14Þ
and on account of (9.6), (9.7), and (9.14),

T ¼ J 1 Re Me Re> : ð9:15Þ

9.2. Free-energy wðpÞ

The spectral representation of Bpen is


X
3
Bpen ¼ bi l i  l i ; with bi ¼ a2i ; ð9:16Þ
i¼1

where ðl1 ; l2 ; l3 Þ are the orthonormal eigenvectors of Bpen and Vpen , and ða1 ; a2 ; a3 Þ are the positive eigen-
values Vpen . Instead of using the invariants IBpen , the free-energy wðpÞ may alternatively be expressed as
 ðpÞ ða1 ; a2 ; a3 Þ:
wðpÞ ¼ w ð9:17Þ
Then, by the chain-rule
 ðpÞ ða1 ; a2 ; a3 Þ X
@w 3  ðpÞ ða1 ; a2 ; a3 Þ @ai
@w 1X 3  ðpÞ ða1 ; a2 ; a3 Þ @bi
1 @w
p ¼ p ¼ : ð9:18Þ
@Ben i¼1
@a i @B en 2 i¼1
ai @ai @Bpen

Assume that bi are distinct, so that the bi and the principal directions li may be considered as functions
of Bpen . Then,
@bi
¼ li  li ; ð9:19Þ
@Bpen
and, granted this, (9.18) implies that
 ðpÞ ða1 ; a2 ; a3 Þ 1 X
@w 3  ðpÞ ða1 ; a2 ; a3 Þ
1 @w
p ¼ li  li : ð9:20Þ
@Ben 2 i¼1
ai @ai

Also, use of (9.16) and (9.20) in (8.4) gives the deviatoric back-stress as
!
X
3  ðpÞ ða1 ; a2 ; a3 Þ
@w
Sen ¼ ai li  li : ð9:21Þ
i¼1
@ai
0

Let

def
X
3
Epen ¼ ln Vpen ¼ ln ai li  li : ð9:22Þ
i¼1

11
Vladimirov et al. (2008) employ a Neo-Hookean free-energy function modified for elastic compressibility (cf., their Eq. (28)).
Here, as an alternate, we use the classical strain-energy function for infinitesimal elasticity, but employ the large deformation
Hencky-logarithmic elastic strain Ee . Anand (1979, 1986) has shown that the simple free-energy function (9.12) is useful for
moderately large elastic stretches. Additionally, use of the logarithmic strain helps in developing implicit time-integration
procedures based on the exponential map (cf., Appendix A).
1854 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

denote a logarithmic energetic strain with principal values ðln ai Þ. Note that since trVpen ¼ a1 a2 a3 ¼ 1,
trðEpen Þ ¼ ln a1 þ ln a2 þ ln a3 ¼ lnða1 a2 a3 Þ ¼ 0; ð9:23Þ
and hence the strain tensor Epen is traceless, Next, consider the following simple defect energy:
h i
 ðpÞ ðEp Þ ¼ BjEp j2 ¼ B ðln a1 Þ2 þ ðln a2 Þ2 þ ðln a3 Þ2 ;
w ð9:24Þ
en en

with B > 0, which, since Epen is deviatoric, is analogous to (9.12). Then using (9.21), (9.22), and (9.23)
Sen ¼ 2BEpen ; ð9:25Þ
we call the positive-valued constitutive parameter B the back-stress modulus.
Then the plastic Mandel stress (4.33) becomes
 
Mpen ¼ Fp> p> p> p p p1
en Sen Fen ¼ R en V en ð2BEen ÞV en Rpen ;

or

Mpen ¼ Rp> p
en Sen R en : ð9:26Þ

9.3. Strength functions

First consider the strength relation


p
jðMeeff Þ0 j ¼ Yðd ; SÞ; ð9:27Þ
appearing in (8.9). Define an equivalent shear stress by
1
s def
¼ pffiffiffi jðMeeff Þ0 j; ð9:28Þ
2
and an equivalent shear strain rate by
pffiffiffi pffiffiffi
mp def p
¼ 2d ¼ 2jDp j; ð9:29Þ
respectively. Using the definitions (9.28) and (9.29), we rewrite the strength relation (9.27) as12
s ¼ Yðmp ; SÞ; ð9:30Þ
and assume that the shear flow strength Yðmp ; SÞ has the simple form
Yðmp ; SÞ ¼ gðmp ÞS; ð9:31Þ
where S now represents an isotropic flow resistance in shear, and
gð0Þ ¼ 0; gðmp Þ is a strictly increasing function of mp : ð9:32Þ
p
We refer to the dimensionless function gðm Þ as the rate-sensitivity function. Next, by (9.32) the func-
tion gðmp Þ is invertible, and the inverse function f ¼ g 1 is strictly increasing and hence strictly positive
for all nonzero arguments; further f ð0Þ ¼ 0. Hence the relation (9.30) may be inverted to give an
expression
   
s s
mp ¼ g 1 f ð9:33Þ
S S
for the equivalent plastic shear strain rate.
An example of a commonly used rate-sensitivity function is the power-law function
 p m
m
gðmp Þ ¼ ; ð9:34Þ
m0

12
pffiffiffi
We absorb inconsequential factors of 2 in writing (9.30).
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1855

where m > 0, a constant, is a rate-sensitivity parameter and m0 > 0, also a constant, is a reference flow-
rate. The power-law function allows one to characterize nearly rate-independent behavior, for which m
is very small. Further, granted the power-law function (9.34), the expression (9.33) has the specific
form
 m1
s
mp ¼ m0 : ð9:35Þ
S
Next consider the strength relation
p p
jMpen j ¼ Y dis ðddis ; d ; SÞ; ð9:36Þ
appearing in (8.12). Define another equivalent shear stress by
1
sen def
¼ pffiffiffi jMpen j; ð9:37Þ
2
and another equivalent shear strain rate by
pffiffiffi pffiffiffi
mpdis def p
¼ 2ddis ¼ 2jDpdis j; ð9:38Þ
13
respectively. Using the definitions (9.37) and (9.38), we rewrite the strength relation (9.36) as
sen ¼ Y dis ðmpdis ; mp ; SÞ: ð9:39Þ
p p
A specialization of Y dis ðm dis ; m ; SÞ which leads to Armstrong and Frederick (1966) type kinematic hard-
ening is
 
B
Y dis ðmpdis ; mp ; SÞ ¼ mpdis ; ð9:40Þ
cm p

where B is the back-stress modulus (cf., (9.25)), and c P 0 is a dimensionless constant. Using (9.40),
(9.39) becomes
 
B
sen ¼ mpdis ; ð9:41Þ
cmp
so that the term ðB=cmp Þ represents a ‘‘pseudo”-viscosity (Dettmer and Reese, 2004). The relation
(9.41) may be inverted to give
 
sen
mpdis ¼ cmp ; ð9:42Þ
B
which shows that mpdis ¼ 0 when either c ¼ 0 or when mp ¼ 0. Thus note that the constitutive assump-
tion (9.42) constrains the theory such that Dpdis ¼ 0 whenever Dp ¼ 0, and this in turn will lead to no
change in the back-stress Sen when Dp ¼ 0, that is when there is no macroscopic plastic flow.

9.4. Evolution equation for S

The evolution equation for the isotropic deformation resistance is specialized in a rate-independent
form as

S_ ¼ hðSÞmp ð9:43Þ
with hðSÞ a hardening function, which we take to be given by (Brown et al., 1989):
(  a
h0 1  SS for S0 6 S 6 S ;
hðSÞ ¼ ð9:44Þ
0 for S P S ;

13
pffiffiffi
Again absorbing inconsequential factors of 2 in writing (9.39).
1856 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

where S , a, and h0 are constant moduli with S > S0 , a P 1, and h0 > 0. The hardening function (9.44)
is strictly decreasing for S0 6 S 6 S and vanishes for S P S .

9.5. Dissipation inequality

Finally, using (9.28), (9.29), (9.37), (9.38), and (9.42), the dissipation inequality (7.18) reduces to
 c 
s þ ðsen Þ2 mp P 0: ð9:45Þ
B

9.6. Summary of the specialized constitutive model

In this section, we summarize the specialized form of our theory, which should be useful in appli-
cations. The theory relates the following basic fields:

x ¼ vðX; tÞ, motion;


F ¼ rv; J ¼ det F > 0, deformation gradient;
F ¼ Fe Fp , elastic–plastic decomposition of F;
Fe ; J e ¼ det Fe ¼ J > 0, elastic distortion;
Fp ; J p ¼ det Fp ¼ 1, inelastic distortion;
Fe ¼ Re Ue , polar decomposition of Fe ;
Dp ¼ symðF_ p Fp1 Þ, plastic stretching corresponding to Fp ;
Fp ¼ Fpen Fpdis , elastic–plastic decomposition of Fp ;
Fpen ; det Fpen ¼ 1, energetic part of Fp ;
Fpdis ; det Fpdis ¼ 1, dissipative part of Fp ;
Fpen ¼ Vpen Rpen , polar decomposition of Fpen ;
Dpdis ¼ symðF_ pdis Fp1
dis
Þ, plastic stretching corresponding to Fpdis ;
T ¼ T> , Cauchy stress;
w, free-energy density per unit volume
of intermediate structural space;
S > 0; scalar internal variable.

(1) Free energy.


We consider a separable free energy
w ¼ wðeÞ þ wðpÞ : ð9:46Þ
With
X
3
Ue ¼ kei rei  rei ; ð9:47Þ
i¼1

denoting the spectral representation of Ue , and with


X
3
Ee ¼ ln kei rei  rei ; ð9:48Þ
i¼1
denoting an elastic logarithmic strain measure, we adopt the following special form for the free-
energy wðeÞ :
1
wðeÞ ¼ GjEe0 j2 þ KðtrEe Þ2 ; ð9:49Þ
2
where
G > 0; and K > 0; ð9:50Þ
are the shear modulus and bulk modulus, respectively.
Further, with
X
3
Vpen ¼ ai li  li ; ð9:51Þ
i¼1
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1857

denoting the spectral representation of Vpen , and with


X
3
Epen ¼ ln ai li  li ; trEpen ¼ 0; ð9:52Þ
i¼1

we adopt a free-energy wðpÞ of the form

wðpÞ ¼ BjEpen j2 ; ð9:53Þ


where the positive-valued parameter
B P 0; ð9:54Þ
is a back-stress modulus.
(2) Cauchy stress. Mandel stresses. Back-stress. Effective stress.
Corresponding to the special free-energy functions considered above, the Cauchy stress is given
by

T ¼ J 1 Re Me Re> ; ð9:55Þ
where
Me ¼ 2GEe0 þ KðtrEe Þ1; ð9:56Þ
is an elastic Mandel stress.
The symmetric and deviatoric back-stress is defined by
Sen ¼ 2BEpen ; ð9:57Þ
p
and the driving stress for D is the effective stress given by
ðMeeff Þ0 ¼ Me0  Sen : ð9:58Þ
The corresponding equivalent shear stress is given by
1
s ¼ pffiffiffi jðMeeff Þ0 j: ð9:59Þ
2
The driving stress for Dpdis is the stress measure

Mpen ¼ Rp> p
en Sen R en : ð9:60Þ
The corresponding equivalent shear stress is given by
1
sen ¼ pffiffiffi jMpen j: ð9:61Þ
2
(3) Flow rules.
The evolution equation for Fp is
9
F_ p ¼ Dp Fp ; Fp ðX; 0Þ ¼ 1; >
>
>
 e  =
ðM Þ
Dp ¼ mp 2effs 0 ; ð9:62Þ
>
>
 >
;
mp ¼ m0 sS 1=m ;
where m0 is a reference plastic strain rate with units of 1=time, and m is a strain-rate-sensitivity
parameter.
The evolution equation for Fpdis is
9
F_ pdis ¼ Dpdis Fpdis ; Fpdis ðX; 0Þ ¼ 1; >
>
 p =
M
Dpdis ¼ mpdis 2senen ; ð9:63Þ
  >
>
;
mpdis ¼ c sBen mp ;
where c P 0 is a dimensionless constant.
1858 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

(4) Evolution equation for the isotropic-hardening variable S.


The internal variable S is taken to obey the evolution equation
9
S_ ¼ hðSÞmp ; SðX; 0Þ ¼ S0 ; >
>
=
(  a
h0 1  SS for S0 6 S 6 S ; >
ð9:64Þ
hðSÞ ¼ >
;
0 for S P S ;

where S , a, and h0 are constant moduli with S > S0 , a P 1, and h0 > 0.

10. Some numerical solutions

In Appendix A, we develop a semi-implicit time-integration procedure, and obtain an (approxi-


mate) algorithmic tangent for the special constitutive theory summarized in the previous section.
The time-integration procedure and the associated algorithmic tangent have been implemented in
the commercial finite element program Abaqus/Standard (2008) by writing a user material subroutine
(UMAT). In this section, we investigate some salient properties of the model and the accompanying
time-integration procedure.

10.1. Simple tension/compression

First, in order to demonstrate that the constitutive theory captures the various isotropic and kine-
matic-hardening effects, we carried out a numerical simulation of a symmetric strain-cycle in simple
tension and compression. The material parameters used in our numerical study are:
G ¼ 80 GPa; K ¼ 175 GPa;
m0 ¼ 0:001 s1 ; m ¼ 0:02;
ð10:1Þ
S0 ¼ 100 MPa; h0 ¼ 1250 MPa; a ¼ 1 S ¼ 250 MPa;
B ¼ 40 GPa; c ¼ 400;
the value of m ¼ 0:02 is intended to represent a nearly rate-independent material.
The axial strain-cycle was imposed between true strain limits of  ¼ 0:02 at an absolute axial true
strain rate of j_ j ¼ 0:01 s1 ; a fixed time step Dt ¼ 0:01 s was used in the numerical simulations. The cal-
culations were performed for: (i) no hardening, h0 ¼ 0 and B ¼ 0; (ii) isotropic hardening only, h0 –0 and
B ¼ 0; (iii) kinematic hardening only, h0 ¼ 0 and B–0; and finally (iv) combined kinematic and isotropic
hardening, h0 –0 and B–0. Plots of the axial stress r versus the axial strain  are shown in Fig. 1(a). With no
hardening, we obtain an elastic-perfectly-plastic response, as expected. When isotropic hardening only is
allowed, we observe an increase in the flow stress, and the stress level at which plastic flow recommences
on strain reversal is equal in magnitude to the stress level from which the reversal in strain was initiated.
In the case of kinematic hardening only, we again observe an increase in the flow stress, but this time the
magnitude of the stress level at which plastic flow recommences on strain reversal is substantially smaller
than the stress level from which the reversal in strain direction was initiated; a clear manifestation of the
Bauschinger effect. When isotropic hardening is combined with kinematic hardening, we see the contin-
ual evolution in the flow stress associated with combined isotropic and kinematic hardening during for-
ward straining, and a clear Bauschinger effect due to the kinematic hardening upon strain reversal.

10.2. Simple shear

Next, we concentrate on numerical solutions related to homogeneous simple shear. With respect to
a rectangular Cartesian coordinate system with origin o and orthonormal base vectors fei ji ¼ 1; 3g, a
simple shearing motion is described by
x ¼ X þ ðmtÞX 2 e1 ; ð10:2Þ
with m a shear strain rate, and C ¼ mt is the amount of shear. Throughout this section we use the mate-
rial parameters summarized in (10.1).
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1859

a 600

400

Axial stress, σ (MPa) 200


No hardening
Isotropic only
0
Kinematic only
Combined
−200

−400

−600
−0.02 −0.01 0 0.01 0.02
Axial strain, ε

b 300

200
(MPa)

100
No hardening
Isotropic only
12

0
Stress, T

Kinematic only
Combined
−100

−200

−300
0 0.01 0.02 0.03 0.04 0.05
Amount of shear, Γ
Fig. 1. (a) Comparison of axial stress r versus axial strain  for various types hardening in a symmetric strain-cycle simulation
in simple tension and compression. (b) Comparison of shear stress T 12 versus shear strain C for various types of hardening in
reversed simple shear.

The specific goals of our numerical solutions related to simple shear are as follows:

 To further demonstrate that the constitutive theory captures the major features of isotropic and
kinematic hardening in the case of simple reversed shear.
 To investigate the convergence and stability properties of the time-integration scheme in simple
reversed shear.
 To demonstrate that the constitutive equations do not suffer from oscillations in the stress response
at large shear strains.
 To demonstrate that the constitutive equations and the time-integration procedure are indeed
objective.
1860 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

First, we wish to further demonstrate that the constitutive theory captures the various isotropic
and kinematic-hardening effects. To this end, we consider simple shear, as described above, for a time
of 5 s at a shear strain rate of m ¼ 0:01 s1 for a total shear strain C ¼ mt ¼ 0:05, and then at a shear
strain rate of m ¼ 0:01 s1 for another 5 s to complete one reversal of strain, while using a fixed time
step Dt ¼ 0:01 s. As for simple tension/compression, the calculations were performed for: (i) no hard-
ening, h0 ¼ 0 and B ¼ 0; (ii) isotropic hardening only, h0 –0 and B ¼ 0; (iii) kinematic hardening only,
h0 ¼ 0 and B–0; and finally (iv) combined kinematic and isotropic hardening, h0 –0 and B–0, and the
results are shown in Fig. 1(b). Again, we observe the important qualitative features of each case in sim-
ple reversed shear, i.e. elastic-perfectly-plastic response in the case of no hardening, a continual in-
crease in the flow stress in cases involving isotropic hardening, and a clear Bauschinger effect in
cases involving kinematic hardening.
Next, we investigated the stability and convergence properties of our time-integration procedure.
Since our integration procedure is semi-implicit rather than fully-implicit, one might be concerned
that there may be a restrictive time step constraint for stability. To ascertain any time step restrictions,
we performed a convergence test. Reversed simple shear to a maximum shear strain of C ¼ 0:05 at a
shear strain rate of jmj ¼ 0:01 s1 was considered while increasingly coarsening the time step. The spe-
cific time steps Dt of

1 103 ; 1 102 ; 2:5 102 ; 1 101 ; 2:5 101 ; and 1 s


were considered; these time steps translate to shear strain increments of

1 105 ; 1 104 ; 2:5 104 ; 1 103 ; 2:5 103 ; and 1 102 ;


respectively. Fig. 2 shows the stress–strain curves corresponding to these cases. For increasing time
steps, no evidence of instability is observed, but of course the accuracy of the solution degrades, espe-
cially in the regions of elastic–plastic transitions. The order of convergence is demonstrated in Fig. 3;
since no exact solution is available, the solution obtained using a time step of Dt ¼ 1 103 s was trea-
ted as the ‘‘exact” solution and used to calculate the pointwise error for the cases corresponding to the
larger time steps. The error E was then obtained by taking the L-infinity norm of the pointwise error.
As seen in Fig. 3, the time-integration procedure is first-order accurate. Thus, while no stability prob-
lems are encountered, the accuracy of the integration procedures deteriorates, especially when large
time increments are taken during periods of rapidly changing elastic–plastic transitions.

300

200
(MPa)

100
Δ t = 1×10−3 s
12

−2
0 Δ t = 1×10 s
Stress, T

−1
Δ t = 1×10 s
Δt=1s
−100

−200

−300
0 0.01 0.02 0.03 0.04 0.05
Amount of shear, Γ
Fig. 2. Comparison of solutions for reversed simple shear for various time steps.
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1861

8.5

8 1
1
7.5
log(E)

6.5

6
−2.5 −2 −1.5 −1 −0.5 0 0.5
log(Δ t)
Fig. 3. Convergence diagram for the time-integration procedures demonstrating first-order accuracy.

We wish next to demonstrate that the constitutive equations do not produce shear-stress oscilla-
tions at large strains – a feature which has plagued numerous previous large deformation constitutive
theories for kinematic hardening. To this end, we performed a monotonic simple shear simulation to a
very large final shear strain of C ¼ 10, at a shear strain rate of m ¼ 0:01 s1 using a fixed time step
Dt ¼ 0:1 s. The shear T 12 and normal components, T 11 and T 22 , of the Cauchy stress T are plotted in
Fig. 4(a). Clearly there are no oscillations in the T 12 versus C response; the shear stress rises and satu-
rates (because of the chosen value of the material parameters) at a shear strain of C
1:5. We have also
carried out a similar calculation with all other material parameters the same as before, but by setting the
dynamic recovery parameter c ¼ 0; this corresponds to the case of ‘‘linear” kinematic hardening. The
variation of the shear and normal components of the Cauchy stress T for this case are plotted in
Fig. 4(b). Here, since there is no dynamic recovery for the kinematic hardening, the stress levels for this
physically unrealistic case get quite large, but the T 12 versus C curve still does not show any oscillations;
the peak in this curve is due to the use of a logarithmic energetic strain lnðVpen Þ and not due to any oscil-
latory behavior in the evolution of the back-stress.
Finally, in order to demonstrate the frame-indifference of the constitutive equations and numerical
algorithm, we have carried out a numerical calculation for simple shear with superposed rigid rota-
tion, which is described by Weber et al. (1990):
x ¼ Q ðtÞ½X þ ðmtÞX 2 e1 ; ð10:3Þ
with
Q ðtÞ ¼ ðe1  e1 þ e2  e2 Þ cosðxtÞ þ ðe1  e2  e2  e1 Þ sinðxtÞ þ e3  e3 ; ð10:4Þ
a rotation about the e3 -axis. In performing our numerical calculation, we used the material parameters
listed in (10.1), and the calculation was performed with x ¼ 0:1p radians per second and m ¼ 0:01 per
second, for t 2 ½0; 20 , so that h  xt 2 ½0; 2p and C  mt 2 ½0; 0:2 . The calculation was carried by using
a fixed time step of Dt ¼ 0:1 s, which corresponds to a shear strain increment of DC ¼ 0:001 and a rota-
tion increment of Dh ¼ 3:6 . Snapshots of the initial and deformed geometry at a few representative
stages of the simple shear plus superposed rotation are shown in Fig. 5. Let the solution for the stress cor-
responding to the motion (10.3) be denoted by T ðtÞ. The result of the shear component of the ‘‘unrotat-
ed” Cauchy stress ðQ ðtÞ> T ðtÞQ ðtÞÞ versus the amount of shear C, obtained from the numerical
calculation is shown in Fig. 6. Also shown in this figure is T 12 versus C for the motion (10.3) with
Q ðtÞ ¼ 1. The two shear stress versus shear strain curves are indistinguishable from each other; the
1862 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

a 400

350

300

Stress (MPa) 250


T12
200
T
11
150 T
22
100

50

−50
0 2 4 6 8 10
Amount of shear, Γ

b 1.5
5
x 10
T
12
T11
1
T
22

0.5
Stress (MPa)

−0.5

−1

−1.5
0 2 4 6 8 10
Amount of shear, Γ
Fig. 4. Simple shear to large shear strain (a) with dynamic recovery c–0; and (b) without dynamic recovery, c = 0. Note that
there are no shear-stress oscillations in either case.

excellent agreement between the two calculations verifies the objectivity of the constitutive equations
and time-integration procedure.

10.3. Cyclic loading of a curved bar

Next, in order to exercise the algorithmic tangent in a case of inhomogeneous deformation, we consider
the cyclic deformation of a curved bar of circular cross-section. A finite element mesh for the curved bar is
shown in Fig. 7(a). The bar, which has a diameter of 50 mm, curves through a radius of 100 mm for 90 , and
then extends straight for another 100 mm. The mesh consists of 2184 Abaqus-C3D8R elements. The nodes
on the face denoted as ‘‘fixed”-face in Fig. 7(a) have prescribed null displacements, u1 ¼ u2 ¼ u3 ¼ 0, while
the boundary conditions on the face denoted as the ‘‘moved”-face were specified as
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1863

Fig. 5. Simple shear with superposed rigid rotation. The deformed geometry at each stage is shown with a solid line, while the
dashed line represents the initial geometry.

u1 ¼ u3 ¼ 0; and u2 ðtÞ ¼ a sinðxtÞ; with a ¼ 20 mm and x ¼ 0:2p s1 : ð10:5Þ


The calculation was run to a final time of t f ¼ 30 s, so that the u2 -displacement on the ‘‘moved”-face
goes through three complete cycles. The material properties used in the calculation were the same as
those in the previous calculations (10.1). The deformed mesh at the point of maximum deflection is
also shown in Fig. 7(a).
In this (and subsequent) multi-element calculations, a variable time-stepping integration proce-
dure was employed. As in the paper by Lush et al. (1989), the calculation was performed with a
slightly modified version of the Abaqus/Standard static analysis procedure. The modification was
introduced to enhance the automatic time-stepping procedure in Abaqus to control the accuracy of
the constitutive time-integration. This was done by using as a control measure the maximum equiva-
lent plastic shear strain increment

Dcpmax ¼ ðDtÞðmpnþ1 Þmax

occurring at any integration point in the model during an increment. The value of Dcpmax was kept close
to a specified nominal value Dcps . From Fig. 2, we note that the accuracy of the solution begins to
1864 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

350
Shear and rotation
Shear only
300

250
Stress, T12 (MPa)
200

150

100

50

0
0 0.05 0.1 0.15 0.2
Amount of shear, Γ
Fig. 6. Comparison of results for T 12 versus C curves from shear with superposed rotation against shear with no superposed
rotation.

degrade for shear strain increments of around 1 103 ; accordingly, we chose Dcps to be half this va-
lue, 0:5 103 . The automatic time-stepping algorithm operated to keep the ratio
Dcpmax
R ð10:6Þ
Dcps
close to 1.0 by adjusting the size of the time increments.14
As the curved bar is cyclically displaced in the ðe2 ; e3 Þ-plane at the ‘‘moved-face”, the section of
the bar near the ‘‘fixed”-face is subjected to reversed-torsion, while the part of the bar towards the
‘‘moved”-face is subjected to reversed-bending, resulting in states of shear, tension, and compres-
sion in various parts of the body. With increasing displacement magnitude at the ‘‘moved”-face,
inhomogeneous plastic deformation initiates and spreads in various disparate regions of the curved
bar. The integrated reaction force in the 2-direction on the ‘‘moved”-face is plotted against the pre-
scribed u2 -displacement in Fig. 7(b). This resulting overall cyclic load–displacement curve is the re-
sult of the complex evolution of the internal variables controlling the combined isotropic and
kinematic hardening at the various sections of the bar which are undergoing plastic deformation.
The overall load versus displacement response is smooth, and one can clearly see the effects of
both kinematic hardening in the prominent Bauschinger effect and isotropic hardening in the con-
tinual increase in the magnitude of the load with cyclic deformation.
This example shows that our semi-implicit integration procedure, when coupled with our heuris-
tic automatic time-stepping algorithm and our approximate algorithmic tangent, is effective in

14
After an equilibrium solution for a time increment Dtn ¼ t nþ1  tn was found, the value of the ratio R was checked to determine
whether this solution would be accepted or not. If R was greater than 1.25, then the solution was rejected, and a new time
increment was used that was smaller by the factor ð0:85=RÞ. If R 6 1:25, then the solution was accepted, and the value of R was
used to determine the first trial size of the next time increment. The following algorithm was used:

if 0:8 < R 6 1:25 then Dt nþ1 ¼ Dtn =R;


if 0:5 < R 6 0:8 then Dt nþ1 ¼ 1:25Dt n ; ð10:7Þ
if R 6 0:5 then Dt nþ1 ¼ 1:50Dtn :

Note that the measure Dcpmax was allowed to exceed the user specified value Dcps by up to 25%. This was done to avoid having to
recalculate increments that came out just slightly above the specified nominal value, but were otherwise essentially acceptable.
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1865

b 150

100

50
Load (kN)

−50

−100

−150
−20 −10 0 10 20
Displacement, u , (mm)
2

Fig. 7. (a) Undeformed and deformed meshes for the cyclic loading of a curved bar. (b) Resulting tip load versus tip
displacement.

obtaining an accurate, stable, and efficient solution in an inhomogeneous large deformation implicit
finite element calculation.

10.4. Bending and spring-back of an aluminum sheet

As our final numerical example, we consider a simple plane-strain sheet forming operation in
which an aluminum sheet is plastically bent about a mandrel of fixed radius and then unloaded.
We use the recent experimental data of Cao et al. (2008) for AA6111-T4 sheets to estimate the mate-
rial parameters in our model. The fit of the model to the experimental data is shown in Fig. 8, and the
estimated material parameters are listed in (10.8):

G ¼ 25:5 GPa; K ¼ 68 GPa;


m0 ¼ 0:001 s1 ; m ¼ 0:02;
ð10:8Þ
S0 ¼ 67:55 MPa; h0 ¼ 1200 MPa; a ¼ 1; S ¼ 156:7 MPa;
B ¼ 1:09 GPa; c ¼ 61:08:
1866 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

400
Experiment
Model
300

True axial stress (MPa)


200

100

−100

−200

−300

−400
0 0.02 0.04 0.06 0.08 0.1 0.12
True axial strain
Fig. 8. Fit of the model to experimental data for AA6111-T4 sheet material. The data is from Cao et al. (2008); note that the kink
in the reversed loading data for the smallest strain amplitude is actually reported by these authors, and is probably an artifact of
the experimental difficulties associated with performing reversed loading experiments on sheet materials.

The initial configuration for our plane-strain sheet forming simulation is shown in Fig. 9(a). We
consider a 1 mm thick sheet with a length of 75 mm, which is to be to bent to a radius of 50 mm
between a pair of matched-rigid dies; accordingly, the radius of the top and bottom dies is pre-
scribed to be 50.5 mm and 49.5 mm, respectively. Due to the symmetry of the problem, we con-
sider only half of the geometry with suitable boundary conditions at the symmetry-plane. The
sheet is modeled using a mesh consisting of 1880 Abaqus-CPE4H plane-strain elements with seven
elements through the sheet thickness, and the dies are modeled using rigid surfaces. Contact be-
tween the sheet and the dies was modeled as frictionless. In the first step of the sheet forming
simulation, the top die is moved downward to bend the sheet completely around the bottom
die. Fig. 9(b) shows the geometry of the initial and fully bent geometries of the sheet. In the sec-
ond step, the top die is moved back upwards, and the sheet is allowed to spring-back. The un-
loaded sprung-back geometry is also shown in Fig. 9(b).
It is widely believed that plasticity models which do not account for kinematic hardening tend to
incorrectly predict the amount of spring-back in simulations of sheet-metal forming operations (cf.,
e.g., Zhao and Lee, 2001). Thus, our large deformation theory which not only includes both kinematic
and isotropic hardening but also accounts for large rotations, should be useful in improved modeling
of spring-back phenomena in sheet forming operations.

11. Concluding remarks

In this paper, we have formulated a large deformation constitutive theory for combined isotropic
and kinematic hardening based on the dual decomposition F ¼ Fe Fp and Fp ¼ Fpen Fpdis . We now show
that a similar kinematic-hardening theory may also be constructed without using the decomposition
Fp ¼ Fpen Fpdis . Recall the relation (6.3) for the evolution of Bpen ,

B_ pen ¼ Dpen Bpen þ Bpen Dpen ; ð11:1Þ

and the kinematical equation (2.28)


D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1867

b 5

0
x2 (mm)

−5

Undeformed
−10
Fully bent
Sprung−back
−15

−20
0 10 20 30 40
x (mm)
Fig. 9. (a) Initial configuration for plane-strain matched-die sheet bending simulations. (b) Undeformed and fully bent sheet
geometries, together with the sprung-back geometry.

 
Dp ¼ Dpen þ sym Fpen Dpdis Fp1
en : ð11:2Þ

The relation (11.2) gives


 
Dpen ¼ Dp  sym Fpen Dpdis Fen
p1
; ð11:3Þ
 p   
mdis
¼ Dp  sym Fpen Mpen Fen p1
ðusing ð9:63ÞÞ;
2sen
 p   
mdis
¼ Dp  sym Fpen Rp> p p1
en Sen R en Fen ðusing ð9:60ÞÞ;
2sen

 p   
mdis
¼ Dp  sym Vpen Sen Ven p1
;
2sen

 p   
mdis
¼ Dp  sym Vpen ðB ln Bpen ÞVp1 en ðusing ð9:57Þ and Epen  ð1=2Þ ln Bpen Þ;
2sen

1868 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

 p 
Bmdis
¼ Dp  ðln Bpen Þ;
2s
en
 
1 p
¼ Dp  cm ðln Bpen Þ ðusing ð9:42ÞÞ: ð11:4Þ
2
Using (11.4) in (11.1) gives
       
1 p 1
B_ pen ¼ Dp  cm ðln Bpen Þ Bpen þ Bpen Dp  cmp ðln Bpen Þ ;
2 2
or

B_ pen ¼ Dp Bpen þ Bpen Dp  cmp Bpen ðln Bpen Þ: ð11:5Þ


Thus, if we set
A  Bpen ;
we obtain the evolution equation

A_ ¼ Dp A þ ADp  cmp Aðln AÞ: ð11:6Þ


Therefore, instead of using the decomposition F p
¼ Fpen Fpdis
and developing a theory which gives rise to
a back-stress Sen because of a defect free-energy wðpÞ ðBpen Þ,

one may develop an alternate theory based
on an internal variable A, with a corresponding defect free-energy w  ðpÞ ðAÞ, and an evolution equation
for A obeying (11.6); indeed, this is the track taken recently by Anand et al. (2008). However, we do
note that the development of an implicit time-integration procedure for a constitutive theory based on
A is not as straightforward as the one developed here for a theory utilizing the Fp ¼ Fpen Fpdis
decomposition.
Finally, as is abundantly clear from the extensive literature on hardening models for metal plastic-
ity (cf., e.g., Chaboche, 2008), the simple theory with combined isotropic and kinematic hardening
developed in this paper is only foundational in nature, and there are numerous specialized enhance-
ments/modifications to the theory that need to be incorporated in order to match actual experimental
data for different metals; we leave such work to the future.

Acknowledgements

This work was supported by the National Science Foundation under Grant Nos. CMS-0555614 and
CMMI-062524.

Appendix A

In this Appendix we develop a semi-implicit time-integration procedure, and obtain an (approxi-


mate) algorithmic tangent for the theory formulated in this paper. The time-integration procedure and
the associated tangent are intended for use in the context of ‘‘implicit” finite element procedures. We
have implemented the procedure described below in the commercial finite element program Abaqus/
Standard (2008) by writing a user material subroutine (UMAT), and it is using this finite element pro-
gram that we have carried out the calculations whose results are reported in the main body of the
paper.

A.1. Time-integration procedure

The constitutive time-integration problem is that given


fTn ; Fpn ; Sn ; ðSen Þn ; ðFpdis Þn g; as well as Fn and Fnþ1 ;
at time t n , we need to calculate
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1869

fTnþ1 ; Fpnþ1 ; Snþ1 ; ðSen Þnþ1 ; ðFpdis Þnþ1 g


at time tnþ1 ¼ tn þ Dt.
The evolution equation for F_ p ¼ Dp Fp is integrated by means of an exponential map (Weber and
Anand, 1990)
 
Fpnþ1 ¼ exp DtDpnþ1 Fpn ; with Dpnþ1 ¼ D
^ p ðððMe Þ Þ ; Snþ1 Þ;
nþ1 eff nþ1 0 ð12:1Þ
the inverse of Fpnþ1 is then
 
Fp1
nþ1 ¼ Fp1
n exp DtDpnþ1 : ð12:2Þ
e p1
Next, using F ¼ FF and (12.1)1 , the elastic deformation gradient at the the end of the step is
given by
 
Fenþ1 ¼ Fetrial exp DtDpnþ1 ; ð12:3Þ
where
def
Fetrial ¼ Fnþ1 Fp1
n ðtrial values are evaluated with plastic flow frozenÞ ð12:4Þ
e
is a trial value of F at the end of the step. The elastic right Cauchy–Green tensor and its trial value at
the end of the step are

Cenþ1 ¼ Fe> e
nþ1 Fnþ1 ; Cetrial ¼ Fe> e
trial Ftrial : ð12:5Þ
Thus, using (12.3),
   
Cenþ1 ¼ exp DtDpnþ1 Cetrial exp DtDpnþ1 : ð12:6Þ
To proceed further we make our first approximation:

(A1) To first order in Dt, we approximate


 : 
exp DtDpnþ1 ¼ 1  DtDpnþ1 : ð12:7Þ
Hence, (12.6) becomes
:   
Cenþ1 ¼ 1  DtDpnþ1 Cetrial 1  DtDpnþ1 ;
:
¼Cetrial  DtDpnþ1 Cetrial  DtCetrial Dpnþ1 ;
 
:
¼Cetrial 1  DtCtrial
e1 p
Dnþ1 Cetrial  DtDpnþ1 : ð12:8Þ
 
Next, consider the term Ce1 p e eG 1 e
trial Dnþ1 Ctrial ; with Etrial ¼ 2 ðCtrial  1Þ denoting a Green strain correspond-
e e eG
ing to Ctrial , we have Ctrial ¼ 1 þ 2Etrial , and hence
   
e1 p :
Ctrial Dnþ1 Cetrial ¼ 1  2EeG p eG
trial Dnþ1 1 þ 2Etrial ;
:
¼ Dpnþ1 þ 2Dpnþ1 EeG eG p eG p eG
trial  2Etrial Dnþ1  4Etrial Dnþ1 Etrial : ð12:9Þ
We now make our second approximation:

(A2) We assume that jEeG


trial j 1, so that (12.9) may be approximated as

e1 p :
Ctrial Dnþ1 Cetrial ¼ Dpnþ1 : ð12:10Þ
Using (12.10), we find that (12.8) reduces to

:  
Cenþ1 ¼ Cetrial 1  2DtDpnþ1 : ð12:11Þ
 
Further, using the symmetry of Cenþ1 we note that the symmetric tensors Cetrial and 1  2DtDpnþ1 com-
mute, and hence share the same principal directions. Thus taking the logarithm of (12.11) we obtain
1870 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

: 1  
Eenþ1 ¼Eetrial þ ln 1  2DtDpnþ1 ; ð12:12Þ
2
where
1 1
Eenþ1 ¼ ln Cenþ1 and Eetrial ¼ ln Cetrial : ð12:13Þ
2 2
Next, we make our third approximation:

(A3) Recalling the series representation of the logarithm of a tensor, we assume that
 :
ln 1  2DtDpnþ1 ¼  2DtDpnþ1
for small DtDpnþ1 . Thus, under the assumptions (A1)–(A3) of small Dt and small jEeG
trial j, (12.6) yields the
important relation

:
Eenþ1 ¼Eetrial  DtDpnþ1 : ð12:14Þ
Next, let
 
def 1
C ¼ 2G I  1  1 þ K1  1 ð12:15Þ
3
denote the elasticity tensor, and let
Menþ1 ¼ C½Eenþ1 and Metrial ¼ C½Eetrial ; ð12:16Þ
respectively, denote the elastic Mandel stress at the end of the time step, as well as its trial value.
Then, operating on (12.14) by C gives
Menþ1 ¼ Metrial  DtC½Dpnþ1 : ð12:17Þ
15
Subtracting ðSen Þnþ1 from the left-hand side and
ðSen Þtrial  ðSen Þn ; ðtrial values are evaluated with plastic flow frozenÞ
from the right-hand side of (12.17), and writing
def
ðMeeff Þtrial ¼ Metrial  ðSen Þtrial ; ð12:18Þ
gives
ðMeeff Þnþ1 ¼ ðMeeff Þtrial  DtC½Dpnþ1 : ð12:19Þ
Next, from (9.62) we have
1 ððMeeff Þnþ1 Þ0 1
Dpnþ1 ¼ pffiffiffi mpnþ1 Npnþ1 ; Npnþ1 ¼ pffiffiffi ; snþ1 ¼ pffiffiffi jððMeeff Þnþ1 Þ0 j; ð12:20Þ
2 2s nþ1 2
where
pffiffiffi
mpnþ1 ¼ 2jDpnþ1 j: ð12:21Þ
Using (12.15) and (12.20) in (12.19) we obtain
pffiffiffi
ðMeeff Þnþ1 ¼ ðMeeff Þtrial  2GðDt mpnþ1 ÞNpnþ1 : ð12:22Þ
Since Npnþ1 is deviatoric, the deviatoric and spherical parts of (12.22) give

15
Subtracting ðSen Þn rather than ðSen Þnþ1 from the right-hand side of (12.17) to arrive at (12.19), is a critical step in our time-
integration procedure. It allows us to decouple the update of Fp and Fpdis into to two separate, yet otherwise implicit updates.
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1871

  pffiffiffi
ðMeeff Þnþ1 0
¼ ððMeeff Þtrial Þ0  2GðDt mpnþ1 ÞNpnþ1 ;
trðMeeff Þnþ1 ¼ trðMeeff Þtrial : ð12:23Þ
Using (12.20)2 , (12.23)1 may be arranged as
pffiffiffi 
2 s nþ1 þ GðDtmpnþ1 Þ Npnþ1 ¼ ððMeeff Þtrial Þ0 : ð12:24Þ
Next, defining

def ððMeeff Þtrial Þ0 1


Nptrial ¼ pffiffiffi ; strial def
¼ pffiffiffi jððMeeff Þtrial Þ0 j; ð12:25Þ
2s trial 2
(12.24) may be written as
 
snþ1 þ GðDtmpnþ1 Þ Npnþ1 ¼ strial Nptrial ; ð12:26Þ
which immediately gives
Npnþ1 ¼ Nptrial ;
snþ1 þ GðDtmpnþ1 Þ ¼ strial : ð12:27Þ
Thus the direction of plastic flow at the end of the step Npnþ1
is determined by the trial direction of plastic
flow Nptrial .
Next, the implicit form of the flow strength relation (9.30) is
 
snþ1 ¼ Y mpnþ1 ; Snþ1 : ð12:28Þ
We integrate the evolution equation for S in a semi-implicit fashion as
Snþ1 ¼ Sn þ hðSn ÞðDtmpnþ1 Þ: ð12:29Þ
Since the hardening function h typically does not change rapidly, we do not expect the semi-implicit
integration of the evolution equation for S to have a significant effect on the stability of our algorithm.
Using (12.29) in (12.28), we have
  
snþ1 ¼ Y mpnþ1 ; Sn þ hðSn ÞðDtmpnþ1 Þ : ð12:30Þ
Finally, using (12.30) in (12.27)2 gives the following implicit equation for mpnþ1 :
  
Gðmpnþ1 Þ ¼ s
trial  GðDt mpnþ1 Þ  Y mpnþ1 ; Sn þ hðSn ÞðDtmpnþ1 Þ ¼ 0: ð12:31Þ
Once a solution for pnþ1 m is obtained by solving (12.31), Snþ1 is easily evaluated by using (12.29). The
elastic Mandel stress Menþ1 is obtained from (12.17) using
1
Dpnþ1 ¼ pffiffiffi mpnþ1 Nptrial ; ð12:32Þ
2
as
pffiffiffi
Menþ1 ¼ Metrial  2GðDtmpnþ1 ÞNptrial : ð12:33Þ
p
The update for F is obtained from (12.1)1 using the result (12.32):
 
1
Fpnþ1 ¼ exp pffiffiffi ðDt mpnþ1 ÞNptrial Fpn : ð12:34Þ
2
Next, calculating

Fenþ1 ¼ Fnþ1 Fp1


nþ1 ; ð12:35Þ
and performing a polar decomposition of gives Fenþ1 Renþ1 . Finally, using Renþ1 , and the relations (12.33)
and (9.55) gives the update for the Cauchy stress as
e e e >
Tnþ1 ¼ J 1
nþ1 ðRnþ1 ÞðMnþ1 ÞðR nþ1 Þ with J nþ1 ¼ det Fnþ1 : ð12:36Þ
1872 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

We turn next to updating Sen and Fpdis . The evolution equation for F_ pdis ¼ Dpdis Fpdis is also integrated by
using an exponential map and (9.63) as
  c
ðFpdis Þnþ1 ¼ exp DtðDpdis Þnþ1 ðFpdis Þn with ðDpdis Þnþ1 ¼ mp ðMp Þ ; ð12:37Þ
2B nþ1 en nþ1
the inverse of ðFpdis Þnþ1 is then
 
ðFpdis Þ1 p 1 p
nþ1 ¼ ðFdis Þn exp DtðDdis Þnþ1 : ð12:38Þ
Hence, using Fpen ¼ Fp Fp1
dis ,
p
the energetic part of F at the end of the step is given by
 
ðFpen Þnþ1 ¼ ðFpen Þtrial exp DtðDpdis Þnþ1 ; ð12:39Þ
where
def
ðFpen Þtrial ¼ Fpnþ1 ðFpdis Þ1
n ð12:40Þ
is a trial value of Fpen at the end of the step.
The tensors ðFpen Þnþ1 and ðFpen Þtrial admit the polar decompositions
ðFpen Þnþ1 ¼ ðRpen Þnþ1 ðUpen Þnþ1 ; ðFpen Þtrial ¼ ðRpen Þtrial ðUpen Þtrial : ð12:41Þ
Using (12.41) in (12.39) and rearranging, we obtain
 
ðRpen Þnþ1 ðUpen Þnþ1 exp DtðDpdis Þnþ1 ¼ ðRpen Þtrial ðUpen Þtrial : ð12:42Þ
Next, from (9.57) and (9.60)

Mpen ¼ Rp> p p> p p p


en Sen R en ¼ R en ð2B ln V en ÞRen ¼ 2B ln Uen : ð12:43Þ
Thus the principal directions of ðMpen Þnþ1 , and hence from (12.37)2 those of ðDpdis Þnþ1 , are the same as
those ðUpen Þnþ1 . Hence
 
ðUpen Þnþ1 exp DtðDpdis Þnþ1 is symmetric: ð12:44Þ
Then, because of the uniqueness of the polar decomposition theorem
ðRpen Þnþ1 ¼ ðRpen Þtrial ; ð12:45Þ
 
ðUpen Þnþ1 exp DtðDpdis Þnþ1 ¼ ðUpen Þtrial : ð12:46Þ
Eq. (12.46) implies that ðUpen Þnþ1 and ðUpen Þtrial have the same principal directions. Thus taking the
logarithm on both sides and rearranging we have
lnðUpen Þnþ1 ¼ lnðUpen Þtrial  DtðDpdis Þnþ1 : ð12:47Þ
Multiplying (12.47) through by 2B and using (12.43) gives
def
ðMpen Þnþ1 ¼ ðMpen Þtrial  2BDtðDpdis Þnþ1 ; where ðMpen Þtrial ¼ 2B lnðUpen Þtrial : ð12:48Þ
Finally, using (9.60), (12.37)2 , (12.45), and pre-multiplying (12.48) by ðRpen Þtrial and post-multiplying by
ðRpen Þ>
trial gives

def
ðSen Þnþ1 ¼ ðSen Þtrial  cðDtmpnþ1 ÞðSen Þnþ1 ; where ðSen Þtrial ¼ 2B lnðVpen Þtrial : ð12:49Þ
Thus, the back-stress Sen is updated as
!
1
ðSen Þnþ1 ¼ ðSen Þtrial : ð12:50Þ
1 þ cðDtmpnþ1 Þ

Correspondingly,

ðMpen Þnþ1 ¼ ðRpen Þ>trial ðSen Þnþ1 ðRpen Þtrial ; ð12:51Þ


use of which in (12.37) provides the update ðFpdis Þnþ1 for Fpdis .
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1873

Remark. Due to the use of the exponential map in integrating the evolution equations for Fp and Fpdis ,
the constraint of plastic incompressibility is exactly maintained by our time-integration procedure.
This is easily verified by recalling the identity detðexp AÞ ¼ expðtrAÞ and recognizing the deviatoric
nature of Dp and Dpdis in (12.1) and (12.37), respectively.

A.1.1. Summary of the time-integration procedure

 Given: fTn ; Fpn ; Sn ; ðSen Þn ; ðFpdis Þn g, as well as Fn and Fnþ1 , at time tn .


 Calculate: fTnþ1 ; Fpnþ1 ; Snþ1 ; ðSen Þnþ1 ; ðFpdis Þnþ1 g at time t nþ1 ¼ t n þ Dt.

Step 1. Calculate the trial elastic deformation gradient

Fetrial ¼ Fnþ1 Fp1


n : ð12:52Þ
Step 2. Perform the polar decomposition
Fetrial ¼ Retrial Uetrial : ð12:53Þ
Step 3. Calculate the trial elastic strain
Eetrial ¼ ln Uetrial : ð12:54Þ
Step 4. Calculate the trial Mandel stress and the trial effective Mandel stress
Metrial ¼ C½Eetrial ; ð12:55Þ
ðMeeff Þtrial ¼ Metrial  ðSen Þn : ð12:56Þ
Step 5. Calculate the trial mean normal pressure, the deviatoric part of the trial effective Mandel
stress, the trial equivalent shear stress, and the trial direction of plastic flow
1
ptrial ¼  trðMeeff Þtrial ; ð12:57Þ
3
ððMeeff Þtrial Þ0 ¼ ðMeeff Þtrial þ p
trial 1; ð12:58Þ
rffiffiffi
1
strial ¼ ððMeeff Þtrial Þ0 ; ð12:59Þ
2
e
ððM Þ Þ
Nptrial ¼ peff ffiffiffi trial 0 : ð12:60Þ
2s trial
Step 6. Calculate mpnþ1 by solving
  
Gðmpnþ1 Þ ¼ s
trial  GðDt mpnþ1 Þ  Y mpnþ1 ; Sn þ hðSn ÞðDtmpnþ1 Þ ¼ 0: ð12:61Þ
p
Step 7. Update D
pffiffiffiffiffiffiffiffiffiffi p
Dpnþ1 ¼ ð1=2Þmnþ1 Nptrial : ð12:62Þ
p
Step 8. Update F
 
Fpnþ1 ¼ exp DtDpnþ1 Fpn : ð12:63Þ
Step 9. Update the Mandel stress Me
pffiffiffi
Menþ1 ¼ Metrial  2GðDtmpnþ1 ÞNptrial : ð12:64Þ
e
Step 10. Update F

Fenþ1 ¼ Fnþ1 Fp1


nþ1 : ð12:65Þ
Step 11. Perform the polar decomposition
Fenþ1 ¼ Renþ1 Uenþ1 : ð12:66Þ
1874 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

Step 12. Update the Cauchy stress

J nþ1 ¼ det ðFnþ1 Þ; ð12:67Þ


e e e >
Tnþ1 ¼ J 1
nþ1 ðR nþ1 ÞðMnþ1 ÞðR nþ1 Þ : ð12:68Þ
Step 13. Update S

Snþ1 ¼ Sn þ hðSn ÞðDt mpnþ1 Þ: ð12:69Þ


Step 14. Calculate ðFpen Þtrial

ðFpen Þtrial ¼ Fpnþ1 ðFpdis Þ1


n : ð12:70Þ
Step 15. Perform the polar decomposition

ðFpen Þtrial ¼ ðVpen Þtrial ðRpen Þtrial : ð12:71Þ


Step 16. Calculate the trial back-stress

ðSen Þtrial ¼ 2B lnðVpen Þtrial : ð12:72Þ


Step 17. Update the back-stress Sen
!
1
ðSen Þnþ1 ¼ ðSen Þtrial : ð12:73Þ
1 þ cðDtmpnþ1 Þ

Step 18. Update the plastic Mandel stress Mpen

ðMpen Þnþ1 ¼ ðRpen Þ>trial ðSen Þnþ1 ðRpen Þtrial : ð12:74Þ


Step 19. Update Dpdis
c
ðDpdis Þnþ1 ¼ mp ðMp Þ : ð12:75Þ
2B nþ1 en nþ1
Step 20. Update Fpdis
 
ðFpdis Þnþ1 ¼ exp DtðDpdis Þnþ1 ðFpdis Þn : ð12:76Þ

A.2. Jacobian matrix

In typical ‘‘implicit” finite element procedures utilizing a Newton-type iterative solution method,
one needs to compute an algorithmically consistent tangent, often called the Jacobian matrix. We ob-
tain an estimate for our Jacobian matrix below. From the outset we note that Jacobian matrices are
used only in the search for the global finite element solution, and while an approximate Jacobian
might affect the rate of convergence of the global iteration scheme, it will not impair the accuracy
of our constitutive time-integration algorithm.
Consider the Cauchy stress at the end of the increment:

Tnþ1 ¼ ðdet Fenþ1 Þ1 Renþ1 Menþ1 ðRenþ1 Þ> : ð12:77Þ


Then with D denoting a variation, (12.77) gives
      e1 
DTnþ1  ðDRenþ1 ÞðRenþ1 Þ> Tnþ1 þ Tnþ1 ðDRenþ1 ÞðRenþ1 Þ> þ tr DFenþ1 Fnþ1 Tnþ1
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} |fflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} |fflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
incremental elastic rotation incremental elastic rotation incremental elastic vol: change
 
¼ ðdet Fenþ1 Þ1 Renþ1 DMenþ1 ðRenþ1 Þ> : ð12:78Þ
    
We assume that the variations ðDRenþ1 ÞðRenþ1 Þ>
and tr DFenþ1 Fe1
nþ1 of the incremental elastic rota-
tion and volume change are reasonably well-estimated by a commercial finite element program, such
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1875

as Abaqus/Standard (2008), in which we have implemented our constitutive model by writing a user
material subroutine (UMAT).16. Thus, here we concentrate on evaluating the variation DMenþ1 in (12.78),
which is computed from

DMenþ1 ¼ C DEetrial ; ð12:79Þ
where the fourth-order tensor

def @Menþ1
C¼ ð12:80Þ
@Eetrial
is the important constitutive contribution to the global Jacobian matrix.
Recall that the time-integration procedure gives
pffiffiffi
Menþ1 ¼ Metrial  2GðDtmpnþ1 ÞNptrial : ð12:81Þ
Using (12.27)2 , (12.81) may be written as
pffiffiffi
Menþ1 ¼ Metrial þ 2ðs trial ÞNptrial :
nþ1  s ð12:82Þ
By the product rule
pffiffiffi 
@Metrial pffiffiffi @Nptrial p @s
nþ1 @ strial
C¼ þ 2 ð s
 nþ1  s
 trial Þ þ N  2  : ð12:83Þ
@Eetrial @Eetrial trial
@Eetrial @Eetrial
Since

Metrial ¼ C Eetrial ; ð12:84Þ
we have
@Metrial
¼ C: ð12:85Þ
@Eetrial
Next, since
ððMeeff Þtrial Þ0
Nptrial ¼ pffiffiffi ; ð12:86Þ
2s trial
we have
rffiffiffi 
@Nptrial 1 1 @ððMeeff Þtrial Þ0 ððMeeff Þtrial Þ0 @ s
trial
¼   e : ð12:87Þ
@Eetrial 2 s trial e
@Etrial strial
2
@Etrial
Now,
ðMeeff Þtrial ¼ Metrial  ðSen Þn

¼ C Eetrial  ðSen Þn ð12:88Þ
   
1
¼ 2G I  1  1 þ K1  1 Eetrial  ðSen Þn :
3
Hence,
  
1 e
ððMeeff Þtrial Þ0 ¼ 2G I  1  1 Etrial  ðSen Þn ; ð12:89Þ
3

16
Abaqus uses a hypoelastic constitutive equation for the stress of the type

_
T ¼ TWe _
TþTWe ¼ TWTþTW p
|fflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflffl} ¼ C½DD :
for Wp ¼0
1876 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

and therefore
 
@ððMeeff Þtrial Þ0 1
e ¼ 2G I  1  1 : ð12:90Þ
@Etrial 3
Further,
rffiffiffi
1
strial ¼ ððMeeff Þtrial Þ0 : ð12:91Þ
2
Therefore, by the chain-rule
rffiffiffi > 
@s
trial 1 @ððMeeff Þtrial Þ0 @jððMeeff Þtrial Þ0 j
¼ ð12:92Þ
@Eetrial 2 @E e e
@ððMeff Þtrial Þ0
rffiffiffi  trial > 
1 1 ððMeeff Þtrial Þ0
¼ 2G I  1  1 ; ð12:93Þ
2 3 jððMeeff Þtrial Þ0 j
pffiffiffi   
1 ððMeeff Þtrial Þ0
¼ 2G I  1  1 ð12:94Þ
3 jððMeeff Þtrial Þ0 j
pffiffiffi ððMe Þ Þ
eff trial 0
¼ 2G ð12:95Þ
jððMeeff Þtrial Þ0 j
pffiffiffi
¼ 2GNptrial : ð12:96Þ
Thus, using (12.90) and (12.96) in (12.87) we obtain
  
@Nptrial 2G 1
¼ pffiffiffi I  1  1  Nptrial  Nptrial ; ð12:97Þ
@Eetrial 2s trial 3
and substituting (12.85) and (12.97) in (12.83), we have
    
snþ1 1
C ¼Cþ  1 2G I  1  1  Nptrial  Nptrial
strial 3
pffiffiffi 
nþ1 pffiffiffi p
@s
þ Nptrial  2 e  2GNtrial : ð12:98Þ
@Etrial
Noting that Eetrial enters the equations through s
trial , we have, using (12.96)
@s
nþ1 @ s nþ1 @ s nþ1 pffiffiffi p
trial @ s
e ¼ e ¼ 2GNtrial : ð12:99Þ
@Etrial @ strial @Etrial @ s
 trial

Finally, substituting (12.99) into (12.98), using (12.15), and rearranging, we have
 
~  2G snþ1  @ snþ1 Np  Np
 
C¼C ð12:100Þ
strial @ strial trial trial

where
 
~ I  1 1  1 þ K1  1 with G
~ ¼ 2G ~ def snþ1
C ¼ G: ð12:101Þ
3 strial
Thus, it remains to determine the derivative
@s
nþ1
@s
trial
in (12.100). First, we rewrite the updates for s
nþ1 ; (12.27)2 , and Snþ1 , (12.69), as
 
snþ1
snþ1 ¼ strial  GDtf ;
Snþ1
  ð12:102Þ
snþ1
Snþ1 ¼ Sn þ DthðSn Þf ;
Snþ1
D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878 1877

respectively, where f is the flow function of (9.33). Using the definitions


 
def snþ1 def
X¼ ; and Y ¼ f s
trial g; ð12:103Þ
Snþ1
the system of equations (12.102) may be written as
X ¼ GðX; YÞ; ð12:104Þ
where
8   9
< strial  GDtf sSnþ1 >
> 
=
nþ1
G¼   : ð12:105Þ
: Sn þ DthðSn Þf sSnþ1 >
> ;
nþ1

Differentiating (12.104) with respect to Y (at the solution point) we obtain


@X @GðX; YÞ @GðX; YÞ @X
¼ þ ; ð12:106Þ
@Y @Y @X @Y
from which
 
@X @GðX; YÞ
¼ A ; ð12:107Þ
@Y @Y
where17
 1
@GðX; YÞ
A ¼ I : ð12:108Þ
@X
Straightforward calculations show
2 3
@f
GDt@@fs GDt @S 
@GðX; YÞ 6 tnþ1 tnþ1 7 @GðX; YÞ 1
¼4 5; and ¼ : ð12:109Þ
@X @f @f @Y 0
DthðSn Þ@ s DthðSn Þ@S
tnþ1 tnþ1

Thus, we obtain that


@s
nþ1
¼ A11 ; ð12:110Þ
@s
trial
substitution of which in (12.100) completes our estimate for the Jacobian matrix C.

References

Abaqus/Standard, 2008. SIMULIA, Providence, RI.


Anand, L., 1979. On H. Hencky’s approximate strain-energy function for moderate deformations. ASME Journal of Applied
Mechanics 46, 78–82.
Anand, L., 1986. Moderate deformations in extension-torsion of incompressible isotropic elastic materials. Journal of the
Mechanics and Physics of Solids 34, 293–304.
Anand, L., Gurtin, M.E., 2003. A theory of amorphous solids undergoing large deformations, with applications to polymeric
glasses. International Journal of Solids and Structures 40, 1465–1487.
Anand, L., Ames, N.M., Srivastava, V., Chester, S., 2008. A thermo-mechanically-coupled theory for large deformations of
amorphous polymers. Part 1: Formulation. International Journal of Plasticity, doi:10.1016/j.ijplas.2008.11.004.
Armstrong, P.J., Frederick, C.O., 1966. A mathematical representation of the multiaxial Bauschinger effect. Report RD/B/N731,
CEGB, Central Electricity Generating Board, Berkeley, UK.
Bauschinger, J., 1886. Über die Veränderung der Position der Elastizitätsgrenze des Eisens und Stahls durch Strecken und
Quetschen und durch Erwärmen und Abkühlen und durch oftmals wiederholte Beanspruchungen. Mitteilung aus dem
Mechanisch-technischen Laboratorium der Königlichen polytechnischen Hochschule in München 13, 1–115.
Brown, S.B., Kim, K.H., Anand, L., 1989. An internal variable constitutive model for hot working of metals. International Journal
of Plasticity 5, 95–130.

17
Here I is the second-order identity matrix.
1878 D.L. Henann, L. Anand / International Journal of Plasticity 25 (2009) 1833–1878

Cao, J., Lee, W., Cheng, H.S., Seniw, M., Wang, H.-P., Chung, K., 2008. Experimental and numerical investigation of combined
isotropic-kinematic hardening behavior of sheet metals. International Journal of Plasticity, 10.1016/j.ijplas.2008.04.007.
Casey, J., Naghdi, P.M., 1980. A remark on the use of the decomposition F ¼ Fe Fp in plasticity. Journal of Applied Mechanics 47,
672–675.
Chaboche, J.L., 2008. A review of some plasticity and viscoplasticity constitutive theories. International Journal of Plasticity 24,
1642–1693.
Clayton, J., McDowell, D., 2003. A multiscale multiplicative decomposition for elastoplasticity of polycrystals. International
Journal of Plasticity 19, 1401–1444.
Dafalias, Y., 1998. Plastic spin: necessity or redundancy? International Journal of Plasticity 14, 909–931.
Dashner, P.A., 1986. Invariance considerations in large strain elasto-plasticity. Journal of Applied Mechanics 53, 55–60.
Dettmer, W., Reese, S., 2004. On the theoretical and numerical modelling of Armstrong–Frederick kinematic hardening in the
finite strain regime. Computer Methods in Applied Mechanics and Engineering 193, 87–116.
Frederick, C.O., Armstrong, P.J., 2007. A mathematical representation of the multiaxial Bauschinger effect. Materials at High
Temperatures 24, 1–26.
Green, A.E., Naghdi, P.M., 1971. Some remarks on elastic–plastic deformation at finite strain. International Journal of
Engineering Science 9, 1219–1229.
Gurtin, M.E., Anand, L., 2005. The decomposition F ¼ Fe Fp , material symmetry, and plastic irrotationality for solids that are
isotropic-viscoplastic or amorphous. International Journal of Plasticity 21, 1686–1719.
Håkansson, P., Wallin, M., Ristinmaa, M., 2005. Comparison of isotropic hardening and kinematic hardening in thermoplasticity.
International Journal of Plasticity 21, 1435–1460.
Kröner, E., 1960. Allgemeine kontinuumstheorie der versetzungen und eigenspannungen. Archive for Rational Mechanics and
Analysis 4, 273–334.
Lemaitre, J., Chaboche, J.L., 1990. Mechanics of Solid Materials. Cambridge University Press, Cambridge.
Lion, A., 2000. Constitutive modelling in finite thermoviscoplasticity: a physical approach based on nonlinear rheological
models. International Journal of Plasticity 16, 469–494.
Lush, A.M., Weber, G., Anand, L., 1989. An implicit time-integration procedure for a set of internal variable constitutive
equations for isotropic elasto-viscoplasticity. International Journal of Plasticity 5, 521–549.
Menzel, A., Ekh, M., Runesson, K., Steinmann, P., 2005. A framework for multiplicative elastoplasticity with kinematic hardening
coupled to anisotropic damage. International Journal of Plasticity 21, 397–434.
Prager, W., 1949. Recent developments in the mathematical theory of plasticity. Journal of Applied Physics 20, 235–241.
Ristinmaa, M., Wallin, M., Ottosen, N.S., 2007. Thermodynamic format and heat generation of isotropic hardening plasticity.
Acta Mechanica 194, 103–121.
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. Journal of the Mechanics and Physics of Solids 48, 581–607.
Vladimirov, I.N., Pietryga, M.P., Reese, S., 2008. On the modelling of non-linear kinematic hardening at finite strains with
application to springback – comparison of time integration algorithms. International Journal for Numerical Methods in
Engineering 75, 1–28.
Wallin, M., Ristinmaa, M., 2005. Deformation gradient based kinematic hardening model. International Journal of Plasticity 21,
2025–2050.
Weber, G., Anand, L., 1990. Finite deformation constitutive equations and a time integration procedure for isotropic,
hyperelastic–viscoplastic solids. Computer Methods in Applied Mechanics and Engineering 79, 173–202.
Weber, G.G., Lush, A.M., Zavaliangos, A., Anand, L., 1990. An objective time-integration procedure for isotropic rate-independent
and rate-dependent elastic–plastic constitutive equations. International Journal of Plasticity 6, 701–744.
Zhao, K.M., Lee, J.K., 2001. Material properties of aluminum alloy for accurate draw-bend simulation. Journal of Engineering
Materials and Technology 123, 287–292.

Das könnte Ihnen auch gefallen