You are on page 1of 317

Numerical Analysis and

Simulation of Plasticity
J.C. Simo
Division of Applied Mechanics
Department of Mechanical Engineering
Stanford University
Stanford, CA 94305, USA

HANDBOOK OF NUMERICAL ANALYSIS, VOL. VI


Numerical Methods for Solids (Part 3)
Numerical Methods for Fluids (Part 1)
Edited by P.G. Ciarlet and J.L. Lions
1998 Elsevier Science B.V. All rights reserved
183

Contents
PREFACE

187

CHAPTER I. The Classical Models

191

1. Notation and brief summary of some standard results


2. Classical rate-independent plasticity: Local evolution equations
3. The rate form of classical plasticity: Elastoplastic tangent moduli
4. Some specific models of classical plasticity
5. General rate-independent multisurface plasticity
6. Dissipation: Interpretation of the model of associative plasticity
7. Rate-dependent plasticity: The viscoplastic regularization
8. Example: Viscoplastic regularization of J2-flow theory
9. The weak formulation of dynamic plasticity
10. Contractivity, uniqueness, and dissipativity of the elastoplastic flow
CHAPTER II. Integration Algorithms
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.

229

Local integration of rate-independent plasticity: An overview


Motivation: The perfectly plastic model of J2-flow theory
Backward difference and implicit Runge-Kutta methods: Basic results
Generalized backward difference return mapping algorithms
Generalized implicit Runge-Kutta return mapping algorithms
Algorithms for the computation of the closest-point projection
The consistent algorithmic elastoplastic moduli
Examples: Closed-form return mapping algorithms
Practical accuracy assessment: Iso-error maps
Accuracy analysis of return mapping algorithms
Extension of return mapping algorithms to viscoplasticity
Return mapping algorithms for general models of viscoplasticity
The algorithmic initial boundary value problem
Nonlinear stability analysis: Uniqueness and dissipativity
Spatial finite element discretization: An illustration
Application: A class of mixed methods for incompressibility
Illustrative numerical simulations

CHAPTER III. Nonlinear Continuum Mechanics


28.
29.
30.
31.
32.
33.

191
196
200
204
211
213
216
218
221
224

230
234
240
244
247
249
253
254
261
264
269
271
273
277
283
291
296
303

Basic kinematic results in nonlinear continuum mechanics


Stress tensors and alternative forms of the equations of motion
Objective transformations and frame invariance
Elastic constitutive equations and isotropy group
Volumetric/deviatoric uncoupled finite elasticity
Isotropic elasticity formulated in principal stretches
185

304
311
313
317
322
327

186

J.C. Simo
34.
35.
36.
37.
38.
39.
40.
41.

Multiplicative plasticity at finite strains: Basic concepts


Elastic response and free energy for multiplicative plasticity
Plastic flow evolution equations for multiplicative plasticity
Volumetric/deviatoric uncoupled finite plasticity: J2 -flow theory
Rate form and variational inequality for multiplicative plasticity
Variational formulation: Weak form of the momentum equations
The total and incremental weak forms of momentum balance
Initial boundary value problem: Dissipativity and a priori estimate

CHAPTER IV. The Discrete Initial Boundary Value Problem


42.
43.
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.

The Galerkin projection: The spatially discrete problem


The linearized (semidiscrete) initial value problem
Matrix form of the semidiscrete initial value problem
Mixed finite element discretization: An illustration
Exponential return mapping algorithms for multiplicative plasticity
Exponential return mappings for isotropic elastic response
Implementation of exponential return mapping algorithms
Linearization: The exact algorithmic tangent moduli
A generalization of J 2-flow theory to finite strains
Two-stage projected exponential return mapping algorithms
Closed-form of the two-stage projected IRK for elastic isotropy
Global time-stepping algorithms for dynamic plasticity
Remarks on the implementation of return mapping algorithms
Representative numerical simulations

CHAPTER V. The Coupled Thermomechanical Problem


56.
57.
58.
59.
60.
61.
62.
63.

Integral, local and weak forms of the general conservation laws


Constitutive equations for multiplicative thermoplasticity
Formal a priori stability estimate and conservation laws
Time integration of the coupled problem: General considerations
Monolithic and staggered schemes: Product formula algorithms
Model problem: The coupled system of linearized thermoelasticity
Generalization: A staggered scheme for nonlinear thermoplasticity
Representative numerical simulations

REFERENCES

330
335
336
341
346
351
355
357
361
363
365
369
373
383
387
390
394
398
402
408
411
414
416
433
435
439
447
451
456
460
469
475
489

Preface
Efforts aimed at understanding the behavior of metals under loading beyond the
elastic regime provided the starting point, at the turn of this century, for the development of a body of constitutive theories for deformable solids commonly referred
to as classical plasticity. By the mid-fifties, the foundations of the subject were wellestablished within the context of the infinitesimal theory as summarized, for instance,
in the classical book of HILL [1950] and the review article of KOITER [1960]. Because
of its inherent nonlinear character, only a few elementary problems in classical plasticity are tractable via analytical methods. The application of the theory to problems of
engineering interest, metal forming in particular, required drastic simplifying hypotheses involving, for instance, the assumption of rigid plasticity. A revolution in the field
took place with the advent of the computer and the development of the finite element
method. This made possible the numerical solution of complex problems by utilizing
the full theory of plasticity without introducing unduly restrictive assumptions.
The objectives of this article are to provide a comprehensive overview of both the
continuum mechanics and the numerical analysis relevant to the large scale simulation of problems involving plastic deformations. Large scale simulations of this type,
ranging from crash analysis of vehicles to metal forming and forging processes, are
becoming routine and constitute a significant portion of current scientific computing
in solid mechanics. The goals of this exposition are to present in a unified setting
the classical models of plasticity, emphasizing the unifying mathematical structure,
and address in detail the algorithmic issues involved in the solution of the resulting
(highly nonlinear) initial boundary value problem. An overview of the topics covered
in this article is given below.
Chapter I provides a concise formulation of rate-independent plasticity and (ratedependent) viscoplasticity, with attention restricted to the case of linearized kinematics. Most of the results described there are fairly classical. Emphasis is placed on a
formulation of the initial boundary value problem suitable for the subsequent algorithmic treatment. This includes, in particular, the formulation of the loading/unloading
conditions of rate-independent plasticity as Kuhn-Tucker optimality conditions, the
interpretation of viscoplasticity as a (penalty) regularization of the rate-independent
theory as well as the formulation of plasticity as a variational inequality. The chapter
concludes with an overview of the properties of the initial boundary value problem
that play a central role in the analysis of the discretized problem.
Chapter II addresses the central topic of interest in this article, namely, the design and analysis of algorithms for plasticity and viscoplasticity. The key observation
187

188

J.C. Simo

exploited in the design of numerical schemes is the interpretation of the constitutive equations of rate-independent plasticity as a differential algebraic system. This
point of view allows the generalization of classical schemes, known as return mapping algorithms and typically restricted to first-order accuracy, to methods possessing
higher-order accuracy. A number of results presented in this chapter provide new insights into the algorithmic treatment of plasticity. These include the formulation of
higher-order return mapping algorithms, a precise accuracy analysis that exploits the
geometric structure of classical plasticity and a detailed analysis of numerical stability
that exploits the contractive structure of the continuum problem.
Chapter III undertakes the generalization of the classical models of plasticity described in the first chapter of this article to the finite deformation regime. To make the
presentation self-contained, a comprehensive review is given of those aspects of nonlinear continuum mechanics relevant to the problem at hand. Motivated by micromechanical considerations, touched upon only briefly in the course of the exposition,
the extension of classical plasticity described in this chapter relies on a multiplicative
decomposition of the deformation gradient. The rate form of the resulting theory is
shown to be in a one-to-one correspondence with certain formulations of finite strain
plasticity based on a rate formulation of the elastic constitutive equations, which have
been widely used in past algorithmic treatments of the subject. The chapter concludes
with a detailed presentation of the total and incremental initial boundary value problems.
Chapter IV describes a generalization of the algorithms presented in Chapter II
to the full finite deformation problem. For the isotropic theory, which includes the
important case of J2-flow theory as a particular case, it is shown that the classical
and generalized return mapping algorithms carry over to the finite deformation regime
with essentially no modification. The exposition emphasizes the design of algorithms
aimed at preserving key qualitative features inherent to the continuum models, such as
frame invariance and certain conservation laws. A number of representative numerical
simulations are presented to illustrate the effectiveness of these techniques in a wide
class of problems.
Chapter V is concerned both with the extension of the continuum models of classical plasticity to the full thermomechanical regime and the numerical approximation
of the resulting coupled initial boundary value problem. The different time scales
involved in the coupled thermomechanical problem have motivated the widespread
use of staggered schemes involving the sequential solution of a mechanical phase
followed by a thermal phase. Approximations of this type, however, typically lead to
schemes that are, at best, only conditionally stable. The numerical analysis relies on
an interpretation of staggered algorithms as fractional step methods constructed via an
operator split of the problem of evolution. This interpretation is shown to lead to the
design of alternative staggered schemes which retain the computational convenience
of conventional methods while achieving unconditional stability. The performance of
these schemes is again illustrated in a number of representative simulations.
I am grateful to a number of colleagues, former students and visitors at Stanford
University for their collaboration on the subject of this article over the past years. In
particular, Dave Bamnett and Thomas J.R. Hughes at Stanford, Robert L. Taylor and

Preface

189

Jerrold E. Marsden at U.C. Berkeley and Peter Wriggers at Darmstad for many fruitful
discussions. I am particularly indebted to Francisco Armero and Christian Miehe for
our joint work on the topics covered in Chapter V; without their collaboration, many
of the results described there would not have been possible. I am also indebted to Tod
Laursen for generously contributing to the simulations involving contact and friction
described in Chapter IV, to Steve Rifai for his help with a number of figures drawn by
computer, as well as to Sanjay Govindjee for useful remarks on the stability analysis
of return mapping algorithms. Finally, my sincere thanks to Carlos Agelet, Oscar
Gonzalez, Alecia Chen and Charles Taylor for their careful reading of the manuscript.
It is a pleasure to acknowledge the support of the Air Force Office of Scientific Research and the Lawrence Livermore National Laboratory while writing this manuscript,
as well as the support of the National Science Foundation.

CHAPTER I

The Classical Models of


Rate-Independent Plasticity and
Viscoplasticity
This chapter summarizes the equations governing classical rate-independent plasticity, and its viscoplastic regularization. The presentation is restricted to an outline
of the mathematical structure of the field equations relevant to the numerical solution
of initial boundary value problems and the analysis of numerical algorithms.
To set the stage for the basic theory described below and the numerical analysis
results presented in Chapter II, a summary of the notation adopted along with some
basic results on continuum mechanics is first given. Attention is restricted here to
linearized kinematics and the purely mechanical theory. Further details can be found
in standard textbooks, e.g., SOKOLNIKOFF [1956] or GURTIN [1972]. The basic structure of rate-independent plasticity is then outlined within the classical framework of
response functions formulated in stress space, as in HILL [1950] or KOITER [1960].
Special attention is given to the proper formulation of the loading/unloading conditions in the so-called Kuhn-Tucker form. These are the standard complementarity
conditions for constrained problems subject to unilateral constraints. This form of
the loading/unloading conditions is, in fact, classical and has been used by several
authors, e.g., KOITER [1960], MANDEL [1965] and MAIER [1970]. Because the algorithmic elastoplastic problem is typically cast into a strain driven format, throughout
this discussion the strain tensor is viewed as the primary (driving) variable. This is
the standard point of view adopted in the numerical analysis literature, starting from
the pioneering work of WILKINS [1964]. Alternative stress space frameworks have
been explored by several authors; e.g., JOHNSON [1977] and SIMO, KENNEDY and
TAYLOR [1988].
1. Notation and brief summary of some standard results
The scheme of notation adopted in this work is fairly standard in modern textbooks on
continuum mechanics, such as CHADWICK [1976], GURTIN [1981] or CIARLET [1988],
and follows widely accepted conventions as well as standard abuses in notation. A very
191

192

J.C. Simo

CHAPTER I

brief summary of the conventions used herein is given below.


Throughout this work italic letters are used to designate scalars and real-valued
functions and lowercase bold face letters are used to designate vectors and vector fields
in Rn , while both lower- and uppercase bold letters are used to designate tensors and
tensor fields of any order. Second-order tensors are viewed as linear transformations
on IRn, n
2, elements of the vector space L(R', Rn). Fourth-order tensors are
viewed as linear transformations on the Cartesian product RI x RI. We designate
by {eij, 1 < i < n, the standard basis in R n and assume throughout the standard
summation convention on repeated indices. Accordingly,
a = aiei,

A = Aijei

ej,

D = Dijklei 0 ej ek ei,

(1.1)

where the symbol 0 denotes the standard tensor product of two vectors, i.e., the linear
transformation defined by the relation (a b)d = (b d)a, for any a, b and d in ]Rn.
Here a b = aibi is the standard Euclidean dot product with associated norm denoted
by lal = a a. We use the symbols
1 = ijei ej

and

I = [ik6jl + il 8jk]ei (ej 0 ek 0 el,

(1.2)

to designate the second-order and fourth-order symmetric tensors, respectively.


The vector space of second-order tensors, with dimension n x n, is equipped with
an inner product and an induced norm, also denoted by a dot and two vertical bars,
respectively, defined by
A B = tr [AT B] = AijBij

and

IAI =

A,

(1.3)

for any A, B E L(IRn, Rn). Finally, we denote by S c L(In, R n ) the subspace of


symmetric second-order tensors with dimension n(n + 1)/2.
Following a standard convention, GL+(n) C L(IR n , Rn) will stand for the subset
of positive definite, n x n real matrices. This set is a compact Lie group under
multiplication. Finally, the subset of symmetric, positive definite, n x n real matrices
will be denoted by S+ C GL+ (n).
(A) Local form of the momentum equations. Let 1 < ndim < 3 be the space dimension, 1[ = [0, T] C R+ the time interval of interest, and Q c R" d?m the reference
placement of a continuum body. We assume that 2 is an open and bounded set with
smooth boundary aQ and closure 2 = Q U Q2.
We denote by po: 2 -- R+ the reference density and by u: x I -t IRndi the
displacementfield at time t E Il of material points x E 2 in the reference placement,
with displacement gradient Vu(x,t) = uije 0 ej and velocity field v(x,t) =
it(x,t). In what follows we use the conventions (),j = (.)/xzj and (-) = 0(-)/at
for spatial and time derivatives, respectively. The infinitesimal strain field in the body
at time t is the symmetrized displacement gradient
E[u] = sym[Vu] =

[Vu + (VUt,)T],

(1.4)

The classical models

SECTION

with component expression E[u] =


In addition, we denote by
or:

(ui,j + uj,i)ei

with ar(x, t)=

ijei

193

ej

relative to the standard basis.

ej

(1.5)

the (infinitesimal) stress tensor field on the body at time t. We assume that the boundary
a2 of the body is divided into two parts Pr and r,, such that
FUr=

a2,

and

Fr,P=
0,

(1.6)

where the displacement field and the traction vector are respectively prescribed as
u = i on r

and

on = t on r, x 1I.

(1.7)

Here t(., t) and t(-, t) are given functions and n: a2 - S 2 is the outward unit normal
field on a0. Under these conditions, the local form of balance of linear momentum
for a prescribed body force field f: 2 x I ]R ndim takes the form
9

po = div[tr] + f

(1.8a)

div[ in 12xIE,

where div[or] = aij,jei denotes the spatial divergence of the stress field. Recall that
balance of angular momentum results in the symmetry condition or = rT is assumed
at the outset. This system of partial differential equations is supplemented by the
boundary conditions specified by (1.7), subject to the restrictions (1.6), together with
the initial data
u(*,t)jto = uo(-)

and

v(.,t)lt=o = v 0 (.)

in 2,

(1.8b)

where uo(.) and vo(-) are prescribed functions in Q. Equations (1.8a,b) together with
the boundary conditions (1.7) yield an initial boundary value problem for the displacement field u(x, t) when the stress field o-(x, t) is related to the displacement field
u(x, t) through a constitutive equation. This problem is linear only if the constitutive
relations are linear.
EXAMPLE 1.1. The simplest model for a constitutive equation is provided by a hyperelastic material, in which the stress response is characterized in terms of a stored
energy function W: 2 x S
R such that
It
(x, t) = 0aW(x, E(x, t)),

i.e.,

aij = aW/laij.

(1.9)

The second derivative of the stored energy function defines the linearized elasticity
tensor
C(, t)

= 82W(x,E(,

t)),

i.e.,

Cijkl =

(1.10)
(2W.aEij1aE,

194

J.C. Simo

CHAPTER I

which possesses the symmetries Cijkl = Cklij = Cijlk = Cjilk. For the infinitesimal
theory, C is positive definite restricted to S in the sense that
C4 C = (ijCijklkl i> aC 2,

(l.l)

for some a > 0 (depending on x e Q), and any C E S. This condition, also known as
point-wise stability (see, e.g., MARSDEN and HUGHES [19831, Chapter 3) is equivalent
to demanding that the stored energy function W be a convex function of the infinitesimal strain tensor e e . If W does not depend on x E 2 [that is, aW = 0] the
material is said to be homogeneous. Finally, if W is rotationally invariant the material
is said to be isotropic. If, in addition, C is constant the material is said to be linearly
elastic and one has the classical result
C = A1
X1

+ 2I

= nl 1+

2[I

- 1

1],

(1.12)

where A,p are the Lame constants and ;r = A + 2p/ is the bulk modulus.
REMARK 1.1. Convexity is an unacceptable restriction for the full (geometrically)
nonlinear theory, see, e.g., CIARLET [1988]. A restriction on C which is significantly
weaker than positive definiteness is the strong ellipticity condition
(a

b) C(a X b)

ola121bl2 ,

(1.13)

for some a > 0 (depending on x E Q) and any a, b E Rnd m. It can be easily shown
that (1.11) implies (1.13) but not conversely; see MARSDEN and HUGHES ([1983],
Chapter 3). The strong ellipticity condition, on the other hand, is equivalent to the
requirement that wave speeds in the material be real, i.e., the so-called Hadamard
condition on the acoustic tensor.
(B) Weak form of the momentum equations. The existence theory and the numerical
implementations of linear elasticity are most easily formulated in terms of the weak
form of the local equations (i.e., the virtual work principle). For example, the numerical
solution of the preceding initial boundary value problem by finite element methods
relies on the weak formulation, see CIARLET [1988].
The weak form of the momentum equations is constructed as follows. Let St be
the solution space for the displacement field at (frozen) time t E 1idefined as
St = {u(.,t)

Wl'P(2)ndim: u(.,

t) =

Ut(,

t) on ru},

(1.14)

where the appropriate exponent p 2 in the Sobolev space W P() nd im is dictated by


technical considerations. For elasticity, the choice of p is dictated by growth conditions
on the stored energy function; see, e.g., CIARLET [1988] or MARSDEN and HUGHES
([1983], Chapter 6). For instance, p = 2 is the appropriate choice if the stored energy

195

The classical models

SECTION 2

is quadratic in the strains. We shall let V denote the space of admissible displacement
variations (test functions) associated with St and defined in the standard fashion as
V =

7 E Wi'P(2)dd

7 = 0 on rF}.

(1.15)

If, with a slight abuse in notation, the symbol (, .) is used to denote the standard
L 2 (2)-inner product of either functions, vectors, or tensors depending on the specific
context, the weak form of the momentum equations (1.8a) takes the form
(pov, ri) = (pOU,

(pov, ) +

o)
E[])= (f,

V)rV,

(1.16a)

subject to the initial conditions


(u(.,0),r) = (uo, )

and

(v(.,O),i ) = (vo, 77),

Vqre V.

(1.16b)

Formally, one arrives at this result by taking the L 2-inner product of the local form
(1.8a) with any admissible variation and using integration by parts. Result (1.16a)
holds regardless of any constitutive assumption.
By specialization of the weak form via a specific choice of test function, one obtains
a basic result known as the mechanical work identity. Consider the case in which the
essential boundary conditions are time-independent, i.e., au/at = 0 on Fr. Under
this assumption, for fixed but otherwise arbitrary time t GEI, the velocity field v(x, t)
is an admissible test function, i.e., v(-, t) E V. Setting (.) = v(-, t) in (1.16a) and
making use of the elementary identity

(poi, v) =

(pov, v)

yields the following fundamental result

dK(v) + Pint(U, v) = Pext(v) for all t E ,

(1.17)

where K(v) is the kinetic energy of the system, Pit(o, v) is the stress power and
Pxt(v) is the external power of the applied loading, respectively defined by
K(v) = (pov, v)
Pint(a,v) = (, E[v])

kinetic energy,
stress power,

Pext()

external power.

= (f, v) + (t, v)r

(1.18)

Both the weak form (1.16a) of the momentum equations and the preceding result are
independent of the specific form adopted for the constitutive equations.

196

J.C. Simo

CHAPTER I

2. Classical rate-independent plasticity: Local evolution equations


With the preceding notation in hand, we summarize below the governing equations
of classical rate-independent plasticity within the context of the three-dimensional
infinitesimal theory. Throughout our discussion, if no explicit indication of the arguments in a field is made, it is understood that the fields u, E, ' and so on, are
evaluated at a point x E 2 and at current time t G , where the time interval 1 of
interest is often taken as the entire 1R+ for convenience. In addition, we shall denote
by r - E, (x) = E(x, ), where T E (-oo, t], the strain history at a point x E Q up
to current time t E R+. Typically, one assumes that this mapping is C . Frequently,
we shall omit explicit indication of the spatial argument and write Tr - (r) or simply
use the symbol e,, for r C (-o, t].
From a phenomenological point of view we regard plastic flow as an irreversible
process in a material body, typically a metal, characterized in terms of the history
of the strain tensor and two additional variables: the plastic strain eP, and a set of
nint > 0 internal variables generically denoted by e and often referred to as hardening
parameters. Accordingly, in a strain driven formulation, plastic flow at each point
x E Q up to current time t C ]R+ is described in terms of the histories
T C (-o,

t] -

{E(x,

T),

P(X,

r),

T)}.

(2.1)

In this context the stress tensor is a dependent function of the variables (e, EP) through
the elastic stress-strain relations, as discussed below. This leads to a strain space formulation of plasticity. Although we regard (2.1) as the independent "driving" variables,
the response functions in classical plasticity, i.e., the yield condition and the flow rule,
are formulated in stress space in terms of the variables
r E (-o, t] H~{a(x,

),q(x, )},

(2.2)

where the stress tensor ot and the internal variables q are functions of (EP ) . In
the following discussion of classical plasticity we shall adopt this point of view and
formulate the response functions in stress space. Nevertheless, implicitly we always
regard (2.1) as the independent variables.
The basic assumptions underlying the formulation of phenomenological models of
classical plasticity, leading to a set of local evolution equations for the plastic strain
eP and the internal variables , can be summarized as follows.
(A) Additive decomposition of the strain tensor. One assumes that the strain tensor
E(x, t) =-E[u(x, t)] can be decomposed into an elastic and plastic part, denoted by
e and eP, respectively, according to the relation
E= Ee +

EP

i.e.,

Eij = E

+ Cu
2.j.

I Within a thermodynamic framework, q are viewed as the "fluxes" conjugate to the affinities .

(2.3)

197

The classical models

SECTION 2

Since is regarded as an independent variable, and the evolution of EP is defined


through the flow rule (as discussed below), Eq. (2.3) should be viewed as a definition
of the elastic strain tensor

Ce

P.

(B) (Elastic) stress response. The stress tensor and the stress-like hardening variables
(o, q) are related to the elastic strain and the strain-like hardening variables (e, E)
by means of a free energy function T : 2 x S x Rflnit -- R according to the potential
relations
o = kae

(x,

e, )

and

q= -oT1(x, e,

).

(2.4a)

The free energy function


is independent of x if the material is homogeneous.
Furthermore, in most applications, l is uncoupled and has the form
!(E,

) = W(Ee)

+ 7-/(S),

(2.4b)

where W(.) is the elastic stored energy function and () is the potential function
for the (stress-like) hardening variables. For linearized elasticity, both W(-) and 7-(.)
are quadratic forms in the elastic strains e and the strain-like hardening variables ,
respectively, i.e., W = Ie
C e e and 7-t = 1 He where C is the tensor of elastic
moduli and H is the matrix of plastic moduli both assumed constant. Equations (2.4a)
and (2.3) then imply
r

=C[

q = -HE,

P],

i.e.,
i.e.,

oij

Cijkl(Ekl qi = Hijj.

k),

(2.5)

We observe that Eqs. (2.4a) and the decomposition (2.3) are local. Therefore, although
the total strain is the (symmetric) gradient of the displacement field, the elastic strain
is not in general the gradient of an "elastic" displacement field. Note further that EP
and, consequently, Ee are assumed to be symmetric at the outset, i.e., EP E S . Thus,
the notion of plastic spin plays no role in classical plasticity.
The essential feature that characterizes plastic flow is the notion of irreversibility.
This basic property is built into the formulation as follows.
(C) Elastic domain and yield condition. We define a function f:S x

called the yield criterion and constrain the admissible states (, q) E


stress space to lie in the set E defined as
E = {(o, q) E $ x

]lin':

f(a, q) < O}.

IR
x RTin in

Rnint

(2.6)

One refers to the interior of E, denoted by int(E) and given by


int(E) = {(a, q) E S x R'nt: f(a, q) < 0(,

(2.7)

198

J.C. Simo

CHAPTER

as the elastic domain; whereas the boundary of E, denoted by E and defined as


aE = {(a, q) CE x

l1'Ti:

f(or, q) _ 0},

(2.8)

is called the yield surface. For fixed internal variables q the admissible stress fields
lie in a set denoted by E. Note that states (, q) outside E are nonadmissible, and
are ruled out in classical plasticity.
(D) Flow rule and hardening law Loading/unloadingconditions. The key notion of
irreversibility of plastic flow is introduced by means of the following (nonsmooth)
equations of evolution for (P, )
P=

yr(,q)

(2.9)

= yh(a,q),

and

which are referred to as the flow rule and hardening law, respectively. Here
IR n t are prescribed functions which define
R
S and h: S x lTn'
r:
x
the direction of plastic flow and the type of hardening. The parameter y > 0, referred
to as the consistency parameter, is assumed to obey the following Kuhn-Tucker complementarity conditions
-Rni'

Y > 0,

f(a, q) < 0,

and

In addition to conditions (2.10), y


yf(

, q) = 0

yf(a, q) =0.

(2.10)

O0satisfies the consistency requirement

if f(a, q) =0.

(2.11)

In the classical literature on plasticity, conditions (2.10) and (2.11) go by the name
of loading/unloadingand consistency conditions, respectively. These conditions play
a fundamental role in convex programming (see, e.g., KARMANOV [1977] or CIARLET
[1989], Chapter 9) and, in the present context, replicate our intuitive notion of plastic
loading and elastic unloading.
The Kuhn-Tucker complementarity conditions provide a compact statement of the
different regimes possible in a model of classical plasticity. To illustrate the alternative
situations that can arise in a complex loading program, consider first the case in which
(a, q)
int(E) so that, according to (2.7) f (a, q) < 0. Therefore, from condition
(2.10)3 we conclude that
yf = 0 and f <0

y =0.

From (2.9) it then follows that P = 0 and


rate form of (2.5) leads to
dr = C = Cue.

(2.12a)
= O.
0 Thus, (2.3) yields E = e, and the

(2.12b)

In view of (2.12b) we refer to this type of response as instantaneously elastic. Now


suppose that (a, q) G aE which, in view of (2.8) implies that f(a, q) = 0. Condition

The classical models

SECTION 2

199

dE,s (f = 0)
[NON-ADMISSIBLE]

f>0

FIG. 2.1. Illustration of the elastic domain in stress space.

(2.10)3 is then automatically satisfied even if -y > 0. Whether y is actually positive


or zero can be concluded from condition (2.11). Two situations can arise.
(i) First, if f(a, q) < 0, from condition (2.11) we conclude that

yf = 0 and

<0

==

y =0.

(2.13)

Thus, again from (2.9) it follows that P = 0 and := 0. Since (2.12b) holds and
(a, q) is on aE, one refers to this type of response as unloadingfrom a plastic state.
(ii) Second, if f(a, q) = 0 condition (2.11) is automatically satisfied. If y > 0
then &P 0 and i y 0, a situation which is referred to as plastic loading. The case
y = 0 (and f = 0) is termed neutral loading.
To summarize the preceding discussion we have the following possible situations
and corresponding definitions for any (a,, q) E E.
f <0

(cr, q)

int(E)

==,

y=0

elastic response,

0=and 7y > 0

plastic loading.

We observe that the possibility f > 0 for f = 0 has been excluded from the analysis
above. Intuitively, it is clear that if f(a, q) > 0 for some (a, q) E aE at some time
t E R+, then the condition f < 0 would be violated at a neighboring subsequent time
(see Fig. 2.2). A formal argument is given in the following.
THEOREM 2.1. Let T - {a-,, q} for T E (-o, t] be the history in stress space up
to current time t C R+. Set

f(t) = f(aot, qt),

(2.15)

and assume that (at, qt) is on IE so that f (t) = O.Then the time derivative of f(t)
cannot be positive, i.e.,
If f(t) = 0 at t e R+, then f(t) < O.

(2.16)

200

J.C. Simo

CHAPTER I

f(r)

f(t) > 0

77Y

f(t) =

---

f(t) <o

FIG. 2.2. Illustration of the consistency condition f(t) = 0 => f (t) < 0.

PROOF. Assuming that f(.) is smooth the result follows from elementary considerations. In fact, for C > t by Taylor's formula
f()

- +t]f(t) + 01

= f(t)

t 2,

(2.17)

0 at -- t. Now since we must have


0< and f(t) = 0, dividing (2.17) by [C - t] leads to the inequality

where, by definition,

f()

OIl

/(t) + 9-

- t12/[( - t] -*

(2.18)

t] < 0.

K - t]

The result then follows by taking the limit of (2.18) as [ - t] - 0.


The consistency condition (2.11) is used to relate the plastic multiplier y > 0 to
the current strain rate in the rate form of the constitutive equations.
3. The rate form of classical plasticity: Elastoplastic tangent moduli
The constitutive theory outlined above can be cast into an evolution equation that
relates the stress rate to the total strain rate via a constitutive matrix of elastoplastic
tangent moduli. The derivation of this rate equation makes crucial use of the consistency condition (2.11) as follows. The time derivative of f at (, q) C aE is
first evaluated by making use of the chain rule together with the rate form of the
stress-strain relations (2.5), and the flow rule and hardening law in (2.9), leading to
f =af

C +aqf

= af- C[ - &P] +
= a,f - C-

[,f'.Cr

aqf

ii

aqf. Hh] < 0.

(3.1)

The classical models

SECTION 3

FIG. 3.1. Plastic loading at (a, q) aE takes place if the angle 0 defined by cos
Ca, f]j/ 2[ C0] 1/2 is such that 0 < 7.

201

C/[raaf

*af

To carry the analysis a step further, an additional assumption on the structure of


the flow rule and hardening law in (2.9) is needed to preclude "excessive softening
response". Explicitly, the following hypothesis is made.
ASSUMPTION 3.1. The flow rule, hardening law, and yield condition in stress space
are such that the following inequality holds
G = [af

. Cr + aqf. Hh] > 0,

for all admissible states (, q)

(3.2)

E.

It will be shown below that this assumption always holds for associative plasticity
either in the hardening or perfectly plastic regimes. An interpretation of Assumption 3.1 will be given in the context of a one-dimensional problem in a subsequent
section. With such an assumption in hand, it follows from (2.11) and (3.1) that
f = 0 and

= 0

where ]x[=[x-xi]
+
conclude that

-y = G-l]af .CE[,

(3.3)

denotes the ramp function. In view of (3.2) and (3.3) we also

forf= 0 andf= O, y

>afCfE>O.

(3.4)

This relation provides a useful geometric interpretation of the plastic loading and
neutral loading conditions in (2.14) which are illustrated in Fig. 3.1. Plastic loading
or neutral loading takes place at a point (a, q) E aE if the angle in the inner product
defined by the elasticity tensor C between the normal a,f(a, q) to E at (, q) and
the strain rate E is less than or equal to 90 .
The computation of the tangent elastoplastic moduli is completed by using the
relation
a = C[E - EP] = C[E -

],

(3.5)

202

J.C. Simo

CHAPTER I

which follows from (2.5) and (2.9). Substitution of (3.3) into (3.5) then yields the rate
of change of o in terms of the total strain rate as
a=

(3.6)

ceph,

where the tensor CeP of elastoplastic tangent moduli is given by the expression
Cep~

CP{

C-G

-l

[c ,

[Cr]

if
'Y
~O,

(3.7)

(3)

if y > 0.

Note that ceP is generally nonsymmetric for arbitrary r(o, q), except in the case
r(a,q) =

(3.8)

f(o, q),

which has a special significance and is referred to as an associative flow rule.


The derivation leading to expressions (3.3) and (3.4) relies crucially on Assumption 3.1. This assumption is also necessary in order to establish the equivalence between the Kuhn-Tucker complementarity conditions and the classical loading/unloading conditions in strain space, which are essentially equivalent to (3.4).
REMARK 3.1. While the elasticity tensor C is always positive definite (point-wise stable), the matrix H of plastic moduli need not be. The case H = 0 corresponds to
perfect plasticity whereas the case in which H is either positive or negative definite
is referred to as hardening and softening plasticity, respectively. For either hardening
or perfect plasticity, Assumption 3.1 always holds provided that the flow rule is associative. This conclusion is the direct result of the positive definiteness of C since
(3.2) reduces to
G = af

Caf

+ aqf Haqf > QCEIaf

+ ZHlaqf 2 > 0,

(3.9)

where aOE > 0 is the ellipticity constant and aH > 0 both for perfect and hardening
plasticity. Recall that positive definiteness of C holds for any symmetric tensor C E S,
in particular, for ( = daf.
The Kuhn-Tucker form of the loading/unloading conditions admit an alternative, but
entirely equivalent formulation which will prove particularly useful in the formulation
of return mapping algorithms. Define the trial elastic stress rate as
tr

(3.10)

= C,

and declare a process to be plastic whenever


f(o,q)=0 and

f(.

'

,q)

r >0.

(3.11)

203

The classical models

SECTION 3

trial := C:

FIG. 3.2. Interpretation of the loading/unloading conditions in terms of the trial elastic stress rate dr.

TABLE 3.1
Classical rate-independent plasticity.
(i) Uncoupled, linear elastic stress-strain relations:
=
C=

EW( e

= C[E -

2eeeW(ee)

P]

and

q =-aH(~

H = a?2E()

and

=-H~

(both constant)

(ii) Elastic domain in stress space (single surface):


E = {(o, q) E

x Rnin': f(oa, q)

0}

(iii) Flow rule and hardening law:


(iiia) General nonassociative model
p=

yr(,

q)

and

E=

yh(a, q)

(iiib) (Particular) associative case

E = ya-f(oa, q)

and

4 = yaqf(o, q)

(iv) Kuhn-Tucker loading/unloading (complementarity) conditions:


y

0,

f(-,

q) < 0

and

-yf(o,q) =0

(v) Consistency condition:


(f = 0) =

yf(a,q) = 0

The fact that this condition is equivalent to the Kuhn-Tucker conditions follows at
once from (3.10) and (3.3) by noting that, for f (a, q) = 0,

f =0 =

= G-'af

&.tr

(3.12)

204

J.C. Simo

CHAPTER I

Since G > 0, it follows as a result of Assumption 3.1 that


ry>0 0= a,f &r >0,

for

(3.13)

= f = 0,

and the equivalence between (3.11) and the Kuhn-Tucker conditions follows. The
simple geometric interpretation of tr = CE should be noted, and is illustrated in
Fig. 3.2. We observe that Ce is the rate of stress obtained by "freezing" the evolution
0 hence the
of plastic flow and internal variables (i.e., by setting P = 0 and = O)
denomination "trial elastic stress" rate.
The notion of a trial elastic state arises naturally in the context of the algorithmic
treatment of the elastoplastic problem. In the computational literature, use of the algorithmic counterpart of the trial stress rate condition goes back to the pioneering work
of WILKINS [1964] on the now classical radial return algorithm for J 2-flow theory.
The notion was subsequently formalized independently by MOREAU [1976, 1977] who
coined the expression "catching-up" algorithm. We remark that the preceding form of
the loading/unloading conditions is equivalent to the so-called strain space form of
the loading/unloading conditions discussed in NAGHDI and TRAPP [1975] or CASEY
and NAGHDI [1981, 1983a,b].
For the convenience of the reader and subsequent reference we summarize the basic
governing equations of general rate-independent plasticity Table 3.1.
4. Some specific models of classical plasticity
We illustrate the general structure of classical plasticity outlined above with a number
of representative examples.
(A) One-dimensionalplasticity. Consider a one-dimensional bar with unit cross section A = 1 occupying a closed interval Q = [0, L]. The stress, total strain, and plastic
strain then become one-dimensional fields {r, E, EP}. Consider further ni,nt = 2 internal hardening variables ~ = (, ) and specify the uncoupled free energy function
T(E', E) = W(Ee ) + 7-t() using
W(Ee) = lE'EE,

-()=

. Ht

where H =

Here E > 0 is the Young modulus and K, H are referred to as the isotropic and kinematic hardening moduli, respectively. The (generalized) elastic stress-strain relations
are
= E(e-EP),

q =-Kf

and

q=-H.

(4.1)

The yield criterion in stress space is specified as


f(, q) =

+q- y

0.

(4.2)

The classical models

SECTION 4

205

elastic domain
in stress space

I
-cry

cry

FIG. 4.1. Elastic domain in stress space for a one-dimensional model of plasticity, with linear isotropic
hardening, corresponding to K > 0 and H = 0. The ordinate is = -q/K > 0.

Accordingly, the interior of the elastic domain int(E) = {(or, q): f(a, q) < O} in
stress space is an open set centered at r = q with half length (ay - q) > 0. The
case in which K > 0 and H = 0, illustrated in Fig. 4.1, is referred to as isotropic
hardening. The rate equations for the internal variables (, ) (the hardening law)
dictate the evolution of the center of the (one-dimensional) elastic domain and its
"size" in time. An associative flow rule yields
gP = ya f((, q) = ysign[ - ],
1 ra~i(,4i= ?~ _,,.~-,)
qf(a,
f= q) = ' -sign[a - ]

(4.3)

The first component of (4.3)2 defines a linear isotropic hardening law. The idealized
stress-strain curves for the case H = 0 and both K = 0 (perfect plasticity) and
K > 0 (isotropic hardening) are illustrated in Fig. 4.2. The second component of
(4.3)2 is referred to as the kinematic hardening law, often credited to PRAGER [1956]
with further improvements by ZIEGLER [1959], and is motivated by the experimental
observation that the center of the yield surface experiences a motion in the direction
of the plastic flow. This hardening behavior is closely related to a phenomenon known
as the Bauschinger effect. The stress-like internal variable , which defines the center
of the yield surface, is known as the back-stress.
The incremental stress-strain relations are given by the rate equations

u= EP

where E

(.

E[H+K]

E+[H+K]

if3' =O,
if

> 0,(4.4)

which are easily obtained by specialization of (3.7). Ee P is the one-dimensional elastoplastic tangent modulus.

206

J.C. Simo

CHAPTER I

Cry I

--oy

FIG. 4.2. Idealized one-dimensional stress-strain curve for perfect plasticity and (nonlinear) isotropic strain

hardening corresponding to a yield criterion f(cr, q) = Icrj + g(q) - uy.


a

HI>0

ftening)

ftening)
FIG. 4.3. For one-dimensional plasticity, Assumption 3.1 places a limit on the amount of allowable softening
in the model.

REMARK 4.1. In the present one-dimensional context, Assumption 3.1 has a simple
interpretation. Substitution of (4.3) into (3.2) along with a straightforward computation
reveals that

G = af.

Cr + a3qf. Hh = E + [H + K] > 0.

(4.5)

The physical significance of this condition can be easily appreciated by inspection


of Fig. 4.3. Condition (4.5) places a restriction on the amount of allowable softening
in the sense that [H + K] > -E A> E eP > -oc where Ee P is defined by (4.4).
It should be noted that the classical condition of nonnegative second-order work
density (see HILL [1958]) implies that E + H > 0, which is much stronger than
the restriction implied by Assumption 3.1. Even if this assumption holds, models of
plasticity exhibiting softening response present considerable mathematical challenges
which do not arise in conventional hardening models. For instance, in the presence of

The classical models

SECTION 4

207

softening response the initial boundary value problem can be shown to always exhibit
solutions possessing strong discontinuities, see SIMI, OLIVER and ARMERO [1993]
and references therein. These topics are currently an active area of research.
(B) General quadratic model for three-dimensionalplasticity. Let A E R9X 9 be a
symmetric (positive definite) matrix and, as before, let ay be the flow stress. Consider
a yield criterion of the form
f(a, q) = (a) + q - .A

((ot) =

0,

o,

(4.6)

q = -K.
As before I stands for a strain-like hardening variable which models isotropic hardening, with evolution defined below and dual variable q = -KE, where K is the
(isotropic) hardening modulus. If K < 0 we speak of strain softening response. The
evolution of eP and is defined by the associative rules

P= yaf = yV

and

=yaqf = ,

where V(rr) =

Aa
(4.7)

with y > 0. Observe that = /P A-l&P. The elastoplastic moduli are obtained by
specialization of the general expression given in (3.7). Using the preceding relations,
an easy manipulation yields the result

Cep=

C
C

CV&(r)

CV(r)

if Ty=0,
if y > 0.

(4.8)

Observe that this expression gives symmetric elastoplastic moduli, in agreement with
the associative character of the model under consideration.
REMARK 4.2. The quadratic form S(r) defined by (4.6) is obviously a convex function

which is homogeneous of degree one, in the sense that 0((a) =


By Euler's theorem, this property implies the result
a,&(t)

r = q(o'),

for all

E S,

(a) for any ( > 0.

(4.9)

which is easily verified. Practically all of the yield criteria used in plasticity enjoy the
preceding property.
REMARK 4.3. The extension of the preceding model to account for other types of

hardening is straightforward. For instance, consider a yield criterion f(ao, q) of the


form
f(a, q) = (o -

) + q-

y ~<0,

(4.10)

208

J.C. Simo

CHAPTER I

where q defines the center of the yield surface, and define the hardening potential as
7-/() = - H where
(

={j

so that q = -H.

HI'

and H =[

(4.11)

The associative hardening evolution equations

= aq'yf(a, q)

'HV4(oa - O) and

=-

K,

(4.12)

then furnish a generalized model of combined kinematic/isotropic hardening of the


type described in the context of the preceding one-dimensional example. All the results
described above carry over to this more general setting without essential modification.
(C) Plane strain and three-dimensional J2 -flow theory. The preceding general
quadratic model includes a number of plasticity models widely used in practice. In
particular, by suitably restricting the quadratic form that defines the yield condition,
one recovers the anisotropic criterion of HILL [1950], the general anisotropic yield
condition of TSAI and Wu [1971], and the von Mises-Huber yield criterion both in
plane strain and plane stress.
As an example, a widely used extension of the classical Prandl-Reuss equations of
perfect plasticity that incorporates hardening is obtained by considering linear isotropic
elastic response, with elasticity tensor C = AXl 1 + 2/I, and the following pressure
insensitive, isotropic, Huber-von Mises yield condition
f(a, , q) = Idev[ ] -

l- [Oy-]

(4.13)

0,-

where dev[a] = ao - tr[o ]l is the deviatoric part of the stress tensor and ry is the
one-dimensional flow stress. Here q is the back stress and q is the isotropic (stresslike) hardening variable, assumed to be related to the strain-like hardening variables
= ({, E) via the relations
q=-K'( )

and

(4.14)

q=-sH4,

which arise from the hardening potential 7-() = K() + HItg 2 . It is implicitly
assumed that tr[] = 0, i.e., the back-stress is a deviatoric tensor. Denoting by n =
(dev[a] - q)/dev[r] - qJ the unit normal field (in S) to the von Mises cylinder, the
associative evolution equations can be written as
gP = af

= 7n,

= 7qf
= -'n

and

= ?aqf
a = By.

(4.15)

Relation (4.15)1 is known as the Levy-Saint Venant flow rule. Equation (4.15)2 along
with (4.14)2 define the evolution equation for the back stress as q = 72Hn; a relation
known as the Prager-Ziegler kinematic hardening rule already discussed above; K"(E)

209

The classical models

SECTION 4

and H are referred to as the isotropic and kinematic hardening moduli, respectively.
Since IP I = y, relation (4.15)3 implies
,(xt)

3P(x, T)

d,

(4.16)

which is in agreement with the usual definition of equivalent plastic strain. Alternatively, one may use the notion of equivalent plastic work; e.g., see NAGHDI [1960],
KACHANOV [1974] or MALVERN [1969], to characterize hardening. In applications
to metal plasticity it is often assumed that the isotropic hardening is linear so that
K(() = K 2 . The following form of combined kinematic/isotropic hardening laws
is widely used in computational implementations
H = (1 - f)H,

K'(()

= /3H,

p E [0, 1]

and

H = constant.

(4.17)

More generally, nonlinear isotropic hardening models are often considered in which
a saturation hardening term of the exponential type, as in VOCE [1955], is appended
to the linear term, i.e.,
K'() = PHf + (Koo - Ko) [1 - exp(-6S)],

(4.18)

where H > 0, Koo > Ko > 0 and 6 > 0 are material constants. Finally, the
elastoplastic tangent moduli in plastic loading are obtained by particularization of
(3.7) as
CeP=

1 X1+2P [Ip- 1&


L

n
1 + H+K"J

for y > 0.

(4.19)

(D) Plane stress J2 -flow theory. In this example classical J2-flow theory with combined kinematic and isotropic hardening is cast into the general format of the quadratic
model described above under the assumption of plane stress. Recall that the plane
stress subspace is defined by Sp = {o E S: c13 = 23 = 33 - 0}. Since Sp is
isomorphic to Ri3, we introduce the following vector notation:
=

[11

22 c12]T

and

q =

.
[qll q22 ql]T
2

(4.20)

Note that 33 = -(qll + q22) since the back stress tensor qij is deviatoric. Following
standard conventions, we collect the components Eij of the strain tensor E S in
vector form as
E = [1

22 212]T

and EP = [EP EP2

P2]T

(4.21)

The standard convention of multiplying the shear strain component E12 by a factor of
two allows us to write the stress power as rijij = aTg. To formulate the plane stress

210

J.C. Simo

CHAPTER I

version of J 2-flow theory directly in Sp, observe that the von Mises yield condition
can be written as
f(O,

, q) =

P=

2
-1
0

/[TY

- q] < O,

(4.22)
2
0

06

0,

where = r - q is the relative stress. Noting that the mechanical dissipation can be
written in terms of the vector notation introduced above as D = TkP + qT + q J,
the specialization to plane stress of the basic equations of three-dimensional J2-flow
theory takes the form

= Y7P3,

= -yP,3

and

where we have replaced y > 0 by y/


take the form
O=

q =-H

EP],

CIE -

=y /23Tp/3,
3
/3TP/3. The generalized stress-strain relations

and

q = -K'(().

(4.23)

Here C and H are the elastic and plastic moduli in plane stress which, with the
preceding conventions, take the following form for the isotropic case:

C =

and

(- 0

H= 2 H

,1

(4.24)

where E > 0 is the Young's modulus and v E [f, ] is the Poisson's ratio. The
elastoplastic moduli can be easily obtained either by specialization of the general result
in (3.7) or via a direct computation. To facilitate comparison with future algorithmic
developments, the explicit result is quoted below under the assumption of plastic
loading:

Cep

CP/ 3CP/
CP
C
/ T[P(C + H)P + 2K"( )P]
D

(4.25)

Equations (4.22) and (4.23) are now in a form which is ideally suited for the application
of the general return mapping algorithms discussed in the next chapter.
REMARK 4.4. For the case of isotropic elasticity the constitutive matrices C, H and

the projection matrix P have the same characteristic subspaces and, therefore, can be
simultaneously diagonalized (H is already diagonal); see SMO and TAYLOR [1986]
for the explicit result. Since P and C have the same eigenvectors, it follows that P,
C and H commute.

211

The classical models

SECTION 5

REMARK 4.5. It should be noted that the strain components 33, 33, E3P3 and q33 do
not appear explicitly in the formulation. These are dependent variables obtained from
the basic variables (, EP, o), the plane stress condition, and the condition of isochoric
plastic flow. For the case of isotropic elasticity we have
E33 =

(El + E22),

EP3 = (

+P2)

The total strain

E33

and

q33

-(qll

then follows simply as

+q22)

(4.26)

E33 = E3 + EP3.

5. General rate-independent multisurface plasticity


In many application of interest, the boundary aE of the elastic domain E is nonsmooth.
A classical example is furnished by the Tresca yield criterion of metal plasticity, see
HILL [1950], which can be visualized as a hexagon on the 7r-plane. Yield criteria
widely used in solid mechanics, such as the Drucker-Prager condition or the so-called
cap models introduced in DIMAGGIO and SANDLER [1971] and SANDLER, DIMAGGIO
and BALADI [1976], are constructed as the intersection of several yield surfaces leading, therefore, to a nonsmooth boundary aE. An overview of possible applications of
these in geomechanics is given in DESAI and SIRIWARDANE [1984]. Multisurface plasticity also arises in a class of phenomenological models for metal plasticity known as
comer theories, see, e.g., CHRISTOFFERSEN and HUTCHINSON [1979], which attempt
to replicate experimentally observed results not explained by classical J 2-flow theory.
A modified J 2-flow theory that partially accounts for these effects is given in SIMO
[1987]. Multisurface plasticity also plays an important role in formulations of crystalline plasticity due to the presence of multiple slip systems, see the review article
of ASARO [1983].
In this more general setting of multisurface plasticity, the elastic domain E is a
convex set defined in terms of smooth convex functions f,:S x IRin't -i R, with
, C { 1,... , m}, as the constrained set
E= {

= (, q)

S x Inint: f(X) < 0, forp = 1,..., m}.

(5.1)

It is implicitly assumed that the constraints f,(Z) are qualified; a condition which
holds if either the constraints are affine or if there exists a point E S x IRnT ' such that
f () < O0for = 1,... ,m, see CIARLET ([1988], p. 345) for further details. This
technical requirement is satisfied by all elastic domains used in classical plasticity.
For convenience, in what follows we shall refer to Z = (, q) E as the generalized stress. The smoothness hypothesis on the functions f,(Z) along with the
convexity assumption imply the following standard result (see, e.g., CIARLET [1989],
p. 243)
f,(T) - f,(-) >, [T - _]

Vf,(),

T,.E c

x R'i'",.

(5.2)

J.C. Simo

212

CHAPTER I

FIG. 5.1. Illustration of the convexity property for a smooth one-dimensional function f: IR -- R

(ndim =

).

An illustration of this property is given in Fig. 5.1. The associative form of the
flow rule and the hardening law for multisurface plasticity is given by the following
evolution equations which go back to KOITER [1960] (see also MANDEL [1965]):
r=l

'yPa.f',(oq),

Ep=

=aq),

(5.3)

yU > 0,

f,(a, q)

0 and

yf,(

, q) = 0,

for

=1,...,.

P=1

As before, y* > O0are the plastic consistency parameters and relations (5.3)3 define
the Kuhn-Tucker form of the loading/unloading conditions. A deeper motivation for
these evolution equations is deferred to the following section.
The elastoplastic tangent moduli in the incremental relation = Cepe are obtained
by a generalization of the consistency argument described in Section 3. This generalization requires a precise definition of the active set of constraints during plastic
loading. For this purpose, let
JIt

E {1,

= {t

*,

m}: f(a,

q) = 0}

(5.4)

denote the set of indices containing the constraints which have the possibility of being
active at (a, q) aE and set mtr = dim[Jt]. The subset J C Jt' of active constraints
is clearly defined by the condition

E Jt':

f?(r, q)- O.

(5.5)

Our goal is to provide an explicit characterization of this set. Using the chain rule and
the (generalized) stress-strain relations, it follows that
f,

= af4

C[i -

P] -

aqfo Hi.

(5.6)

The classical models

SECTION 6

213

Now define the mtr(m tr + 1)/2 coefficients G,, = Gm by the relation


Gm , = a,f, Caaf, + aqf,

Haqfv,

/'

E Jtr.

(5.7)

Observe that the matrix G = [G,,] is symmetric and positive definite for either
perfect or hardening plasticity, with inverse denoted by G- 1 = [GIm]. For softening
plasticity, the requirement that this matrix remain positive definite is analogous to
enforcing Assumption 3.1. Inserting the evolution equations (5.3) into (5.6) and setting
otr = C

gives
tr

= afl

G>f
,,'

for

J.

(5.8)

By enforcing the consistency condition (5.5), the plastic multipliers are explicitly
computed as

Y7=
veJ

tr

af, .

tr for

Jtr,

(5.9)

and the set J of active constraints becomes J = {/ E Jt: 7y > 0}. From this result
and (5.9), the elastoplastic tangent moduli are computed as

cep = C -- E

GI"'[Ca,f,]

[Caf].

(5.10)

l~EJ vEJ

The preceding developments have an algorithmic counterpart. The interested reader


is referred to SIMO, KENNEDY and GOVINDJEE [1988] for further details.
REMARK 5.1. The set Jtr depends on the specific point under consideration. The additional assumption that the constraints are independent at each point on the boundary
aE, understood in the sense of linear independence of the vectors a,f, is crucial for
expression (5.10) to remain valid. Without this hypothesis the multipliers y/' > 0 need
not be unique and the matrix [G,,] is generally noninvertible. The importance of these
conditions does not appear to have been appreciated in the literature on plasticity. For
a discussion of these issues in the general context of nonlinear optimization and, in
particular, the role played by the technical hypothesis that assumes the constraints to
be qualified, see KARMANOV [1977] and CIARLET ([1988], Chapter 9).
6. Dissipation: Interpretation of the model of associative plasticity
The associative form of the flow and the hardening laws arise when the functions
r(oa, q) and h(a, q) coincide with the gradients of the yield criterion. A more fundamental interpretation of this choice arises when the notion of dissipation is introduced.
Classical plasticity defines a dissipative dynamical system in the sense that a local
function of the state variables, called the dissipation function, does not decrease with
the evolution of plastic flow. Given a general elastic domain, among all possible

214

J. C. Simo

CHAPTER I

choices of flow rule, the associative form is that leading to maximal dissipation in the
system. In essence, this is the contents of a classical result going back at least to VON
MISES [1925] (see HILL [1950]) and known as the principle of maximum dissipation.
This result is described below after motivating and defining the appropriate notion of
dissipation for classical plasticity.
We shall be concerned with a general model of plasticity in which hardening processes are characterized by nint internal (stress-like) variables q(x, t). It proves expedient to introduce the following notation:
E P = (EP, ),

_ = (, q),

and

E[u] = (E[U],O).

(6.1)

We consider a general free energy function 1(x, e, E), convex in (es, 5) and a general
elastic domain defined by a closed, convex set IE C S x RIIn ~ with boundary OlE not
necessarily smooth. The specific examples described in the preceding section provide
an illustration of this general setting. Here, attention is focused on the general case of
multisurface plasticity corresponding to
E=

0, for ti= l,...,m},

'i: f()

(6.2)

where f,: x in' - 1 are smooth convex functions.


Within this general framework, the local internal dissipationfunctionD is defined
as the difference between the stress power and the rate of change of free energy, i.e.,
E[v] - dT(Ee, ,),

D=a

in

x .

(6.3)

Carrying out the time differentiation via the chain rule and using the generalized
stress-strain relations along with e = E[V] - EP yields the expression
D=

+q

- EP

(6.4)

The local form of the second law requires that the dissipation inequality D) > 0 hold
in any admissible process in the material body. Since the internal variables EP are
not present in elasticity, it is clear that D) O0for an elastic material, a property
which motivates the term "perfect materials" coined in TRUESDELL and NOLL [19651.
It follows that classical elasticity trivially satisfies the dissipation inequality.
EXAMPLE 6.1. A more interesting example is furnished by associative plasticity with
elastic domain E defined by (6.2) where the convex functions f,(Z) are of the form
fl(I) =

(7) -rryf,

p = 1,2,. ., m.

(6.5)

Here cry > 0 are constants and the functions 5,(-) are convex, homogeneous of
degree one, so that VO,(Z)
= 5,(L') where Va, = (,,
q,). Using this

SECTION 6

215

The classical models

property and the associative flow rule (5.3), the dissipation function becomes
m

v=

yVO

= E 1' [(b - orY) + ay,]

A=1

/=1

+ ayl]

(6.6)

L=1
t
Hence D =
A
yl uyuo> ) 0 since yr/ > 0 and 2=l yf. = 0, as a result of the
Kuhn-Tucker conditions. The model, therefore, obeys the dissipation inequality.

Let Z denote the actual stress field in 2 for prescribed strain rates E[v] and EP.
The evolution of plastic flow is said to obey the local principle of maximum dissipation
if, for any (x, t) E Q x II, the following inequality holds:
[T-X] -EP = ( -

P + (p - q)

0VT
O
= ()
(r,p)E E.

(6.7)

The significance of this postulate lies in its intimate connection with the classical
model of associative plasticity, as the following result shows.
THEOREM 6.1. The principle of maximum dissipation is equivalent to the associative
flow rule with loading/unloading conditions formulated in Kuhn-Tucker form, i.e.,
inequality (6.7) is equivalent to
T

ip

= Z'Y"Vf.(Z),
tI=l

(6.8)
y"

f (Z) I 0, and

0,

yfb() = 0,

/z=l

with E defined by (6.2) and an arbitraryconvex free energy function @(x, e, E).
PROOF. Suppose that maximum dissipation holds and assume that the elastic domain
is given by (6.2). According to standard results in optimization, this means that the
actual state is a stationary point of the (unconstrained) Lagrangian
(T, y

=-T

f,(T).

EP +

(6.9)

p1=1

The standard optimality conditions for this saddle-point problem are (see, e.g., CIARLET [1989], Chapter 9)
aT(T, Y')T=

= -EP + E y7Vfpi(Z) = 0,
fi=l

(6.10)

JC. Simo

216

CHAPTER I

along with the Kuhn-Tucker complementarity conditions. Hence maximum dissipation


implies the associative flow rule (6.8). Conversely, if (6.8) holds, contracting the flow
] yields
rule with [T -

i[T-] *E =Z yatVfp(L) *([T

Using the convexity assumption on the functions f(E)


Tucker conditions we conclude that

aUVfi,(z

) [T - Z] 4

/1=1

(6.11)

VT E .

and enforcing the Kuhn-

-[ f()]
=1

(6.12)

= 7~f,(T) < 0,
since y" > 0 and f,(T) < 0 for any admissible T
(6.11) we arrive at the dissipation inequality (6.7).

E. By combining (6.12) and

7. Rate-dependent plasticity: The viscoplastic regularization


The inclusion of rate-effects in classical plasticity leads to the so-called viscoplastic
models considered subsequently. We restrict our discussion to those mathematical
aspects which are essential for our subsequent algorithmic developments. In particular,
emphasis is placed on the interpretation of viscoplasticity as a penalty regularization of
the local dissipation function. This point of view, pioneered in the work of DUVAUT and
LIONS [1976], is the most useful one in an algorithmic sense. For general treatments
concerned with the physical aspects underlying models of viscoplasticity we refer to
PERZYNA [1971] and LUBLINER [1972, 1973], among others.
The construction of viscoplastic models described below, inspired in the approach
advocated in DUVAUT and LIONS [1976], is very closely related to modern algorithmic
treatments of plasticity based on the notion of closest-point projection of the generalized stress onto the convex domain E described in the next chapter. For simplicity,
attention is restricted to the case of a quadratic free energy function defined as
(E'e ,)

= 2e

* Ce +

2a.HE,

(7.1)

where both the elasticity tensor C and the ni,,n x nin matrix of plastic moduli H are asT
T
sumed to be constant and strictly positive definite on S and ]R'" x lR` , respectively.
In other words, attention will be restricted to hardening plasticity. It therefore follows that the complementary Helmholtz free energy function, defined via the standard
Legendre transformation as
7(Z) = 2
2( )

+C-lo-+
q*-H q,

(7.2)

217

The classical models

SECTION 7

is a strictly convex function on S x Rn int. Under the preceding hypothesis 2(T()


defines a norm on the finite-dimensional vector space S x Inin' equivalent to the
standard Euclidean norm.
The key feature exhibited by viscoplastic models is that, in sharp contrast with rateindependent plasticity, admissible (generalized) stresses Z are no longer constrained
to lie in the elastic domain E. Stress states may lie outside of the elastic domain as a
result of rate effects and, as time progresses, may relax to the yield surface, often at
an exponential rate if the loading is held constant. Plastic flow then becomes a ratedependent process. To incorporate this fundamental effect in the model one proceeds
as follows.
Step 1. Consider an arbitrary stress state Z E $ x Rni nt, not necessarily lying in E,
and define the functional J:S x ]Rni by the constrained minimization problem
J(Z) = min{/22(

-T),

for all T

E}.

(7.3)

It follows that J(Z) gives the distance measured in the complementary Helmholtz free
energy (.) between a given point 7 and the convex set E. Clearly, J(Z) > 0 for
any Z E x ]Rni t, and J(r) = 0 iff E IE.In view of the preceding considerations
we have
J(')

= V/23(-

),

X cIE,

(7.4)

where X. denotes the closest-point projection of Z onto E in the norm 2(.). By


the projection theorem (see, e.g., CIARLET [1989], Chapter 8), this point is unique.
Step 2. Let y+:IR property
y+ (x ) >

IR+ U {0} be a nonnegative, C'-convex function with the

and y+(x) =

x = 0.

(7.5a)

Consider the following viscous regularization of the dissipation function D = E .EP:


-D

= -z.

EP +

(J(r)),

1-+

(7.5b)

where E (0, o) is a regularization parameter with dimensions of time, referred to


as the characteristicrelaxation time in what follows. A possible choice for y+(-),
widely used in nonlinear optimization, is
7)+'

( X_{x
=

2
0

if x

if

O,
0,
< 0,

so that

d(

+()

={

ifx 0
ifX 0
if X<0

(7
(7.6)

218

J.C. Simo

CHAPTER I

By standard results in convex optimization, see, e.g., LUENBERGER [1984], it follows


that the problem of minimizing minus the regularized dissipation -D over all unE S x IRn'tn is simply the penalty regularization of the classical
constrained stresses
constrained principle of maximum dissipation, with dissipation function D.
7.1. The optimality conditionsassociatedwith the unconstrainedmaximization of the regularized dissipation D, yield the following viscoplasticflow rule
THEOREM

= {g(J(17))G-'

[a - 1

if J()

>

(77)

if J(17) = 0,

where g(x) = dy+(x)/dx and G = DIAG[C, HI are the generalized moduli.


PROOF. The result follows from a computation of the gradient of J(-) that uses relation
(7.4). Further details are omitted.

In summary, as a result of the preceding construction a model of associative viscoplasticity is completely defined by specifying the elastic domain E, the free energy
function or, equivalently, the complementary Helmholtz free energy function _(X)
and the function g: i -- ]R+ that characterizes the rate-dependent flow rule. Observe
that the Kuhn-Tucker optimality conditions of rate-independent theory are now replaced by the conditions in (7.7). The two-step construction outlined above will be
illustrated in the context of the classical model of J2 -flow theory described in the
preceding section.
8. Example: Viscoplastic regularization of J2-flow theory
Consider the model of J2-flow theory described in Section 4 incorporating both
isotropic and kinematic hardening, assumed to be linear for simplicity. In the present
context, the complementary Helmholtz free energy function is given by

2 (lH)
where

-=

I2 trj]| 12

(8.1)

> 0 is the bulk modulus, bM> 0 is the shear modulus, K is the isotropic

hardening modulus, and H is the kinematic hardening modulus. The yield condition
is the von Mises yield criterion

f(E) = dev[

[
[-

y--

< 0.

(8.2)

Our goal is to carry out explicitly the two steps in the construction outlined above
leading to the regularized flow rule in (7.7).

The classical models

SECTION 8

219

FIG. 8.1. Geometric interpretation of the closest-point projection of a stress point onto the elastic domain
for J2 -flow theory in the absence of hardening mechanisms (perfect plasticity). The notation s = dev[ar]
is used.

Step 1. Given Z = (a, q, q), the projection X. = (, , q,) onto E and the
distance J(Z) to E are computed from the optimality conditions of the Lagrangian
(27, X,) = S(

- 2,,) + A.f('*).

(8.3)

Figure 8.1 provides an illustration of the significance of 7 and J(Z) in the simpler
case of perfect plasticity. Since
af(

) = -af(E.)

= n, =

dev[,]

- q/

dev[u*.] -

(8.4)

with In. I = 1, a straightforward calculation determines the optimality conditions as


dev[a,,] - dev[o] = -2/ A. n.,
*-

= 2H
3 An*,

q.-q
= -V-K

(8.5)

A,

along with A,, > 0 and A*f('*.) = 0. We assume the nontrivial case A* > 0 so that
f(2*) = 0. Clearly, the point 7.*is completely determined from the multiplier A*
and the unit vector n,.
(A) Determinationof n* and A* > 0.
[Idev[a] -

From Eqs. (8.5)1,2 we conclude that

] n = [dev[a*] - q*1 + A*(2/ +

H)] n,

(8.6)

where n is the unit vector also given by expression (8.4)3 but with (ar, q*) replaced
by the given (a, q). This relation determines n, as
n

dev[a]-

dev[] Idev[o,] - 41'

(
(8.7)

220

J.C. Simo

CHAPTER I

Using this result, relations (8.6), (8.5)3 along with the von Mises criterion (8.2) yield
f(*)

= f(Z) - (2t +

(8.8)

K)X,.

H+

Therefore, by enforcing the condition f(17) = 0 one arrives at the following explicit
expression for the multiplier A,:
(8.9)

)* = f (Z)/(2p+ K + H).

The projected point ,

is then defined by relations (8.5) along with (8.7) and (8.9).

(B) Computation of the distance J(Z). From the optimality conditions (8.5), the
complementary Helmholtz free energy function of the difference
- Z,, is easily
evaluated as
.7

2;(

- 1.) = (2

+ K + ~H)A2..

(8.10)

Inserting (8.9) into this expression yields

J(Z) = /2 - (

) = ()/

/(2

+ K + H).

(8.11)

Step 2. Adopting for simplicity expression (7.6) for the function y+(), the viscoplastic regularization of the dissipation can be written as

- EP -

D,r =.

[f(_)] 2 ,

for f(7) > 0,

(8.12)

where 7 > 0 is a viscosity coefficient (with dimensions [stress] x [time]) defined in


terms of the relaxation time r as

1 = (2 + K + H) .

(8.13)

Finally, for the present model problem the general viscoplastic flow rule takes the
following explicit form:
ep=

1(f(Z))n,

=--g(f())n,
37

where g(x) =

and

.1(8.14)

= -g(f())
37

(8.14)

[x + z[x] is the ramp function.


We observe that the evolution equations (8.14) can be obtained from their counterparts (4.15) in the rate-independent theory merely by providing the constitutive

The classical models

SECTION 9

221

equation y = g(J())/rl for the consistency parameter and omitting the KuhnTucker conditions. This idea appears to have been suggested by PRAGER [1956], see
PERZYNA [1971]. From well-known results in penalty methods for constrained optimization (see, e.g., LUENBERGER [1984] and BERTSEKAS [1982]), one concludes that
the rate-independent model can be recovered as the limit of zero viscosity ( - 0);
an obvious property from a physical standpoint.
REMARK 8.1. Early algorithmic treatments of plasticity often exploited the view of
viscoplasticity as a penalty regularization of classical plasticity as a means of devising numerical schemes for the rate-independent problem, as in ZIENKIEWICZ and
CORMEAU [1974], CORMEAU [1975], HUGHES and TAYLOR [1978] and PINSKY, ORTIZ
and PISTER [1983]. This approach, however, has the disadvantage of inheriting the
characteristic ill-conditioning experienced by penalty methods for very high values of
the penalty parameter and are considered no longer efficient.
REMARK 8.2. From an algorithmic standpoint the advantage of the class of models
described above, which are based on the regularization suggested in DUVAUT and
LIONS [1976], is that the characteristic ill-conditioning experienced by penalty methods can be completely circumvented. As observed in SIMO, KENNEDY and GOVINDJEE [1988], the key idea is to compute the viscous (rate-dependent) solution from
the rate-independent solution and not vice versa. The numerical approximation to
the rate-independent problem is obtained via return mapping algorithms described in
subsequent chapters.
9. The weak formulation of dynamic plasticity
The weak form of the momentum equations, briefly described in Section 1, along
with the constitutive equations formulated as a variational inequality comprise the
weak formulation of the initial boundary value problem for classical plasticity and
viscoplasticity. The goal of this section is to provide an account of those mathematical
aspects involved in the weak formulation of the problem which are relevant to its
numerical approximation. Throughout this section attention is restricted to the case of
a quadratic free energy function defined by (7.1). All the results are easily extended
to an arbitrary convex stored energy function.
The weak form of the constitutive equations for associative plasticity, the situation
of interest here, is simply the global formulation of the maximum (plastic) dissipation inequality (6.7) over the entire domain Q2 of the body. Denote by (ij,pi) the
components of T = (r, p) relative to the standard basis and let
T = {T = (r,p):

$ X T/x
"int: Tij G L 2(Q) and pi

L 2(0)}.

(9.1)

By eliminating the generalized plastic strains via the constitutive relations, inequality
(6.7) for the rate-independent problem can be written as
(r-a).,

e[v]-(r-a).C-

- (p-q) .H-'<0

V(T,p) EE.

(9.2)

222

J.C. Simo

CHAPTER I

trial := (C: E[t], 0)

.E
FIG. 9.1. Geometric interpretation of the variational inequality A(

{E: f()

tr

- T, ,T-

E)

0, for E =

< 0%

Motivated by this expression, one defines the bilinear forms a(-, ) and b(-, ) as
a(,r)

C-' lQ,

b(q,p) =

q HpdQ.

(9.3a)

Since the elastic moduli C are positive definite on and the plastic moduli H are
assumed positive definite on ni t, it follows that the bilinear forms a(., ), and b(-, )
are coercive. Consequently, the bilinear form A(., ) : T x T defined by
A(,

T) = a(o, r) + b(q, p)

(9.3b)

induces an inner product on I. The square of the induced norm, A(, E), of any
generalized stress field Z E IE is precisely twice the integral over the body of the
complementary Helmholtz free energy function E(L). With this notation in hand, the
global version of the maximum dissipation inequality (9.2) for the rate-independent
problem takes the form of the following variational inequality:
A(~trwhere

, Tttr=

0,

(Ce[Ui], 0).

E EnT,
(9.4)

Recall that Ztf is the trial stress rate about X! E E, obtained by freezing plastic flow
(i.e., by setting P = 0 and = 0).
The geometric interpretation of inequality (9.4)1 is illustrated in Fig. 9.1. The actual
stress rate Z is the projection in the norm defined by the bilinear form A(-, .) of the
trial stress rate tr onto the tangent plane at Z E aE. Convexity of E then implies
that the angle measured in the inner product defined by A(-, ) between [tr -_I]
and [T - ] is greater than or equal to 90 for any T E E; a condition equivalent to
(9.4)1.

223

The classical models

SECTION 9

The weak form of the momentum equations (1.16a) together with the dissipation
inequality (9.4) give rise to the following weak formulation of the initial boundary
value problem for rate-independent plasticity:
= (, q) E E n T such that

Problem Wt: Find t E Il H4u C St and t e I1-

(p(v - it), n) = O V E V,
(pov, 1 ) + (,e E[]) - (f,r) - (t,7)r = 0

A('t -

0 Vr

, T-

E)E

Vr

(9.5a)

V,

n T,

subject to the initial conditions


(u(-, O), 7) = (Uo, 7)

(v(.,O),

and

) = (vo, q), 7, V
7

V.

(9.5b)

REMARK 9.1. For the rate-dependent problem inequality (9.4) is replaced by the weak
formulation of the maximum dissipation inequality associated with the regularized
is
dissipation (7.5b). The dissipation over the entire domain
J Dd =A(

,y+

(J())d

> 0,

(9.6)

and the assumption of maximum dissipation implies the following inequality for viscoplasticity:

A(

tr -

,T-

r [y+(J(T)) - y+(J(Z))] d,

(9.7)

for all r c T not necessarily contained in E.


The weak form (9.5a) of the dynamic elastoplastic problem involves the velocity
and the generalized stress field. In the subsequent analysis it is convenient to use the
notation
X = (v, Z) C Z

where Z = V x [TnE
I]

(9.8)

is the space of admissible velocities and admissible generalized stress fields. The
space Z is equipped with a natural inner product induced by the kinetic energy and
the complementary Helmholtz free energy function, which is denoted by ((., .)) and
defined by
((XI, X2)) = ((POrl, r12)) + A(T, T 2 ).

(9.9)

The associated natural norm is denoted by


= |Il
-|
|| ) and is interpreted
as twice the sum of the kinetic and Helmholtz energies of the elastoplastic body. It

224

J.C. Simo

CHAPTER I

*)) is the natural norm for the

will be shown below that the norm


elastoplastic problem.

REMARK 9.2. The variational formulation of plasticity given by equations (9.5b) goes
back to the pioneering work of DUVAUT and LIONS [1976] and has been considered
by a number of authors, in particular, JOHNSON [1976a, 1978]. The assumption of
hardening plasticity, i.e., the presence of the bilinear form b(., ) defined by (9.3a) 2,
is crucial for the functional analysis setting outlined above to remain applicable.
REMARK 9.3. The situation afforded by perfect plasticity is significantly more complicated than hardening plasticity since the regularity implicit in the choice of St
in (1.14) no longer holds. The underlying physical reason for this lack of regularity is the presence of strong discontinuities in the displacement field, the so-called
slip lines, which rule out the use of standard Sobolev spaces. For perfect plasticity
the appropriate choice appears to be St = BD(S2), i.e., the space of bounded deformations [displacements in L 2 (2) with strain field a bounded measure] introduced
by MATTHIES, STRANG and CHRISTIANSEN [1979], and further analyzed in TEMAM
and STRANG [1980]. See MATTHIES [1978, 1979], SUQUET [1979, 1981], STRANG,
MATTHIES and TEMAM [1980], and the recent comprehensive summary in DEMENGEL
[1989] for a detailed elaboration on these and related issues. Very little is currently
known for the case of softening plasticity, other than the solutions must exhibit strong
discontinuities (see SIMO, OLIVER and ARMERO [1993] and references therein). These
discontinuities are intimately related to the formation of shear bands.
10. Contractivity, uniqueness, and dissipativity of the elastoplastic flow
Two properties of the initial boundary value problem are examined below in some
detail which are of fundamental importance in the formulation and analysis of algorithms. First, the dissipative character of the problem, which translates into an a priori
stability estimate for the solutions. Second, the contractivity property of the solutions
relative to certain norm, called the natural norm of the problem, which motivates the
notion of algorithmic stability employed in the subsequent analysis of return mapping
algorithms.
The following contractivity property inherent to problem Wt identifies the norm
II III induced by the energy inner product (9.9) as the natural norm and plays,
therefore, a crucial role in the subsequent algorithmic stability analysis. Let

Xo = (vo, o)

and

Xo = (iio,

(10.1)

)EZ

be two arbitrary initial conditions for the problem of evolution defined by (9.5a) and
denote by
tC]Ik-X=z(v,L)CZ

and

t E:1jX-

= (0,

(10.2)
.)

The classical models

SECTION 10

225

the corresponding flows generated by (9.5b) for the two initial conditions (10.1),
respectively. For the quasistatic problem, the following result is implicitly contained
in MOREAU [1976, 1977]; see also NGUYEN [1977].

II induced by the inner product (9.9) the

THEOREM 10.1. Relative to the norm ||1


following contractivity property holds:
IIllx(-,t)-(-,t)lll <IllX0-xoll,

VtE .

(10.3)

PROOF. By hypothesis, the flows in (10.2) satisfy the variational inequality (9.5a) 3. In
particular,

A(, s-A(X,

) < A(Lr, S-

5),1

) < -A(tr,

(10.4)
)

Adding these inequalities and using bilinearity, along with definitions (9.3a)l and
(9.4)2 yields
A(-

, L-

) < A(

,H,
L -

) = (e[v-i],

- ).

(10.5)

Now observe that, for fixed (but arbitrary) t E 1, v(., t) - (., t) E V. Using the
linearity of (-,-) and the fact that the flows (10.2) satisfy the momentum equation
(9.5a) 2 gives
(a, E[V-

]) - (,

ey -

]) = -(po

-dK(v -

),

(10.6)

so that the right-hand side of (10.5) equals minus the rate of change in the kinetic
energy of the difference v - . Combining (10.5) and (10.6), along with the definition
for the natural norm, yields
dK(v -

) + ((E-

which implies dlHX -

)) =

allXX il

< 0,

(10.7)

tl[j/dt < 0. Therefore, for any t E 1[,

XIlixo- IIx0
- x0ol1 =

iIx(
~

,- ) - y(', T)II

dr

0,

(10.8)

which proves the result.


An immediate consequence of contractivity is the following uniqueness result for
the initial boundary value problem of infinitesimal elastoplasticity.
THEOREM 10.2. The stress and velocity fields in the solution of the initial boundary
value problem of elastoplasticity are unique.

226

CHAPTER I

J.C. Simo

PROOF. Suppose that X = (v, 27) and X = (, ) are two solutions of the system
associated with the same initial data Xo. Inequality (10.3) then yields 0 < lIX-Xl I
0 for all t C , which implies that X = X since IlI III is a norm.
The contractive property of the elastoplastic problem relative to the natural norm
IlI Il motivates the algorithmic definition of nonlinear stability employed subsequently.
REMARK 10.1. An identical contractivity result holds for classical viscoplasticity. The
corresponding initial boundary value problem is obtained by replacing inequality
(9.5a) 3 with (9.7). The flows t C I - 27 ECT and t E II E
C
associated
with the two initial conditions (10.1) then satisfy

A(,2 -)

< A(

-A(Z,2 - 2;)

tr,

--

+
-

<-A(,

[g(J()) - g(J(2))] drQ


)-7 /

[g(J(Z))-

(10.9)
g(J(2;))] dQ.

Adding these inequalities yields again (10.5) and the contractivity result for viscoplasticity follows exactly as in the proof of result (10.3).
A fundamental property of the elastoplastic flow closely related to contractivity is
provided by the notion of dissipativity, which translates into an a priori energy-like
stability estimate for the solutions of the initial boundary value problem. The total
internal free energy in the continuum body at any time t C II is given by the integral
Vint(, (E ) = /

TI(x, E,

,) d2.

(10.10)

By integration over the reference placement of the local (instantaneous) mechanical


dissipation SD at any (x, t) cE 2 x II one obtains the total dissipation in the continuum
Dr2 at time t as
DQ = Pint(v) - dtVint

(10.11)

DQ2

where Pi,, is defined by (1.18)2. The inequality


> 0 for all times t GE1 is accepted
as a basic principle (see TRUESDELL and NOLL [1965] for a detailed discussion).
A mechanical system for which D -0 for all t G II is said to exhibit no dissipation.
Otherwise, the system is called dissipative. From the foregoing discussion it is clear
that classical plasticity and viscoplasticity are dissipative systems while elastodynamics is not.
If the external loading is conservative with potential Vext: St

, the power Pext

of the applied loads satisfies the relation


Pext(v) = -

Vext()

= -VVext(U)
V.
dt~~~~~~~~~~~~~~~1.2

(10.12)

227

The classical models

SECTION 10

The total potential energy in the system is then given as the sum of Vext and Vit. The
total energy E in the system, obtained as the contributions of the potential energy and
the kinetic energy K defined by (18.1)1, then satisfies the following a priori decay
estimate.
THEOREM 10.3. The total energy E(u, v, se, E) = Vext(u) +Vint(E, 5) +K(v) decays
at the rate

d E(u, v, e,) = -D

<0

Vt

E1.

(10.13)

Thus, the total energy is a nonincreasingfunction along the flow.


PROOF. This result is an immediate consequence of the mechanical work identity
(1.17), since
dE(uV, e,
~dt
dt

)=

[K(v) + Vext(U)]

nt(e,

.)-et(u)]+ dt

Fd?-Pi(, ()v)+ d~,(ee,

(10.14)

The result follows from the definition of DQ in (10.11).


REMARK 10.2. For elastodynamics Dr 0 and (10.13) reduces to the familiar law
of conservation of the total energy. Inequality (10.13) is a formal a priori energy
estimate on the solution of the initial boundary value problem that captures its intrinsic dissipative character. The contractive and dissipative properties of the flow
are key qualitative features of the dynamics that should be inherited by any meaningful algorithmic approximation. Recent work in numerical analysis of dynamical
systems is aimed at constructing numerical algorithms that preserve key qualitative
feature of the dynamics. See SIMO and TARNOW [1992] and the review in SCOVEL
[1991] for the Hamiltonian case; FOIAS, JOLLY, KEVREKIDIS and Trri [1991], STUART
and HUMPHRIES [1993], and SIMO and ARMERO [1993] for representative numerical
treatments of dissipative systems.

CHAPTER II

Integration Algorithms for Classical


Plasticity and Viscoplasticity
The numerical solution of nonlinear initial boundary value problems in solid mechanics involving inelastic response is based on an iterative solution of a discretized
version of the momentum balance equations arising from finite element/finite difference procedures. In most of the computational architectures currently in use, the weak
form of the stress divergence term is evaluated via numerical quadrature and the evolution equations defining the inelastic constitutive response are enforced locally, as
a system of ordinary differential equations at each quadrature point. The integration
in time of this local system, typically stiff and subject to algebraic constraints, may
be regarded as the central problem of computational plasticity as it corresponds to
the main role played by constitutive equations in actual computations. A key feature
of this problem, illustrated in Fig. 10.1 and justified below, is that it can always be
regarded as strain driven in an algorithmic context in the sense that the state variables
are computed for a given deformation history.
The role played by the local integration of the plastic flow evolution equations is
illustrated in Fig. 10.1.
The evolution equations of classical elastoplasticity, as summarized in Table 3.1,
define a stiff differential algebraic system subject to unilateral constraints. By applica-

Strain Increment Aen(x)

({ n(z) En(),

An(Z)

>

RETURN
MAPPING
ALGORITHM

n+l(Z), EPn+l(Z), "n+ ()

FIG. 10.1. The role of elastoplastic return mapping integration algorithms. The stress tensor a is viewed
e
as a dependent variable, defined in terms of e by means of elastic strain relations, which (a) is computed
only at quadrature points and (b) is computed for prescribed total strain.

229

230

J.C. Simo

CHAPTER II

tion of a suitable integrator in time, the numerical integration of this system reduces
to a constrainedoptimization problem governed by discrete Kuhn-Tucker optimality
conditions. Below we describe the structure of this discrete problem, the fundamental
role played by the discrete Kuhn-Tucker conditions, and the geometric interpretation
of the solution as the closest-point projection in the energy norm of the trial elastic
state onto the elastic domain. In particular, because of the contractive and dissipative
properties of the initial boundary value problem, conventional higher-order methods
do not necessarily lead to improved numerical performance. These and related issues
involved in the design of numerical algorithms are motivated within the context of
the classical model of J2-flow theory, subsequently generalized to viscoplasticity. The
chapter concludes with a complete nonlinear stability analysis of the algorithmic problem, which exploits in a crucial manner the contractive and dissipative properties of
the continuum problem, along with a brief exposition of mixed finite element methods
for the elastoplastic problem.
The developments in this chapter provide a unified treatment of a number of
existing algorithmic schemes which, starting with the classical radial return algorithm of WILKINS [1964], have been largely restricted to J2 -flow theory. Representative algorithms include the extension of the radial return method in KRIEG and
KEY [1976] to accommodate linear isotropic and kinematic hardening, the midpoint return map of RICE and TRACEY [1973], the extension of these ideas to
the plane stress problem in SIMO and TAYLOR [19861, and alternative formulations of elastic-predictor/plastic-corrector methods given in KRIEG and KRIEG [1977].
To a large extent, return mapping algorithms of the type described below, or
"catching-up" algorithms in the terminology of MOREAU [1977], have replaced
older treatments based on the elastoplastic tangent modulus, as in ARGYRIS [1965],
MARCAL and KING [1967], NAYAK and ZIENKIEWICZ [1972] or HINTON and
OWEN [1980]. Because of the significant industrial applications, these methods
have received considerable attention over the last decade in the engineering literature.
The now standard return mapping algorithms for classical plasticity are based on
the backward Euler method. Attempts to generalize these techniques to higher-order
methods have found mixed success, see, e.g., ORTIZ and POPov [1985], SIMO and
GOVINDJEE [1988] and the follow-up work in CORIGLIANO and PEREGO [1991]. The
new schemes described below generalize these algorithms within the context of the
extremely convenient backward difference methods or ideas emanating from projected
implicit Runge-Kutta methods widely used for stiff problems.
11. Local integration of rate-independent plasticity: An overview
To appreciate the role played by numerical integration of the evolution equations of
classical plasticity and viscoplasticity, an informal outline is given below of the key
steps involved in a numerical solution of the elastoplastic initial boundary value problem. The local form of the problem of evolution to be integrated numerically is summarized next to set the stage for the general class of algorithms described in the subsequent sections. The feature that differentiates this problem from a conventional system

SECTION

Integration algorithms

231

of ordinary differential equations is the presence of an algebraic constraint, which is


defined by the Kuhn-Tucker conditions, leading to a differential algebraic system.
(A) The role of return mapping algorithms. The weak form of the momentum equations, supplemented by the history-dependent constitutive equations, comprise the
initial boundary value problem for classical plasticity. Current solution strategies in
computational inelasticity adopt a local formulation of the evolution equations that
define the constitutive model as a system of ordinary differential equations at each
point x e Q2. Solution schemes based directly on a weak formulation of the constitutive equation, although not common, are also possible; see, e.g., JOHNSON [1976b,
1977] or SIMO, KENNEDY and TAYLOR [1988]. The iterative solution procedure for
the elastoplastic problem involves two phases as described below.
Phase I. Spatial discretization of the domain of interest Q by a finite element
triangularization Tt. The solution space St and the space V of test functions are
approximated via finite-dimensional spaces to arrive at the discrete counterpartof the
weak form of the equilibrium equations. In doing so, one is confronted with evaluating
the finite element approximation to the weak form of the divergence term. The key
observation is that, at this stage, the stress field is prescribed via the computation
performed in Phase II.
(i) Consider the contribution of a typical element Qe computed via numerical
quadrature. From the quadrature rule it follows that knowledge of the stress field
is required only at the quadraturepoints of the element. This information is provided
by the local return mapping algorithm outlined below and allows the evaluation of
the element contribution to the weak form.
(ii) The contributions of all the elements are assembled via a standard assembly
algorithm into a global residual vector which depends nonlinearly on the displacement
field. If the norm of this residual vector is not below a specified tolerance, the system
is linearized about the current iterate and a correction to the displacement field is
computed.
Phase II. Time discretization of the interval of interest = Un= 0 [t", t,+l]. Within
a typical time step [t, tn+l] one is given as initial data, at the quadraturepoints of
a typical element Q2, the internal variables (eP , ~n). The key observation is that, at
this stage, the displacement field is prescribed at the current iterate found in Phase I.
(i) Consider a specific quadrature point x. of a typical element Q2e and evaluate
the current strain field E[uh+l] at x. E ?e via the chosen finite element interpolation,
for prescribed nodal displacements.
(ii) Evaluate at x. the stress field ,n+l (xz) and the internal variables via the return
mapping algorithms described below. The problem then reduces to the integration for
prescribed strain field of the differential algebraic system defined by the flow rule and
the hardening law at the quadrature points of each element in the triangulation.
Phase I is accomplished by the standard finite element method, see, e.g., CIARLET [1978] or JOHNSON [1987], and typically remains unchanged even if models
different from those described in the preceding chapter are employed. Phase II, on
the other hand, depends crucially on the particular constitutive model and constitutes

232

J.C. Simo

CHAPTER II

the central objective of this chapter. A conceptual algorithm that combines these two
phases proceeds as follows:
Within each interval [t, t+l] the nodal displacements ds of the triangularization
and the internal variables (Pn, E,) at the quadrature points of each element Q2e are
given. We then proceed with the following scheme to determine the nodal displacements dn+l at time t,+l.
Step 1. Initialization:Define an initial guess of the nodal displacement field d+
and compute the associated stress field at quadrature points (Phase II).
Step 2. Residual evaluation and convergence test: Compute the finite element approximation to the weak form for given iterate dn+ l and test the norm of the residual
for convergence (Phase I).
Step 3. Linearization and displacement update: If convergence has not been attained,
linearize the residual about current iterate d+l to compute an increment Adk+l.
Update the solution as d+l = d+l + d+l(Phase I).
Step 4. Local integrationof plasticflow: Compute the stress field ak+l at the quadrature points of each element in the elements of the triangularization via a local return
mapping algorithm. Compute algorithmic moduli for Step 3 above (Phase II). Return
to Step 2.
Convergence of the preceding incremental solution procedure can be easily established under the assumption of convexity. These issues have been recently addressed
in MARTIN [1988] and COMMI and MAIER [1989].
(B) The local strain driven differential algebraic system. To provide a concise statement of the local problem of evolution to be treated numerically, it proves convenient
to make use of the compact notation introduced in the preceding chapter and summarized below for convenience:
X= (,

EP

q),

(EP, ),

E=(e, 0),
and

G =DIAG[C,H].

(11.1)

For simplicity, attention will first be restricted to the case of a single surface elastic
domain E = {2: f(X) < O}. Let
N
X--

U [tntn+II
n=O
1r0

be a partition of the time interval of interest and xa: G


a fixed, but arbitrary point
in the reference placement, identified with a quadrature point in a finite element
triangularization. Assume that at time t, E and at points x E 2 (to be subsequently
identified with the quadrature points in a Galerkin finite element discretization) one

Integration algorithms

SECTION 12

233

is given the internal variables E P = (eP, ,) and the strain history {ek(x): k =
0, 1,...,, n + 1}. The problem to be addressed is the integration (for prescribed E) of
the system
= G[E - EP],

(11.2)

EP= fyVf(),
f(Z) < 0,

y > 0,

and

-yf( )= O.

The simple appearance of this problem is deceiving. The key feature that renders
matters nontrivial is the appearance of the additional algebraic constraint defined by the
Kuhn-Tucker conditions. Problems of this type are referred to as differential algebraic
systems. The monograph of BRENAN, CAMPELL and PETZOLD [1989] and the book of
HAIRER and WANNER [1991] provide excellent introductions to this extensive subject.
A first approach to the solution of (11.2) is to eliminate the multiplier y > 0 by
differentiating the constraint. This yields

ly

Vf(). GE
. GE
Vf(Z)
-f(V) GVf()

if Vf. GE > 0,

(11.3)

i.e., for plastic loading, and y = 0 otherwise. Substitution of this result then yields the
conventional rate problem already considered in the preceding chapter, which can be
treated by conventional methods. This is, in fact, the approach taken in early treatments
of elastoplasticity. The well-known drawback is that the algorithmic solution fails to
satisfy the constraint. Matters worsen considerably as time progresses.
Since the constraint need only be differentiated once to eliminate the multiplier, this
differential algebraic system is said to have index one; see BRENAN, CAMPBELL and
PETZOLD [1989]. Two basic techniques for the solution of stiff differential algebraic
problems are
(i) Linear multistep methods and, amongst them, the so-called backward difference
methods first applied to differential algebraic systems in GEAR [1971]. The lowestorder backward difference method is the backward Euler method: a very stable scheme
(it is simultaneously A-stable, B-stable and L-stable; see HAIRER and WANNER [1991]).
Backward difference methods are widely used for stiff differential algebraic systems.
(ii) Implicit Runge-Kutta methods. These are one-step methods which, in sharp contrast with linear multistep methods, are not subject to the so-called Dahlquist barriers
which limit to two the order of accuracy of A-stable linear multistep methods. In fact,
there exist implicit Runge-Kutta methods of arbitrarily high order which are A-stable;
see, e.g., HUNDSDORFER [1985] and HAIRER, NORSETT and WANNER [1993]. In spite
of their excellent stability properties, implicit Runge-Kutta methods are cumbersome
to implement and rather expensive compared to linear multistep methods.
As alluded to above, the key issue in the numerical integration of (11.2) is related
to the enforcement of the constraint. To gain insight into the issues involved in the
numerical approximation of (11.2), we consider a model problem that plays a central
role in classical plasticity.

234

J.C. Simno

CHAPTER 11

12. Motivation: The perfectly plastic model of J2-flow theory


To motivate the general class of algorithms presented in the subsequent sections, three
specific schemes are described in detail below within the setting of the perfectly plastic model of J2-flow theory. The goal here is to identify a number of key features
specific to the problem at hand, and to gain insight into a number of possible algorithmic strategies. Moreover, the early work of WILKINS [1964] and MAENCHEN and
SACKS [1964] on this classical model of perfect plasticity, leading to the development
of the radial return algorithm, constitutes the point of departure for modem treatments
of classical plasticity. It is for this reason that this setting is adopted as a model
problem to motivate alternative algorithmic schemes.
Set s = dev[oa], R =
of J2 -flow theory as

avy, and n = s/Isl to write the perfectly plastic model

s = str _ 2yn,

(12.1)

'y

f()=

sl - R

0,

and

yf(ao)=0.

Here, tr = 2 /i dev[] where dev[] is the strain rate deviator and t > 0 is the shear
modulus. Consider the following strain history, linearly increasing in time
dev[e(t)] = et for t > 0, with e = constant.

(12.2)

In what follows we will compare the performance of alternative algorithms developed


below to the exact solution of the differential algebraic system (12.1) for the strain
history (12.2). We remark that this solution also satisfies the full initial boundary value
problem under the simplifying assumptions of homogeneous deformation, negligible
inertia forces, zero external loads, and suitable displacement boundary conditions.
(A) The exact solution. Assume plastic loading so that f = 0 and st` * n > 0. Since
f = 0 we enforce the consistency condition f(a) = 0 and conclude that -y = eon > 0.
The system (12.1) then yields the incremental problem
s =

tr - (r

n)n

where

tr = 2/te.

(12.3)

Define the unit vector mo = e/lel and set = cos- (m 0 n). It is easily concluded
from (12.3) that the evolution of s takes place on a circle in the plane defined by
mo and the unit normal no = n(t)lt=o (see Fig. 12.1). Taking the dot product of
(12.3) with m 0 yields, after use of the identity sin 2 ( 0) = 1 - cos 2 (0), the differential
equation
--

2/ lel sin(O),
(12.4)

19(0) = 00 =

os - ' (no mo).

Integration algorithms

SECTION 12

235

nf
s,$so

S.=
1_111~m
n+l

(t)

(it)

FIG. 12.1. Perfectly plastic J2 -flow theory. (i) The radial return method, an example of a backward difference
method, and (ii) The two-stage mid-point rule, an example of an implicit Runge-Kutta method.

Integration by quadrature yields the closed-form expression


tan (9)

= tan(

90)exp [Is - so /Rl ,

(12.5)

where str = so + (2/te)t and so = s(O) is the initial stress at time t = 0 assumed to
be such that Isol = R. Setting to = m - cos(9o)no, the stress field s(t) becomes
s(t) = 2-tlel [cos(9 -

o)no + sin(0 - i9o)to].

(12.6)

Two key properties of this solution should be noted:


(i) As Is tr - sol

as time t
s(t)

oo we have

0. Equivalently, n becomes parallel to mo

coc. In view of (12.6), the stress field


-so

= Rmo

as Is

- sol -

o.

(12.7)

(ii) Recall that the instantaneous dissipation is D = - &P. Using the flow rule,
this gives D = 'ys *n = Ry, since by assumption Is ] = R for t E [0, o). Noting that
y = e

n = leln mo, the total dissipation in a monotonic plastic loading process

for t C [0, T] becomes


D[O,TJ =

hence,

Ddt = Rlel

9(T)

cos(d) dO = Rlel [sin (o)

lim D[O,T] = D[o,o] = Rjel sin(0o).

- sin ((T))],
(12.8)

Since the complementary Helmholtz free energy during plastic loading remains fixed
at its initial value ,2/2pu, the stress power is converted entirely into mechanical
dissipation leading, therefore, to a strongly dissipative process.

236

J.C. Simo

CHAPTER

From the preceding properties one concludes that soo defined by (12.7) is a steadystate equilibrium point that attracts all the trajectories associated with the loading e(t)
regardless of the initial condition so at time t = 0. This property is consistent with
both the dissipative and contractive properties of the elastoplastic flow, described in
the preceding chapter.
(B) ALGO 1: Backward Euler method. Consider the approximation of (12.1) by
the lowest-order backward difference method, the backward Euler scheme, and set
Ae = Ate to obtain the algebraic system
Sn+1 = s

(12.9

+ 2 Ae,

Sn+l = Sn+

-2/1 Ay nn+l,

where nn+l = sn+l/Sn+l and the multiplier Ay > 0 is to be determined by enforcing the constraint f(Un+l) =
Sn+lI
- R = 0. To do so, observe that (12.9)2 can be
written as
[In+ll-I 2[tA'y]n,+i = s+ 1.
Since A-y > 0,
conclude that

Isn+1 = R, and f(en+l


1)

nn+l =Sn+
l
Stn+ II

and

2tAy

(12.10)
=

Is,+l - R > 0, from this expression we

St+II-

R>0.

(12.11)

Substitution of these expressions into (12.9) yields the following result:


Sn+l= [R/Ist+ll]s+l,

(12.12)

which defines Sn+l as the radial scaling of the trial stress sn+l onto the von Mises
circle sI = R; see Fig. 12.1. This is the classical radial return method of WILKINS
[1964] and MAENCHEN and SACKS [1964]. This algorithm is of course consistent, i.e.,
Is, - s(tn)I

0 as At --+ 0 and first-order accurate. In spite of its restriction to

first-order accuracy, two features make this algorithm remarkable:


(i) Let s, = so be the initial data, denote by sat = Rnat the stress computed

with the radial return method defined by (12.12) in a single step At = t and let
Oat = cos
I (nAt
9
lim O
At = 0

At-oo

m 0 ). From (12.12) it is easily concluded that


and

lim Sat = Rmo = s.

(12.13)

At-oo

Therefore, the radial return method yields the exact solution as At - oo.
(ii) Let DAt be the total dissipation in a single step At during plastic loading. In
view of expression (12.8) and (12.13)1 one concludes that
lim DAt = Rlelsin(i9o) - At-oo
lim=

At-oc

Rel sin()]
sin(00).

(12.14)

237

Integration algorithms

SECTION 12

Therefore, the radial return method yields the exact value of the total dissipation in
the process as At - oo.
These properties imply that the long-term behavior of the radial return method is
identical to that exhibited by the exact solution, since the exact steady-state equilibrium
point s, is also an attracting point for the algorithmic solution. This example illustrates that the radial return method inherits the dissipative and contractive properties
of the continuum problem, a fact rigorously proved for the general problem below.
(C) ALGO 2: Two-stage mid-point rule. To improve upon the first-order accuracy of
the backward Euler method, consider the approximation of (12.1) via the implicit midpoint rule. To account for the constraint this single-step method is cast into the form of
a two-stage projected implicit Runge-Kutta method, as follows (see Fig. 12.1). Set
Sn+1/2 = Sn + 2[2
2

e],

Sn+ = Sn +

2/Ae,

(12.15a)

and consider the two-stage formulae


Sn+1/2 = Sn+1/2S'+

t + -

A'Y 2tnn+l/2

where nn+l/2 = Sn+/2

y2 2unn+/2

(12.15b)

iSn+1/21

with the multipliers Ayl, Ay 2 to be determined by enforcing the constraints


f(o'+1/2) =

sn+l1/21 -

R = 0 and f(on+l) = 1sn+1I - R = 0.

(12.16)

This ensures that the intermediate and final stages lie on the manifold lE = {s: Is R = 0}. Since the first stage is identical to the backward Euler method with half the
step-size At, the preceding analysis gives
nn+l/2 = Sn+l/2/lsn+l/21

and

2/A/

= I Sn+1/21 - R,

(12.17)

provided that f(s~l/2 ) > 0. Otherwise, set A-yl = 0, nn+l/2 = s


t+l/lss+iand
proceed to the second stage to determine AY2. For the case f(sn+l/2) > 0 we determine A'Y2 as follows. Take the dot product of the second stage in (12.15a) with
nn+l/2 to obtain
[(Sn+l -

n)

nn+l/2 - 2]LAy 2] = (Sn+l - Sn)

nn+l1/2

Then observe that the second stage in (12.15a) along with (12.17) implies that
is parallel to (sn+l + sn), since
sn+l + Sn =2[sn +

(12.18)
nn+l/2

(2,uAe)] - 2/LAy 2 nn+l/ 2

= [21Sn+1/2 - 2Y

2]

nn+t/2 -

(12.19)

238

J.C. Simo

Finally, use the identity (s+l - s,,)

CHAPTER 11

(sn+i + sn)

[sn+-l
-

-]s,,

2]

to conclude

that the right-hand side of (12.18) vanishes, thus arriving at the following expression
for the multiplier Ay 2 in the second stage:
2/tAy2

(sn+l - Sri)

trtr
[
Str+1/
2 /IS+l/21

(12.20)

In summary, Eqs. (12.17) and (12.20) provide an explicit solution of the two-stage
method (12.15b). This scheme is obviously consistent, second-order accurate and ensures that the final stage s,+ l is on the admissible domain. It would appear, therefore,
that the final stage s,+I computed with this scheme gives an improvement over ALGO
1. Remarkably, this is not the case if one is interested in the long-term behavior of
the system as the following observations show.
(i) Again take Sn = so as the initial data and denote by sat the algorithmic
solution computed with the two-stage mid-point rule in a single step At = t. Property
(A) of the backward Euler method implies that limsAt/2 = Rmo. However, since
SL+1 +1/2 = Sn Sn+1/2, the second stage of the algorithm yields
lim sat so = R2 cos(2090)

At

while s,

so = R 2 cos(19().

(12.21)

This means a 100% error in the angle between the final stage sat and the initial data
so. The intermediate stage sat/2, however, yields the exact solution.

(ii) Similarly, denoting again by DAt the total dissipation in a single step At computed with the two-stage mid-point rule via the exact formula (12.8), in view of (12.21)
one obtains
(12.22)

lim DAt = Riel [sin(o0) - sin(-00)] = 2RLel sin(90o),

i.e., a 100% error relative to the exact solution.


The formulation of this algorithm as a one-stage method was originally introduced in
RICE and TRACEY [1973] in the context of J2 -flow theory and subsequently generalized
in ORTIZ and PoPov [1985].
(D) ALGO 3: Projected mid-point rule. The second stage in ALGO 2 involves an
oblique projection of the trial state st+,+ onto the von Mises surface. However, such
an oblique projection cannot be guaranteed to exist for arbitrarily large time steps if
a general model of plasticity is considered. An alternative scheme that circumvents
this difficulty, while retaining the second-order accuracy of the preceding algorithm,
can be constructed as follows (see Fig. 12.2).
Step 1. Define the first stage of the method exactly as in ALGO 2; i.e.,
str+/2 =
S+1/2

s,
_

[2IAe]

n+-l/2 -- ATY 2n+/2

with ni+~l/2
=n2[

S+1
-

n+1/21

,~n+1/2' (12.23)

SECTION 12

239

Integration algorithms

eI

II

(t)
(tt)

FIG. 12.2. Perfectly plastic J2-flow theory. (i) Illustration of the projected mid-point rule, (ii) Long term
asymptotic behavior as At - oc.

AYi 1 > 0,

f(oa,+l/2) = ISn+l/2l -

R < 0,

and

AyIf(On+ll/2) = 0.

As before, this ensures that the intermediate stage lies in the elastic domain E =
{s: FSI - R O}.
Step 2. Extrapolate the solution Sn+l/2 in the first step via the mid-point rule formula s,+l = [2Sn+1 /2 - Sn], set s+, = sn+l and define s,+l as the closest-point
projection of this extrapolated value onto the elastic domain; i.e.,
St+

sn+l = n+l -A'y2

Ay72 > 0,

with n,+l =

Sn+1/2S -

2nn+

f(-.+l) =

,+

sn+lI

Is+,I - R < 0,

and

(12.24)

Ay 2f(On+l) = 0-

Again, this ensures that the final stage s,+l lies in the elastic domain E.
For the von Mises yield criterion, both steps can be solved in closed form. The
solution to Step 1 is given by formulae (12.17), under the assumption of plastic
loading, while the solution to Step 2 is given by
rl+l l/ and

2,tA

(12.25)

In contrast with ALGO 2, stage 2 in the extension of ALGO 3 to general models


of plasticity is guaranteed to always have a solution for arbitrarily large time steps.
However, the long-term behavior of ALGO 3, although improved relative to ALGO 2,
is not optimal when compared with that exhibited by ALGO 1. Figure 12.2 shows the

240

J.C. Simo

CHAPTER II

asymptotic solution obtained by letting At --+ oo. Since the justification of this result
uses the same analysis as in ALGO 2, further details are omitted. ALGO 3 was introduced in SIMO [1992] improving upon earlier work of SIMo and GOVINDJEE [1988].
In summary, if one is interested in short term accuracy, the second-order accurate
ALGO 3 is optimal. If, on the other hand, long-term is all that matters, the less accurate
ALGO 1 becomes optimal. These results, although obtained in a rather simple setting,
are representative of the actual performance to be expected in more complex situations.
In KRIEG and KRIEG [1977], for instance, iso-error maps obtained for a wide range of
strain increments show consistently superior performance of the radial return method
over schemes possessing higher accuracy, as measured by the local truncation error.
Similar results, obtained by SCHREYER, KULAK and KRAMER [1979], YODER and

WHIRLEY [1984] and others, are reviewed below.


REMARK 12.1. The striking conclusion to be drawn from the preceding analysis is
that, for dissipative systems, higher accuracy as measured by the local truncation error
does not necessarily imply improved long-term performance. For the incompressible
Navier-Stokes system, for instance, SIMO and ARMERO [1992] have recently shown
that the backward Euler method preserves the dissipativity of the dynamics, as characterized by the presence (in two dimensions) of absorbing sets and a universal global attractor, see, e.g., TEMAM [1988] for a comprehensive overview of infinite-dimensional
dissipative systems. In fact, for the incompressible Navier-Stokes equations, the backward Euler method again exhibits optimal long-term properties while the higher-order
accurate mid-point rule does not.
13. Backward difference and implicit Runge-Kutta methods: Basic results
As a prelude to the generalization of the ideas presented in Section 12 a brief summary
will be given of a number of stability results on classical backward difference and
implicit Runge-Kutta methods.
Backward difference methods are a particular class of linear multistep methods
widely used in the numerical integration of stiff ordinary differential equation systems, originally introduced in CURTISS and HIRSCHFELDER [1952] and popularized in
the work of GEAR [1971] on stiff differential algebraic systems. The brief overview
given below is aimed at summarizing the main properties relevant to the algorithmic
treatment of plasticity. Consider the standard initial value problem

z = f(Z,t),
Zt=o = Z0.

Let {z,,+l-j: j = 1, .. , s} denote s given approximations to the solution z(t) of


(13.1) at times t,+l-j, j = 1, .. ., s. Recall that a linear multistep method of order s
defines the algorithmic approximation z,+l to the solution z(t,+l) by the formula
S

EjZn+,-j
j=0

= AtEZjf(zn++ljtnlj)
j=0

(13.2)

Integration algorithms

SECTION 13

241

The coefficients (aj, /3j) E R x IR define the specific method. Associated with (13.2),
one defines the companion one-leg method by the formula
ajZn+lj

Atf( E

j=0
o

jZn+l-j,

j=o

tn+l-j

(13.3)

j=o

The two methods (13.2) and (13.3) are identical for linear problems. For nonlinear
problems, there is a one-to-one mapping between the solutions computed via (13.2)
and (13.3) (DAHLQUIST [1975]). It is conventional to introduce the following notation
for the polynomials defined by the coefficients of the above methods
s

p(() = Z

s- j

and

)=

3j(c - j

j=o

for

E C.

(13.4)

j=o

The algebraic equation p(() -/ a() = 0 for p C C is known as the characteristic


equation of the method. In general this equation has k < s roots i(z) E C with
multiplicities mi(/p) > 1 such that EkI= mi(u) = s. The set

S=

cC: (i(L)

1 if mi(l)

1 and I i(p)I < 1 if mi(I) > 1}

(13.5)

is known as the (linearized) stability region of the method for reasons summarized
below. An example of a linear multistep method is the trapezoidal rule (for cl =
-a2 = 1, 31 = 322 =),
with the implicit mid-point rule as the companion one-leg
method.
An s-step backward difference method is a one-leg method (or a multistep method
for that matter in view of the definition given below), with coefficients defined according to the formulae:

o= O,

30 = 1,

ojZn+l-j =

DJz,+l,

and

fj = 0,

j > 1,
(13.6)

where Dj denotes the backward difference operator in time, as defined by the following recurrence relation:
DJ+lZn+l= DjZ,+

- DJz,

where Dz = z.

(13.7)

From the preceding formulae, one concludes that the one-step and two-step backward
difference methods are given by the following explicit expressions:
s = 1: Zn+ - Zn = Atf(zn+l, tn+1),
S=

2: 2Zn+l

2z, + 2Zn- =

tf(z+lI

tn+l).

(13.8)

242

J.C. Simo

CHAPTER

The feature that in the present context renders backward difference methods particularly attractive is the need for only one evaluation of the right-hand side of (13.1).
It will be shown below that this property results in an extremely convenient procedure for the enforcement of the algebraic constraint in classical plasticity. An s-step
backward difference method has order of accuracy p = s and possesses the following
linearized stability properties.
(i) A-stability (DAHLQUIST [1963]). Recall that a method is A-stable if the algorithmic solution {z,, E C: n E N} in the complex plane of Dahlquist's model equation
=z )z, A G C, has the property
lim IZnl = 0

(13.9)

for Re(A) < 0.

A-stability is equivalent to the requirement that the left-half complex plane be contained in the stability region S defined by (13.5). It can be shown that both the firstand second-order backward difference methods are A-stable. Higher-order backward
difference methods cannot, of course, be A-stable since Dahlquist's second barrier
(DAHLQUIST [1963]) limits to p = 2 the maximum attainable order of accuracy of
an A-stable linear multistep method. Backward difference methods progressively lose
their good stability properties with increasing order of accuracy. It can be shown that
for s > 7 the methods are unconditionally unstable.
(ii) L-stability. Recall that a one-step implicit Runge-Kutta method is said to be
L-stable if it is A-stable and the algorithmic solution R([t) to Dahlquist's model
problem z = Az, known as the stability function of the method, has the property
limd0 R( ) = 0, where = AAt. To generalize this property to linear multistep
methods one writes the algorithm for Dahlquist's model problem as
Z,+ = A(Pt)Zn

with Z, =

[zn

zn-!

...

(13.10)

+t

where A is the amplification matrix whose characteristic polynomial coincides with


the characteristic polynomial p() - cr( ) of the method. Let ps[A(/)] be the
spectral radius of A(Xi). Then, an A-stable linear multistep method is L-stable if
lim.o ps [A(/l)] = 0. L-stability implies the asymptotically annihilating property
lim o,, zn+l/z = 0. Both the backward Euler method (s = 1) and Gear's two-step
backward difference method (s = 2) can be shown to be L-stable.
Strictly speaking, the preceding two notions of stability are formulated within the
realm of linear problems. A possible extension of the notion of numerical stability to
nonlinear problems is based on the concept of contractivity. Suppose that z(t) and
z(t) are two solutions of the initial value problem (13.1) corresponding to two initial
data zo and o, respectively. The initial value problem is said to be contractive if,
relative to a certain norm 1- called the natural norm of the problem, the following
estimate holds

Ilz(t 2 ) -

(t2 )j

J|z(t)-

(tl)||,

Vtl,t 2

et with t2

t,.

(13.11)

Integration algorithms

SECTION 14

243

This notion is especially relevant to elastoplasticity since this problem is contractive


relative to the norm induced by the complementary Helmholtz free energy function,
a fact proved in the preceding chapter.
(iii) B-stability. Suppose that the initial value problem (13.1) is contractive in the
sense of the estimate (13.11). Let {zn} and {,} be two sequences generated by a
given algorithm for initial data zo and io, respectively. The algorithm is said to be
B-stable if it inherits the contractivity property (13.11) relative to the natural norm

* Il,i.e.,

IIZn+-

-Z

Vn e N.

(13.12)

It can be shown that the backward Euler method is B-stable but other backward
difference methods, and in general all linear multistep methods, are not. The notion of
B-stability was introduced by BUTCHER [1975] in the context of implicit Runge-Kutta
methods. A complete characterization of contractive Runge-Kutta methods is due to
BURRAGE and BUTCHER [1979, 1980] and CROUZEIX [1979]. It can be shown that,
for Runge-Kutta methods, B-stability implies A-stability, but not conversely.
(iv) G-stability. The notion of B-stability is too restrictive for linear multistep
methods and, historically, was motivated by the concept of G-stability introduced
in DAHLQUIST [1975]. An s-step linear multistep method defined by (13.2) is said
to be G-stable if the companion one-leg method, defined by (13.2), inherits the contractive property (13.3) relative to an algorithmic norm induced by a suitable s x s
positive definite matrix G. Accordingly, if the natural norm is induced by the inner
product (., .), the algorithmic G-norm is given by
S

=1S

IZ

(13.13)

j(Zn+l-i, Zn+l-j)
gi

i=l j=1

for some positive definite matrix G = [gij], and the G-stability condition for a linear
multistep method takes the form
I1Zn+1- Zn+1IlG < IIZn - ZnlG,

Vn E .

(13.14)

The backward Euler method (s = 1) is obviously G-stable since it is B-stable and


therefore contractive relative to the natural norm. The two-step (s = 2) backward
difference method can be shown to be G-stable with
G =

12];

(13.15)

see HAIRER and WANNER ([1991], p. 332) for a direct verification of this result.
Remarkably, in spite of the broader scope of the notion of G-stability, a fundamental
result of DAHLQUIST [1978] shows that G-stability is equivalent to A-stability (under
the mild assumption that the polynomials p(() and a(() have no common divisor).
Therefore, G-stable linear multistep methods can be at most second-order accurate.

244

J.C. Simo

CHAPTER 11

14. Generalized backward difference return mapping algorithms


The first generalization of the now classical return mapping algorithms for classical
plasticity described below is based on the use of backward difference methods. From
our preceding discussion, two schemes are of interest for the problem at hand: (i) The
backward Euler method (s = 1), a first-order accurate, B-stable and therefore A-stable,
L-stable scheme and (ii) Gear's two-step method (s = 2), a second-order accurate,
G-stable and therefore A-stable, L-stable scheme. Backward difference methods of
order s > 2 are no longer A-stable and therefore of little interest, while methods of
order s > 7 are unconditionally unstable. Accordingly, only schemes of order s = 1
and s = 2 are considered in what follows.
Suppose that at time t the history {E ,... , EP+l_s} of the internal variables is
given. Then, by applying a backward difference method for prescribed strain field
E,+l = (en+l, 0), one arrives at the nonlinear algebraic problem

DjEP+ = AyVf (En+')

where En+t = C [En+ - E+],

(14.1)

j=l

supplemented by the following discrete counterpart of the Kuhn-Tucker conditions


A-y > 0

f(n+l) < 0,

and

A-yf(nl) = 0.

(14.2)

As in the continuum case, the Kuhn-Tucker conditions (14.2) define the appropriate
notion of loading/unloading. These conditions may be reformulated in a form directly
amenable to computational implementation by introducing the trial elastic state. Consider first the case s = 2 and rewrite (14.1)1 as
(E+l

EP )

(E

- En

) = A/Vf(+l).

(14.3)

Multiplying this equation by and noting that it is permissible to redefine the multiplier Ay as 2A-y for the case s = 2, the algorithmic flow rule becomes
s = 1: En,+ - E =

Vf(-E,+),
(14.4)

s=2:E+

- E

=-Vf(X,,+l)
A

+ (E -

p
En

Then define the trial elastic state by the formula

= EP, + (s-1)[EP - E n I]
+E4= G
3+
l)Enl - E i
t 1 = G[En+1 - E]

E,+n

f-12.14
for s = 1,2.

(14.5)

Integration algorithms

SECTION 14

245

With this definition the algorithm can be written in a unified form, valid both for
s = 1 and s = 2. Relations (14.1) and (14.2) then imply
.n+l = t.+1

- AYGVf( n+l),
1

f(n+l) < 0,

Ay > 0,

and

(14.6)
Ayf(

) +l

= 0.

From a physical standpoint the trial elastic state is obtained by freezing plastic flow
during the time step. Observe that only function evaluations are required in definition
(14.5). From an algorithmic standpoint, a basic result is the fact that loading/unloading
conditions can be characterized exclusively in terms of the trial state, provided that the
yield criterion is convex. This is the key result needed for the computational statement
of the loading/unloading conditions.
THEOREM 14.1. (i) If f: S x RInn -- Rt is convex then f(G +l)
>
f(Z+l), moreover; (ii) loading/unloadingis decided solely from f (Z+l) according to the conditions
f(n+l) <0

elastic step

Ay = 0,

f(LZ+1) >0

plastic step X Ay > 0.

(14.7)

PROOF. (i) The convexity assumption on f(.) implies


f(E 1+,) - f((n+l) > [+
From (14.6) we have
(14.8) yields

(14.8)

n+l]Vf(n+).

n+1 = Z'+ -AyGV7f(

+l). Substituting this relation into

GVf (En+)]

f (~+,) - f (n+,) > y [Vf (En+,+)

0,

(14.9)

since Ay > 0 and G is assumed positive definite.


(ii) First, if f(Z+) < 0, then by the last result it follows that f(E,+1) < 0. The
discrete Kuhn-Tucker condition A-yf(Zn+l) = 0 then implies Ay = 0. Thus,
El

= En +

(S-1

) (EP-EP

for either s = 1 or s = 2, and the process is elastic. On the other hand, if f (,r+ 1) > 0,
then etr , cannot be feasible, that is,
', + E'+. Thus, we must have A-y + 0. Since
Ay cannot be negative it follows that Ay > 0. The discrete Kuhn-Tucker condition
tAf( En+,) = 0 then implies f(,+ 1) = 0, and the step is plastic.
The geometric interpretation associated with algorithm (14.6), under the assumption
that the elasticity tensor C is constant, is entirely analogous to that associated with
the continuum problem and provides the key to its numerical implementation.

246

J.C. Simo

CHAPTER II

FIG. 14.1. Geometric illustration of the backward difference closest-point projection algorithm for perfect
plasticity and s = 1 (backward Euler).

THEOREM 14.2. The solution xZ,+1 of algorithm (14.6) is the closest-pointprojection


of the trial state ~n+l onto the boundary aE, of the elastic domain in the norm

induced by the metric G. Equivalently, the minimization problem

[Zn+t -

min

]. G - l [nr'+

]: for

E,

(14.10)

yields the solution Z+1 of the backward difference algorithm (14.6) for prescribed
x +I computed via (14.5).

PROOF. The result follows merely by noting that the Lagrangian associated with the
constrained minimization problem (14.10) is
(,

7y) =

[r+-

G-yfQ()

] +

(14.11)

The corresponding Kuhn-Tucker optimality conditions are given by

G -.

+ + n+] + ATyVf(En+I) =0,


(14.12)

f(,n+i) < 0,

Ay > 0,

and

Ayf(~,+i) = 0,

which coincide with the algorithmic equations (14.6).


To summarize the preceding developments, the steps involved in the local integration of the constrained differential algebraic system in a typical interval [t, tn+l ] via
a projected backward difference method are summarized in Table 14.1.

SECTION 15

247

Integration algorithms
TABLE 14.1

General backward difference return mapping algorithm.


(1) Given the history {EP,..., EPn+l_s} (s = 1, 2) for the generalized internal
variables (at quadrature points) compute:
Et+

= EP + 3(s- 1) [

Ep -

EPI]

for either s = 1 or s = 2.

(2) For prescribed strain En+l = (E+l,0) compute the trial state via the
generalized stress-strain relations and evaluate the yield criterion:
tr+ = G [En+l-E

pt r

and

f+=

f(

+l)

(3) Test the trial state and perform a closest-point projection:


If fr+
I < 0 then set E,+ = Er+l and EP+I = Ep t
Otherwise, compute the closest-point projection:
Find E.n+ C E such that J(E+l)= min{J(X): E C E},
where J(E) =
- ES] G- [E+r + .
Update generalized strains as EP+ I = En+ -G-lEn+l

15. Generalized implicit Runge-Kutta return mapping algorithms


The second generalization of the classical return mapping algorithms for classical
plasticity described below uses ideas from projected implicit Runge-Kutta methods.
Although it is possible to construct one-step methods with arbitrarily high order of
accuracy by considering several intermediate stages, attention will be restricted to
schemes possessing at most second-order accuracy. The goal is to extend ALGO 3,
introduced in the context of J2 -flow theory, to general models of classical plasticity.
Suppose that one is given at time t, the generalized internal variables E p =
(EPn, n), the strain field e,, and a strain increment AE so that En,+ = E,, +Ae. Motivated by the structure of ALGO 3 in Section 12, consider the following interpretation
of the generalized mid-point rule:
Step 1. By definition, the generalized stress at time t, is ,n = G[En - EP]. Set
AE = (Ae, 0) and define the intermediate step at time t+ by the formulae
.Etr+ = E,

n+@ = n+

A'Y > 0,

+ 9GAE,

- A'yGVf (n+),

f (E.+) < 0,

and

(15.1)

A-yf(E,+4) = 0.

Here 9 E (0, 1] is an algorithmic parameter.


Step 2. Extrapolate the generalized stress E,+o at time t,+, to time t,+l by the
generalized mid-point rule formula
T+I

= [n+9 - (1 -

(15.2)
()1n]

J.C. Simo

248

CHAPTER II

FIG. 15.1. Geometric interpretation of the return mapping algorithm based on the projected generalized
mid-point rule.

TABLE 15.1
Projected generalized mid-point return mapping algorithm.
(1) Given the generalized internal variables Ee = (e', -n)
points) and the strain increment AE = (Ae, 0), compute:
E

= GEn,

Eent+

= Ee + OAE,

and

r+,,

(at quadrature
= GEe ' 9

(2) Perform a return mapping (s = 1) to compute the intermediate stage:


<
If f(~Et+o9)
0 then set
,n+
t o = Et7+. o
Otherwise, compute the closest-point projection:
Find .Z,+o e E such that J(E,+O) = min{J(LZ):
te E},
where J(E) = [Ent+ - E] *G-' [9+,
J.
(3) Define tr+l
[Tn+, -(1-iO)n ]/O by linear extrapolation, and perform
a return mapping (s = 1) to compute the final stage:
If f(0tr+) S O set E,+ =
and Ee+l =
t+l. rGOtherwise, compute the closest-point projection:
Find r7,~+ t E such that J(F,+) = min{J(/): E E E},
where J(_E) = [tr +1 - ] G-[Etr+l - ].
Update generalized strains as En+ = G-1,+l.

The linearly extrapolated value Z,+ does not, in general, lie within the elastic domain
E. In fact, the convexity assumption on f(-) implies that
f(

,n+l)

[f(Z+o) - (1 -

)f(Zn)]/.

Hence, f(Zn+l) > 0 in sustained plastic loading since f(Z,) = 0 and f(Z+;)
Step 3. Define the final stress state
extrapolated value

7,+l; i.e.,

(15.3)

= 0.

,+1 exactly as in Step 1, with trial state the

tr=

249

Integration algorithms

SECTION 16

-n+l,

,+l=Et+r --

Aj' > 0,

AyGVf (7, +1),

f(Ln7+l) < 0,

and

(15.4)

Aj'f(.n+I) = 0.

This ensures that X,+1 lies in the elastic domain E = {Z: f(27) < O}.
A detailed accuracy and stability analysis of this algorithm is deferred to the last
section of this chapter. Here, we remark that second-order accuracy is attained if 19=
and B-stability holds for 9 . The scheme can be interpreted in terms of closestpoint projections exactly as the backward difference methods described above (see
Fig. 15.1). A summary of the steps involved in the implementation of the algorithm
is given in Table 15.1. For the subsequent extension of this scheme to the finite strain
theory we have replaced EP by the elastic strain e = e - EP and introduced the
alternative set of generalized history variables E e = ( e , _-)*
16. Algorithms for the computation of the closest-point projection
The key step in the implementation of the class of algorithms described in the preceding two sections lies in the computation of the closest-point projection of the trial
state onto a convex set, the elastic domain E, in the metric induced by the generalized moduli G- 1. This is the standard problem in convex optimization. Two possible
solution strategies are outlined below.
(A) The general closest-pointprojection algorithm. This procedure, first introduced
in SIMO and HUGHES [1987], boils down to a systematic application of Newton's
method to the system of equations (14.6) to compute the closest-point projection from
the trial state onto the yield surface. A geometric interpretation of the iteration scheme
is contained in Fig. 16.1. To accommodate the algorithms summarized in Tables 14.1
and 15.1, it is implicitly understood that 9 = 1 for the first- and second-order backward
difference methods, whereas 19 (0, 1] for the projected generalized mid-point rule.
Step 1. Initialization. Assume plastic loading, i.e., f(n+o9) > 0
k = 0, Z(

- Ay > 0. Set

,9,

and Ay( ) = 0.

Step 2. Residual evaluation and convergence test. For current values


Ay(k) at iteration k > 0 evaluate the residual and the yield condition as
R(k)
-G-Tstr
n+19G-

_(k) -

-is

(k)y

_V
(k)
n+

(k) and

),

(16.1)
~ff = (ere+)

f n+If < TOL, where TOL is a prescribed tolerance, convergence has been attained.
Then set (,+9,AT)

= (

k)+,

-(k))
and terminate the algorithm.

250

J.C. Simo

CHAPTER II

FIG. 16.1. Geometric interpretation of the closest-point projection algorithm in stress space. At each iterate
(.)(k) the constraint is linearized to find the intersection (cut) with f = 0. The next iterate (.)(k+l),
located on level set f(kL+ l) = 0, is the closest-point of that level set to the previous iterate ()(k) in the
metric defined by the complementary Helmholtz free energy function.

Step 3.

Linearization. If fko > TOL, linearize the residual about the current iterate

(,(k9,Ay(k)) to obtain the constrained linear problem


-DEk (

(k (0,

(k))Vf(

)]

(16.2)

)+ * D

Vf (+f
where d(+k

o + n+f =

= [G- + D(A-y(k))V2f(ZUnk+)]_

is the tensor of algorithmic moduli.

Step 4. Solution of the linear system and update. Solve the linear problem by taking
the inner product of (1 6 .1)1 with Vf(4k,+ ) and using (16.2)2 to obtain
f(k)
D(Ay(k)) =

Vf(
f+

n+0
n+

k))

n+

_ n+f

n+

(k) R(k)
n+f
7+)

(16.3)
DP(")

~(kr)

[R

- D (-

Then update the current iterate by setting


(k+l) =

, (k)

r(k

and

A? ( k+ l) =n_ A/y(k)
D(An+,(k)).and
3 n+, +
+D

Increment the counter k = k + 1 and return to Step 2.

(16.4)

Integration algorithms

SECTION 16

251

Once X,+,g and A-y are computed, the internal variables are updated by inverting
the stress-strain relations, as indicated in Tables 14.1 and 15.1. Observe that the
backward difference formulae enter only in the computation of the trial state E+, for
the internal variables. The rest of the algorithm is independent of the specific scheme
adopted for the computation of the trial state. For purposes of comparison with the
alternative scheme described below, it is useful to summarize the basic characteristics
of the preceding Newton iteration.
(i) This scheme is an implicit procedure that involves at each step the solution of
a local system of N x N equations, where N = 4ndim(ndim + 1) + nint.
(ii) The normality rule (i.e., the associative character of the plastic flow) is always
enforced at the final (unknown) iterate.
REMARK 16.1. Convergence of the method is guaranteed since the problem is convex.

The scheme outlined above is trivially modified to include a line search procedure that
renders the iteration globally convergent, see, e.g., LUENBERGER [1984]. Given the
small size of the problem, Newton's method becomes extremely competitive relative
to alternative iteration schemes.
(B) The cutting plane algorithm. The main drawback associated with the closestpoint iteration procedure described above is the need for computing the first and
second gradients of the yield criterion. This task may prove to be exceedingly laborious
for complicated plasticity models. The objective of the iterative algorithm described
below, originally proposed in SIMo and ORTIZ [1985] and further analyzed in ORTIZ
and SIMO [1986], is precisely to circumvent the need for computing these gradients.
The algorithm falls within the class of convex cutting plane methods for constrained
optimization, see LUENBERGER ([1984], Sections 13.6 and 13.7).
Assuming plastic loading, i.e., f( ,+ ) > 0 so that Ay > 0, the key idea is to
approximate the closest-point projection by a gradient flow constructed as follows.
Step 1. Initialization.Set k = 0 and

(O)9

Z.t

Step 2. Steepest descent. Consider the hypersurface S = {(E, z): z = f ()}.


Given Ek+ the intersection of the hyperplane Pk +l tangent to S at (k, with the
normal section of S at EX(k) is the straight line of steepest descent on the tangent
hyperplane pk+1 The parametrization

x() =

z(,)

n+
f((k)
J-

(GVf (Y

Z()-- sn+)

+ Vf

),n+

Vs\z n+
+)

.) _

k](k)

9c

(16.5)

defines a point (,(), z()) on the line of steepest descent on Pk+l forC ( [, oo).

252

J.C. Simo

CHAPTER II

Elastic
Domain
tt
r

... -swa

FIG. 16.2. Geometric interpretation of the cutting plane algorithm in stress space. At each iterate ()(k) the
constraint is linearized about ()(k). The intersection of the plane normal to f(k) = 0 with the level set
f(k+l) determines the next iterate ()(k+l).

Step 3. Convergence check and update. The intersection of the steepest descent line
defined by (16.5) with the plane z = 0 is obtained for the following value of the
parameter ~ which defines the updated iterate:

) ((k

[Vf(Z+)

(k+1)and
(n+).
GVf(~o)]=

U((+). =

(16.6)

If f(Lk+ ))
> TOL, where TOL is a prescribed tolerance, set k = k + 1 and return to
9
Step 2. Otherwise, set n+9 = nk+, and terminate the algorithm.
Again, once Xn+Z is computed by this algorithm, the generalized internal variables are updated via the inverse stress-strain relations. It should be noted that the
convergence of this algorithm towards the final value Zn+ is obtained at a quadratic
rate. A geometric interpretation of the algorithm is given in Fig. 16.2. The path that
returns the trial state onto the elastic domain can be viewed as a sequence of straight
segments on the space S x ni't of generalized stresses. To some extent, the characteristics of this algorithm are opposite to those of the closest-point projection algorithm
as summarized below.
(i) This scheme is an explicit procedure that involves only function evaluations.
No equation system needs to be solved to update the sequence of iterates.
(ii) Normality is enforced at the initial (known) and not at the final state. The
algorithm is clearly consistent.
REMARK 16.2. The simplicity of the cutting plane algorithm leads to a very attractive
computational scheme for large scale simulations. However, our computational experiments indicate that, in sharp contrast with the closest-point projection algorithm,
significant errors can result for large time steps. This suggests that the cutting plane
algorithm is best suited for explicit transient simulations, where the allowable time
step is severely restricted by the Courant condition, while the closest-point-projection

SECTION 17

Integration algorithms

253

scheme is appropriate for implicit calculations where large time steps are typically
used. Further evidence on the suitability of the former scheme for explicit rather than
implicit transient simulations is provided by the fact that an exact closed-form linearization of the algorithm, leading to the notion of algorithmic elastoplastic moduli
described below, does not appear to be possible.
17. The consistent algorithmic elastoplastic moduli
The closest-point projection algorithm described above can be exactly linearized
in closed form, leading to the notion of consistent-as opposed to continuumelastoplastic tangent moduli. The former are obtained essentially by enforcing the
consistency condition on the discrete algorithmic problem, whereas the latter notion
results from the classical consistency condition on the continuum problem. This distinction, first made in SIMO and TAYLOR [1985], is essential if the global algebraic
problem arising from a temporal and spatial discretization is to be solved by Newton's method. Use of the algorithmic elastoplastic moduli produces the exact Hessian
for the discrete problem, thus preserving the quadratic rate of convergence of Newton's method, while use of the continuum elastoplastic moduli destroys this quadratic
convergence.
REMARK 17.1. In the iterative solution of the initial boundary value problem discussed subsequently, the algorithmic moduli are always computed at time t+l for
the backward difference methods summarized in Table 14.1, whereas for the implicit
Runge-Kutta scheme in Table 15.1 the algorithmic moduli are always computed at
time t+o. To accommodate these two cases in a unified setting, it will be understood
in what follows that En,+ = (, + 9As, 0) with = 1 for backward difference
methods.
The derivation of the algorithmic moduli mimics that of the continuum elastoplastic
moduli, but performed on the algorithmic problem. Differentiation of the generalized
elastic stress-strain relations and the discrete (algorithmic) flow rule yields
dXn+

= Gn+o [dE,+o - dEP+o],

dEn+g = Ay [V 2 f (E+O)] d~7+o + d(A-y)Vf(X'n+).

(17.1)

Denoting by G+e the generalized elastic algorithmic moduli defined as


G~+o
= [G
ATy[V2f(Xn+o9)]],
relation
(17.1)
can be+ written
as2)

(17.2)

relation (17.1) can be written as

dn+g = G,+ [dEn+o - d(A-y)Vf(

-)].

(17.3)

J.C. Simo

254

CHAPTER 11

To determine d(Ay) one differentiates the discrete consistency condition f (Zn+)


to obtain Vf(Zn+) d~Z+o = 0. Using this condition on (17.3) gives
d(AT)
d(A

= Vf(,+o) . G+ dE,+
Vf(y) n
* Gn Vf(Ln+ti)
Vf( ,,)
- G.+gVf(7,+Eo)

=0

(17.4)

Finally, let P = DIAG[I, O] denote the projection operator defined by the relation
P[] = a and, likewise, P[E] = e. Expression (17.3) then implies
dn+ = Cn+9 [dEn+o - d(Ay)8af(,+

)]

where C,+og = [PG,+gP],

(17.5)

since a,f(Z) = PVf(2U). Combining (17.5) with (17.4) yields the result

c,
Cn9=

n+i

n+L97

(17.6)
nn+o =

This derivation shows that to obtain the algorithmic tangent moduli all that is needed
is to replace the elastic moduli C in the expression for the continuum elastoplastic
moduli by C,+ = PG,+gP, where G,+o is defined by (17.2).
18. Examples: Closed-form return mapping algorithms
As an illustration of the preceding ideas, the case of J2-flow theory incorporating kinematic and isotropic hardening will be examined in detail. In particular, new secondorder accurate radial return algorithms are described based on the backward difference
return mapping algorithms presented in the previous section. For three dimensions,
the simplicity of the von Mises yield condition-a hyper-sphere in stress deviator
space--enables one to obtain essentially a closed-form solution of the closest-point
projection algorithm. To illustrate the applicability of the general scheme, the case of
plane stress is considered subsequently.
(A) Radial return mapping algorithms for J2-fiow theory. Recall that for J2 -flow
theory the plastic internal variables are EP = (EP,1 , ), with conjugate generalized
stresses E = (s,q,q), where tr[EP] = 0 and s = dev[acr]. To arrive at a compact
formulation of the algorithm, it proves convenient to define a relative stress as the
difference between the stress deviator s and the back stress . Accordingly,
3 = s-q,

whereq=-_2H

and

q=-K'(().

(18.1)

Integration algorithms

SECTION 18

255

Note that a general nonlinear isotropic hardening function K(() is assumed. It is


therefore convenient to use S in place of q as the independent variable in the return
mapping algorithm and write the von Mises yield criterion as
f(,

11 -

[y

+ K'(()] < 0,

and let n =

P/IP1.

(18.2)

Consider the application of the backward difference scheme in Table 14.1 to the
present example. Within a typical time step [t, t+l], the trial state for prescribed
en+l = dev[en+l] and given En is then defined by the following formulae:
Sn+I = 2[en+l - (ep + (s - 1) (eP
n+l = -H[n

C.+1 = n + 3 (S

(s - 1)(&n )(-

en1))],
(18.3)

n-l)],

n-)

In the present context, the general equations defining the return mapping algorithm
reduce to
Sn+

= Sn+

- 2Aynn+l,1

qn+l = qn+l + 3HA-y nn+l,

(18.4)

+ AT.
+l

n+ = n

An extension of the argument described above shows that the solution of (18.4) reduces
to the solution of a scalar equation for the consistency parameter Ay. To see this,
observe first that 3n+l can be written as
3n+

= (Sn+-

+1) - [2

H]

'Ynn+l.

(18.5)

Next, to arrive at the algorithmic counterpart of the consistency condition we note


that by definition, 3 n,+1 = 1n,+l
lnn+l. Hence, from (18.5) the unit normal nn+l is
determined exclusively in terms of the trial elastic stress tn+ as
n,+l,=

t+l

/l1t+l i where 03+1 =

+, -

r+l

(18.6)

By taking the dot product of (18.5) with nn+l we obtain the following scalar
(generally nonlinear) equation that determines the consistency parameter Ay:
[ay + K'((n+1 +

(A'Y) =-

-A

[2u +

Ay)]
1

H] = 0.

(18.7)

The solution of Eq. (18.7) may be effectively accomplished by a local Newton iteration
procedure since g(A-y) is a convex function, and convergence of the Newton procedure

256

J.C. Simo

CHAPTER II

TABLE 18.1

Consistency condition. Determination of Ay.


(1) Initialize: A(00 ) = 0,n.
and ()+ =

(2) Iterate on k: Do until g(A-y(k))j < TOL


(2.1) Compute iterate A-y(k+l):

(ayk)

Dg(A))
Ay(k + )

yy +K' ((l)]

+P$
I -A
-t

[2s+ 2H]

H +2K1,1(l)

=A-2p 1

A( k ) _g(A)

Ug(Ay(k))

(2.2) Update equivalent plastic strain:

A,+l) = ,tr +

(k+l

is then guaranteed to occur. Details pertaining to the local Newton procedure are
summarized for convenience in Table 18.1.
REMARK 18.1. If the kinematic/isotropic hardening law is linear, then Eq. (18.7) is
amenable to closed form solution that results, for s = 1, in the generalization of the
radial return algorithm in KRIEG and KEY [1976]. Set
fn

= 13+l

[Y + K+l]

(18.8)

where oy > 0 is the flow stress in pure tension, K is the isotropic hardening modulus,
and H > 0 is the kinematic hardening modulus. Substitution of (18.8) into (18.7)
yields
A7 =

(18.9)
2/ + 2K + H'

The update is completed by substituting (18.9) into formulae (18.4).


For convenience, a step-by-step description of the algorithm discussed above has
been summarized in Table 18.2. The geometric interpretation of this algorithm is
shown in Fig. 18.1. Next, we proceed to compute directly the consistent algorithmic
elastoplastic tangent moduli by linearization of the two-step return mapping algorithm.
These moduli relate incremental strains and incremental stresses and play a crucial
role in the overall solution strategy of the boundary value problem.
We conclude this example by providing the exact linearization of the radial return
algorithm leading to a closed-form expression for the elastoplastic tangent moduli.

SECTION 18

Integration algorithms

257

TABLE 18.2

Radial return algorithm. Nonlinear isotropic/kinematic hardening.


(1) Compute trial elastic stress for prescribed en+l = dev[en+l]:
Sn+l = 2

-e
-[(en+
2 [n

tn+r =-

++ =

3 (51)(

- 1)(n

3(

3l

(e

)]

-n-1)

-n-).

(2) Check yield condition with 3n+l = s +l - 4t+l

fn+ =

0-n+4

[nY +

)]

(+

If fn+l > 0 then perform the following steps:


1n+I+II and compute A-y as

(3) Set n+ +l =

A'y = defined by Table 18.1 (nonlinear hardening)


A

ft+

2y(K + H)

(linear hardening).

(4) Update back-stress, plastic strain, and stress:


qn+l =/n+l
+1 .tr

HAfnn+l

- 2Ayn+ l

eP+l = e+l

- sn+1/2p

+n+I=r,*+

A.

(5) Compute consistent elastoplastic tangent moduli:

eCn+

lS 1

+ 2Cn+l [I -

g 1] -21n+lnn+l nn+,
3I

2/A'y
(n

2/.t
2p, + 2 (K"(, + 1 ) + H)

Rather than particularize the general result (17.6) to this specific example, we proceed
directly by differentiating the algorithmic stress-strain relations defined by n+l =
K(tr[,+i])1 + sn+l - 2[Aynn+l to obtain
don+l = C dEn+l- 2
=

c-

2/-tnn+

[d(A-y)nn+l

Ay dnn+l]

-+ - 2UAy aE n+l

dEn+l,

(18.10)

J.C. Simo

258

CHAPTER II

J.C. Simo

258

CHAPTER

al
-1

II~~~~~~~~~~~~~..

FIG. 18.1. Geometric interpretation of the return mapping algorithm for the von Mises yield condition and
isotropic/kinematic hardening.

where C = il1+2,t(I- 11) is the elasticity tensor. To carry out the computation
further use is made of the following result.
THEOREM 18.1. The derivative of the unit normalfield n(/3) = /3/l/3 is given by the
formula

3n/a3 = [I - n

(18.11)

n]/31.

PROOF. The result easily follows with the aid of the directional derivative. First we
note that for an arbitrary vector h C R 6 we have
d /

+ hl =

h/l

= n

(18.12)

By the chain rule it then follows that


d I
d

=0

n(3 + h) = h - (n

h)n/131 =

[I-n

n]

h3

(18.13)

so that (18.11) holds.


As in the general case, the term ATy/e,,+l appearing in (18.10) is obtained by
differentiation of the scalar consistency condition (18.7). Accordingly,
aeysn+l
tu

Ko

(n+l)+

in

Ha(. o

,.

(18.14)

Substitution of (18.11) and (18.14) into (18.10) produces, after some manipulation,

the expression summarized in Table 18.2. This expression should be contrasted with
Eq. (4.19) for the "continuum" elastoplastic moduli. As a result of the radial return

Integrationalgorithms

SECTION 18

259

algorithm the shear modulus pI enters in the "consistent" tangent moduli scaled down
by the factor ,,+l. Observe that
1 so that, for large time steps, st+, may
I+l
lay far out of the yield surface and (,+l may become significantly less than unity.
In addition, since <(n+1 = Ay + (n+ - 1, we have the bound Ay - 1 < 5(+l < Ay.
Therefore, for large time steps, the algorithmic tangent moduli may differ significantly
from the "continuum" elastoplastic tangent. However, as At -- 0 and Ay - 0, the
algorithmic and continuum tangent moduli coincide. This result is a manifestation of
the consistency between the algorithm and the continuum problem.
REMARK 18.2. According to the algorithm in Table 18.1, the values ()
are calculated based on the converged values (-), at time step t = t. The (nonconverged)
values ()k,+l at the previous iteration play no explicit role in the stress update.
REMARK 18.3. The adaptation of the projected generalized mid-point rule in Table 18.2 to J2-flow theory is immediate from the preceding results. The intermediate
stage is computed via the algorithm in Table 18.2, with s = 1 and e+l = e, + Ae
replaced by e,+, = e, + 9Ae. The final stage is computed again via the same algorithm, with s = 1 and the trial state replaced by the linearly extrapolated values as
indicated in Table 15.1.
(B) Backward difference algorithmsfor plane stress J2-plasticity. For the plane stress
case, a simple radial return would violate the plane stress condition and thus is
no longer applicable. The basic idea in the algorithm proposed in SIMO and TAYLOR [1986] and described below is to perform the return mapping directly in the
constrained plane stress subspace so that, by construction, the plane stress condition
is identically satisfied. In the plane stress subspace, however, the von Mises yield
condition is an ellipse and not a circle. Thus, when solving the discrete problem
(14.6) one is confronted with the general problem addressed in Section 16; namely,
the computation of the closest-point projection in the complementary Helmholtz free
energy.
Recall that for plane stress J2 -flow theory with combined kinematic/isotropic hardening the generalized strain-like internal variables are EP = (EP, ~, 6). Using the
backward difference formulae, it proves convenient to define trial generalized strains
via the expressions
EPn+l =

3 (s-

1) (P

4n+ = 4n + '(S tr+ =

)(4n -

n-

n 1)I

(18.15a)

(S - 1)(4n -n-I),

where s = 1 for the backward Euler method and s = 2 for Gear's two-step backward
difference method. The trial elastic state
tn+1 is defined for a prescribed strain
increment E,n+l via the generalized stress-strain relations as
=
tr'n+1 =

C [EI

and

q+l =-HHP +

(18.15b)

J.C. Simo

260

CHAPTER II

where C and H are given by (4.24) for the isotropic case, and qt+l =-K'( ) is
the internal variable controlling isotropic hardening. In addition, from Eqs. (4.22) the
generalized backward difference return mapping algorithm takes the following form:
an+l =+1

A'y CP3n+1,

n+ t + AYHP/3n+l,

Qn-

=+
,+n

where /3,+1 =

(18.16)

/n A-Yfn+l,

n,+l - qn+l is the relative stress and P along with f are defined by

(4.22) as
2
+1

/fn
P+l 3n+1

and

-1
0

-1
2

0
0

(18.17)

Instead of applying directly the closest-point projection algorithm described above, it


proves more efficient to combine (18.16)1 and (18.16)2 to obtain, assuming the nontrivial case of plastic loading (Ay > 0), the following system of equations involving
only the relative stress and the multiplier AA:
rn+ =
=3+,, [I + Ay(C + H)P]

(18.18)

f 2 (y)

= i n+2-

[K'(

l)] 2

0,

where fn+l is defined by (18.17)1. This system can now be solved by Newton's
method for Ay > 0 and ,n+.
REMARK 18.4. The algorithmic elastoplastic tangent moduli consistent with the preceding integration algorithm are developed by linearization of the preceding formulae.
This is an exercise in the application of the chain rule from which only the relevant
results are quoted. Differentiation of the system (18.18) gives the linearization of the
consistency parameter as
d(Ay)

nfiPCn+l d,+l

+l

=
/3n+l[P(Cn+

+Cn+ilCH)P+

(18.19a)
2K"P]

n+I

where C,+1 are the effective algorithmic tangent moduli given by


C,+1 = [C - 1 +

y(I + C-1H)P]-.

(18.19b)

Substitution of (18.19a) into the differentiated version of the algorithmic stressstrain relation (18.15b) 1 yields the desired result.

SECTION 19

Integration algorithms

261

REMARK 18.5. For the case of isotropic elasticity, the system (18.18) has a particularly

simple form since the matrices P, C and H all commute. By introducing a spectral
decomposition, this system can be reduced to four scalar equations exhibiting minimal
coupling. The interested reader is referred to SIMO and TAYLOR [1986] for a detailed
account of these developments. More recently, SIMO and GOVINDJEE [1988] noted that
for linear kinematic hardening and certain forms of isotropic hardening the system
(18.18) can be further reduced to a quartic equation for the consistency parameter Ay
which can then be solved in closed form.
Although the preceding examples are restricted to J2 -flow theory, the same methodology applies with essentially no modification to other classical models of plasticity.
In particular, first-order accuracy for Drucker-Prager and other cap models for geomaterials, as described in SANDLER and RUBIN [1979], LORET and PREVOST [1986],
SIMo, JU, PISTER and TAYLOR [1988] and others, are extended in a straightforward
fashion to achieve second-order accuracy via backward differentiation formulae.
19. Practical accuracy assessment: Iso-error maps
A practical assessment of the accuracy of the proposed class of algorithms can be
made by examining iso-error maps, developed on the basis of a strain controlled
homogeneous problem. The procedure has been employed by a number of authors,
e.g., KRIEG and KRIEG [1977], SCHREYER, KULAK and KRAMER [1979], IWAN and
YODER [1983], ORTIZ and PoPov [1985], ORTIZ and SIMO [1986], SIMO and TAYLOR [1986] and LORET and PREVOST [1986]. The construction of iso-error maps will
be outlined taking as an example the return mapping algorithm for plane stress described above. For simplicity, attention is restricted to the conventional backward Euler
method (s = 1). A rigorous accuracy analysis is presented in the following section.
Three points on the yield surface are selected which are representative of a wide
range of possible states of stress. These points, labeled A, B, and C on Fig. 18.2
correspond to uniaxial, biaxial, and pure shear stress, respectively. To construct the

-a1

= 30,000 Ksi

= 0.3

1/Vjllsl-R=0

R = av
1r2= 0
YT

FIG. 18.2. Plane stress yield surface. Points for iso-error maps.

262

J.C. Simo

CHAPTER II

E2Y

&FI/EiY

FIG. 19.1. Iso-error map corresponding to point A on the yield surface.

iso-error maps we consider for each selected point on the yield surface a sequence
of specified normalized strain increments. The stresses corresponding to the (homogeneous) states of strain prescribed in this manner are then computed by application
of the algorithm. At each point the normalization parameters are chosen as the elastic
strains associated with initial yielding. Without loss of generality, the calculation is
performed in terms of principal values of the strain and stress tensors; i.e., it is assumed that el 2 = 0. Results are reported in terms of the relative root mean square of
the error between the exact and computed solution, which is obtained according to
the expression
6=

"(-1

aC)) (O
(

"*)) x 100.

(19.1)

Here,
is the result obtained by application of the algorithm, whereas * is the
exact solution corresponding to the specified strain increment. The exact solution for
any given strain increment is obtained by repeated application of the algorithm with
increasing number of sub-increments, The value for which further sub-incrementing
produces no change in the numerical result is regarded as the exact solution.
The iso-error maps corresponding to points A, B, and C are shown in Figs. 19.1
to 19.3, respectively. The values reported here were obtained for a von Mises yield condition with no hardening and a Poisson ratio of 0.3. Observe that Figs. 19.2 and 19.3

SECTION 19

Integration algorithms

263

2y
E25

a E,/E1
FIG. 19.2. Iso-error map corresponding to point B on the yield surface.

ae2
E2y

FIG. 19.3. Iso-error map corresponding to point C on the yield surface.

exhibit a symmetry which may be expected from the location of points B and C
on the yield surface. From these results, it may be concluded that the level of error
observed is roughly equivalent to that previously reported in the literature for other

264

J.C. Simo

CHAPTER

return mapping algorithms. As a rule, good accuracy (within 5 percent) is obtained


for moderate strain increments of the order of the characteristic yield strains. It is also
noted that exact results for any strain increment are obtained for radial loading along
both symmetry axes, as expected. Iso-error maps for higher-order methods, such as
the two-step backward difference scheme (s = 2) and the projected mid-point rule
(9 = ), confirm the superior accuracy of these methods for small time steps over
the backward Euler method but reveal a significant degradation in performance for
moderate and large time steps. These observations are consistent with the analysis presented in Section 12 which identifies the backward Euler return mapping algorithm
as the optimal scheme for very large time steps (i.e., optimal long-term dissipative
behavior).
20. Accuracy analysis of return mapping algorithms
The accuracy analysis of general return mapping algorithms described below exploits
the following geometric framework. We regard the boundary BE of the elastic domain
as a smooth hypersurface in S x IRn'mI with Riemannian metric induced by G-1 (and
endowed with the standard Levi-Civita connection). From this point of view the
equations of rate-independent plasticity possess an intrinsic geometric interpretation
independent of the parametrization. Within this context, higher-order rates of change
of the plastic flow at a point are naturally measured in terms of higher-order covariant
derivatives. To illustrate the main ideas involved, we choose as a model algorithm
the generalized projected mid-point rule in Table 15.1 and show that this scheme is
second-order accurate only if 9 = . A completely analogous analysis, the details
of which are omitted, confirms that the two step (s = 2) backward difference return
mapping algorithm in Table 14.1 is also second-order accurate.
(A) The local geometry of the plasticflow. For convenience, let us denote in what
follows the local inner product induced by G as (, )G = () G - ' (.) and use the
notation I - IG = V , )ci for the associated norm. Then define the normal field N at
a point X E aE by the expression
GVf(z)
N = IGVf(Z)

so that INIG = 1.

(20.1)

For given strain rate history t - Et, the map t - r = GEt defines a corresponding
rate in stress space. In general t. will not be tangent to the yield surface aE. The
r
projection of 7Zt
onto the tangent plane at Zt aE, denoted by P(t,), is given by

() r=- (r,
-

Nt)GNt
_F__r~

if (tr , Nt)G > 0,


otherwise,

(20.2)

SECTION 20

265

Integration algorithms

so that, in view of the definition for the elastoplastic tangent moduli, the evolution
equation describing the plastic flow t
Et can be written as
r

= ?(

(20.3)

) = (GEt).

Plastic loading will be assumed throughout so that ('7, Nt)G > 0. The vector field
t - t is therefore tangent to the yield surface. To compute its time derivative denoted
by t, observe that (20.3) implies (7, Nt)G = (t, Nt)G since (Nt, Nt)G = 0 as
a result of the unit length constraint (20.1). Using this observation, time differentiation
of (20.3) then gives

_t

tr _ (A,

N)GN

(tr,

Nt)G

, -t (,

Nt)GNt,

(20.4)

where Xtr = GEt. In general, however, Xt is not tangent to the yield surface alE. The
covariant derivative of t, denoted by D/Dtt, is precisely the projection (in the
G-inner product) of t onto the tangent plane at Zt and defines, therefore, a vector
field tangent to the yield surface (see, e.g., THORPE ([1979], p. 45) for an introductory
exposition of these ideas). Using again the fact that (Nt, Nt)G = 0 one obtains

-t,4 = iX,

(r,

Nt),

(Zr,Nt)GNt,

(20.5)

since the normal component of the acceleration is given from (20.4) as t,


-(t,,
Nt)c. From (20.1) the time derivative of the unit normal field is

Nt

tt-

(Dtt,

Nt)G

Nt

where f2t

GV 2 f(L't)
IGVf( ()t

Nt)G =

(20.6)

The bilinear form (, Ot )G at Xt induced by S2t is easily shown to be self-adjoint


and is known as the Weingarten map. In summary, expression (20.6) along with (20.4)
translate into the classical result
(t,

Nt)G = -(t, Nt)G

S)G
t
= KN(L't)

l tl,

(20.7)

where rtN(Xt) is known as the normal curvature of the surface aIE at the point Xt
in the direction of t; see THORPE ([1979], p. 62). It follows from (20.7) that the
normal acceleration of the plastic flow at a point on the yield surface is proportional
to the normal curvature of the yield surface at that point.

266

J.C. Simo

CHAPTER

(B) The discrete local problem: Accuracy analysis. In the two-stage implicit algorithm summarized in Table 15.1 the generalized stress at each stage is defined via
a closest-point projection of two trial states. To assess the accuracy of this scheme
assume the nontrivial case of plastic loading and consider a generic stage written in
the form
- YhGVf( h)

Zh = -h

with Yh > 0 so that f(Zh) = 0.

(20.8)

Here Zh and Z'~ are interpreted as the final and trial states, respectively, of either the
first- or second stages of the algorithm. The subscript h [0, At] is used to emphasize
that the computed algorithmic solution in each of the stages depends parametrically on
the time step. We assume that h' h=O = hIh=O and define the algorithmic moduli
Gh = [G- 1 + hV2f (h)]

(20.9)

where Yh > 0 is computed by enforcing the condition f (Zh) = 0 via the closest-point
projection algorithm in Section 16. The same argument described in Section 17 for
the derivation of the algorithmic elastoplastic moduli now yields
Vf(ah)

Gh(G - 1dxh)

Vf(oh) Gh Vf (zVh)

dh

(20.10)

= Gh (G - dh) - dYh GhVf(Lh).

Z' is prescribed, (20.8) defines an algorithmic flow


Assuming that the mapping h
h - Zh with initial data Zh Ih=O = lh h=O Observe that by setting

h ,

Nh

GhVf(Zh)

IGhVf (Zh) IGh

and

Oh = -

GhV 2 f (Zh)

where I 1Gh is the algorithmic norm associated with the inner product (,
by Gh, result (20.10) can be written as
dZh = Gh(G-'

(20.11)

IGhVf (oh) lIh

dU) - (Nh, (G 1 dZy))GhNh.

)h induced

(20.12)

This algorithmic differential equation is of the form dZh = I h[Gh(G-' d ' hr)], which
is identical to (20.3) with the Riemannian metric G - l replaced by the algorithmic
metric Gh . Using the notation d(-)o = d(.)Ih=o/dh we have the following basic
results.

Integration algorithms

SECTION 20

267

THEOREM 20.1. The first derivatives of the functions h, Gh and Zh are


dYo = (No, dsr)G/lGVf (o)lG,

dGo = -G-'oG-(No, d)G,

(20.13)

dlo = dX0f - (dl'4, NO)G No.


PROOF. Relation (20.13)1 follows from (20.9) and (20.10) by noting that ?YhIh=0 = 0.
Relation (20.13)2 follows from (20.13)1 and (20.9). Relation (20.13)3 follows from
the preceding two and (20.10).
The second-order accuracy assessment of the one-parameter family of algorithms
in Table 15.1 hinges on the following result.
THEOREM 20.2. The second derivative of the mapping h
d2 o = d2 r

(d2

F-+ Zh

at h = 0 is given by

, No )GNo

- 2(No, dT)GdNo - (dL'o, dNo)GNo,

where dNo =

(20.14)

od~o - (od~o, NO)G No-

PROOF. This is an involved but otherwise straightforward application of the chain rule
to expression (20.10) that uses the relations in the preceding lemma. The details are
omitted.
EXAMPLE 20.1. We consider here the conventional backward Euler return mapping
algorithm, obtained as a particular case of either the backward difference methods in
Table 14.1 for s = 1, or the projected implicit Runge-Kutta methods in Table 15.1
for 9 = 1. Within a typical interval [t,, t,+l], this algorithm is recovered from (20.8)
by setting
Z

XT = n + G[En+l-En]

and

h =

+l.

(20.15)

Since d
= G
=
, consistency of the algorithm (i.e., first-order accuracy)
follows immediately from (20.13)3, which becomes identical to (20.3). Noting that
d2 ,r = GE = , , expression (20.14) reduces to
d2

tr -(',

Nn)GNn

2(t,

Nn)GNn

n, n)GNn,

(20.16)

which differs from expression (20.4) for the exact flow at time t = t, by a factor
of two in the third term. Hence, the conventional backward Euler return mapping
algorithms are not second-order accurate. Observe, however, that the curvature term
(d2Z, NO)G = -(d2o, Qo d0o)G is second-order accurate.
Consider next the application of the preceding results to the two-stage scheme in
Table 15.1 with 0 < < 1. One comment may be appropriate here. The result below

268

J.C. Simo

CHAPTER II

may seem obvious at first glance but it is not. Many different options are possible in
the design of the second stage. The one recorded in Table 15.1 is precisely that one
capable of retaining second-order accuracy.
THEOREM 20.3. The two-stage, implicit,projected mid-point rule method in Table 15.1
achieves second-order accuracyfor ?9=

PROOF. The proof involves the sequential application of the result in the last theorem
to each of the two stages of the algorithm.
Stage 1. The first stage of the algorithm is recovered from (20.8) for
0 = n,

.n+ 9G[[En+
l - E]

and

Zh = n,+3.

(20.17)

As a result, dEor = VGE = 9


which, in view of (20.3), gives dL 0 = 0L,' when
inserted in (20.13)3. Since d2So r = 09L, application of (20.14) then yields
d2 z 0 = o[E -

n Nn)GN]

- 202(1X, Nn)GNn - 09(,, N,)GNn.

(20.18)

Therefore, the first stage is first-order accurate with the exact solution at time t = t,+o.
Stage 2. The second stage of the algorithm in Table 15.1 is recovered from (20.8)
by setting

Zo = Xn.+o,
th=

[,+

(20.19)
- (1 -

),,]/0

and

Xh = 7,+1.

The composite two-stage method is consistent since the preceding formulae and
(20.13)3, along with the condition (n, Nn)G = 0, imply
d Jt = 1 dh

E(n) so that dZo =


= i7,

..

(20.20)

Next observe from (20.19) that d2L'r


2
= d2 _f,+O/dh 2 Ih=O, which is given by expression (20.18) computed in the first stage. Therefore, since dZ 0 = En, result (20.14)
specialized to the second stage takes the form
d2o=

Nn)GNn] -

[:

(,

- a(n,

Nn)GN,

which agrees with (20.4) if and only if

2(,

Nn)GNn

(20.21)
=

[]

REMARK 20.1. A similar analysis shows that the algorithm ALGO 2 described in
Section 12 for the Prandl-Reuss model is also second-order accurate. The extension

SECTION 21

Integration algorithms

269

of this scheme to the general case appears to be questionable since it involves the
computation of an oblique projection. By contrast, ALGO 3 and its generalization
outlined in Table 15.1 involves the computation of two closest-point projections which,
by the classical projection theorem, are guaranteed to exist if E is convex. From the
point of view of implementation, it should be emphasized that the two stages are
carried out with an identical algorithm. The only difference between them lies in the
definition of the trial values. Finally, we remark that the accuracy analysis of backward
difference methods involves the same ideas although it is considerably simpler. Further
details are omitted.
21. Extension of return mapping algorithms to viscoplasticity
The general class of algorithms developed in Sections 3 and 15 can be easily extended to viscoplasticity by regarding this latter model as a viscous regularization of
the rate-independent model, in the sense described in the preceding chapter. By adopting the regularization procedure described in Section 6, which can be viewed as an
extension of the technique first proposed in DUVAUT and LIONS ([1976], p. 234), the
severe ill-conditioning exhibited by early treatments of viscoplasticity (e.g., HUGHES
and TAYLOR [19781 and the references in Section 6), can be entirely circumvented.
We consider first the extension of the general backward difference return mapping
algorithms and then examine the important case of J2 -flow theory as a specific example.
To motivate the key idea underlying the algorithmic treatment of viscoplasticity
introduced in SIMO, KENNEDY and GOVINDJEE [1988], we shall consider first the
case of linear viscosity. The general evolution equations (7.7) yield the following
viscoplastic regularization of the rate-independent problem (11.2):

P=

.]

G-[7T --

with

(21.1)

= G[E - EP].

Recall that > 0 is the relaxation time and , is the closest-point projection of
X onto the convex elastic domain E in the complementary Helmholtz free energy,
as described in Section 6. Consider an algorithmic approximation of (21.1) via the
s-stage backward difference method in Table 14.1 with either s = 1 or s = 2. Setting
E+,
---

E + 3-

n+l = G [E,,+-

1) (En - En-l,

(21.2)

En+l]

and denoting by Z,+1. the closest-point projection of the (unknown) generalized


stress field .,+1 onto E, the backward difference approximation to (21.1) can be
written as
n+l =

t['+ 1
~+;~~~~
7~~--

At,
-n+1*],
,-

At

[1 -

3(s -

1)]At.

(21.3)

J.C. Simo

270

Multiplying both sides of this equation by


equivalent result

Zn+l

+l* +

T/AJ[,

T/Ats

- nl*]j

CHAPTER II

and collecting terms yields the

(21.4)

-I + , * as T/At, -- 0 and one recovers the rate-independent solution


defined as the closest-point projection of Z+1 onto E. Denoting this projection by
F+1,, it follows that E,+l, = n+,1 * and (21.4) reduces to

Clearly, .n+

n+l
n = t+1*

+ l+i+ -/Ats [[r-

7+/At*]

(21.5)

The algorithm implied by this result reduces to the following three-step method that
relies crucially on the solution of the rate-independent problem:
Step 1. Compute the trial state exactly as in the rate-independent problem via formulae (21.2).
Step 2. Compute the rate-independent solution
projection algorithms described in Section 16.

+,* via the general closest-point

Step 3. Compute the regularized, rate-dependent, viscous solution by applying the


explicit formula (21.5).
The remarkable property that makes this algorithm rather useful from a practical
standpoint is the well-conditioning of formula (21.5) for any value of the relaxation
time rT [0, oc). In particular, one can set = 0 in (21.5) to recover exactly the
rate-independent limit. Observe that for s = 2 the algorithm is second-order accurate,
retains the property of A-stability and is G-stable. This method appears to be new.
REMARK 21.1. The linearization of the preceding algorithm is trivial. Let C? denote
the algorithmic elastoplastic tangent moduli consistent with the closest-point projection
algorithm described in Section 16. From (21.5) it immediately follows that

= Ce+
-

'/At

1 + T/At8

-[C-

6*]

(21.6)

are the algorithmic elastoplastic moduli associated with the three-step method outlined
above.
21.2. An identical construction applies to each of the two stages in the
generalized projected mid-point rule algorithm summarized in Table 15.1.
REMARK

Integration algorithms

SECTION 22

271

22. Return mapping algorithms for general models of viscoplasticity


The algorithmic treatment introduced above for linear viscosity, the only case considered in DUVAUT and LIONS [1976], is easily extended to accommodate the general
model of viscoplasticity, with evolution equation now defined by
p=

g(J())G

,],

(22.1)

where the function g(.) has the property g(z) = 0 if and only if x < 0, and g(x) > 0
for x > 0. The algorithm (21.5) now reads
At5

[ ,,

.+,] - g(J(+l))

[,+

1l -

+l*]

= r 0,
=

(22.2a)

where the distance J(,n+l) between En+1 and the convex set E remains to be
computed to complete the algorithm. Remarkably, this computation involves only the
solution of a scalar equation, as the following result shows.
THEOREM 22.1. The determination of J(Z,+l) is accomplished by solving the following nonlinear scalar equation:
J(LZ+l) [T/At,

+ g(J(n+l))]

(22.2b)

J(ln'+l),

prescribedfrom the solution of the rate-independentproblem.

with J(T+l)

PROOF. Adding and subtracting

= T/Ats

+ g(J(L'n+t))] (r.+i

n,+l in (22.2a) yields


- Xn41*) = k-(L ~

At,
[At,

n l*)

n l*,

(22.3)

Recalling that the complementary Helmholtz free energy is given by .E(Z) =


G- 1Z, Eq. (22.3) implies that

(22.3

At,
The result follows from definitions (7.3) and (7.4) for the distance J(Z).

We remark that the three-step algorithm outlined above, with (21.5) now replaced
by (22.2a,b), inherits all the properties described in the case of linear viscosity. In
particular, it is well-conditioned for any E [0, oo) including r = 0. An identical
construction also applies to each of the two stages in the projected version of the
generalized mid-point rule algorithm given in Table 15.1.

J.C. Simo

272

CHAPTER II

The preceding strategy is illustrated below in the specific context of J2-plasticity


with combined linear isotropic and kinematic hardening. The algorithm presented
below provides a unified treatment of a constitutive model widely used in applications
which includes both the (linear) viscous case as well as the inviscid limit.
EXAMPLE 22.1. Consider the model of J 2-flow theory treated in Section 18, with
quadratic potential K(() = Kt2 for isotropic hardening and a constant kinematic
hardening modulus H. As noted earlier, under this assumption, the radial return algorithm described in Section 18 then becomes amenable to a closed-form solution. Let
= (a, /, q) and denote by In+l, = Zt[+l, the rate-independent solution computed with the algorithm in Table 18.2. Consider for simplicity linear viscosity, with
characteristic relaxation time r > 0, and rewrite the inviscid radial return algorithm
in the following form which allows a literal application of formula (21.5). Set

where n,+l =
given by

n+/

N+ti,

+t

Vf(Zn+1 ) =I
3

nf+1l

and observe that the generalized elastic moduli G are

1+

G=
G= [O

(22.5)

O
H1
0T

o
0T

(22.6)

In terms of this compact notation, the rate-independent solution given in Table 18.2
takes the form
tr

r+l-y* AyGNU
GNtr+l

_r

n+l*

where y
A7* = 2/ + (K + H)
where

(22.7)

(22.7)

Inserting this expression into formula (21.5) gives


n+l =

AT *G
A
y*GNr+,
+1 /At
In~l
1 + L/AtA'y(.

ntr+l = tr

At GN

'

N ~+

(22.8)

By comparing (22.7) and (22.8) we arrive at a remarkable result from a computational standpoint. To incorporate linear viscosity in J2-flow theory only two trivial
modifications are needed in the algorithm in Table 18.2 (with K" = K constant):
(i) Redefine the consistency parameter Ay by the formula
AY (=1
(1 +

)[2 + (K + H)](22.9a)

Azt,,r

'1

(22.9a)

SECTION 23

Integration algorithms

273

(ii) Redefine the coefficient ~n+l in the linearization of the algorithm as

+I

[2/t + (K

H)] -

(1 -

(22.9b)

p2+1).

It should again be emphasized that if expressions (22.9a,b) are adopted in place of


those given in Table 18.2, the rate-independent model can be recovered exactly without
introducing any ill-conditioning in the algorithm merely by setting r = 0.
23. The algorithmic initial boundary value problem
A discretization in time using the generalized return mapping algorithms described in
Sections 14 and 15, along with a suitable time discretization of the inertia term in the
weak form of the momentum equations, reduces the initial boundary value problem of
dynamic plasticity to a boundary value problem within a typical time step [t., tn+l].
The goal of this section is to provide a concise statement of this problem suitable for
the nonlinear stability analysis addressed in detail subsequently.
The first step in the weak formulation of the initial boundary value problem for dynamic plasticity is the reformulation of the return mapping algorithm as a variational
inequality. Before doing so, we recall briefly the functional setting for the problem
introduced in Section 9 under the assumptions of (i) hardening plasticity, with generalized plastic moduli H assumed constant and positive definite, and (ii) linear elastic
response with point-wise stable, constant elastic moduli C. The space T of generalized
stress states is defined by
T=

= (,p):{T ?

L 2 (2) and Pi
nintht
Tij

L 2 (2)}.

(23.1a)

The space T is equipped with the natural energy inner product induced by the symmetric bilinear form defined by twice the complementary Helmholtz free energy function
as

A(,

T) =/

C.-

d2 + Jq . H-lpdQ = a(, r) + b(q, p).


(23.1b)

The solution space St for the displacement field in the body S2 is the set defined as
St = {u(.,t) E Hl(2))ndi: u(.,t) = (.,t) on Fu}.

(23.2)

The associated vector space of admissible displacement variations V consists of vector


fields on 2 satisfying the homogeneous form of the essential boundary conditions,
i.e.,
V =

H()ndim

'

= O on Fu}.

(23.3)

274

J.C. Simo

CHAPTER

Assuming the essential boundary conditions are time-independent, then for fixed time
t E II the velocity field v(., t) is in V. As in Section 9, we use the notation X =
(v, IF) Z where Z = V x [T E]. The space Z is equipped with the natural inner
product ((., .)) induced by the sum of twice the kinetic energy and the bilinear form
A(-, ), i.e.,
((Xl, X2)) = (Pol,72) + A(T, T 2)

VX,X2

C Z.

(23.4)

The associated norm is denoted by II II = 7-))


and provides the natural norm
for the elastoplastic problem since, relative to this norm, the initial boundary value
problem for dynamic plasticity is contractive.
Let II C IR+ be the time interval of interest and consider an arbitrary partition
II = U: o[t,
t+l]. Our first objective is to develop the variational form of the two0
step generalized return mapping algorithms in Table 15.1 within a typical time step
[tn, tn+l] C II for given initial conditions X, = (v,, XZ) E Z and u, C S,. Let
u,+l C Sn+l be the algorithmic approximation to the displacement field at time t,+l
and let Au = u,+l - u, E V be the displacement increment. Consistent with the
local algorithm in Table 15.1, define the trial stresses at the intermediate and final
stages as
Z+.14 = (a, + CE[Au], qn)
and

(23.5a)

Z+

= [n+

- (1

-)]/

respectively. The solutions E7,+ and IE,+l of the intermediate and final stages,
defined locally as the closest-point projections onto E in the metric G - l , can be
characterized as the optimality conditions of the variational inequalities
A(

+1

- -.n+

) <

0
(23.5b)

and

A(Z.+

- Zn+, T -

n+ 1) < 0

for all admissible stresses T E E n T. It should be noted that this characterization


of the two-stage algorithm is rather general and holds even when the boundary ]E
of the elastic domain is nonsmooth. The following result, a generalization of (14.10),
furnishes the algorithmic counterpart of (6.8).
THEOREM 23.1. If the elastic domain is the convex set defined by (6.2), i.e.,
E=

E$ x

nit: f()

{X
< O, for / = 1, ... , .m},

(23.6)

then the local optimality conditions associated with (23.5b) yield the two-stage
algorithm

SECTION 23

Integrationalgorithms

275

En+@ = X+"
-

A-yPVf(X+ +),
p=I

(23.7a)
17n+ = Z,'

-a7ey+ Vff(Z+ )
,1=l

where A-yL > 0 and Aj"

>

0 obey the Kuhn-Tucker conditions

m
and

EA'y1f(n+o) = 0
L=

m
ZAyif,(X,+l) = 0.

(23.7b)

ft=1

PROOF. Suppose that the optimality conditions (23.7a) hold. Taking the G-inner product of (23.7a)l with (T - 7n+o) and using the convexity assumption on the functions
f,(.) along with condition (23.7b) 1 yields
(T --En+9) G-' (1X.+

- 17,+o)

-n+0) * Vf,'(+,)

= -(T
/z=l

A'y [ff(T)

f (n+)]: =

A'yf,(T) < 0,

(23.8)

since Ay,
0 and f,,(T) < 0 for any T E E. Integrating (23.8) over Q gives
(23.5b) 1. An identical argument holds for (23.5b) 2. The converse result is proved
exactly as in (14.10) by replacing the single constraint in the Lagrangian
defined
by (14.11) with the m constraints associated with the convex set (23.7a).
REMARK 23.1. For the viscoplastic problem a similar argument shows that the variational inequalities characterizing the algorithmic treatment in Section 23 become
A(17~'nr+ go - 7+0:,T-

X,+

< Lt)

/[g(J(T))

-g(J(1X+n

))] dO,
(23.9)

At+

Xn+1T-

n+1)

<

Al

[g(J(T))-g(J(17n+,))]d?,

for all admissible stresses T E T which are no longer constrained to lie in E.


The unknowns to be determined in the incremental algorithmic problem are Ln+1 E
T, u,+l E S,+l and v,+l E V for prescribed forcing function f(., t) and prescribed
boundary conditions t(-, t), t(., t) in [t, t,n+l]; see the statement of the problem
given in (9.5b). As pointed out above, for the continuum viscoplastic problem the
generalized stresses 1 E T need not lie within the elastic domain E; a fact also
reflected in the algorithmic inequality (23.9). On the other hand, inequalities (23.5b)

J.C. Simo

276

CHAPTER 11

ensure that in the algorithmic version of rate-independent plasticity both Z and


Zn,+l, as well as the intermediate stage n+0o, are in E n T.
Viewing the return mapping algorithm as a generalization of the conventional midpoint rule suggests enforcement of the algorithmic counterpart of the momentum
equations precisely at the intermediate stage. This leads to the following algorithm
which is consistent with the generalized mid-point discretization of the initial boundary
value problem:
Stage I: Find Au and X,+o = (v,+o,
I (poAu,

n+o1)

such that

) - (vn+O9,i ) = 0,

t( Po(un+o - v),) + (+.o, L [7]) = (fn+,


A(X+

- Z+

) + (n+l, r)r,

(23.10)

) < 0,

T - Zn+

for all (, T) E Z. The variational equations (23.10) determine the velocity and
generalized stress Xn+2 = (vn+o, Z+) e Z, along with the displacement increment
Au E V. Finally, the displacement and velocity fields are updated consistent with a
generalized mid-point rule approximation while the generalized stress is computed by
enforcing consistency.
Stage IIa: Update

Vn+ =

Un

v+9 +

and v, via linear extrapolations as

I-

jVI'I

;1

v n,

Vn,(23.1
=

a)

Un+l = Un + Au.

Stage IIb: Compute Zn+l for given Z"+l via the return mapping algorithm
A(E+, -Zni'.T-

n+i)

VT

EnV,

(23.1 lb)
where Z+, = [Xn+

- (I - )X.n]/V.

Computational issues involved in the implementation of this problem are deferred to


Chapter IV where they are addressed within the context of the full nonlinear theory.
A brief account of mixed finite element methods suitable for the spatial discretization
of the elastoplastic problem is given in the last three sections of this chapter together
with sample numerical simulations.
REMARK 23.2. For viscoplasticity, Eq. (23.10)3 is replaced by (23.9). In general, the
stress a,,+l does not satisfy the equilibrium equation at t+l unless the forcing terms
are linear in time since the momentum equation (23.10)2 is enforced at the generalized
mid-point configuration at time t,,+o.

SECTION 24

Integration algorithms

277

24. Nonlinear stability analysis: Uniqueness and dissipativity


This section provides a complete nonlinear stability analysis of the preceding algorithmic problem and addresses some of the issues involved in a numerical implementation.
The notion of stability used in the analysis is dictated by the contractive property of
the continuum problem. In short, this notion implies attenuation of arbitrarily large
perturbations in the initial data relative to the natural norm of the problem, which is defined by the complementary Helmholtz free energy. The nonlinear approach described
below should be contrasted with other approaches found in the literature which typically employ linearized notions of stability (see, e.g., HUGHES [1983] and references
therein).
Let Xn E Z and n G Z be two initial conditions at time t. The preceding
algorithm then generates two sequences {Xn}ncN and {Xn,}nEN. Nonlinear stability
holds if the algorithm inherits the contractive property of the continuum problem
relative to the natural norm III 111; a property also known as A-contractivity or
B-stability. Equivalently, finite perturbations in the initial data are attenuated by the
algorithm relative to the natural norm if the B-stability property holds. To carry out the
stability analysis the following preliminary result for the second stage of the algorithm
is needed. This property is an immediate consequence of the projection theorem (e.g.,
CIARLET ([1988], p. 268).
t
THEOREM 24.1. Let A7+ 1 =
+l - ~-+1, where Znr+l C T and +1
T
are arbitrary. Then, the second stage of the return mapping algorithm, defined by the
variationalinequality (23.11lb), does not increase the energy norm in the sense that

A(An+i, A,,+l ) - A(A~r+,, I An'+l)

where AI7+

= 7n+1 -

A'[+,)

+l,r.rn+l,
A,,-

- A.

-<-A(An+

(24.1)

< 0,

n+1l is the difference in the solutions of (23.11b).

PROOF. Applying inequality (23.1 lb) to n+l and Zn+, with T E EnT respectively
chosen as T = L',+1 and T = En+1, yields
A(-t+l -

1,v'+

A (nt+ 1- n+1,

+1-I

+1) < 0,
(24.2)

0-Yn+-

n+ ) <

0.

Adding these two inequalities gives, after adding and subtracting a term, the result
A(A_+ I- - Al
+

[An+l -

1,

+1])

+ + An+

I]

(24.3)

0.

Estimate (24.1) then follows from the bilinearity property of A(., .).

[]

278

J.C. Simo

CHAPTER 11

The B-stability property of the two-stage algorithm defined by (23.10) and


(23.1 1a,b) is contained in the following result.
THEOREM 24.2. The algorithm defined by the variationalproblem (23.10) [Stage I]
and the updateformulae (23.11a,b) [Stage II] is B-stablefor > , i.e., the following
inequality holds:

Illxn - xnlll < IIixo -

olil,

Vn e N, if

>.

(24.4)

PROOF. (i) First, consider Stage I of the algorithm defined by (23.10) and introduce
for convenience the following notation for any 09C [0, 1]:

AZn+, =

n+g -

+o

and

Av,,+ = v~,+

- v,+.

(24.5)

By hypothesis, the two sequences generated by the algorithm satisfy the problem
using the definition of
(23.10). Choosing successively T = Z,+g and T = ~Z+,
E+7t
and the notation introduced above yields
a(Ce[Ai4 ao,+

c&,+) + A(

,+, AnZ+9)

0,

(24.6)
-a(VCE[Au], c,+ - &,+o) - A(Zn

Zn+o, An+)) < 0.

Adding these two inequalities and using (24.5) yields the result
A(A27,+9 - An,, A2]n+O) < Ja(CE[Au - Au],

n,,+ - a,+).

(24.7)

Since the difference Au - Ai is in V, use of definition (23.11a)1 along with the


algorithmic weak form of the momentum equations yields
a(Ce[Au - Aiu],

- &+,9)

= (a,+g[ E[Au-Aii]) - (,+v,

t
=

(P0(Av,+

E[u - A])

- Av), Au- Au)

-t(Av +l - Av )AAu -Aii),

(24.8)

where we have used again the identity 9[Av+o - Av,] = Av,+l - Av, implied by
the linear extrapolation formulae (23.11 a). On the other hand, the variational equation
(23.10)1 along with a simple algebraic manipulation yield the local relation

t [Au - Au] = Av,+o = Av,+l/ 2 + (0a-

)[Av,+

Avj.

(24.9)

Inserting this expression into (24.8), and using the identity

K(Av,+t) - K(Av,) = (p0 Avn+1/ 2, AVn+ -Avn)

(24.10)

SECTION 24

Integration algorithms

279

gives, after a straightforward manipulation, the result


a(CE[Au] - AU],

c,,+B

n+T)

= -[K(Avn+l) - K(Avn)] - (219- 1)[K(AVn+ - Av,)].

(24.11)

Combining (24.7) and (24.11) one obtains the following estimate:


K(Av,+l) - K(Avn) + -A(AXn+o - A.,
- (V -

A+o)

)2K(Av,+l - Avn).

(24.12)

(ii) Now consider Stage II of the algorithm defined by formulae (23.11a,b). The
extrapolation formula (23.1 1b) 2 that defines the trial state implies the identity
An+

[An+l + AXt'T+] + ( -

along with AZ,+o - Aln =

9[At+,

2)

- AX',]

[AZ"+1

(24.13)

- ALn']. Using the bilinearity of

A(., ) we

arrive at
1A(AE~n+o - AZ,

AZn+9)

[A(A.tr+l, A.-r+)

- A(An

An)].

(24.14)

Combining (24.12) and (24.14) gives the estimate


[2K(Avn+,) + A(A,+
< -(9-

1 , A.+l 1

)] - [2K(Av) + A(An, AZ)]

An+l- AV)
) [2K(-

+ A(A2?+

- A7, A+

- AX)].

(24.15)

Since the right-hand side of (24.15) is nonpositive for 9 0 if follows that the
linearly extrapolated state is contractive for > . To complete the proof we need to
estimate A(AX,+i, AZ,+ 1 ) in terms of A(A
A2+l) using the second stage.
A,'+,
This is precisely the result stated in the preceding theorem. Thus, adding inequalities
(24.15) and (24.1) gives
IIIx + -

-- IIX
x+1112
-

1112

= [2K(Avi+l) + A(An+,, AEl)]


< -(0-

2) [2K(Avn+l-

- A(ATl+1 -

.+

Avn)

--

- [2K(Av)

+ A(A?+l1

A,

+ A(A,

An)]

7+

- A2)]

+, - An+l
(0
for 9
1 )A~],~+i
(~~~~24.6
A~~~~n~l,

(24.16)

280

J.C. Simo

CHAPTER II

which in conjunction with a straightforward induction argument implies (24.4) and


completes the proof of the contractivity result. [1
An argument entirely analogous to that given above also shows that the incremental
algorithmic problem for viscoplasticity is A-contractive relative to the natural norm
II - Ill. In summary, the analysis for dynamic plasticity and viscoplasticity described
above proves nonlinear stability in the velocity and the stress field X = (v, Z),
provided that '>9 2.
24.1. Uniqueness of the solution to the algorithmic problem defined by
(23.10) and (23.11a,b) follows immediately from the preceding contractivity result.
Suppose that X, and X, are two solutions of the problem for the same initial data.
Setting Xo = Xo in (24.4) gives 0 < I IX,,-Xn
l
I11 0, which implies IIX,
l-X,,
I =0
and therefore X, = , since * 11Iis a norm.

REMARK

The dissipative property of the initial boundary value problem for dynamic plasticity,
encoded in the a priori stability estimate (10.13), is also inherited by the algorithmic
problem defined (23.10) and (23.11a,b) provided that 9 2. To prove this result
observe first that the internal energy Vint(e, ~), defined by (10.10), coincides with the
total complementary Helmholtz free energy function SE2 (I) under the assumption of
a quadratic free energy function, i.e.,
= Vint(Ee , ) ,Q(7)

= A(,

),

(24.17)

where A(-, ) is the bilinear form defined by (23.lb). The total energy of the system
then becomes a function of X e Z defined by the expression
E(X) = K(v) + _En()

+ Vext(u)

(24.18a)

where K(v) is the kinetic energy. For simplicity, attention will be restricted to dead
loads, with potential energy given by
Vext(U)= (f,u) + (t, u)r.

(24.18b)

Next, consider the general model of multisurface plasticity with elastic domain defined
by (23.6) and assume that the functions f,: S x REin -- R are of the form
fl(1) = ,(17) -

yf,

for

= 1, 2,. . , m,

(24.19)

where , (.) is convex, homogeneous of degree one. As shown in Example 6.1,


from Euler's theorem for homogeneous functions it follows that the model obeys the
dissipation inequality Do > 0 in 11.Under these hypotheses we have the following
result.

Integration algorithms

SECTION 24

281

THEOREM 24.3. For > 2 the algorithmic initialboundary value problem defined by
(23.10) and (23.1 la,b) inherits the dissipativeproperty in the sense that the following
estimate holds
E(Xn+l) - E(Xn)

(24.20)

0 for any n E N

where E: Z -- IR is the energy function defined by (24.18a).


PROOE First write the change in complementary Helmholtz free energy as
'(-n+l )

(n)
(
)

EDX
(n+)

- -

+ ['Q(nr+l) -

(+l

)]

72 (xn)].

(24.21)

The first term on the right-hand side of (24.21) is estimated merely by noting that
EQ (Z) is a quadratic functional. Definition (24.17) together with an elementary identity then yields
[Qn

2Q(n+,li)]
l= A(n+l,+

< A (,+l,

t-

)-

.EQ(an+

1tr)

(24.22)

17+ - Znr+1)

For the second term on the right-hand side of (24.21), an elementary manipulation
together with the extrapolation formulation in the first phase of Stage II of the algorithm, defined by (23.1 la), yields the following expression for the second term on the
right-hand side of (24.21):
,F-(

)-

(n)

= A(X.r+, +

+ -

En,

,)

= A(Xn+o9, En+l - n,) - (29 - 1) E(


~<WA(wn+,

.,+

- n),

-_+
-

(24.23)

r+, +
+l =
, + 9(CE[Au], 0), setting ,+Z =
provided that 9 2. Since
[17+o - n7+g] in the first term of (24.23) and using bilinearity gives the following
estimate valid for > 1:

+l- -

1 ()

()

A(,n+, n+
.
+ (Z+go,e[Au]).

,+g)
(24.24)

For > , the change in kinetic energy during the time step is easily estimated by
means of the extrapolation formula (23.1 la), the same identity repeatedly used above

282

J.C. Simo

CHAPTER 11

and Eq. (23.10)1 in Stage I leads to

- v,v,,+v ) - (20- I)K(v ,+,)


1

K(v+l) - K(v,,) =
1

(24.25)

Au,+9).
- vr,
V(t(vn+V

<

Finally, we observe that under the assumption of dead loads, the change in the potential
energy of the loading can be written as

) - Vext(Un)

Vet(U7,+l

(f,

Au)

+ (t,

(24.26)

Au)F.

According to (24.18a), the total energy with the time step [tT, tn+l] is obtained by
adding the contributions arising from (24.21), (24.25) and (24.26). By inserting (24.22)
and (24.24) into (24.21) and cancelling terms using the momentum equation (23.10)
in Stage I we arrive at the following estimate valid for 19 > 1:

E(xn+l

- E(Xn)

A(Zn+i, n+l -

'+l) +

A(Z.n+,

,+ -

(124.27)

The proof is completed by noting that the algorithmic flow rule (23.7a), arising in the
application of the algorithm to multisurface plasticity, together with Euler's theorem
and the Kuhn-Tucker conditions imply

A(Zn~v,

Zn~l-v

rn+ )

(AjYVfp('ivni), Znvi)

=-E

=- hoJ_Af(+
-

and similarly A(Zn+i, -n+l - E+!)

) d?

= 0; hence the result.

<

0,

(24.28)

REMARK 24.2. The preceding result can be easily generalized to the case of an arbitrary free energy function T (, ~), not necessarily quadratic, but convex in its two
arguments. The required modifications are (i) the closest-point projection and the linear extrapolation formula must be written in terms of the elastic strains and (ii) the
proof is carried out directly in terms of Vit(Ee, ~) rather than the complementary
Helmholtz free energy function En(Z). Since this is the setting adopted in the full
nonlinear theory, further details will be omitted.

SECTION 25

Integration algorithms

283

25. Spatial finite element discretization: An illustration


It is by now well established that displacement based finite element methods may
lead to grossly inaccurate numerical solutions in the presence of constraints such
as incompressibility, or nearly incompressible response, see, e.g., HUGHES ([1987],
Chapter 4) for a review and an illustration of the difficulties involved in the context of linear incompressible elasticity. As first noted in the NAGTEGAAL, PARKS and
RICE [1974], the classical assumption of incompressible plastic flow in metal plasticity is the source of similar numerical difficulties. Finite element approximations based
on mixed variational formulations have provided a useful framework in the context
of which constrained problems can be successfully tackled. A large body of literature exists on the subject, which has its point of departure in the pioneering work of
HERRMANN [1965], TAYLOR, PISTER and HERRMANN [1968], KEY [1969], and NAGTEGAAL, PARK and RICE [1974]. Review accounts of several aspects of this exponentially
growing area can be found in several textbooks, e.g., CIARLET ([1978], Chapter 7)
and more recently in GIRAULT and RAVIART ([1986], Chapter III), JOHNSON ([1987],
Chapter 11), ZIENKIEWICZ and TAYLOR ([1989], Chapter 12). A comprehensive review
of the subject is given in BREZZI and FORTIN [1991].
The goal of this section is to provide an illustration of the implementation of a
particular class of mixed methods, which retain the simplicity and computational convenience afforded by strain driven return mapping algorithms and, at the same time,
properly account for nearly incompressible response. These methods, often referred
to as assumed strain methods, has gained considerable popularity in recent years.
Direct precedents of this methodology can be found in the reduced and selectivereduced integration techniques originally introduced mostly in an ad hoc fashion in
ZIENKIEWICZ, TAYLOR and Too [1971], DOHERTY, WILSON and TAYLOR [1969] for
the nearly incompressible problem. The seminal work on the subject appears in the
paper of NAGTEGAAL, PARKS and RICE [1974]. This approach is subsequently generalized in HUGHES [1980] where a related methodology commonly referred to as the
B-bar method is suggested. A unified approach that identifies these techniques as a
particular class of mixed finite element methods, both for linearized and finite strain
elasticity and plasticity, is described in SIMo, TAYLOR and PISTER [1985]. Within this
context, for instance, B-bar methods are shown to arise from finite element approximations constructed on the basis of a three-field variational formulation. The fact that
general assumed strain methods can be made consistent with a three-field variational
formulation of the Hu-Washizu type was first pointed out in SIMO and HUGHES [1986].
To develop the class of assumed strain methods considered in this section, the algorithmic flow rule and hardening law, as well as the Kuhn-Tucker form of the loading/unloading conditions, are enforced locally rather than via the variational inequality
described in the preceding sections. Accordingly, the internal variables {Pn+ n+l}
are determined locally at the quadrature points via the closest-point projection algorithm with the strain field regarded as the primary "driving" variable. This point of
view proves particularly useful in the generalization of the techniques described herein
to the finite strain regime. To emphasize this latter aspect, we shall adopt throughout
this presentation a general stored energy function W(ee) not necessarily quadratic in

284

J.C. Simo

CHAPTER I

the elastic strain tensor. Following a common usage in the finite element literature,
in the exposition given below tensor notation is replaced by the following matrix and
vector notation:
(i) Rank two tensors in three-dimensional Euclidean space are mapped into column vectors according to the following convention which distinguishes the stress field
from the strain fields:

EP=

(22

2.2

2eP2

eP3

E[1]=

2eP2

3,3

and
2,1

U,2

o'=

0233l

(25.1a)

. 12

'

-n

2eP 3

U1,3 + U3,1

0 13

2eP 3

U2,3 +

023

3,2

(ii) Rank-four tensors are mapped onto matrices. Of particular importance is the
elasticity tensor C = V2 W(e) which in matrix notation takes the explicit form
-C1 11

C1 1 2 2

C1 1 3 3

C1

C2222

C2233

C2212

C2213

C2223

C33 33

C3 3 12

C3 3 13

C3 32 3

C1212

C1312
C1313

C2312
C2313

C =_~

11 2

sym

C1 1 1 3

C1 1 2 3

(25.lb)

C2 3 2 3

(iii) According to the preceding conventions, the rank-four unit tensor I and the
rank-two unit tensor 1 become
I = DIAG[l,1, 1, 1, l, 11

and

1=[1

1 0

O]T .

(25.1c)

(iv) Contraction between rank-two tensors is replaced by dot product, and application of a rank-four tensor to a rank-two tensor reduces to a matrix transformation.
Similar conventions apply to two-dimensional problems including plane strain, plane
stress and axisymmetry.
For simplicity, attention will be restricted to the quasistatic perfectly-plastic case
(i.e., no hardening) and the three-dimensional problem ndim = 3 will be assumed
throughout. The class of mixed finite element approximations of interest is constructed
from the following weak formulation of the elastoplastic problem:

a "+

/2'

- E[7/]- pof r/)


d

(E[Un+l] -

C o [-

'+l

n+l)

t~("--~.iol0,sfll~dn 1 .q~dP=
t
d =

df2 = 0,

+ VW(En+l

En+)] d2 = 0.

(25.2a)
(25.2a)
(25.2b)
(25.2c)

SECTION 25

Integration algorithms

285

These variational equations hold for any admissible displacement test functions 7 E
V, any stress field r E ;T = L 2 (S')6 , any strain field
E V = L2(S2)6, and
are supplemented by the local algebraic constraints that define the return mapping
algorithm, i.e.,
En+I = En + AyVf (VW(E+

AY > O,

f(VW(En+ - Pn+,)

0,

(S- 1)[Pn

EP _])),

(25.2d)

(25.2e)

Ayf (VW(En+ - E)n+l)) = 0.


Observe that in this formulation of the return mapping algorithm for single-surface
perfect plasticity the (assumed) strain tensor En+l E V,, and not the stress tensor is the
field that plays the role of the independent variable. The treatment of viscoplasticity
follows the same lines as rate-independent plasticity and, therefore, further details will
be omitted.
(A) Discontinuous stress and strain interpolations. The finite element formulation
discussed below is based on discontinuous interpolations of stress and strain over a
typical element S2?C RIndi m of a discretization
U=di 2e. Our goal is to recover
a displacement-like finite element architecture. We start by introducing a conforming
finite-dimensional approximating subspace Vh c V for the space of test functions,
along with a finite element space Th for the admissible stress fields defined as
Th =

rh f2

-,

R6: Irhjf2 = S(X)ce for c, E Rm}.

(25.3)

Here, S(x) is a (6 x m) matrix of prescribed functions, generally given in terms of


natural coordinates. Explicit examples will be given below. Similarly, one introduces
a strain finite element approximating subspace Th defined as
h = {Ch:

-- JR26 : ChI,2 =

F(x)ae for a, E Rm ,

(25.4)

where F(x) is a (6 x m) matrix of prescribed functions. In general, S(x) 4 F(x).


The specific form taken by these prescribed functions will be specified subsequently.
In what follows, for notational simplicity, the superscript h will be omitted. Since the
approximations (25.3) and (25.4) are discontinuous over the elements, the variational
equations (25.2b,c) hold for each element 2?. By substituting (25.3) and (25.4) into
(25.2b) and solving for the element parameters a, GCR one obtains a mapping
un+l
E+1 = ~[Un+l] defined as

e[un+l]lC2e = F(x)H-1

ST(x)E[Un+lI] d,

(25.5)

J.C. Simo

286

which defines the element strain field E,+l


n
u,+ll2e in a typical element S2, where
H=J

CHAPTER

in terms of the displacement field

ST(zx)F(x)d.

(25.6)

Similarly, by substituting (25.3) and (25.4) into (25.2c) we find the following finite
element counterpart of the elastic constitutive equations for the stress field within a
typical element F2,:
r,,+

S(X)H-T

d2.

rT(X)VW(E:[u,tI

(25.7)

It follows that, as a result of the discontinuous (assumed) strain and stress interpolations defined by (25.3) and (25.4), one obtains a discrete approximation to the
symmetric gradient operator [-.] defined by (25.5), which we shall denote by [.] in
subsequent developments, along with an algorithmic constitutive equation defined by
(25.7). Remarkably, the following result shows that only the discrete gradient operator
e [.] appears in the final expression for the discretized weak form (25.2a). The algorithmic constitutive equations (25.7) do not enter explicitly in the variational formulation
of the problem and are needed only in the stress recovery phase.
THEOREM 25.1. The momentum balance equation in the assumed strainfinite element
method outlined above has identicalform as in a displacement model, with gradient
operatorE[-] replaced by the discrete gradient operator [-] defined by (25.5), i.e.,

G(un+jleP,,;n
r; ,X

W(t[un,+

-P

+1)

E[q] dV

- Gext(ij) =0,

(25.8)

for all r E Vh, where Gext(.) is the virtual work of the external loads defined by
Gext() =

Pof

r2dQ+ J

t.

dF.

(25.9)

PROOF. Making use of (25.5), (25.6) and (25.7), the variational equation (25.2a) restricted to a typical element Qe can be rewritten as follows:

e[rt] a,+ dQ
=J

= /

[[i] [S()H-T /
VW([u

n+

l] - EP+)

FT(1x)VW([u+i]

[F(x)H - '

-E+i)

ST (x)e[J]

dQ d

dS

dQ

SECTION 25

Integration algorithms

J~Vw(([un+1 ] - En+J)

287

[/] dQ,

(25.10)

which implies the result in the theorem.


Expression (25.8) becomes completely determined once e+ is defined in terms
of En+-s for either s = 1 or s = 2, and the discrete strain field en+l = [Un+l].
This is accomplished by solving at each quadrature point the local equations (25.2d,e)
that define the closest-point projection algorithm. For this purpose the general closestpoint projection iteration described in Section 16 can be employed, reformulated in
strain space, with the strain field evaluated by means of the discrete gradient operator
as En+l = [un+l]. For completeness, a step-by-step summary of the computational
procedure for the case of ideal plasticity is contained in Table 25.1. A derivation of
the expression for the consistent discrete tangent stiffness matrix is given below.

TABLE 25.1

Closest-point projection algorithm in strain space.


(1) Compute the trial stress
If f(0+

= VW(E[Un+l
)
- En

(s3 - 1)[en -

1) < 0. The quadrature point is elastic.

Set P
= EP and terminate the algorithm
Otherwise, the quadrature point is in the plastic regime
Set A-y() = 0 and proceed to Step 2.
(2) Return mapping iteration algorithm (closest-point projection)
(2a) Compute stress residuals for k = 0, 1,...
W = VW (E[u-n1J
n+ll

'(k)

rn+1

e (k))
n+l1

'(
&(k)

( nV3,r+l

n+l

(2b) Perform the following steps while IIr(kl 11> TOL:

Compute consistent (algorithmic) tangent moduli


C(k

2W ([U+l]

[C-l +A

C(k)-

_ En(k)

()Vf(&kn+l)

Update plastic variables with kth-increments

Ay(k+

l)

p(k+l)

n+I

,(k)

= A(k) +!+l

rvf
&)

T~k) (k)

+
+
n
[vf(k) ]T-(k)
Vf(k)

ep (k) +C(~d(
n+1
n-sl

k l[At7(k) Vf(l

n+

-(k)]
- rn'lj

288

J.C. Simo

CHAPTER

REMARK 25.1. The algorithm summarized in Table 25.1 is performed at each quadrature point. Consistency and loading/unloading are therefore established independently
at each quadrature point of the element. For the von Mises yield criterion, either in
three dimensions or in plane strain, convergence is attained in one iteration and the
algorithm reduces to the radial return method of WILKINS [1964].
(B) Linearization. Consistent tangent operator. It remains to determine the expression for the tangent stiffness operator associated with the reduced residual (25.8). Here,
the notion of consistent tangent moduli, introduced in Section 17 and obtained by exact linearization of the closest-point projection algorithm, plays an essential role. The
closed-form expression for the linearization of the residual (25.8) is easily obtained
in view of the following observations:
(i) The discrete gradient operator g[u.+l] defined by (25.5) is a linearfunction of
the displacement field u,+l.
(ii) The plastic strain En+1 and the consistency parameter A7 are nonlinearfunctions of en+l = [un+l] and the plastic strain Ep defined by the algorithm in Table
25.1. Since EP+,-. for either s = 1 or s = 2, is a fixed (given) history, the only
remaining independent variable is displacement field u+ l.
From the preceding two observations it follows that the reduced residual defined by
(25.8) becomes a function of the displacement field u,+l. Setting dun+l = Au and
differentiating expression VW([u,+ - P+l) in the reduced residual (25.8) gives
- EPn+l])]

d [VW ( [,+

(25.11)

Cn+l ([Au] - d+,),

where Cn+l := V 2 W(g[Un+ -EP +) is the tensor of elastic moduli. As in Section 17,
we define the algorithmic elastic moduli C,+l by the expression
Cn+

= [C+ ,1 +AyV

f(VW( [u+I]- E+],))]

(25.12)

Proceeding exactly as in the derivation described in Section 17, de,+1 is determined by


differentiating the return mapping algorithm and enforcing the algorithmic consistency
condition, to arrive at the following result:
Cn+ ([u+]

--+de+ 1

where nn+l:

+l

n+ ()= n i,+) [A],

Cf+l]f+

(25.13)

[Vfn+i]TCn+i Vf+i
The linearization of the reduced equilibrium equation (25.8) in the direction of the
incremental displacement Au G V is now easily obtained via a straightforward application of the chain rule. The result is easily shown to take the following form:
[I

dG (u+ , E,+ 1; 77)


(Cnl -nn+l

nA'n+!)[Au] d.

.
(25.14)

Integration algorithms

SECTION 25

289

Observe that result (25.14) has an identical structure as a displacement model with
the gradient operator e[-] replaced by the discrete gradient operator ~[].
(C) Matrix expressions. To illustrate expressions (25.8) and (25.14) for the residual
and tangent operator in a familiar context, let Na(x), a = 1,..., nnode, denote the
shape functions of a typical element e2C In di m with nnode nodes, so that
nnode

uh

Na()

da.

(25.15)

a=l

Denoting by Ba the matrix expression of the symmetric gradient of the shape functions
Na, the preceding interpolation together with (25.5) imply
nnode

nnode

E[UI]h] = ZBada and

g[uh]jnc =

a=l

EBada,

(25.16a)

a=l

where
Ba = F(x)H- 1

ST(X)Ba dr.

(25.16b)

The contribution of node a in element Qf2to the momentum equation (25.10) is


then given by

ra

=j

Bai{VW([Un+] - En+,)} d,

(25.17a)

where EPn+ is computed from the algorithm in Table 25.1. Finally, the contribution to
the tangent stiffness matrix associated with nodes a and b of element Q2, is obtained
from (25.14) as
kab =

B (Cn+l - n+l nf+l)Bb d,

(25.17b)

where the algorithmic elasticity tensor Cn+l is defined by (25.12).


(D) Variational consistency of assumed strain methods. The foregoing arguments
show that the preceding three-field mixed variational formulation is equivalent to a
generalized displacement method in which the standard discrete gradient matrix Ba
in (25.15)2 is replaced by an assumed matrix Ba defined in the present context by
(25.16a,b). We shall be concerned here with the converse problem, and consider the
conditions for which an assumed strain method with Ba given a priori, not necessarily defined by (25.16a,b), is variationally consistent. The question of variational
consistency is relevant because of the following two reasons:
(i) The convergence analysis of assumed strain methods is brought into correspondence with the analysis of mixed methods for which a large body of literature

J.C. Simo

290

CHAPTER

exists. In particular, the convergence of certain assumed strain methods, such as the
B-method for quasiincompressibility discussed in the next section, is readily settled
once the variational equivalence is established.
(ii) By exploiting the variational consistency of the method, one can develop expressions for the stiffness matrix which are better suited for computation, a point
which will be illustrated in the next section.
To simplify our presentation, attention is focused in what follows on linear elasticity. The results, however, carry over to the nonlinear situation by straightforward
linearization. Thus, consider an assumed strain method in which strains and stresses
are computed according to the following expressions
Ehls 2 = Bde

and

(25.18)

oh e= CBde,

where B is given a priori, and we have employed the following matrix notation
(25.19a)

B = [BIB 2 ... Bnod]


(de)

= [(del

. * (dn,. d.)

(de)

(25.19b)

Assumption (25.18) leads to admissible variations Ch E


(hs

= Bae

and

rh|

e
= CBa
,

and 'rh

CETh given by

(25.20)

im, where ndim < 3 is the spatial dimension of the problem.


for arbitrary ae E RT" " XdeXd
Substitution of (25.18) and (25.20) into the Hu-Washizu variational equations using
(25.2) yields

[I

Ge =ae.

BTCB

dQjd -G.XIJQe

0=ae.

BTC[B
T
- B] dQ] de,

o = ae .J

T [

Cde +

Cde]

(25.21a)
(25.21b)

dS2.

(25.21c)

The third equation is satisfied identically. The second equation on the other hand, is
satisfied if the following condition holds

BTCB

dF =

BTCB

dQ.

(25.22)

This orthogonality condition, first derived in SIMO and HUGHES [1986], furnishes the
requirement for an assumed strain method to be variationally consistent. Note that
the right-hand side of (25.22) yields an equivalent expression for the stiffness matrix
which is better suited for computation than the standard expression for assumed strain

SECTION 26

291

Integrationalgorthms

methods furnished by the left-hand side of (25.22) since B is usually a fully populated
matrix whereas B is sparse. Further details on implementation are discussed below.
26. Application: A class of mixed methods for incompressibility
As an application of the ideas developed above we consider a generalization of the
ideas presented in NAGTEGAAL, PARKS and RICE [1974] within the context of a threefield mixed finite element formulation, following the approach described in SMO,
TAYLOR and PISTER [1985]. The resulting class of mixed finite element methods can be
cast into a format suggested in HUGHES [1980] and leads to methods currently widely
used in large scale inelastic computations, see, e.g., HALLQUIST [1984]. Because of
its practical relevance an outline of this methodology is given below along with two
alternative implementations.
The basic idea is to construct an assumed strain method in which only the dilatationalpart of the displacement gradient is the independent variable. The main
motivation is the development of a finite element scheme that properly accounts for
the incompressibility constraint emanating from the volume preserving nature of plastic flow. The situation is analogous to that found in incompressible elasticity.
(A) Assumed strainand stressfields. According to the preceding ideas, one introduces
a scalar volume-like variable E L 2(Q) and then considers the following assumed
strain field
E = dev[e[u]] +

091,

(26.1)

ndim

where, as usual, dev[] = () - tr()l/ndim denotes the deviator of the indicated


argument. Similarly, one introduces a pressure-like variable p
L 2 ( Q2) so that the
assumed stress field takes the form
r = dev[VW(E - eP)] + pl.

(26.2)

For subsequent developments, it proves convenient to rephrase (26.1) and (26.2) in


an alternative form in terms of projection operators as follows. Define the rank-four
tensors
Pdev

= I---1

ndim

1 and

Pvol,

ndim

1.

(26.3)

The matrix P defines an orthogonal projection which, therefore, satisfies the following
standard properties:
PdevPvol = PvolPdev = 0

and

Pdev + Pvol = I,

(26.4)

together with P2ev = Pdev and PVo1 = Pvo1. It follows that Pdev and Pvo 1 are orthogonal

projections that map a second rank tensor into its deviatoric and spherical parts,

292

J.C. Simo

CHAPTER

respectively. In terms of these projections, the assumed strain and stress fields (26.1)
and (26.2) can be written as
91

E = Pdev [[U]] + -

ndim

a = Pdev [VW(e -

and

P)] + pl.

(26.5)

In the present context, the variational structure of the assumed strain method takes the
following form. We regard {un+ 1, n+t, p,n+l} as the set of independent variables in
the variational problem and, as before, assume that the plastic strains EPn+ are defined
p
locally in terms of X,+l and the history of plastic strain En+
_,8 for either s = 1 or
s = 2, by the local equations (25.2d,e). Then, we have the following result.
TIEOREM 26.1. For the assumed strain and stress fields defined by (26.5) the weak
forms (25.2a,b,c) reduce to:
(E[r] dev[VW(E +l - En+l)] + pdiv[t]) dQ2 - Gext()=
s q(div[un+l] -

,ldQ

" +n + di

0,

0,

(26.6a)
(26.6b)

tr [vw(,+l -Ep+)]

dQ

0,

(26.6c)

for any rl E V, q E L 2(Q) and 0 E L 2((Q).


PROOF. The proof follows by application of properties (26.4). In particular, from
expressions (26.3) and (26.5) we observe that
Pdevl = 0

and

Pdev [E[Un+t] -En+l]

(26.7)

= 0.

By making use of these relations along with properties (26.4) it follows that
r

([Un,,

- En+,+)

= (Pdev7)

([Un+l]

- En+l) + (PvoIr)

= T * (Pdev[E[Un+l] - En+l]) +
=
ndir

(E[Un+l] - En+l)

I (tr[r])1
ndim

(E[u,+l] - e,+l)

(tr[r]) (tr [e[u,+1] -en+l])

= q(div[un+l] - 9n+1),

(26.8)

where q = (l/ndir) tr['r]. This proves the variational equation (26.6b). A similar
argument holds for (26.6a) and (26.6c). [

293

Integration algorithms

SECTION 26

As in the preceding section, we consider discontinuous interpolations of the assumed


stress and strain fields, and introduce the following approximating subspace both for
the volume and the pressure fields:
Tvol = {

E L 2 (/2):

Oe
Fn(
Fr(x)&e;

(26.9)

i},

where FT(x) = [Fl(x),..., m(x)] is a vector of m-prescribed local functions.


Setting

H=

(26.10)

F(x)FT(x) dQ,

the substitution of these interpolations into the variational equations (26.6b,c) then
yields
+] ==rT(x)H

+ di

Pn+l =

)H

Jr(x)

J
F

'

df2,
F(x)div[un+l
1

-- tr [V(E+
~x)
dim

l) ] df2.

(26.11a)

(26.1 lb)

On the basis of these finite element approximations to the volume element zn+l and
the pressure field Pn+, two alternative implementations of the method are possible.
To alleviate the notation in the exposition of these to approaches, we shall omit superscript h. All the fields involved are understood to be approximated via the preceding
finite element discretization.
(B) First implementation: Assumed strain method. Within the present variational
framework, the first implementational procedure hinges on the following alternative
expression for the term involving the pressure field in the weak form (26.6a):

I tr [VW(n+l - p+)]jdv[7] dQ2,

pdiv[rq]d = /2
k

fsiv~rlldn=

En+ = dev[E[un+]] +

div[un+lt]

(26.12a)

ndim

T(x)H - '

ndim

(dv[un+l]),

(26.12b)

F(x)div[un+,]d.

(26.12c)

This result is easily verified by proceeding exactly as in the proof of expression


(25.8). To complete the implementation, we use matrix notation and set
n.ode

nnode

div[u,+l]

adn+,

=
a=l

and

bda+l,

div[u+l] =
a=l

(26.13)

294

J.C. Sino

CHAPTER

11

where the vector ba is defined by

bT

= FT(X)H-

F(x)bTdQ.

(26.14)

Now define assumed strain nodal matrices Ba via the expression:


Ba

[Ba

Ba

Bao + B

1
Tdim

where B=al

a dimn

( ba and

( ba

(26.1.5)

Noting that properties (26.4) imply the relation


B,{dev[VW(en+I

P+)]} = Ba{dev[VW(En+i ~ - P)

},

(26.16)

by making use of (26.4) it follows that the contribution of node a of element Q to


the finite element approximation (26.6a) to the equilibrium equation can be written as
re =

B{

VWE([u+]

- En+I)

(26.17)

d2,

which is the same result obtained in the preceding section, see Eq. (25.17a). Therefore,
the tangent stiffness matrix is again given by expression (25.17b).
(C) Second implementation: Standard mixed method. An alternative implementation
of the present assumed strain method can be developed directly from the mixed formulation (26.6a,b,c). As will be shown below, the main advantage over the previous
implementation is that computations are performed with the standard discrete gradient
matrix B,, which is sparse, instead of the Ba-matrix which is fully populated. The
main steps involved in the present implementation are as follows:
(i) In addition to computing div[un+l], one also computes P,+I by using (26.12b).
(ii) The contribution to the momentum balance equation of node a of a typical
element Qe is now computed by setting
B a {p

re =

+l

(iii) Setting CP = C

+ Pdev [W([unIFj] - eP+ )] } dR,


-

Tl,

(26.18)

the contribution to the tangent stiffness

matrix of nodes a, b of a typical element Q2 is computed according to the expression


as
kah =J

2
Ba [PdevCePPdev] Bb dQ + T(ndim
1

J {BaT [PdevCePl] bb + ba [iT

ba

b](1T C ePl) d2

ePdev] Bb} d2.

(26.19)

SECTION 26

295

Integration algorithms

This result is easily verified by noting that Ba = PdevBa + (l/ndim)l ba and using
properties (26.4).
The advantages of this second implementation become apparent in the common case
encountered in most applications for which the elasticity is linear, with uncoupled
pressure/deviatoric response, and plastic flow is isochoric. The reason is that the
pressure p,+l is trivially evaluated once 9,f+I is computed from (26.1 lb),, since
Pn+l = Ki9n+1

where K = bulk modulus.

(26.20)

Note that no additional computations are involved since only the evaluation of ba
is needed, which is also required in the first implementation. Furthermore, since the
deviatoric response is uncoupled from the bulk response, no coupling terms arise in
the calculation of the stiffness matrix. Consequently, in this situation we have
CeP1 = 1

PdevCePl = 0,

(26.21)

and result (26.19) reduces to


keab =

BT[CeP

/1 ( 1]Bbd

[ba (bb]dQ

(26.22)

Finally, because of the linearity of the pressure/volume elastic response, the calculation
of the second term in (26.22) reduces to a mere rank-one update. This observation
follows from (26.15)1 and definition (26.10) of H, since

;J,

[ba bb] d
= n[

(x) 0

ba

df2TH-'

F(x)

bb dQJ.

(26.23)

This is the implementation adopted in the numerical simulations described below.


EXAMPLE 26.1. The finite element method outlined in this section includes, as a particular case, classical mixed methods for which a rather complete accuracy, stability
and convergence analysis currently exist. Two standard examples are:
(i) Four-node quadrilateral element with bi-linear isoparametric interpolation functions Na for displacements, and F = [1] constant over Q2e. Essentially, this is the mean
dilatation formulation advocated by NAGTEGAAL, PARKS and RICE [1974]. This is a
widely used element known not to satisfy the BB condition. However, application
of a pressure filtering procedure renders the discrete pressure field convergent; see
PITKARANTA and STERNBERG [1984].

(ii) Nine-node element with bi-quadratic isoparametric interpolation functions Na


for displacements, and F = [1 x y]T, where (x, y) are defined in terms of natural coordinates by means of the standard isoparametric mapping. This is an optimal element
known to satisfy the BB condition, see BREZZI and FORTIN [1991].

J.C. Simo

296

CHAPTER

The convergence properties of the class of assumed strain methods described in


this section are the direct consequence of their variational consistency, and follows at
once from the convergence characteristics of the corresponding mixed finite element
formulations.
27. Illustrative numerical simulations
A number of numerical simulations are presented that illustrate the performance of the
return mapping algorithms and the practical importance of consistent tangent operators
in a Newton solution procedure. These simulations exhibit the significant loss in rate
of convergence that occurs when the elastoplastic "continuum" tangent is used in place
of the tangent consistently derived from the integration algorithm. The overall robustness of the algorithm is significantly enhanced by combining the classical Newton
procedure with a line search algorithm. This strategy has been suggested by a number of authors, e.g., see DENNIS and SCHNABEL [1983], or LUENBERGER [1984]. The

specific algorithm used is a linear line search which is invoked whenever a computed
energy norm is more than 0.6 of a previous value in the load step (see MATTHIES and
STRANG [1979]). Computations are performed with an enhanced version of the general
purpose nonlinear finite element computer program FEAP, developed by R.L. Taylor
and the author, and described in ZIENKIEWICZ and TAYLOR [1989]. Convergence is

measured in terms of the (discrete) energy norm, which is computed from the residual
vector R(d+l) and the incremental nodal displacement vector Ad,+ as
AE(d(')

= [Ad+)] T R(d) +l).

(27.1)

Alternative discrete norms may be used in place of (27.1), in particular the Euclidean
norm of the residual force vector. In terms of the energy norm (27.1) our termination
criteria for the Newton solution strategy takes the following form
AE(d() )

<

10-9AE(d()l ).

(27.2)

While it would appear that this convergence criterion provides an exceedingly severe
condition difficult to satisfy, it will be shown through numerical examples that criterion
(27.2) is easily satisfied with a rather small number of iterations when the consistent
tangent operator is used.
All the numerical simulations described below employ a four-node quadrilateral
with bilinear isoparametric interpolations for the displacement field and discontinuous
pressure assumed and volume fields assumed to be constant within a typical element.
The results for the stress field and the internal variables reported below are computed
via an L 2-projection of the values at the quadrature points of a typical element onto the
nodes that uses the bilinear interpolation functions. Related "smoothing" procedures
are discussed in ZIENKEWICZ ([1977], Section 11.5, and references therein), and are
often used as a device for filtering spurious pressure modes of this particular element
(e.g., LEE, GRESHO and SAM [1979]).

Integration algorithms

SECTION 27

297

(A) Thick wall cylinder subject to internalpressure. An infinitely long thick-walled


cylinder with a 5 m inner radius and a 15 m outer radius is subject to internal pressure.
The properties of the material are E = 70 MPa, v = 0.2. The isotropic and kinematic
hardening rules are of the exponential type, defined according to the expression:
K() = K

- [K

- Ko] exp[-b] + H,

~=
~~~({~)

~K'
H() = (1 - 3)K((),

~K ~(27.3)
(),

EC[0,1].

The values i3 = 0 and 3 = 1 correspond to the limiting cases of pure kinematic


and pure isotropic hardening rules, respectively. The simulation is performed with
classical backward Euler radial return method described in Section 18 (corresponding
to s = 1), with the parameters in (27.3) chosen as

Ko = 0.2437 MPa,
6 = 0.1,

K,

H=0.15MPa,

= 0.343 MPa,
(27.4)

3 = 0.1.

The internal pressure is increased linearly in time until the entire cylinder yields.
The finite element mesh employed in the calculation is shown in Fig. 27.1(a). The size
of the time step is selected as to achieve yielding of the entire transverse section in
two time steps involving plastic deformation. The stress distribution in the transverse
direction (i.e., all) at time t = 0.0875, corresponding to an elastic-plastic solution,
is shown in Fig. 27.1(b). The calculation is performed with both the "continuum"
and the "consistent" elastoplastic tangent, and the results are shown in Table 27.1. In
spite of the better performance exhibited by the "consistent" tangent, no substantial
reduction in the required number of iterations for convergence is obtained except in

(a)

(b)

STRESS 1
Min = -2.85E-01
Max = -1.24E-03

I111! IHHH HIIHN HTHII HI

PUMPZ

-2.45E-01
-2.04E-01
-1.64E-01
-1.23E-01
-8.24E-02
-4.18E-02

Current View
Min=-2.85E-01
X = 5.OOE+00
Y = .OOE+00
Max = -1.24E-03
X = 1.50E+01
Y = .OOE+00

FIG. 27.1. Thick-wall cylinder. (a) Finite element mesh consisting of 40 elements (Q/PO). (b) Contours of
the stress component Arl at time t = 0.0875 (backward Euler return mapping).

298

J.C. Simo

CHAPTER

TABLE 27.1

Simulation (A). Iterations for each time step.


Step

State
Continuum

el
2

el-pl
6

el-pl
9

pl
10

pl
6

Consistent

FINITE ELEMENT MESH (780 nodes, 722 elements)

FIG. 27.2. Plane strain strip with a circular hole. Finite element discretization. Stabilized mixed method
(Qi/PO) with backward Euler return mapping.

the fully plastic situation. This is due to the extreme simplicity and well-posedness of
the boundary value problem at hand: essentially one-dimensional. The next example
will confirm this observation.
(B) Perforatedstrip subject to uniaxialextension. We consider the plane strain problem of an infinitely long rectangular strip, of dimensions 10 x 12m, with a circular
hole of radius R = 5 m in its axial direction, subjected to increasing extension in a direction perpendicular to the axis of the strip and parallel to one of its sides. The elastic
properties of the material are taken as E = 70MPa, v = 0.2, and the parameters in
the saturation type of hardening rule (27.3) are K 0 = Kc = 0.243 MPa, H = 0, and
a = 1 (perfectly plastic behavior). Loading is performed by controlling the vertical displacement of the top and bottom boundaries of the rectangular strip. The finite element
mesh is shown in Fig. 27.2. For obvious symmetry reasons, only 1 of the strip is considered. The contours of shear stress component r12 are shown in Fig. 27.3 after 5 steps
of At = 0.0125, computed with the classical radial return algorithm (backward Euler).

SECTION 27

Integration algorithms

299

STRESS 4
Min = -3.72E-02
Max= 1.74E-01
-7.04E-03
2.32E-02
5.34E-02
8.36E-02

1.14E-01
1.44E-01

Current View
Min = -3.72E-02
X = 6.18E-01
Y = 4.96E+00
Max = 1.74E-01
X = 6.05E+00
Y= .OOE+00

FIG. 27.3. Plane strain strip with a circular hole. Contours of shear stress after 5 steps of At = 0.0125
(backward Euler return mapping).

STRESS 5
Min = .OOE+00
Max= 6.08E-02

|0zqle -ii

5.50E-02
,,
5.52E-02
5.54E-02
5.56E-02
5.58E-02
5.60E-02

<

Current View
Min = .00E+00
X = 4.84E+00
Y = 2.66E+00
Max= 6.08E-02
X = 6.16E+00
Y= 3.41E+00

FIG. 27.4. Plane strain strip with a circular hole. Elastic-plastic boundary obtained in single time step of
At = 1.5 with backward Euler return mapping.

300

CHAPTER II

AC. Simo
TABLE 27.2

Simulation (B). Iterations for each time step.


Step

State
Continuum
Consistent

el
2
2

el-pl
13
5

el-pl
23
5

el-pl
23
4

el-pl
22
5

FINITE ELEMENT MESH (765 nodes, 704 elements)


FIG. 27.5. Plane stress strip with a circular hole. Finite element mesh. Standard Galerkin finite element

method.

The calculation is performed with both the "continuum" and the "consistent" tangent
operators, and the number of iterations required to attain convergence is summarized
in Table 27.2. The superior performance of the "consistent" tangent is apparent from
these results. This example also provides a severe test for the global performance of
the Newton solution strategy. Although the calculation is completed successfully with
a time step of At = 0.0125, for twice this value of the time step the iteration procedure diverges. However, when the Newton solution procedure is combined with a line
search procedure, as described in MATTHIES and STRANG [1979], global convergence
is attained in a single time step of size At = 1.5. The elastic-plastic interface is shown
in Fig. 27.4.
(C) Extension of a strip with a circular hole in plane stress. The geometry and finite
element mesh for the problem considered are shown in Fig. 27.5. A unit thickness is
assumed and the calculation is performed by imposing uniform displacement control

SECTION 27

301

Integration algorithms

STRESS 5
Min = 6.82E-02
Max= 2.11E+00
1

.;

~9.99E-01

9.99E-01
9.99E-01
1.OOE+00
1.OOE+00
1.OOE+00

Current View
Min = 6.82E-02
X = .OOE+00
Y = 8.06E+00

X Max= 2.11E+00
X = 4.71E+00
Y = 1.68E+00

FIG. 27.6. Plane stress strip with a circular hole. Elastic-plastic interface after two time steps of At = 0.06.
Backward Euler return mapping algorithm.

on the upper boundary. For obvious symmetry considerations only one-quarter of the
specimen need be analyzed. A total of 164 4-node isoparametric quadrilaterals with bilinear interpolation of the displacement field are employed in the calculation. It should
be noted that for plane stress problems no special treatment of the incompressibility
constraint is needed. A von Mises yield condition with linear isotropic hardening is
considered together with the return mapping algorithm outlined in Section 18 for s = 1
(classical backward Euler method). The elastic constants and nonzero parameters in
hardening law (27.3) are E = 70, v = 0.2, K0o = K
= 0.243, H = 2.24, and /3 = 1.
The problem is first solved using prescribed increments of vertical displacement on
the upper boundary of 0.04 followed by three subsequent equal increments of 0.01.
The resulting spread of the plastic zone is shown in Fig. 27.6. Note that spread of
the plastic zone across an entire cross-section is achieved in the third load increment.
Only seven iterations per time step are required to reduce the the H-energy norm
of the residual to a value of the order 1 x 10 - 3 3 at an asymptotic rate of quadratic
convergence. No line search is required for this step-size. Finally, as in the preceding
example, the robustness of a solution procedure that combines Newton's method with
line search and the exact algorithmic moduli is demonstrated by solving the problem
in a single time step of size At = 1.5. Although the entire specimen is in the fully
plastic regime during the first two Newton iterations, the solution procedure is able
to produce a solution of the problem for this extremely large time step.

302

J.C. Simo

CHAPTER 11

STRESS 3
Min = -435E-02
Max= 3.61E-01
1.43E-02
7.21E-02

1.30E-01
1.88E-01
2.46E-01
3.03E-01

Current View
Min = -4.35E-02
X = 2.78E+00
Y = 4.16E+00
Max= 3.61 E-01
X = 5.00E+00
Y= .OOE+00

FIG. 27.7. Plane stress strip with a circular hole. Shear stress contours for a solution obtained with two
time steps of At = 0.06. Backward Euler return mapping.

STRESS 5
Min= 2.15E-01
Max= 1.22E+01
-

;
-

9.90E-01

9.92E-01
9.94E-01
9.96E-01
9.98E-01
1.OOE+00

Current View
Min = 2.15E-01
X = .OOE+00
Y = 6.75E+00
Max= 1.22E+01
X = 1.00E+01
Y = 7.50E+00

FIG. 27.8. Plane stress strip with a circular hole. Elastic-plastic interface for a single time step of At - 1.5.
Backward Euler return mapping.

CHAPTER III

Nonlinear Continuum Mechanics and


Multiplicative Plasticity at Finite
Strains
While the formulation and numerical analysis of classical plasticity are now well
established, extensions to the finite strain regime have been the subject of some controversy only settled, at least partially, in recent years. Up to the mid-eighties, conventional schemes for finite strain plasticity exploited a formulation of the elastic stress
relations directly in rate-form, based on the notion of a hypoelastic material (see
TRUESDELL and NOLL [1972]). See, for example the review articles of NEEDLEMAN
and TVERGAARD [1984] and HUGHES [1984]. More recently, computational approaches
directly based on mechanically sound models employing a multiplicative decomposition have received considerable attention in the literature. The idea of exploiting the
multiplicative decomposition in a computational setting is suggested in ARGYRIS and
DOLTSINIS [1979, 1980], within the context of the so-called naturalformulation,but
is subsequently abandoned in favor of hypoelastic rate models; see ARGYRIS, DOLTSINIS, PIMENTA and WiUSTENBERG ([1982], p. 22). Independently, the first successful
computational approach entirely based on the multiplicative decomposition appears in
SIMO and ORTIZ [1985] and SIMO [1986].
A fairly complete account of the current status of numerical analysis issues in finite
strain plasticity will be given in the following chapter. The objective of this chapter
is to outline the continuum basis and the mathematical structure underlying models
of plasticity at finite strains based on a multiplicative factorization of the deformation
gradient. The material in this chapter is organized as follows.
First, to make the exposition self-contained, a number of basic results on nonlinear
continuum mechanics used in the subsequent treatment of nonlinear plasticity are
briefly summarized. This account is intended to provide only a brief overview of
those aspects of the nonlinear theory relevant to the problem at hand and should
not, by any means, be considered exhaustive. Detailed expositions of the subject can
be found in the classical treatises of TRUESDELL and TOUPIN [1960], TRUESDELL
and NOLL [1965], and the more recent accounts of GURTIN [1981], OGDEN [1984],
303

304

J. C. Simo

CHAPTER III

and CIARLET [1988]. For detailed expositions of the geometric structure underlying
continuum mechanics see MARSDEN and HUGHES [1983], and SMO, MARSDEN and
KRISHNAPRASAD [1988].

A detailed formulation of micromechanically based constitutive models of finite


strain multiplicative plasticity is considered next. The last part of this chapter then
summarizes the weak formulation, in the Lagrangian setting, of the quasilinear initial boundary value problem arising in nonlinear solid mechanics and describes the
formulation of dynamic plasticity at finite strains. Topics treated there include a (formal) a priori stability estimate that demonstrates the dissipative character of the initial
boundary value problem along with a description of the structure of the incremental
or rate version of this problem.
28. Basic kinematic results in nonlinear continuum mechanics
We shall denote by Q C IRnd ' , 1 < ndim 3, the reference placement of a continuum
body with particles labelled by X E 2. For present purposes it suffices to regard Q2
as an open bounded set in RIndim with smooth boundary a2. A smooth deformation
of the continuum body is an embedding of the reference placement into R 3, i.e., an
injective, orientation preserving map qo: - I i' m. The set S = Wo(Q) is referred
to as the current placement of the continuum body, with points designated by x C
S. The deformation gradient F: 2 - GIL+(ndim) is the Frechet derivative of the
deformation o. We use the standard notation
F(X) = DWp(X)

with J(X) = det [F(X)] > 0

(28.1)

and, following a standard convention, often omit the explicit indication of the argument X. The orientation preserving condition J(X) > 0 models the local impenetrability of matter. The right and left Cauchy-Green tensors are mappings from Q2 into
S+ respectively defined as
C= FTF and

b= FFT.

(28.2)

According to the polar decomposition theorem, a special case of the singular value
decomposition, at any X G Q2 the deformation gradient can be decomposed as
F(X) = R(X)U(X) = V(X)R(X),

(28.3)

where R(X) G SO(ndim) is a proper orthogonal tensor, called the rotation tensor,
and U(X) E $+, V(X)
S+ are symmetric positive definite tensors called the right
and left stretching tensors, respectively. We shall denote by IA (A = 1,.. ., nldim) the

principal invariants of either C or b. For din = 3 the principal invariants are defined
by the standard expressions
I, = tr[C],

= (I12
I7

- tr[C] 2 )

and

13 = det[C].

(28.4)

Nonlinear continuum mechanics

SECTION 28

305

Since both C and b are symmetric and positive definite at each X E Q2, a standard
result in linear algebra implies the existence of the spectral decompositions
ndim

C=

ndim

A2N(A) N(A)

and

S2n(A)

b=E

A=l

n(A),

(28.5)

A=1

where A > 0 are the eigenvalues of either C or b, while N(A) and n(A) are the
respective unit eigenvectors as defined by the eigenvalue problems
CN(A) =

A2N(A)

and

bn(A) = A2 n(A)

(A = 1,..

.,

ndim).

(28.6)

The polar decomposition theorem together with the unit length normalization of the
eigenvectors imply the relation

b = RCRT

so that n(A) = RN(A) with IN(A) I = In(A) I = 1.

(28.7)

For ndim = 3 the triad {N(1), N(2 ), N( 3 )} defines the so-called Lagrangian axes or
principal directions at a point X E Q2 in the reference placement. On the other hand,
the triad {n(l), n(2 ), n(3)} defines the so-called Eulerian axes located at x = qo(X).
The spectral decompositions of the deformation gradient and the rotation tensor then
take the form
fndim

F =

ndim

AAn(A)

N(A)

and

n(

R = E

A=l

A)

N(A).

(28.8)

A=1

The eigenvalues AA (A = 1,.. ., ndim) are also known as the principal stretches at a
point, which are directed along the principal material directions N(A). The squares
of the principal stretches (i.e., the eigenvalues of either C or b) are the roots of the
characteristic polynomial defined, for ndim = 3, as
p(A2 ) =,6

_ I114 + 22 _ 3.

(28.9)

Finally, from (28.6) the spectral decompositions of the right and left stretching tensor
are given by
3

.
A=1

3
AAN

(A )

N()

and

V =

An(A

)n(A)

(28.10)

A=1

The preceding decompositions play a crucial role in the closed-form numerical implementation of the polar decomposition discussed below.
First, the roots of the characteristic polynomial can be computed in closed form via
the well-known solution for a cubic equation, as summarized in Table 28.1. Second,

306

IJ.C. Simo

CHAPTER II

the rank-one matrices of principal directions can also be computed in closed form via
the following result. (The case ndim = 3 is assumed throughout).
THEOREM 28.1. The rank-one tensors of principal directions associated with b are
given by:
Case 1. Three different roots,

A3 :

Al LC

[B

where A f B

#) A2

AJ

C denotes a cyclic permutation of the indices { 1,2, 3}.

Case 2. Two equal roots, A: = A1


3 n
L3)[1--A

A2

A3 :

(3)

Case 3. Three equal roots, A: =

n(3)] = [A2

A2].

A2 = A3: b = A2 1.

Al

(28.11b)

(28.1Ic)

PROOF. Consider first Case 1. The spectral decomposition of b implies


3

[b

Al ] =

(A2

-2

(A

)n(A)

n(A)

(28.12a)

2)n(A) O

n A).

(28.12b)

A=A#B

and
[b-

A21]

=
A=l
AOC

Result (28.11a) then follows by multiplying (28.12a,b) and noting that n(A) n(B) =
AB. Similarly, for Case 2, the spectral decomposition of b now gives
[b - A2 1] = (A2 - A2 )n( 3 )

n(3)

(28.13a)

and
[b - 321] = (A2 - A2 ) [1-n()

(3)]
X

(28.13b)

which implies (28.11b). Case 3 is obvious since b is a spherical tensor.


Note that the formula for the principal direction associated with a single root remains
uniformly valid in the two cases above since squaring (28.13a) gives (28.11la). The

307

Nonlinear continuum mechanics

SECTION 28

preceding formulae can be used to obtain explicit expressions leading to a closedform algorithm for any isotropicfunction of C and b; see MORMAN [1987] and SIMO
and TAYLOR [1991]. These expressions, although explicit, depend on the number of
repeated eigenvalues. However, for the case of the square root, the preceding formulae
can be combined into a unified, singularity-free expression which encompasses all
three different cases. In fact, U can be written as (see HOGER and CARLSON [1984a,b],
and TING [1985])
U =

-- i

3[ [-c
-

2 +

(i - i2 )C + ili31],

(28.14)

where iA (A = 1,2, 3) are the principal invariants of U. A similar expression can


be derived for the inverse tensor U - as is given in Table 28.1 where the closedform algorithm for the polar decomposition is summarized. The derivation of these
formulae involves a systematic use of the Cayley-Hamilton theorem.
(A) Motions. The Lagrangian description. Next, we summarize a number of results
pertaining to the motion of continuum bodies. Recall that a motion of a continuum
body is a one-parameter family of configurations indexed by time. Explicitly, let I1C
R+ be the time interval of interest. Then, for each t G II,the mapping pt : Q2 -- ]Rndim
is a deformation which maps the reference placement Q onto the current placement
St = pt(Q) c IR3 at time t. We write
x =

t(X) =

(28.15)

(X,t)

for the position of X cE Q at time t. In the Lagrangian (or material) description of the
motion, the coordinates that define the position of particles in the reference placement
Q of the continuum body are taken as the independent variables. Accordingly, the
material velocity, denoted by V(X, t), is the time derivative of the motion while the
material accelerationis the time derivative of the material velocity, i.e.,
V(X,t) = atqp(X,t)

and

A(X, t) = atV(X,t) = aW(X,t).

(28.16)

For fixed time Vt = V(., t) and At = A(, t) define the material velocity and
acceleration fields at time t E II,respectively. Thus, the motion, the material velocity,
and the material acceleration fields are associated with material points X E 2, and
hence parameterized by material coordinates. In components we have
Va(XA, t) =

at

(a(XA,t)

and

Aa(XA,t)=

at2

(28.17)

where {XA} = (XI, X2 , X3 ) are the Cartesian coordinates of a material point X G Q2


relative to an inertial frame, and {za} = (, X2, X3) are the coordinates of a point
x
St. When the material coordinates {XA} are the independent variables one
speaks of the Lagrangian or material description of the motion. Note that the mapping
t
I 1- qt(X)]x=fixed gives the trajectory of the material point X in the time
interval II.

308

J.C. Simo

CHAPTER III

TABLE 28.1

Algorithm for the polar decomposition.


(i) Compute the squares of the principal stretches 12 (A = 1, 2, 3) (the eigenvalues of C) by solving
(in closed form) the characteristic polynomial as follows:
Define the coefficients
b = 12 - I/3,

c -27I+

-I3.

If (bl < TOL) then set


XA = -C/

Otherwise, if (bl > TOL) then set


m=2-b/3,
n=

3c
mb'

t= arctan [1-/n2/g] /3,


XA = m cos [t + 2(A - 1)r/3].

The eigenvalues are then


IA =

A + 11/3.

(ii) Closed form evaluation of the stretch tensor U.


Compute the invariants of U
il = 11+ 12 + 13,
i 2 = 1112 + 1113 + 1213,
i3 = 111213.

Set
D = ili2 - 3 = (

U=- l

+ 12 )(1
1 + 13)(12 + 13) > 0.

- C2 + (i 2 -

-[C23

iU+i21

2)

C+

31],

.]

(iii) Closed-form evaluation of the rotation tensor R.


R= FU 1 .
/- n2) = arctan (/1 - n2/n) should be used instead of the arccosine
function dacos (n) in either FORTRAN or C implementations to avoid the ill conditioning near the
origin.

* The function datan2 (n,

Nonlinear continuum mechanics

SECTION 28

309

(B) The Eulerian description of the motion. The Eulerian or purely spatial description of the motion is obtained from the material description by changing the independent variables from fixed material coordinates (of a particle) to current positions
in Euclidean space. At any time t G the spatial velocity and acceleration fields,
respectively denoted by v(x, t) and a(x, t), are obtained via the change of variables
V(X, t) = v (o(X,t),t)

and

A(X,t) = a(qo(X,t),t).

(28.18)

Since Vt is a deformation for all t, the condition J = det[Dot] > 0 in 2 insures that
Sot is invertible and the preceding relations become
v=Voqo,

and

a=Ao ot 1.

(28.19)

The material time derivative of a spatial field, denoted by a superposed dot, is the
time derivative holding the particle fixed. The material time derivative of the spatial
velocity field is the spatial acceleration field. By the chain rule, one has the well-known
relation
i = a

atv

+ (v V)v.

(28.20)

The tensor Vv(x, t) with components aVa(xb, t)/aXb is referred to as the spatial velocity gradient. Here the notation at (.)(x, t) implies taking the partial time derivative
of the spatial field holding the current position x fixed. In general if a (x, t) is a
spatial tensor field its material time derivative, denoted by &(x, t), is defined by the
formula
a& = [ (a o

(28.21)

t)] o p'.

With a slight abuse in notation, a superposed dot will often be used to designate both
the material time derivative of a spatial field and the time derivative of a material
field. Thus, for a material field (.)(X, t), the notations at(.) and (.) are equivalent.
(C) Classicalrate of deformation tensors. The time rate of change of the deformation
gradient F = D t is obtained from (28.1) as atF = F = GRAD[V]. One refers to
GRAD[V], with components av, /aXB, as the materialvelocity gradient.By the chain
rule

F = GRAD[V] = GRAD[V o ot] = (Vv o ot) Do, = (Vv o

tp,)F.

(28.22)

This relation implies the expression Vv = FF - o -' for the spatial velocity
gradient. The symmetric part of the spatial velocity gradient, denoted by d, is known
as the rate of deformation tensor whereas its skew-symmetric part, denoted by wv, is
referred to as the spin tensor. Since the Lagrangian description of the motion is almost

310

JC. Simo

CHAPTER

TI

universally used in solid mechanics, we append a composition with the motion to the
usual definitions of d and iv, and set
d = I [Vv + (Vv)T] o Wt

and

:ii
=

[v

- VVT]

Pt

(28.23)

The spatial vorticity field w is defined as twice the axial vector of the Eulerian spin
tensor i o pt 1. Recall that the axial vector of a skew-symmetric tensor is uniquely
characterized by the relation ih = w x h, for any h G IR3. This definition yields
the standard result w (x, t) = curl[v(x, t)] for the spatial vorticity field. The following
relation will be useful in our algorithmic treatment of plasticity. From (28.22) and
(28.2) one concludes that
+ FTF = FT [(Vv

= TF

t)T + Vv o Pt] F = 2 FTdF.

(28.24)

The result C = 2FTdF justifies the name material rate of deformation tensor often
given to 2C.
(D) The rotated rate of deformation tensor.

The local configuration obtained by

applying the rotation tensor to a neighborhood Ox of a point x in the current configuration St is known as the rotated configuration. Such a configuration is local in
the sense that the "rotated neighborhoods" do not "fit together" unless the deformation of the body is homogeneous. The rotated rate of deformation tensor is defined as

DR = RTdR. An alternative expression can be derived using the polar decomposition


as follows:
C = 2URTdRU <

DR = RTdR = C1 /2CC-L/2.

(28.25)

The rate of change of the rotation tensor is given in terms of a skew-symmetric tensor
Q according to the conventional expression
= RRT

with

(28.26)

+ hT= 0.

Since n(A) = RN(A), the vector field 1? associated with the skew-symmetric tensor

field Q, defined by the relation ha = x a for any a i R 3, gives the relative angular
velocity between the Eulerian and Lagrangian triads. This result can be verified as
follows. Denote by W the angular velocity of the Lagrangian triad, so that
N(A) = W x N(A)

and set

= RW.

(28.27)

Time differentiation of the relation n(A) = RN(A) and use of (28.27) together with

(28.26) and the well-known property Q(al x a2) = Qal x Qa2, for any Q E SO(3),
then gives
i:(A) =

? X n(A) +

R(W x N(A)) = ( +Q-) x n(A)

(28.28)

Nonlinear continuum mechanics

SECTION 29

311

It is clear from (28.27) that 2 gives the velocity of the Lagrangian axes expressed
in the Eulerian triad and, therefore, measures a physical quantity that has little to do
with the (spatial) vorticity field.
REMARK 28.1. Likewise, the skew-symmetric tensor h does not coincide with the
spin tensor iv. In fact, a direct computation gives
F = RRTRU + RUU lRTF = [

+ (R&U -)RT]F.

(28.29)

Combining (28.23)2 and (28.29) results in the expression


w = D + Rskew[UU-']RT ,

(28.30)

where skew[-] indicates the skew-symmetric part of the rank two tensor []. The
physical significance of this result agrees with the fact that i, and h measure different
physical objects. While the spin tensor gives the instantaneous angular velocity of the
triad defined by the eigenvectors of the rate of reformation tensor d, the tensor w
measures the relative velocity of the Eulerian axes n(A) (obtained by "freezing" the
Lagrangian axes). Expression relating wiiand i2 to different kinematic tensors can be
found in MEHRABADI and NEMAT-NASSER [1987].
29. Stress tensors and alternative forms of the equations of motion
We denote by o the Cauchy stress tensor, a symmetric tensor field defined on the
current configuration St of the body. In addition, we denote by P the nonsymmetric
nominal stress tensor, also known as the first Piola-Kirchhoff tensor. This tensor is
relevant to a Lagrangian description of continuum mechanics. We have the standard
relations
= PFT

and

r=Jo

o t.

(29.1)

The stress field , parameterized by material coordinates but defined on the current
placement, is known as the Kirchhoff stress tensor field and differs from the Cauchy
stress tensor field by a factor of J and a composition with the motion. The convected
second Piola-Kirchhoffstress tensor field S, parameterized by material coordinates
and defined on the reference placement, is defined by
S = F-'P= F-lTF
T.

(29.2)

In index notation one has the component expressions Tab FaASABFbB = PaAFbA.
All the stress tensors introduced above are conjugate to associated rate of deformation
tensors through the following important stress-power relations
7Pint = r d = P P= 1S -.

(29.3)

312

J.C. Simo

CHAPTER

In index notation this relation reads Pint = Tab dab = PaAFaA = SABCAB
29.1. The nonsymmetric stress tensor defined by the relation T = RTP =
US and called the Biot stress, is a stress measure preferred by a number of authors,
see, e.g., OGDEN [1984]. To find its conjugate strain measure one introduces the polar
decomposition to obtain
REMARK

P F=P

(RTRU + RU)

- PF T

+ RTP

+ sym[T] U,

(29.4)

where sym[. ] indicates the symmetric part. Since r is symmetric and ? = RRT is
skew-symmetric, it follows that r
= 0 and P. F = sym[T] U. Therefore, the
(symmetric part) of T is conjugate to the right stretching tensor. The skew-symmetric
part of the Biot stress is a reaction force in the sense that it does not enter in the
expression for the stress power. Other stress tensors and conjugate strain measures
can be introduced by a formalism due to Hill; see OGDEN [1984].
The two forms of the local equations used in numerical treatments of continuum
mechanics problems are the Lagrangian and the Eulerian description. In the Lagrangian
formulation-the primary description used in solid mechanics-the motion itself is
the primary variable in the problem, and the density function p0: 2 - R+ in the
reference placement is a part of the given data. In the Eulerian description the spatial
velocity field and the current density Pt: St x --- R+ are the primary variables in
the problem.
(A) Lagrangian description. Let f2? denote the boundary of Q. Assume that the
deformation is prescribed on r, c a0 while the nominal traction vector tN is
prescribed on the part of the boundary FT C
2, with unit outward normal N,
as

so =

on Fr x

and

tA = PN = t

on FT x ,

(29.5)

where r,.n rT = 0 and F U rFT = an2. The local equations of motion in the
Lagrangian description then reduce to the first-order system

at(p~o,)

poV

in Q2 x I,

(29.6)

at(poV) = DIV[P] + B
where B(., t) is the nominal body force per unit of reference volume and DIV[.] is the
divergence operator in material coordinates, i.e., DIV[P] = PA/XA. Conservation
of mass in the Lagrangian description is merely the statement that atpo = 0 in Q2.

Nonlinear continuum mechanics

SECTION 30

313

(B) Euleriandescription. In the purely spatial or Eulerian description, the counterpart


of Eqs. (29.6) is formulated on the current placement St = Wt(2) of the continuum
body, with the motion qo and its time derivative V replaced by the spatial density Pt
and the spatial velocity field v as the independent variables. Equation (29.6)1 is then
replaced by balance of mass, obtained by time differentiation of the basic relation
po = J(Pt o ot) and use of the chain rule leading to
atpt + (Vpt) v + ptdiv[v] = 0 in St x .

(29.7)

In arriving at this result, use is made of the well-known relation t J = J div[v] ocpt for
the time derivative of J = det[Dot]. Observe that, according to (28.21), the first two
terms in (29.7) comprise the material time derivative of the spatial density function
pt(x) = p(x, t). The Eulerian form of the equation of motion (29.6)2 is obtained
from (29.6)1 by the change of variables X - x = Wot(X). Using the well-known
Piola transformation DIv[P] = Jdiv[to] o Wot and expression (28.20) for the spatial
acceleration field, the continuity equation (29.7) and spatial version of the momentum
equation (29.6) yield the following first-order system written in conservation form:
atpt = -div[ptv]

Or(ptv) = -div[ptv v] + div[a-] + bt

in St x II,

(29.8)

where bt = (B/J)o o - 1 is the body force per unit of volume in the current placement
and div[-] is the divergence operator in spatial coordinates, i.e., (div[a])a = aab/aXb.
REMARK 29.2. The spatial or Eulerian form of the equations is not suited for the

formulation of the initial boundary value problem in solid mechanics in the presence
of anisotropic constitutive response (this notion is briefly described below). The reason
being that it is not possible to characterize the constitutive response of an anisotropic
solid solely in terms of information on its current placement St. Such a characterization
is possible only if the material is isotropic. The assumption of isotropy has wide
applicability in fluid mechanics, the linear viscous Navier-Stokes model being the
classical example, but is unduly restrictive in solid mechanics. For this reason, the
Eulerian description of the motion is abandoned in all of the subsequent developments.
30. Objective transformations and frame invariance
As a first step in the generalization of the elastoplastic constitutive equations to the
finite strain regime we consider the formulation of classical finite strain elasticity. The
principle of objectivity plays a central role in this subject.
Given a motion Wpt: Q2 - RIdim with t e I, a superposed rigid body motion of the
current placement St = Wt((2) is a map It: St --E]Rdi-m defined as

x E St

H- XZ

= ibt(x) = r(t) + Q(t)x cG ndi,

(30.1)

where t - r(t) E I Tndim and t H Q(t) E SO(ndim) are arbitrary functions of time. In
what follows we consider the case ndim = 3. The superposed motion is called rigid

314

J.C. Simo

because for any two given points xl, x 2


have
x+ - x

= Q(t)[I

:1]

St, since Q(t)

- -2 1= X -

CHAPTER Ill

SO(3) is orthogonal we

2,

(30.2)

212 = (x 1 - x2 ) (xl - x 2) is the square of the Euclidean distance.


Thus, (30.1) preserves distances and, therefore, defines an Euclidean isometry.
A spatial tensor field is said to transform objectively under superposed rigid body
motions if it transforms according to the standard rules of tensor analysis.

where Ixl -

EXAMPLE 30.1. The total motion obtained by composition of (30.1) with the given

motion is
zt = iP+(X, t) =

o o(X, t) = r(t) + Q(t)o(X, t).

(30.3)

The deformation gradient for the total motion then becomes


F+ =D

= Q(t) D~9(X, t) = Q(t)F.

(30.4)

Therefore, the spatial velocity gradient is given by

V+v+ = F+ (F+)- = Q(t)VvQT(t) + Q(t)QT(t),

(30.5)

which does not transform objectively due to the additional skew-symmetric term
Q(t)QT(t). However, its symmetric part, the rate of deformation tensor, does transform objectively since, from (30.5) and (28.23)1, we have
d + = Q(t)dQT(t).

(30.6)

Note that the spin tensor, defined by (28.23)2, transforms according to the nonobjective
rule
i+ = Q(t)IiVQ T (t) + Q(t)QT (t).

(30.7)

REMARK 30.1. Convected objects, that is tensor fields on the reference configuration,

remain unaltered under spatial superposed rigid body motions. For example from
(30.4) we have
C + = (F+)TF+ = FTQT(t)Q(t)F = C.

(30.8)

Likewise, C+ = C is unaltered by spatial superposed rigid body motions. To understand this result, Fig. 30.1 may prove useful.
Given an objective spatial tensor field its material time derivative will not, in general, retain the objectivity property. It is precisely this lack of objectivity of the material

315

Nonlinear continuum mechanics

SECTION 30

FIG. 30.1. Reference and current placements. Illustration of superposed rigid body motions.

time derivative of objective spatial fields that motivates the introduction of objective
rates. To see how this lack of objectivity arises, consider the Cauchy stress tensor
and use the definition of the material time derivative together with the chain rule to
evaluate as
dr = [at(a O Pt)] o t 1 = 0a t

+ (V - V)U.

(30.9a)

In component form this result reads


(5'ab(Xt) =

aOb(X, t)

at

aO-ab(X, t) 0Pc(X, t)

ax,

(30.9b)

at

Now assume that ar transforms objectively so that o + = Q(t)aQT(t). From (30.9a)


it then follows that
a + = Q(t)rQT (t) + [Q(t)QT (t)] a+ - a+ [(t)Q T (t)]

(30.10)

which is clearly nonobjective. Objective rates are essentially modified time derivatives of the Cauchy stress tensor constructed so that objectivity is preserved. A large
body of literature has been concerned with the development of objective rates which,
remarkably, extends to recent dates. Concerning the practically infinite number of proposals made we remark, following TRUESDELL and NOLL ([1965], p. 404), that "...
Despite claims and whole papers to the contrary, any advantage claimed for one such
rate over another is pure illusion."

J.C. Simo

316

CHAPTER III

In fact, one can show that any possible objective stress rate is a particular
case of a basic geometric object known as the Lie derivative; see TRUESDELL
and TOUPtN [1960], MARSDEN and HUGHES ([1983], Chapter 1), So and MARSDEN [1984] or ARNOLD [1989]. A few widely used proposals found in the literature
are recorded below.
(i) The Lie derivative of the Kirchhoff stress tensor, also known (up to a factor
of J) as the Truesdell stress rate, is defined as
T
vr = {Fat[F-17F-T]FT } = {F[atS]F }.

(30.11)

Technically speaking this definition gives v o oLt. However, in the present context
the notation in (30.11) is preferred since in solid mechanics one is almost exclusively
interested in a Lagrangian description of the motion. Using the well-known result
at(F-l) = -F- PF-t for the derivative of the inverse of a matrix, the definition
of the material time derivative and the spatial velocity gradient imply
(30.12)

- (Vv)r - r(Vv)T.

Er =

One can easily show that (30.12) is objective (in fact, vr meets the much stronger
condition of covariance, see MARSDEN and HUGHES [19831).
(ii) The Jaumann-Zaremba stress rate of the Kirchhoff stress is essentially a corotated derivative relative to spatial axes with instantaneous velocity given by the spin
tensor. One finds
V=
-

Wr

(30.13)
(.

(iii) The Green-Mclnnis-Naghdistress rate of the Kirchhoff stress is defined by an


expression similar to (30.11), but with F replaced by R, i.e.,
r= {ROt [RTrR] RT}.

(30.14)

Recalling that R = DR, we find

r =

r + r

f.

(30.15)

~ ii unless
Although (30.15) is similar in structure to (30.13), we remark that
U = 0 (i.e., an instantaneously rigid motion), see (28.30). This objective rate was
brought into prominence in DIENES [1979] and is further discussed in JOHNSON and
BAMMANN [1984].
For a catalog of the many proposed objective stress rates, up to the early sixties,
see TRUESDELL and TOUPIN ([1960], Section 48, p. 151).

Nonlinear continuum mechanics

SECTION 31

317

31. Elastic constitutive equations and isotropy group


Attention will be restricted here to classical nonlinear elasticity; see TRUESDELL and
NOLL ([1965], 17-19) for a detailed discussion in a general context. The constitutive
equation for a hyperelastic material is defined in terms of a stored energy function
W(X, F(X, t)), depending locally on the deformation gradient, such that
P(X,t) = aFW(X,F(X,t)),

(X, t) E FQ x .

(31.1)

The stored energy function W(X, F) is said to be objective or frame invariant if


the following condition holds. Let t - Wot be an arbitrary motion, and consider a
superposed rigid body motion of the form (30.1), with deformation gradient F + =
Q(t)F, where Q(t)
SO(3). If no additional restrictions are placed, in general
W(X, F + ) need not coincide with W(X, F). The requirement of objectivity is the
condition
VQ(t) E S0(3).

W(X, QF) = W(X, F)

(31.2)

Restriction (31.2) then implies that the stored energy function can depend on F only
through C = FTF, i.e.,
(31.3)

W(X, F) = W(X, C).

From this reduced constitutive function along with (31.1), one obtains the following
classical constitutive equations for elasticity
S = 2acW,

P = 2F[acW] and

r=2F[acW]F T .

(31.4)

For clarity, explicit indication of the argument has been omitted. We shall often follow
this practice in subsequent developments. Objectivity of the stored energy function
is closely related to the balance of angular momentum; i.e., to the symmetry of the
Cauchy stress tensor. In fact, it is easily shown that W(X, F) is objective (i.e., (31.2)
holds) if and only if the balance of angular momentum condition PFT = FPT holds
(with P given by (31.1)).
(A) Hyperelastic rate constitutive equations. The rate form of the hyperelastic constitutive equations play a central role in the incremental formulation of plasticity. Some
of the key results are summarized below. First, time differentiation of relation (31.4)1
gives

=C
S

i.e.,

SAB

CABCD1CCD,

(31.5)

where C(X, C) is the material elasticity tensor given by


C = 4 0ccW(X, C),

i.e.,

CABCD = 4CABCCD
acAacCD

(31.6)

318

J.C. Simo

CHAPTER III

Recalling that vr = [FSFT], along with the relation C = 2FTdF, expression


(31.5)1 leads to
(vr)ab = FaASABFbB = [FaAFbBFcCFdDCABcD]dcd= cabcddcd,

(31.7)

where with components Cabcd is the spatial elasticity tensor related to C by the
(push-forward) transformation
Cabcd = FaAFbBFcCFdDCABCD.

(31.8)

Note that (31.7) results in the spatial rate constitutive equation


v

d,

i.e.,

(31.9)

(CvT)ab = CabCddCd

Also note that analogous expressions can be derived for any objective stress rate other
than the Lie derivative. The derivation of hyperelastic rate equations for other stress
and strain measures constitutes an exercise (often nontrivial) involving the application
of the chain rule (see, e.g., OGDEN [1984], Chapter 5, and the remark below). Here,
we shall simply quote one further result useful in numerical implementations. From
(31.1), we find
P = AF,

i.e.,

PaA = AaAbBFbB,

(31.10)

where A is called the first elasticity tensor, and is given by


A = aFW(X,,F),

i.e.,

AaAbB

-OFaFbB

(31.11)

Alternatively, starting from (31.4)2 we arrive at (31.10), along with the following
important relation connecting A and c
1
F
Aa BcD = FBb [Cabcd + Tacbd]

(31.12)

REMARK 31.1. Starting from the properly invariant hyperelastic relations (31.4), one

can derive spatial rate constitutive equations of the form (31.9) which are also properly
invariant. This equation can be expressed in terms of any objective rate; for instance,
in terms of the Jaumann stress rate, on account of the relation
Evr = - - (d + -)

(d +

~)T

=7

-dr- rd,

(31.13)

Eq. (31.9) can be written as


V
7 =

ad where

aabcd = Cabcd +

ac~bd + 6bdTac(

(31.14)

Nonlinear continuum mechanics

SECTTON 31

319

Conversely, given any rate constitutive equation of the form (31.14) the question
arises as to whether a stored energy function exists such that -r is given by (31.4)3.
The answer to this question is, in general, negative. In addition to the full symmetry
of the moduli a a set of compatibility relations must hold for a rate equation of the
V
form r = ad to be derivable from a stored energy function. We refer to TRUESDELL
and NOLL (1965], Chapter IV) for a detailed account of the relevant results.
Rate equations of the form (31.14) which are not derivable from a stored energy
function lead to the notion of hypoelasticity. We recall two basic results.
(i) In general, hypoelastic materials produce nonzero dissipationin a closed cycle.
See TRUESDELL and NOLL ([1965], p. 401) for the precise statements.
(ii) The assumption of r = ad with a being the constant isotropic tensor of the
infinitesimal theory of elasticity is in general incompatible with hyperelasticity (SIMO
and PISTER [1984]). It should be noted that such an assumption is typically made in
phenomenological theories of plasticity.
EXAMPLE 31.1. To illustrate the preceding ideas in a concrete setting, consider the
following stored energy function (SIMo and PISTER [1984], CIARLET [1988])

W= A(J2 - 1) - (A +

)log[J] + ml'(tr[C]- 3),

(31.15)

where A, u > 0 can be interpreted as Lame constants. Using the relation acJ =
JC - 1, the constitutive equations (31.4) become
s= 1A(J 2 -1)C

I+,(1 - C- 1),

(31.16)
= X(J2

1)

- 1).
l(b

In view of relations (31.6) and (31.8) the spatial elasticity tensor is given by the
explicit expression
c = A J21 1 + 2u[ -

(J2

1)A//]I,

(31.17)

where I with components labcd = [acSbd + badSbc]/2 is the rank-four symmetric


unit tensor. The following four important properties should be noted. (i) As J -- 0
or J -- 00oo
we have W -- oo. (ii) Both W = 0 and r = 0 for F = 1, i.e., at
the reference state. In addition, c reduces to the elasticity tensor of the linear theory.
(iii) W can be written as W = U(J) + ,/(tr[C] - 3) where
U"(J) =

A(1

+ 1/J2) + l/J

> 0,

for J C (0, oo).

(31.18)

It follows that, U(J) is a convex function of J. (iv) The stored energy W is a polyconvex function of F. Thus the only known global existence results for elasticity
based on the existence of minimizers of the total potential energy for poly-convex
stored energy functions apply to this model (see, e.g., CIARLET [1988], Chapters 4, 7,
and MARSDEN and HUGHES [1983], Chapter 6).

320

J.C. Simo

CHAPTER III

(B) The notion of isotropy group. Isotropic elastic response. The isotropy group
at a point in the reference placement characterizes the invariance properties of the
material constitutive response under superposed rigid body motions on the reference
configuration. The notion of isotropy should not be confused with the notion of frame
indifference. Whereas the former notion refers to a particular property which the
material response may enjoy, the latter notion is a fundamental principle of mechanics,
which holds for all possible response functions.
Let X EC be a point in the reference placement Q of an elastic body. Consider a
superposed rigid deformation on Q, i.e., the isometry
X E Q2-

= b(X) = r + QX

which maps Q onto Q+. Since p+ o


F + = FQT

= fpt,

so that C + = QCQT

for Q

SO(3),

(31.19)

the chain rule implies

and

b+ = b.

(31.20)

It follows that under superposed rigid body motions of the reference state 2 the
deformation gradient and the Cauchy-Green tensors transform according to (31.20).
It is clear that, in general, the values of the stored energy function at X CE 2 evaluated
at C and C + will be different. The isotropy group at X CE 2 is precisely the set of
proper orthogonal transformations which leave the stored energy function unchanged;
i.e.,
Gx = {Q E SO(3): W(X, QCQT ) = W(X, C)}.

(31.21)

It can easily be shown that Gx C SO(3) is indeed a group. Note that Gx is associated with each point X GE 2, unless the material is homogeneous (i.e., unless W
is independent of X). Furthermore, the isotropy group Gx is defined relative to a
particular reference configuration. If Gx = SO(3) the material is said to be isotropic
(relative to 2, at the point X E Q). Otherwise, the material is said to be anisotropic.
The condition of isotropic response places strong restrictions on the admissible
forms of the response function. Here we shall recall only one important result which
will be used below.
From our preceding discussion, a function f: -- R of symmetric tensors is
isotropic if and only if
f(QHQT ) = f(H)

VQ E SO(3) with H E S.

(31.22)

We may think of f as the free energy or any other response function (e.g., the yield
condition as discussed below). Depending on the context, H E $+ C $ will denote
the right Cauchy-Green tensor or any other symmetric tensor (e.g., the Cauchy stress
field, a tensor field in S). The following well-known result is a basic fact on isotropic
tensor functions.

Nonlinear continuum mechanics

SECTION 31

321

THEOREM 31.1. Representation theoremfor isotropicfunctions. A function f : - R


S through its principalinvariants;
is isotropic if and only if f (H) depends on H
such that
i.e., if and only if there exists a function f : R3 - I>R
f(H) = f(1 1 (H),I2(H),13(H)) VH E 5,

(31.23)

(I (H) 2 - tr[H 2 ]) and I3(H) = det[H] are the

where I (H) = tr[H], I2(H) =


principalinvariantsof H.

This result is an immediate consequence of the spectral theorem for symmetric


tensors; see, e.g, GURTIN ([1981], p. 230).
EXAMPLE 31.2. For homogeneous isotropic elasticity, the stored energy is a function
only of the principal invariants of C = FTF; i.e.,
W(C) = W(

1 , 2,

(31.24)

13),

where IA (A = 1, 2,3) are given by (28.4). An example of an isotropic stored energy


function was given in the preceding example. Making use of the relations
acI1 = 1,

aci2 = Il

and

acI3 = I 3 C - ,

(31.25)

the constitutive equation (31.4) for the symmetric Piola-Kirchhoff tensor becomes

S=2[aW+

all

a'2

II] 1- 2

a2

C+2

a3

13C-

(31.26)

Using the relation r = FSFT one obtains the following constitutive equation for the
Kirchhoff stress tensor:
r =2

awi

-I31+
1

ai13j

rawaw h]1 bai

+
1

DI
a2

W-b2,

(31.27)

where b = FFT is the left Cauchy-Green tensor.


REMARK 31.2. For isotropic response, and only for this case, the stored energy function is a function of the left Cauchy-Green tensor. This result is an immediate consequence of the fact that Gx = SO(3). Since the constitutive equations are local it is
permissible to chose Q = R in (31.21), where R is the rotation tensor in the polar
decomposition of the deformation gradient. This gives
W(X, C) = W(X, RCRT ) = W(X, b),
since b = FFT = RUURT = RCRT; hence the result.

(31.28)

322

J.C. Simo

CHAPTER III

REMARK 31.3. The obvious way to automatically insure satisfaction of frame indifference without precluding anisotropic response is to formulate the constitutive response
function in terms of objects associated with the reference configuration. For elasticity
this amounts to formulating the constitutive response in terms of the right CauchyGreen tensor C and the second Piola-Kirchhoff stress tensor S. The formulation
of elasticity in terms of (S, C) is called the convective representation of elasticity.
That use of convected coordinates (the convected representation) automatically insures frame indifference was known to ZAREMBA [1903]; see TRUESDELL and NOLL

([1965], p. 45). Convected coordinates are also used in HILL [1978], where the terminology "embedded system" is favored, and most notably throughout the classical
book of GREEN and ZERNA [1960]. For a discussion of the convected representation
(including the corresponding Hamiltonian structure) see SIMo, MARSDEN and KRISHNAPRASAD [1988].

REMARK 31.4. In a computational setting, the convective representation in terms of


(S, C) (or, equivalently, the use of convected coordinates) has been used by several
authors, see, e.g., NEEDLEMAN [1982], for applications to plasticity theories.
32. Volumetric/deviatoric uncoupled finite elasticity
Within the context of the infinitesimal theory, elastic materials possessing a general
stored energy of the form

W(E) = U(tr[E]) + W(dev[e])

where dev[e] = E -

tr[El,

(32.1)

ndim

are said to exhibit uncoupled volumetric/deviatoric response. An example of such a


material is provided by the linear isotropic elastic solid for which U(tr[E]) = r(tr[E])2
and W(dev[s]) = /i(dev[E] dev[E])2 , where nr> 0 is the bulk modulus and p > 0 is
the shear modulus. The generalization to the geometrically nonlinear of the uncoupled
stored energy function (32.1) first appears in FLORY [1961] for isotropic materials and
is systematically developed in SIMo, TAYLOR and PISTER [1985] both for elastic and
plastic materials. The idea is to introduce the volume preserving and volumetric parts
of the deformation gradient F = Dop, respectively defined by
J = det[D(]

and

F = J-/dm

Dq.

(32.2)

By the properties of the determinant it is clear that this definition yields det[F] = 1.
Expression (32.1) is then generalized by considering an uncoupled stored energy function of the form

W(C) = U(J) + W(C),

(32.3)

Nonlinear continuum mechanics

SECTION 32

323

where C is the volume preserving right Cauchy-Green tensor and E denotes the
volume preserving Lagrangian strain tensor respectively defined by
= TF

and

E=

(C-1).

(32.4)

As in the geometrically linear theory, the functions U and W define the volumetric
and volume preserving contributions to the stored energy function, respectively. An
important application of uncoupled models with stored energy function of the form
(32.3) arises in the context of metal plasticity, as described in a subsequent section.
The stress response is obtain by evaluating the partial derivatives of the stored energy
function (32.3) with the help of the following result:
THEOREM 32.1. The partial derivatives of C relative to C and J are given by
6
[ac
] =J- 2 /ndmi [I-

C-]

and 3cJ =

JC

'.

(32.5)

ndim

Moreover, the material deviator of any rank-two (convected) tensor field H on the
reference configuration 2, defined by the relation
[H C]C- ,

DEV[H] = J2 /ndim [ac6]H = H -

(32.6)

ndim

has the following properties:


dev[FHFT ] = FDEV[H]F T ,
DEV[H]

C - l = dev[FHFT ]

(32.7)
.

1 = 0,

where dev[h] = h- (/ndim) tr[h]l is the usual deviatoricpart of any (spatial) tensor
field h in the current configuration S = (Q2).
PROOF. Consider a one-parameter family of right Cauchy-Green tensors of the form
C, = C + EH,

with H = HT and E > 0.

(32.8)

Clearly, for E > 0 sufficiently small we have det[C,] > 0. Setting J, = det[C
and using the standard formula for the derivative of the determinant it follows that

dE

d J=o 2J2tr[C- d
Jd2

C ] = Jtr[C-IH].

(32.9)

324

J.C. Simo

Differentiation of the defining relation

C'

|d

- --

j=J-/ndiH

de =0

ndim
2

/dim
J-

[I

1C1

CHAPTER I1
2

/ndimC,

and use of (32.9) then yields

tr [C-H]C]

C] H;

(32.10)

ndim

a relation which holds for any H = HT. By the chain rule, it follows that

r e=

[acc

dE

CE = [cC]H.

(32.11)

This result, together with (32.9) and (32.10), implies relation (32.5) in the statement
of the theorem. Relation (32.7) follows from (32.10) and (32.11) while properties
(32.7) are a direct consequence of (32.6) and the definition of the right Cauchy-Green
tensor. El
Observe that definition (32.6) for the deviator of a convected (contravariant) stress
tensor is consistent with the geometric setting of the convective representation of
nonlinear elasticity. Recall that within this context the right Cauchy-Green tensor is
interpreted as a (flat) Riemannian metric on the reference configuration 2, the trace
of a contravariant tensor is given by its contraction with the metric and its deviatoric
part is defined by (32.6).
Using the results in the preceding theorem, the stress-strain relations associated with
the uncoupled stored energy function (32.3) can be easily computed by specialization
of the general hyperelastic stress-strain relations S = 2DcW(C), cast in terms of the
second Piola-Kirchhoff stress tensor and the right Cauchy-Green tensor, by making
use of the following relation implied by the chain rule:
cw(C) =,U'(J)

acJ + [acc]O~w(C).

(32.12)

First, by combining (32.12) with (32.5) the constitutive equation for the second PiolaKirchhoff stress tensor becomes
S = JU'(J) C-I + J-2/ndim2DEV[DcW(C)].

(32.13)

Second, from the relation r = FSFT and properties (32.7), expression (32.13) yields
the following constitutive equation of the Kirchhoff stress tensor:
r =Jpl+dev[2FaaCW(C)F T

and p =U'(J),

(32.14)

where F is the volume preserving part of the deformation gradient defined by (32.2)2.
Thus the uncoupled stored energy function (32.3) produces the uncoupled stressstrain relations (32.14) in which the hydrostatic pressure response, defined by the
term p = U'(J), is uncoupled from the deviatoric stress response.

SECTION 32

Nonlinear continuum mechanics

325

To complete the foregoing developments, it only remains to compute the elastic


moduli associated with the uncoupled stored energy function (32.3). The elastic moduli
in the convective description, associated with the volume preserving part W(C),
denoted by C, are defined as

C = 4~a4
i.e.,

aw(Ca)
CIJKL = 40
J(32.15a)

4O4CW(C),

while spatial moduli associated with W(C) are defined via the standard push-forward
transformation as
Cijkl = FiIFjJFkKF1LCIJKL.

(32.15b)

The components of both C and c obey the usual symmetry conditions associated
with the existence of the potential function W(C). It proves convenient to cast expression (32.15a) directly in terms of partial derivatives relative to the volume preserving
right Cauchy-Green tensor C, which is the argument of the function W. This task
is accomplished by application of the chain rule. Using the results in the preceding
theorem, a somewhat involved but otherwise straightforward manipulation yields the
following result:
2
C = Cdev -

DEV[S]

ndim

+ 2 J2/ndim

di

C-

2
_-

ndim

C-l

C-

W. C) [IC-

(a

DEV[S]

C
],

(32.16)

ndim

where DEV[S] = J-2/ndimDEV[a0W], IC-, is the fourth-order tensor with components


(I-I)IJKL =- [CIKCJL-

(32.17)

CIL CjK],

and Cdev is the fourth-order tensor of deviatoric tangent moduli defined by the expression
Cdev=J-4/ndm"'[O3W
Cdev - 4Jc4/(ndim)-+

ndim

- [
C

[c

WI Cc-,C-1
]

[C-l [ CaW]C- [aLCa W]Cc -l]c.

(32.18)

The preceding notation is motivated by the property CdevC = 0, which is an easy


consequence of expression (32.18), and shows that Cdev is a fourth-order deviatoric
tensor. The complete expression for the elastic moduli is obtained by adding to C

326

J.C. Simo

CHAPTER II

the contribution O2ccU(J) associated with the volumetric part of the stored energy
function. Using (32.5)2 the result can be written as
C = J 2 U"(J)C lC-l + JU'(J) [C-' 0 C-'- 2Ic , + C.

(32.19)

The corresponding expressions for the elastic moduli in the spatial description are easily obtained from the preceding results by using the transformation (32.15b) together
with properties (32.7). From (32.16) one obtains
2

C = Cdev -

dev[r]

dev[T]

ndim
T

ditr[2a

1-

ldim

'

1)

(I-

(32.20)

where
(32.21)

(Cdev)ijkl = FilFjJFkKFlL(dev)zJKL.

Similarly, expression (32.19) together with properties (32.7) yield the following result
for the spatial elastic moduli:
c = J 2 U"(J)1 1 + JU'(J)[1 X 1 - 21] + 6.

(32.22)

The example below provides an illustration of the preceding results in the spatial
description.
32.1. The simplest example of stored energy function of the form (32.3)
is given by the following extension to the compressible regime of that Neo-Hookean
elastic material:
EXAMPLE

U(J) =

(J

1 - logJ)

and

W(C) =

(tr [I] - ndim)

(32.23)

The derivates of the compressible part U(J) of the stored energy function are
U'(J) =

J-

and

U"(J) =

ll

+ J2

(32.24)

Since U"(J) > 0 for J E (0, oo) it follows that U(.) is a convex function for > 0
with U(1) = 0 and U"(1) = . Noting that aaW(C) = 1 expression (32.14) reduces
to
'r =

(j2

1)1 +

dev [FFT ] .

(32.25)

Nonlinear continuum mechanics

SECTION 33

327

Finally, since a-s W(C) = 0, the tensor c of spatial elastic moduli defined by (32.22)
becomes
C =

[J21

- (J2 -

1)I] + 2

tr [FFT](I-

I
ndim

Tndim

--

dev[r] 1 -

ndim

1 0 dev[T].

(32.26)

ndim

Observe that c particularized at the stress-free reference configuration (F = 1, J = 1


and r = 0) gives the usual elastic moduli of linear isotropic elasticity.
33. Isotropic elasticity formulated in principal stretches
The stored energy function of an isotropic material can only depend on the principal
invariants of the right (or left) Cauchy-Green tensor. Equivalently, the stored energy
function can be formulated as a function of the principal stretches. This latter representation, although equivalent to the former, turns out to be often much more convenient
in applications. A notable example is furnished by so-called Ogden materials, which
have proven quite useful in the modeling of the response of rubber like materials
(see OGDEN [1984], and references therein). For isotropic finite strain plasticity, the
formulation of the elastic response in terms of principal stretches leads to a remarkable simplification of return mapping algorithms (SIMO [1992]) which are described
subsequently.
The computation of the stress response for an isotropic elastic material with stored
energy
W(C) = W(AX, A2 , A3 ),

(33.1)

given as a function of the principal stretches, relies on the following result.


THEOREM 33.1. The derivatives of the principal stretches are given by the following
formulae:
(i) Three different roots. If A1 7 A2 7 )A3 then
acA2 = N(A) N(A)

(A = 1,2,3).

(ii) Two equal roots. If A :=


a0c

= 1 - N(3)

N( 3 )

= :2

and aC,

(33.2a)
3

then
= N(3)

N(3).

(iii) Three equal roots. If A = A) = A2 = A3 then aCA2 = 1.

(33.2b)

(33.2c)

The proof of these results is straightforward. For the case of three different
roots, for instance, differentiation of the eigenvalue problem (28.6) gives
PROOF.

dCN(A) + C dN(A)

= d(A2)N(A) +

AA dN(A).

(33.3)

J.C. Simo

328

CHAPTER III

Taking the dot product of this relation with N(A) and using the orthogonality condition
A)
(A ) = 0, which follows from the normalization IN(A) I = 1, gives
dN( N
d(A 2 ) = N(A)

dCN(A) = tr [dC(N(A) N(A))].

(33.4)

Since tr[-, ] defines an inner product in the linear space L(3) of 3 x 3 matrices, the
preceding relation along with the directional derivative formula yields (33.2a). The
other relations are proved similarly. [
(A) Elastic constitutive equations. The elastic constitutive equation for the second
Piola-Kirchhoff stress tensor S is computed by the chain rule as

s = W(C) =i
s = 2acw(c):

1IA Ow( ,a1 XA,2 ) 3

O(

(33.5)

A=I

From the results (33.2) one concludes that (33.5) is merely a restatement of the spectral
decomposition of S, i.e.,
ndim

S=

SAN

(A )

with SA =

3 N(A)

AA

A=1

(33.6)
A(

This expression is consistent with the restriction to isotropy, which implies that S and
C commute or, equivalently, that S and C have the same characteristic subspaces.
Using relation r = FSFT Eq. (33.6) gives
aw

ndim

r =

TAn(A)

n(A)

with TA = AA

(33.7)

AA

A=1

This expression furnishes the constitutive equation for the spatial Kirchhoff stress.
Next, a closed-form expression is given for the material elastic moduli.
(B) The materialtangent elastic moduli. Recall that the material elastic moduli C are
defined by the relation S = C C. An easy way to compute C exploits the following
observation (OGDEN [1984]). The time derivative of the eigenvectors of C can be
expressed as
?Zdim

(A)

=W

x N(A) =

>

WABN(B),

B=l
BAA

(33.8)

Nonlinear continuum mechanics

SECTION 33

329

where WAB = -WBA. Inserting this result into the time derivative of the spectral
decomposition of C gives
ndim

) N (A) N(A)

, t (2

A=I
ndim

dim

2 A _AB)WABN(A)

N(B).

A=l B=1
BOA

Similarly, time differentiation of the spectral decomposition (33.6) of S and use of


(33.9) gives

ndim ndimi

E E

S=
= >32

as

2 \(A)

aa

N(A)

A=1 B=
ndim

ndim

+E E

SA -

A=l B=1 2
B4A

SB

A-AX2B) W

AB

N(A)

(33.10)

N()

By inspection of (33.9) and (33.10) it is easily concluded that the tensor C of elastic
moduli is given by the explicit formula
ndi

B N

SA

A=1 B=1
BOA

X N

X [N(A)

N ( B) + N

(B )

0 N(A)]

r 1

AB

(B)

?dim ndim
A=1 B=1

( A)

aB

AA

aA

The expression for the spatial moduli C follows immediately from (33.11) and (31.8).
(C) The spatial elastic moduli in terms of logarithmic stretches. For the purpose of
characterizing the elastic response of models of multiplicative plasticity, as well as
the formulation of appropriate integration algorithms, it proves convenient to write
the stored energy function in terms of principal logarithmic stretches as
w(A,A, A 3) = f(e1,E2,

3)

(33.12)

where EA = log(AA).

In this situation, the closed-form expression for the spatial elasticity tensor c can be
derived from the preceding results as follows. First, using the chain rule we have
1

-B

a
a-B

I
1 a W] =
AB
_AB
w =- -2
A A
w A

2
aEA--B

2AB

a(31
EB

(33.13)

J.C. Simo

330

CHAPTER III

Second, using definition (31.8) for the spatial elasticity tensor together with the relation
FN(A) = AAn(A) and expression (31.8), we arrive at the following result
ndim

dim [
L

a2-

n
0A)

n
(A) (B)

+ g

(33.14a)

A=IB=1

where the nonzero components


are given by (I, J = 1,..., ndim)
giII

- 2 ,T

and g.jjjj

IJKL

in the Eulerian axes n(A) } of the tensor g

= gIjjj = T2 2 - 2ZJ
A - X

for I

J.

(33.14b)

Note that the principal values of the Kirchhoff stress are given from (33.7)2 in terms
of the logarithmic stretches as r = w/SeE.
33.1. From a computational standpoint, the advantage of the closed-form
expression (33.14) for the elastic moduli is that it remains well-conditioned in the
presence of repeated principal stretches. A conceptual algorithm for the computation
of the stress and the elastic moduli for stored energy functions defined in terms of
principal stretches proceeds as follows.
REMARK

Step 1. Given C compute the Lagrangian axes by solving the eigenvalue problem
CN(A) = A2 N(A).
Step 2.

Use the preceding formulae to evaluate S and C (or r and c).

The key step is the computation of the orthogonal matrix Q that orients the Lagrangian triad, i.e., Q = [N(O) N(2) N( 3 )]. Although closed-form solutions are possible, currently we favor an iterative solution using Jacobi's algorithm to avoid possible
numerical difficulties associated with repeated roots. For the symmetric eigenvalue
problem it is well-known that convergence of Jacobi iteration holds regardless of the
multiplicity of the eigenvalues.
34. Multiplicative plasticity at finite strains: Basic concepts
This section addresses the extension to the finite strain regime of the classical model
of infinitesimal plasticity described in detail in Chapter I. A key feature of the model
described below is the replacement of the additive decomposition of the infinitesimal
strain field by a multiplicative decomposition of the deformation gradient into elastic
and plastic parts. This multiplicative decomposition is motivated by a well-understood
micromechanical picture for single crystal metal plasticity. Comprehensive expositions
of the micromechanical description of single crystal metal plasticity can be found in
the review article of ASARO [1979, 1983] and the recent monograph of HAVNER
[1992]. The basic ideas go back to the fundamental work of TAYLOR [19381 and TAYLOR and ELAM [1923, 1925], subsequently expanded upon in HILL [1966], HILL and
RICE [1972], ASARO and RICE [1977], HAVNER [1971, 1982], NEMAT-NASSER [1983]

SECTION 34

Nonlinear continuum mechanics

331

and others. The emphasis in the exposition below is placed on those mathematical
aspects of the model relevant to its numerical treatment, as described in detail in the
following chapter.
(A) Motivation: The key assumptions of the infinitesimal theory. To motivate the
structure of the finite strain theory, it proves useful to recall the key assumptions
made in the formulation of the classical model of infinitesimal plasticity, as described
in Chapter I.
(Al) The total infinitesimal strain field E(X, t) at (X, t) EG
Q x [ is decomposed
e
additively into elastic and plastic parts as E = E
EP.
(A2) The stress response is defined in terms of a free energy function T(Ee, (a) via
the potential relations (2.4), where {(} are a set int (strain-like) internal variables
conjugate to the (stress-like) variables {q'} in the sense that q' := -jTf.
(A3) The stress a and the set of nint stress-like internal variables {qO} that provide
a phenomenological characterization of microstructural hardening mechanisms are
constrained to lie in a convex set E c S x Rnin, called the elastic domain E. In
general, E is defined by (23.6).
(A4) In the simplest version of classical plasticity, the evolution of the internal
variables {eP, ,} is specified by postulating maximum dissipation in the system,
i.e., via the inequality (6.7). This assumption yields the evolution equations (6.8) of
associative plasticity.
A critical examination of these assumptions reveals two basic issues arising in a
generalization of this classical model to the finite strain regime.
First, the infinitesimal strain tensor is not an invariant measure of deformation in the
finite strain regime; i.e., it does not transform properly under superposed rigid body
motions. Assumption (Al) must therefore be revised. Although a number of alternative
options are possible, the use of Lagrangian strains being the obvious one (GREEN and
NAGHDI [1964, 1966]), the option adopted must be motivated by micromechanical
considerations.
Second, a specific stress measure must be adopted in the definition of the convex
set E within the finite strain regime. In this regard it should be noted that the obvious
choice, the Cauchy stress tensor a, restricts the theory to isotropy. To justify this
assertion consider the simplest situation in which nnt = 1 and E is given by
E: = {(a, q) E S x t: f(r, q) < O}.

(34.1)

Given a motion t - Vpt consider the rigid body motion (30.1) superposed on the
current placement St = pt(). Since the Cauchy stress transforms objectively as
A+ = Q(t)oaQT(t), it necessarily follows that f: x I - IRin (34.1) is objective if
and only if
f(Q(t)aQT (t),q) = f(a,q) VQ(t) G SO(3),

(34.2)

332

J.C. Simo

CHAPTER III

Ce

CP-I

Reference Configuration

Inte

1
be
te Configuration

Current Configuration

FIG. 34.1. Illustration of the local multiplicative decomposition of the deformation gradient and the basic
deformation tensors.

which is the statement of isotropy for the function f(., q). In summary, a classical
formulation of the yield criterion in terms of true stresses necessarily restricts the
resulting theory to isotropy, as claimed.
In what follows, we shall address the extension to the regime of finite strains of
the classical model of infinitesimal plasticity based on assumptions (A1)-(A4) and the
preceding observations.
(B) Basic kinematical result for multiplicative plasticity. The generalization of the
additive decomposition in assumption (Al) to finite strains is motivated by the structure of the single crystal model for metal plasticity, and takes the form of a local
multiplicative decomposition of the deformation gradient as

F(X, t) = Fe(X, t)FP(X, t)

V(X, t) E Q x ,

(34.3)

where F(X, t) := D(X, t) denotes the deformation gradient of the motion at X E Q


(see Fig. 34.1). From a micromechanical point of view F P is an internal variable related to the amount of dislocation flow through the crystal lattice, whereas F e models
the lattice deformation, see KRONER and TEODOSIU [1972] and HAVNER [1992]. In
the phenomenological setting considered in LEE and LIU [1967], LEE [1969], KRATOCHVIL [1973] and others, Fe-t is viewed as defining a local, stress-free, unloaded
configuration obtained by releasing the stresses on each neighborhood in the current
placement.

Nonlinear continuum mechanics

SECTION 34

333

Associated with the multiplicative decomposition (34.3) one defines elastic right
and left Cauchy-Green tensors by the conventional relations
Ve

be

and

= FeTFe

= FeFeT.

(34.4)

Observe that Ce is a covariant tensor field defined on the intermediate configuration


whereas be is a contravariant tensor field defined on the current placement St. Local
rate of deformations associated with the plastic and the elastic parts of the deformation
gradient are defined by the relations
LP =

FP

'

and

l e = FeF

- l .

(34.5)

These definitions are motivated by the expression 1 = Vv o t = FF- 1 found for


the (spatial) velocity gradient. Again, we remark that L P is a generally nonsymmetric
tensor field defined on the intermediate configuration while 1e is a nonsymmetric tensor
field defined on the current placement St. The following result provides the relation
between these two fields.
THEOREM 34.1. The plastic velocity gradient L P and the elastic velocity gradient e
are related by the transformation
FeLPFe -

- le

where I = FF

-1

= Vv o got.

(34.6)

This (push-forward) relation implies that L P and le are contra-covarianttensorfields


on the intermediate configuration and the currentplacement, respectively.
PROOF. The formula for the derivative of the inverse of a rank two tensor, together
with (34.5)1 and the factorization (34.3) give
LP = at [Fe-lF]

= F

e- 1

[F

F p -

- 1

l =

[F

e - l

_- F

e -

lFeF

e -

lF]

F p -

(34.7)

- FeFe-]
]Fe

The result then follows from definition (34.5)2.

Expression (34.6) motivates the following definitions in the spatial description for
the plastic velocity gradient, the plastic rate of deformation tensor and the plastic spin
tensor
IP = 1 - le,

dP = sym[lP]

and

Pi= skew[lP].

(34.8)

Recalling the expressions d = sym[l] and ii = sym[l] for the total rate of deformation tensor and the total spin tensor, respectively, the preceding definitions imply the
additive decomposition
d = de + d

with de = sym[le]

and

w-e = skew[le].

(34.9)

J.C. Simo

334

CHAPTER 111

In summary, the multiplicative factorization (34.3) of the deformation gradient along


with definition (34.8) for the plastic velocity gradient in the spatial description result
in the additive decomposition (34.9) of the rate of deformation tensor d into elastic
and plastic parts. The following result characterizes the behavior of these objects under
superposed rigid body motions on the current placement St.
THEOREM 34.2. In a superposed rigid body motion p+(, t) = r(t) + Q(t)p(o,t)
on the current placement St, with Q(t) E SO(3) and r(t) C R rndi (dim = 3), the
following transformations hold:
(i) F + = QF,
(ii)

F e + = QFe and FP+ = FP .

= QQT + QiQT le+ = QQT + QleQT

and /P+

(34.10a)
= QlpQ

T .

(34.10b)

PROOF. (i) Relation F + = QF was shown to hold in Section 31. Relation FPm = FP
is evident since the intermediate configuration is unaffected by spatial rigid body
motions. The multiplicative factorization (34.3) then implies the relation F e+ = QFe.
(ii) Relation l+ = QQT + QiQT was also shown to hold in Section 31. Similarly,
time differentiation of Fe + = QFe gives relation e+ = QQT + QleQT. The last
relation IP+ = QIPQT in (ii) follows from the preceding two and the definition
lp = I l

e.

[]

REMARK 34.1. Result (ii) immediately implies the objective transformations de+ =
QdeQT and dP+ = QdPQT. All these results are to be expected with the exception
of the objective transformation IP+ = QlPQT for the full plastic velocity gradient in
the spatial description. This relation implies, in particular, the objective transformation
i,-p+ = QipQT for the plastic spin. Obviously, tensor fields defined on the intermediate configuration, such as Ce and L P, are unaffected by rigid motions superposed
on the current placement.
REMARK 34.2. A much debated issue in finite strain plasticity is concerned with the
apparent lack of uniqueness inherent to the multiplicative factorization (34.3), which
arises when considering superposed rigid body motions on the intermediate configuration. In one such rigid motion, the deformation gradients transform for any Q c SO(3)
according to
Fe + = FeQT and

FP+ = QFP so that F + = F.

(34.11)

The entire issue depends on an a priori specification of the class of admissible rotations
Q for such transformations. This question is related to a constitutive assumption on the
symmetry group of the material and appears to have little to do with any fundamental
principle in continuum physics. In particular, if FP is completely specified by the
flow rule, as in the case of the single crystal model (see ASARO [1983]), the subgroup
of admissible rotations on the intermediate configuration is the identity. Thus, no
uniqueness issues or invariance questions arise.

Nonlinear continuum mechanics

SECTION 35

335

35. Elastic response and free energy for multiplicative plasticity


As in the infinitesimal theory [assumption (A2)], the free energy function T is assumed to depend locally on the deformation through the elastic part of the deformation
gradient. From a micromechanical standpoint, T represents the stored energy associated with the elastic lattice deformation. Assuming that hardening mechanisms are
uncoupled from the elastic deformation and enforcing at the outset frame invariance
relative to superposed rigid body motions on the current configuration, one is led to
the functional form
T1(Fe, (se) = W(Ce) +

(J )

where Ce = FeTFe.

(35.1)

The dependence of T on the deformation via C e is used by a number of authors,


MANDEL [1972, 1974] in particular, and occurs in the original work of LEE [1969].
The formulation of a model of plasticity incorporating {FP,I } as microstructural
variables and the internal energy (35.1) then proceeds as follows.
(A) Dissipationfunction and constitutive relations. Recall that the stress power in
the finite strain theory is given by Pint = P F = r d. In addition, for the purely
mechanical theory, the local dissipation D is the difference between the stress power
and the rate of change in internal energy, i.e.,
D) =

*rd

- a(Fe,

(a) .

(35.2)

As in the development of the infinitesimal theory, we accept the inequality D 0,


for all possible motions on 2 x II, as a basic principle. The goal is to identify the
appropriate constitutive equations implied by this dissipation inequality. Using the
kinematic relation
C e = 2FeTsym[FeFe-l]Fe= 2FeTdeFe,

(35.3)

which follows from (34.9), the time rate of change of the free energy is evaluated as

atT = aCeW( e)

e +a

aiy)
hint

=2Fe[aeW(Ce)]FeT

* de

().

(35.4)

Inserting this result in (35.2) and making use of the additive decomposition (34.9)1
implied by (34.3), results in the inequality
nint

D= (T-2Fe

[a, W]Fe ) . de +r . dP+ CZ=l

qa~ > 0,

(35.5)

J.C. Simo

336

CHAPTER III

where we have set q = -ab7. A standard argument in constitutive theory (see


TRUESDELL and NOLL [1965]) then produces the following constitutive relations
T = 2Fe[ZeW( e)]FeT

and

q' =-aa().

(35.6)

The dissipation inequality then takes the reduced form


Lint

D= r

dP+

q'fl t

in

x ,

(35.7)

a=1

which is viewed as a restriction to be placed on the evolution equations that describe


plastic flow (corresponding to assumptions (A3) and (A4) in the infinitesimal theory).
(B) Description relative to the intermediate configuration. It is a trivial matter to
express the preceding results entirely in terms of objects defined on the intermediate
(local) configuration defined by FP . To do so, let
= Fe- TF

e -T

(35.8)

denote the symmetric second Piola-Kirchhoff stress tensor relative to the intermediate
configuration. From (35.8) and (35.6) one obtains the equivalent statement of the
constitutive equations
S = 2a-eW(Ce)

and

(35.9)

qf = -a().

Similarly, noting that r dp = -r 1 (since - is symmetric) and using (35.8) together


with result (34.6), the reduced dissipation inequality (35.7) becomes
?int

Dz= [CeS] .LP+

in

x .

(35.10)

a=1]

This form of the dissipation inequality appears in MANDEL [1972], see also ANAND
([1985], Eq. (23)). The generally nonsymmetric stress tensor H := [C eS] conjugate to
=
the plastic rate of deformation L P is restricted by the symmetry condition C eT- C e-1 and first appears in MANDEL [1972] who refers to L P as the plastic distortion
rate.
36. Plastic flow evolution equations for multiplicative plasticity
To complete the formulation of a model of finite strain plasticity it is necessary to
specify the evolution equations for the internal variables. These are either {IP, , } if a
spatial description is adopted, or {L P, , } if a description relative to the intermediate
configuration is adopted. This is the aspect of the theory where the key differences
between existing models arise.

Nonlinear continuum mechanics

SECTION 36

337

(A) Description in the current configuration. Models of plasticity widely used in numerical simulations are typically formulated directly in the current configuration. See,
e.g., the review articles of TVERGAARD and NEEDLEMAN [1981] and HUGHES [1984].
To establish a connection between these models and formulations of multiplicative
plasticity, consider first the case in which the elastic domain is defined in terms of the
spatial Kirchhoff stress as
E = {(r, q) E S x Rnin': f(-r,q') < O for

= 1,2,..., m}.

(36.1)

As pointed out earlier, frame invariance necessarily implies that f, (QrQT,.) =


fil(r, ), thus restricting the yield condition to isotropic functions. The simplest example is furnished by the von Mises criterion (m = 1) described in Chapter I.
The model of associativeplasticity corresponds to the assumption of maximum dissipation at a fixed configuration which, in view of (35.7), translates into the inequality
nint

r-

>

[q -.dq]~

(T,
('] qua) E E.

(36.2)

a=1l

An argument identical to that described in Section 6 then yields the local evolution
equations
m

-zyrf,(r,

dP

q),

/1fb~~~~=l~

>0,

f,(r,

q)

O and

aqvf(r, qf),
q

,=

(36.3)

m=

Ey"f

(,

q) =0.

t/=1

To complete the theory it is necessary to provide an evolution equation for iwjP so that
the evolution of P is defined. Two possible options are:
(i) The restriction of the entire formulation to isotropy. In this situation, the orientation of the intermediate configuration is irrelevant and the plastic spin remains
arbitrary.
(ii) The specification of a constitutive equation for the plastic spin wiiP, a point
brought into consideration in KRATOCHVIL [1973]. This is clearly possible since, according to (34.10), iv-P is objective. A common choice is iiP = 0. Other options are
examined in DAFALIAS [1984] and ANAND [1985].
Interestingly, the plastic flow rule (36.3) can be obtained from the corresponding flow
rule in the infinitesimal theory merely by replacing EP with dP. This form, however, is
not the most convenient one for the development of numerical algorithms. The following result, due to SIMO and MIEHE [1992] in the restricted context of isotropy, yields
the optimal parametrization of (36.3) from the point of view of numerical analysis. Let
,b e = F{at [F-lbeF- T] }FT = be

Ibe - belT

be the Lie derivative of the elastic left Cauchy-Green tensor.

(36.4)

338

J.C. Simo

CHAPTER II1

THEOREM 36.1. The following relation holds: -fbe

= sym[lPbe].

(36.5)

PROOF. By making repeated use of (34.3) and definition (34.5) for L P one obtains the
identity
-I
FeLPFeT= F[(FP

PFP

FP

-T

] FT

= -F[at(FP-) FP -T]FT.

(36.6)

Now observe that (34.3) implies be = FFT


F[FP-FP- T]FT . Combining this
result with (36.6) yields, in view of expression (36.4),
2sym[FeLPFeT ] = -F[

be

t(FP-IFP-T)]FT= -

(36.7)

e.
The result follows by noting that (34.6) implies FeLpFeT = IPb

Consider now the specific constitutive assumption ivP = 0. Using (36.5), the flow
rule (36.3) can be rewritten as
vbe = sym

[
(

a,

bej,
]fp(r,

(36.8)

which is the form used in SIMO [1992] in the development of a class of algorithms
that inherit all the features of the infinitesimal theory. These methods will be described
in the following chapter in a more general setting.
REMARK 36.1. If the additional assumption of isotropy is made, the entire theory
can be described in terms of the variables {b e , 5, } and the Kirchhoff stress tensor
i-; i.e., entirely in the current configuration. Isotropy means full S0(3) invariance
with respect to rigid body motions of the intermediate configuration; hence the stored
energy function depends on be since
W(Q

QT ) = W(Ce) VQ E S0(3) <#=

W(C

) = W(be).

(36.9)

The Kirchhoff stress tensor is then computed either by the well-know relation (see
TRUESDELL and NOLL [1965]) r = 2abW(be), or using the representation theorems
for isotropic tensor functions summarized above. Clearly, r and be are co-axial.
(B) Description in the intermediate configuration. Recall that in a description of
plastic flow relative to the intermediate configuration the basic internal variables are
{LP, ,}, along with the second Piola-Kirchhoff stress tensor S defined by (35.8).
This is the description adopted in most micromechanical investigations of crystalline
plasticity. Here, attention is restricted to the formulation of associative models of
plasticity with elastic domain
=

(_, q') E

X Rni": f

-9, )

< 0 for pI = 1, 2 ....

7mI.

(36.10)

Nonlinear continuum mechanics

SECTION 36

339

In contrast with the situation found in the spatial description, frame invariance no
longer restricts the convex functions f (-, ,) to isotropy. The first step in the derivation of the flow rule is to enforce the symmetry condition S = S and cast expression
(35.10) for the internal dissipation into the equivalent form
nint

D = S. DP +

qfl

>0

where DP = sym [

eLP].

(36.11)

a=l

Definition (36.11)2 appears in SIMO ([1986], Eq. (33.14b)) and gives the symmetrization of LP relative to 0 e, which is consistent with the interpretation of Ce as the
metric tensor in the intermediate configuration. The same definition for D P occurs in
MORAN, ORTIZ and SHIH [1990] where, consistent with the view of C e as a metric,
the plastic spin is also defined as the skew-symmetric part of L P relative to C e. This
leads to the decomposition
LP = Ve- [D p + WP]

where W

= skew[CeLP].

(36.12)

Since in most metals the elastic deformations are typically small in comparison with
the plastic deformations, (36.12) differs only slightly from the usual decomposition
of L P into symmetric and skew-symmetric parts. By adopting the inequality
int

[ S-S]

DP + E

[q -q]

(S,

E,

(36.13)

ol=l

the same argument quoted in the derivation of the flow rule (36.3) now yields the
evolution equations

DP = CE yasff(s,

q'),

Ypaq f

(S, q(),3

(36.14)

0,
ly

fA (Sq)

and

y/ff(SSqO) = 0.
/z=1

The formulation is completed by specifying an independent constitutive equation for


the plastic spin W P. For single crystals, this additional constitutive relation arises
as a result of Schmidt's law; see HILL and RICE [1972], ASARO [1983], HILL and
HAVNER [1982] and the more recent discussions in BOYCE, WEBER and PARKS [1989],
HARREN [1991] and MORAN, ORTIZ and SHIH [1992] among others. For poly-crystals,
the widely used assumption [see (36.12)]
WP =

LP = e-lDp,

(36.15)

340

J.C. Simo

CHAPTER III

specifies the orientation of the intermediate configuration via (36.14) and (36.15)2.
REMARK 36.2. In the work of MANDEL [1974] the yield surfaces defining the elastic
domain E are assumed to take the functional form f ( , q") < 0, supplemented by
the following dissipation inequality
ninth

+ YE [q -

[a-i*

a=1

](a

(+,

(36.16)

These hypothesis result in the replacement of the six-dimensional flow rule in (36.14)
by the alternative nine-dimensional flow rule

ZL

,iq().

Ey
f

(36.17)

/=

However, as noted in LUBLINER [1986], Mandel's derivation of (36.17) from the


principle of maximum dissipation is questionable since the symmetry constraint
~e-l

:Te-1

is not accounted for; see LUBLrNER ([1990], p. 460) for

further discussion and an alternative form of (36.17). Observe that the constraint
e-'l

v=

Te-l

is a restatement of local balance of angular momentum.

REMARK 36.3. The preceding formalism is consistent with the following geometric
interpretation. Endow the local intermediate configuration with the local Riemannian
metric defined by C. Relation (34.6) then asserts that IP is the push-forward to the
current configuration St of a contra-covariant rank-two tensor, L P, in the intermediate
configuration. Similarly, with the definitions given above, (34.6) implies the relations
dP = Fe-TDPFe - I

and

jjP = Fe

TWPFe-I,

(36.18)

which are push-forward transformation rules if D P and W P are viewed as covariant


rank-two tensors on the intermediate configuration. This interpretation agrees with the
view of (36.11)2 and (36.12)2 as "index lowering transformations" with the metric
tensor e. In the same vein, relation (35.8) is the appropriate push-forward transformation rule if the stress field S is viewed as contravariant rank-two tensor field on the
intermediate configuration. This interpretation is consistent with the view of the term
S DP in the dissipation function (36.11)1 as a contraction between a contravariant
and a covariant tensor field. In summary, the description relative to the intermediate
configuration is the convected description of elastoplasticity with metric tensor C e.
This geometric point of view is developed further in SIMO and ORTIZ [1985] and SIMO
[1986].

Nonlinear continuum mechanics

SECTION 37

341

TABLE 36.1

Elastoplastic model of multiplicative plasticity.


(i) Kinematics induced by the multiplicative factorization F= FeFP. Setting
P=l-l

wherel=FF-l
P

it follows that I = FeLPFe

le = FeF,

and

with LP = ibPF

P-

(ii) Constitutive equations for a free energy function


r=

F [2D

W(Ce)]FeT

and

qa =-a-i(

= W(Ce) + 7{()
a)/ac,

where Ce = F TFe is the elastic right Cauchy-Green tensor.


(iii) Evolution equations describing associative plastic flow

= /P

yP

,ylarf(T,q),
la
(T,q),

f(r,q)

f0,

These equations imply

iiP =

<O

and

= 0.

fu(,q)

skew[lP] = 0.

REMARK 36.4. Using the transformation relations defined by (36.18) and (35.8), the
flow rule (36.14)h in the intermediate configuration implies the following flow rule in
the current configuration
m

dP=

n.

where/

= Fe-T [ -Sf . (

q)]

(36.19)

/=1

which furnishes the generalization of the spatial flow rule (36.3)1 to anisotropic response. Relation (36.19)2 becomes a restatement of the chain rule, with n, given
by n = af,(r,
qn) if the yield functions satisfy the relation f (FeTFeT,.) =
f,(S, .). By Noll's rule (TRUESDELL and NOLL [1965]), this covariant property is an
equivalent statement of the isotropy condition.
To render matters as concrete as possible, in what follows we shall be concerned
almost exclusively with a formulation of plasticity relative to the current configuration,
as explained in detail above. For the convenience of the reader and easy reference,
Table 36.1 summarizes the key relations of the elastoplastic model that will be used
throughout our subsequent developments.
37. Volumetric/deviatoric uncoupled finite plasticity: J2-flow theory
A large body of experimental literature supports the fact that plastic deformations
in metals preserve volume. It is for this reason that the assumption of an isochoric

342

J.C. Simo

CHAPTER iI

plastic flow is introduced at the outset in nearly all treatments of metal plasticity,
as exemplified by the classical monograph of HILL [1950]. The goal of this section
is to specialize the theory of finite strain plasticity developed above to this situation
relevant to metal plasticity. The resulting theory is illustrated by constructing an extension of the classical model of J 2-flow in which the elastic response is characterized
by uncoupled volumetric/deviatoric hyperelastic relations. The developments in this
section rely heavily on the results presented in Section 32 for uncoupled models of
elasticity at finite strains.
The multiplicative factorization of the deformation gradient provides a rather convenient mechanical setting for the enforcement of the volume-preserving constraint
on the plastic flow. Within this setting, this condition translates into the kinematic
assumption det[FP] = 1 on the determinant of the plastic deformation gradient. The
following purely kinematic results give the rate of change of the total, the elastic and
the plastic volumes.
THEOREM 37.1. The time derivatives of J = det[F] and Je = det[Fe] are given by
J = Jtr[d] and Je = Jetr[de],

(37.1)

where d = sym[FF-t], d e = sym[FCFe ']. The time derivative of JP = det[FP] is


given by
JP = JP tr[L] = JPtr[dP],

(37.2)

p where L P = PPF
' and dp = sym[FeL PFe 1] is the plastic rate of deformation
tensor in the spatial description.

PROOF. The following standard result gives an explicit formula for the time derivative
of the determinant of an invertible time-dependent matrix:
d det[A(t)] = det[A(t)] tr[A(t)A- (t)]

(37.3)

for any A(t) E GL(n). Relations (37.1) follow by setting A(t) = F and A(t) = Fe
in the preceding formula and noting that a skew-symmetric matrix has zero trace.
Relation (37.2) also follows from (37.3) by setting A(t) = FP , together with the
property
tr FeLPFe -

!]

= tr[P]

IL~~~~~~~~~~~~~(74

for the trace of a matrix, and the definition of d p in terms of L P.

(37.4)

343

Nonlinear continuum mechanics

SECTION 37

An immediate consequence of relation (37.2) is that the condition JP = 1 is equivalent to requiring that tr[dP] = 0 for all t E If.A formalism based on the use of Lagrange
multipliers provides an obvious way to enforce this constraint. Rather than pursuing
this route, we shall adopt the framework described in Section 32 for the treatment of
models of finite elasticity exhibiting uncoupled volumetric/deviatoric response, and
explore the consequences of considering a free energy function of the following form:
(37.5)

= U(J) + W(Ce) + ((~).

In this functional form C e denotes the elastic, volume preserving, right Cauchy-Green
tensor, defined by the relation
Ce =

eTFe

where Fe = Je-I/ndimFe

(37.6)

which ensures that det[Fe ] = det[Ce] = 1. For the moment, no constraints will be
placed on the determinant JP of the plastic deformation gradient.
(A) Hyperelasticstress-strainrelations in the spatialdescription. To derive the elastic stress-strain relation associated with the free energy function (37.5) we mimic the
procedure employed in Section 35 and compute first the time rate of change of BT,.
Time differentiation of definition (37.6)1 gives
C

= 2Je- 2/ndimFeT [Fe-T

eFe-

dim

Jetlel]

(37.7)

ndim

By inserting (37.1)2 into this expression and making use of the relation Ce =
FeTdeFe derived in Section 35, the time rate of change of Ce can be written using
definition (37.6)2 as
Ce =

eTdev [de]Fe

(37.8)

where dev[de] = de - (l/ndim) tr[de]l is the deviator of the elastic spatial rate of deformation tensor. This relation together with (37.1)1,2 implies the following expression
for the time rate of change of the free energy function:
at@= [JU'(J)] tr[d] + dev[Fe2aeW(Ce)FeT]

~,~1(~).

de

(37.9)

a=1

By making use of the additive decomposition d = d e + dP and inserting expression


(37.9) into the dissipation inequality D = r d -

tg > 0, the same arguments

344

J.C. Simo

CHAPTER

employed in Section 35 yield, in the present context, the constitutive relations


r = [JU'(J)]1 + dev [F e28w
f(Ce)
j

eT]
(37.10)

and

q' = -8(

,),

together with the following expression for the reduced dissipation inequality:
D = dev[l] dP +

q>

O.

(37.11)

UY=

An important feature of this result should be noted. Since dev[T] dP = dev[-r]


dev[d] it follows that neither the trace of the Kirchhoff stress tensor nor the trace
of the plastic rate of deformation tensor, which describes the evolution of plastic
volume according to the relation tr[dP] = dlog(JP)/dt implied by (37.2), appear
in the reduced dissipation inequality. In other words, the Kirchhoff pressure field
7t = (1/ndim) tr[-r] does not contribute to the plastic dissipation in the material. This
result is possible only if tr[dP] = 0 during plastic flow or, equivalently, if JP = 1
throughout the entire deformation process. Therefore, we arrive at the remarkable
conclusion that the free energy function (37.5) automatically implies the isochoric
condition on the plastic flow.
(B) Associative evolution equationsfor isochoric plastic flow. To complete the formulation of the theory it only remains to formulate appropriate evolution consistent
with the condition of isochoric plastic flow. We shall do so by adopting the framework
of associative plasticity, as described in Section 36. Since plastic flow is to be independent of the pressure field, the yield criterion formulated in the spatial description
can depend only on the deviatoric part dev[-] of the Kirchhoff tensor. This conclusion
is one of the key hypotheses made in classical treatments of metal plasticity, see HILL
([1950], Chapter I). Accordingly, the elastic domain is assumed to be defined as
E := {(dev[r],q&) E

x Rni"': f,(dev[r],q) < 0

for/ = 1, 2,..., m},

(37.12)

and subject to the usual conditions which require the functions f(-, .) to be convex
and assume that the m-constraints are qualified. Yield criteria which are independent
of the hydrostatic pressure field, of the type involved in this definition of E, are
referred to as pressure insensitive. The evolution equations of associative plasticity
emanate from the principle of maximum dissipation which, in the present context,
results in the following inequality implied by expression (37.11):
hlint

[dev[r] - dev[;]] . d p + E
a=1

[q

- j]

V(dev[r], qc)

EI.

(37.13)

345

Nonlinear continuum mechanics

SECTION 37

Once again, a standard argument in convex yields the following associative evolution
equations for isochoric plastic flow:
m

dP =

y7 af, (dev['r], qa),

y:

aq f (dev[,r], q),
(37.14)

'

>s0,

fS, (dev[r],q)

<0

and

Ey7f,(dev[r],q)

=0.

/v=1

Observe that by exploiting the relation dev[r] =

- (/ndim)

tr[r]l, the evolution

equation for dP can be written in the equivalent form


m

dP =

,y

dev [adev[,] f (dev[r], qa)],

(37.15)

/L=1

which is consistent with the condition tr[dP] = 0 and implies, therefore, the isochoric
constraint on the plastic flow. As in the general theory, the plastic spin has to be specified via an additional evolution equation, the simplest example of which corresponds
to the widely used assumption wP = 0.
EXAMPLE 37.1 (J2-flow theory). The von Mises yield criterion of metal plasticity is
the prototype of a pressure-insensitive yield condition. For isotropic hardening, the
elastic domain takes the form
E = {(dev[-r],q)
= dev[r] +

x
[

+: f(dev[T],q)

- q] < 0},

(37.16)

and the associative evolution equations that describe plastic flow, the so-called LevySaint Venant flow rule, become
d =

and

wheren-

dev[(37.17)
[dev[r]

The plastic multiplier is subject to the standard Kuhn-Tucker conditions y > 0 and
yf(dev[r], q) = 0. We provide below two examples of hyperelastic constitutive equations that complete the extension of this classical model of metal plasticity to the finite
strain regime within the framework of a multiplicative decomposition of the deformation gradient. Because the general evolution equations (37.14) ensure that JP = 1
throughout the entire range of the deformation, it is permissible to replace J by Je
whenever the former variable appears. In keeping with the preceding framework,
however, we have chosen not to perform this replacement.

346

J.C. Simo

CHAPTER III

Model I. The first model is a direct extension of Example 32.1 considered in Section 32, as defined by the free energy function
T=

-_ -

log(J)] + l(tr [Ce]

nfdim) -+X(),

(37.18)

which is clearly of the form (37.5). The function X(() models the specific form of
isotropic hardening in the material via the constitutive equation q = -- '((). The
stress-strain relations defined by (37.10)1 then become
r = i(2 _ 1)1 + tdev[b6 ]

where be = FeFeT .

(37.19)

The free energy function (37.18) was introduced in SIMO [1987].


Model II. The second model is based on a description of elasticity in principal
elastic stretches via the following free energy function
ndim

[log(J)] +/

yMi ()

+ l(),

(37.20)

A=]

where A' > 0, A = 1,..., ndim, are the eigenvalues of the elastic, volume preserving,
deformation gradient Fe. This model clearly falls within the class of free energy
functions (37.5) since (Ae) 2 are the eigenvalues of C e. The stored energy function
part in (37.20) is known as the Henky model, see ANAND [1979] for a discussion
of the applicability of this model. Using the results in Section 33, the stress-strain
relations (37.10) are easily shown to reduce to

,r= r log(J)1 +

log(be).

(37.21)

This model is discussed in detail in Chapter IV within the context of a recently


proposed class of exponential return mapping algorithms.
38. Rate form and variational inequality for multiplicative plasticity
The model of multiplicative plasticity summarized in Table 36.1 provides a hyperelastic constitutive equation for the Kirchhoff stress field r relative to the intermediate
(released) configuration. This model can be cast into a rate constitutive equation for
the Jaumann derivative of the Kirchhoff stress tensor by enforcing the consistency
condition using a technique entirely analogous to that employed in the infinitesimal
theory. The resulting rate constitutive equation arises naturally in the formulation in
the incremental (or rate form) of the initial boundary value problem.
Remarkably, it will be shown below that older models of plasticity, with elastic
response characterized in terms of the hypoelastic constitutive equation -= aid- dPj,
which is postulated at the outset, can be exactly recovered from the rate form of the
model of multiplicative plasticity summarized in Table 36.1. These two formulations of

Nonlinear continuum mechanics

SECTION 38

347

plasticity differ, therefore, exactly in the same manner as the theories of hyperelastic
and hypoelastic material response, see TRUESDELL and NOLL [1965]. In short, the
multiplicative theory of plasticity provides an explicit expression for the incremental
elastic moduli a, involving derivatives of a stored energy function, while the hypoelastic theory regards the moduli a as constitutive parameters to be specified a priori.
As pointed out in the introduction to this chapter, hypoelastic models of plasticity are
widely used in older algorithmic treatments of the subject, see KEY and KRIEG [1982],
TAYLOR and BECKER [1983] and the review articles of NEEDLEMAN and TVERGAARD
[1984] and HUGHES [1984].

The first step in the derivation of objective rate forms of the hyperelastic constitutive
equations is the computation of the material time derivative of the Kirchhoff stress.
Time differentiation of the hyperelastic constitutive equation for the Kirchhoff stress
and use of the kinematic relations in Table 36.1 gives
*= leT eq-T

,T

BFe[c(lVe)]F
-

(38.1)

where C = 4a2ecW(C
e) is the elasticity tensor in the intermediate configuration.
Define the elasticity tensor c in the current placement by the push-forward relation
Cijkl = FeAFeBFkeFcFeDCABCD

Since le = FeFede = sym[FeF

where CABCD = 4 aC Bc

(38.2)

and de = sym[le], using the kinematic relation


e-

l] =

Fe-T [FeTFe + FeTFeF

e- 1

= Fe-T[Ce]F e - l,

(38.3)

Eq. (38.1) can be written as


= le

+leT

+ cde.

(38.4)

Now recall the kinematic relation le = de + ivu e. Denoting by

= .-

e.

+ rjie

the Jaumann derivative computed with the elastic spin tensor, relation (34.7) takes the
form
=

ade

where

aijk = Cijkl + Tik6jl + Tiljk.

(38.5)

In the model in Table 36.1 the plastic spin is specified as ii P = 0 which, in view
of the additive decomposition (34.8) implied by the multiplicative factorization of the
deformation gradient, is equivalent to the condition wv = we . By specializing result
(38.5) to the current model, the following two alternative forms of the rate constitutive
equations are obtained:
V

=ade=a[d- d P]

and

r-=cd-adP.

(38.6)

348

J.C. Simo

CHAPTER III

The second relation in (38.6) is a direct consequence of definition (30.12), result (38.5)
and the additive decomposition l = -IP specific to this model. Expression (38.6)1 in
terms of the Jaumann derivative arises naturally in the interpretation of multiplicative
plasticity as a variational inequality, while expression (38.6)2 in terms of the Lie
derivative arises naturally in the incremental formulation of the initial boundary value
problem. It is clear that any objective rate other than the two appearing in (38.6) can
be used in the formulation of the incremental elastic constitutive equations, although
little is to be gained from such an exercise.
The model summarized in Table 36.1 can be cast in the form of a variational inequality which arises as a result of the underlying assumption of maximum dissipation.
D2()/0~
be the generalized plastic moduli. Assuming that both
Let hd =
the tensor a in (38.6)2 and the nin t x nint matrix [h'a] are invertible, we have the
following result.
THEOREM 38.1. Denoting by (, ) the L 2 ( 2)-innerproduct, the constitutive model in
Table 36.1 is equivalent to the variational inequality
nint

at)-a-'

7, (Vv o

int

(q

4',h-)

oa=1 3p=l

V(r,q ) E E,
where E = ((,, q)

(38.7)

E Sx

f (r, q

R':

0 for

= 1,2,..., m}.

PROOF. Combining the rate constitutive equation (38.6)1 with the associative flow rule
using the relation de = d - dP yields, under the assumptions that a and [had] are
invertible, the result
'v=dP =E

d-a

3,"f(r

q),

/z=1

(38.8)
nit

-- , h:q
P=1

= CY
= /Oq fit(
A

Tqqu
qa)

3y=1

Inequality (38.7) then follows from the convexity property


ninth

[-

r] . -f(7rq)
< f (K , q- ) - f

(7r, q ),

[a

- qn]qf,(r'qC)

(38.9)

and the Kuhn-Tucker conditions. Since the argument is entirely analogous to that used
in the proof of (6.8), further details are omitted.
All that remains to complete the incremental or rate formulation of multiplicative
plasticity is to eliminate the plastic multipliers y > 0 in the flow rule recorded in

Nonlinear continuum mechanics

SECTION 38

349

Table 36.1 leading to closed-form expression for the elastoplastic tangent moduli. To
do so, one proceeds exactly as in the infinitesimal theory, see Section 5. Assuming
that the constraints are qualified one defines the trial set of constraints as
Jtr = { E 1,..., m}: f, (r, q)

= 0}.

(38.10)

The set of active constraints is then J = f{/ E Jtr: f(r, q") = O}. To compute
f, recall that f(,
q ) is isotropic as a result of the restriction imposed by frame
invariance. Consequently a,fu commutes with T and one has the crucial result

arft Te - af,

[T(a,f,) - (af,

Using the incremental stress-strain relation


it follows that
nint

)T

] = 0.

(38.11)

= a[d - dP] and the preceding result,

ninth

i, = arf, a[d - dP] - E

(a q f)h

(38.12)

;.

a=1 p=1

Now define the metric coefficients g = [g] = gT by the relation


nint

g=

nint

f,f-a afv+ E

(a f)h

(aq,fv),

/,V E

J.

(38.13)

a=1 O=1

Since the constraints are qualified and the basis vectors


at each point (, qu) E aE, it follows that the matrix g
ing exactly as in the infinitesimal theory, setting [g4"]
are obtained by enforcing the consistency requirement

gv

y =

f, .ad

,-f,. are linearly independent


= [g,] is invertible. Proceed= g-1 the plastic multipliers
f, = 0 leading to

E Jt.

for

(38.14)

VEJV

The set J of active constraints then becomes J = {i E J: -yA > 0}. By inserting
into the incremental constitutive equation (38.6)1 the flow rule for dP in Table 36.1,
with the multipliers 7y" defined by (38.14), one arrives at the incremental relation
V

aP d

where aeP = a -

g'V[aa,f,] [a ,fr].

(38.15a)

ltEJ vEJ

Clearly, the tensor aep of elastoplastic moduli satisfies the symmetry conditions
aijkl

tr

alij

and
and

tr

aijk = ajlk

ajikl.

(38.15b)

J.C. Simo

350

CHAPTER III

TABLE 38.1

Rate model of plasticity and relation with the multiplicative theory.

-=a[d-dP]and

=_E

+ Tro and d = sym[FF

-i

(i) Elastic rate constitutive equations with r =

h
h3s,
P=

where aijkl = aklij, aijkl = aijlk = ajikl and ha 3 = h3 a .


(ii) Evolution equations describing associative plastic flow

= E

0,

'

al -f 1 (r,

f(r,

qa),

o and

q)

aq f(,

a =

q),

f (, q) =

O.

i=1

p.

These equations imply ivP = skew[/P] = 0 so that Ip = d

(iii) Relation to multiplicative plasticity. Elastic and hardening moduli


aijkl = Ckl l + ik6Sjl + Tijk
where tf = W(Ce) + 7-(~0)

ha3 = a2R/acat,

is the free energy and

Cijkl = FA

Tij = FIAFjB2ace

and

FBF

CA B

akCFID4a e
acABaCD

Kinematic relation: It = FeLPFe - l with LP = FPF-I.

The rate equation (38.5) is phrased in terms of the Lie derivative and not the Jaumann
rate. By inserting expression (38.14) into the flow rule for dP the rate constitutive
equation (38.6)2 in terms of the Lie derivative reduces to
7, = cepd

where CeP = c -

E EgV[a-f

] [af,]

(38.16)

ktEJ vEJ

Here C is the spatial elasticity tensor defined by (38.2). This expression will be used
in the incremental formulation of the elastoplastic initial value problem.
EXAMPLE 38.1. For the convenience of the reader the basic relations that defined
the conventional rate model of classical finite strain plasticity and its relation to the
multiplicative theory are summarized in Table 38.1. To illustrate this relation in a
concrete setting, consider a stored energy function introduced in Example 31.1
W(Ce) = A(det[V
+

e]

+ 1/, (tr Ve]

- 1) -

3),

(,[ +A)log(det[Ce])

2tt[~]

(3.2
(38.17)

Nonlinear continuum mechanics

SECTION 39

where e =

det[

relation aZe Je =
stress becomes
r=

351

e] and A,p > 0 can be interpreted as Lam6 constants. Using the


Jee-l, the constitutive equations in Table 38.1 for the Kirchhoff

A(J e 2 -1)1 +

where be = FeFeT.

(b e - 1)

(38.18)

Finally, using relations in Table 38.1, the spatial elasticity tensor is easily shown to
be given by the expression
c = Aje2 1

1 + 2 [1 -

(Je2

1)A/i] I,

(38.19)

where I with components abcd = [6acSbd + 6ad6bc]/2 is the rank-four symmetric unit
tensor. An explicit expression for the tensor a is easily found by inserting (38.18) and
(38.19) into the relation in Table 38.1.
Now suppose that the additional assumption is made that the elastic volumetric
strain is small in the sense that Je = 0(1). This hypothesis is quite reasonable for
metal plasticity. In addition, for the von Mises yield criterion one easily concludes
that the magnitude of the deviatoric Kirchhoff stresses are of the order of the flow
stress ay. Suppose further that the hyperelastic constitutive equations of multiplicative
plasticity are replaced by the rate constitutive equation in Table 38.1, with the tensor
a defined by via the ad hoc constitutive assumption
a =A

1 + 2 I,

(38.20)

widely used in conventional treatments of plasticity. A formal argument that uses a


straightforward asymptotic expansion then reveals that the tensor a defined by (38.20)
differs from the tensor a associated with the model of multiplicative plasticity with
free energy function (38.17) by terms of order O(y//). Since the ratio vy/ of
the flow stress over the shear modulus is of the order of 10 - 3 for most metals, the
preceding argument provides a partial justification for the conventional (hypoelastic)
rate formulations of finite strain plasticity.
39. Variational formulation: Weak form of the momentum equations
As a first step in the formulation of the initial boundary value problem for elastoplasticity, we consider the weak formulation of the momentum balance equations for
three-dimensional nonlinear continuum mechanics. Since the interest here is nonlinear plasticity, a Lagrangian formulation of the problem is adopted throughout. For
plasticity problems, however, it is conventional and fairly convenient to cast the entire variational formulation in terms of fields, such as the Kirchhoff stress, defined
on the current placement St of the continuum body. The formulation remains nevertheless Lagrangian, since the parametrization of the problem is in terms of material
coordinates {XA} and the motion itself is the primary variable in the problem.
The weak form of the momentum equations, supplemented by local constitutive
equations for the stress field (say the Kirchhoff stress tensor r), gives rise to a highly

352

J.C. Simo

CHAPTER III

nonlinear initial boundary value problem for the motion t c- got of the continuum
body. The two sources of nonlinearity in this problem are: (i) Geometric, arising
from the intrinsic nonlinearity in the kinematic description of the problem and (ii)
Material, arising from the nonlinearity in the constitutive relations. These two sources
of nonlinearity become apparent when the weak form of the momentum equations is
formulated in rate or incremental form.
Motivated by the foregoing description of the admissible deformations in a continuum body, with reference configuration 2 c Tdi m and prescribed deformations i) on
the part of the boundary F,. C aQ, define the configuration space as the set
C=

{yP

C WI'P(2)nd mi :

det[Dp] > 0 in 2 and cplr = c,}.

(39.1)

For nonlinear elastostatics, the choice for the exponent p is dictated by growth conditions on the stored energy function. As before, St = cPt(2) denotes the current
placement of the body under a motion t i got E C. A time-independent displacement field superposed on St which does not violate the prescribed Dirichlet boundary
conditions on Fr is called an admissible (spatial) variation of the motion pot at time
t C I. Admissible variations span a linear space denoted by Vi and defined as
V =

: (Q)

Rdim:

(39.2)

(,(X)) = Ofor X E rF}.

In short, the space V, is the tangent space to the configuration got C at time t I.
Elements in V, are also referred to as spatial test functions. To remove the dependence
of ri E V , on P one defines material test functions, denoted by rlo: Q2 - I d'm, via
the change of variables ro(X) = 7r(cp(X)). Thus, material test functions span the
fixed linear space
Vo = {ro: Wl'P(Q)dim:

o(X) = 0 for X E

j.

(39.3)

It is convenient to think of Vo as given and then construct test functions 7 E V by


setting ri = r o c-' for each r0o C Vo. By the chain rule, it follows that
GRAD[7ro] = (Vri o cp)F,

i.e.,

70a A = 7rabFbA.

(39.4)

With this notation in hand the weak form of the momentum balance equations can be
formally justified by taking the L2(Q2)-inner product of (29.6) with any 7r0 C Vo and
using the divergence theorem. The result can be written as
Gdyn(P, V; 70) =/

po
od

dQ + G(P; o) =

(39.5)

for any (material) test function i0 E Vo. Here, G(P; r0o) denotes the (static) weak
form of the equilibrium equations which is given by
G(P; ro) = J P

GRAD[ro1 d2dQ
- J B ,od-

Jt,

r7odP.

(39.6)

Nonlinear continuum mechanics

SECTION 39

353

Statement (39.5) is considerably more general than the local form (29.6), since the
configurations 'ot are subject to less restrictive continuity requirements. The variational
equation (39.5) is supplemented by the weak form of the kinematic condition att V =0.
EXAMPLE 39.1. For elasticity the nominal stress P is a local function of the motion
defined by (31.1) in terms of the deformation gradient, i.e., P = aFW(Dwt). If this
constitutive equation is enforced locally then the static weak form becomes a function
of the motion t - 'Pt C and the dynamic weak form reduces to
Gdyn((Pt, V;

r0o)

pov

o0d2 +

G(aFW(Dwpt); o).

(39.7)

The dynamic weak form thus becomes a function of (t, V) Z where Z = C x Vo is


referred to as the canonical phase space for elastodynamics. The variational equations
Gdyn(pt,V;

0) =0 and

po(,t - V) 7odQ = 0,

(39.8)

for arbitrary r0o c Vo, then furnish the weak form of Hamilton's canonical equations
for elastodynamics viewed as an infinite-dimensional Hamiltonian system; see SIMo,
POSBERGH and MARSDEN [1991] for a detailed account of this formalism. As alluded
to above, only short-time existence results are known for this quasilinear hyperbolic
system (see MARSDEN and HUGHES [1983], Chapter 6).
For the pure traction problem, the weak form of the momentum equations exhibits
two conservation laws which are a direct consequence of the classical Euler law of
motion. Recall that the local form of the momentum equations are the direct result of
applying Euler's law of motion to a neighborhood O(X) of an arbitrary point X CG
in the reference configuration, see, e.g., TRUESDELL and NOLL [1965]. The global form
of balance of momentum, on the other hand, are obtain when Euler's equations are
applied to the entire reference configuration Q. Explicitly, let Next and Mreac denote
the resultant force of a prescribed system of external loads and the resultant force of
the reaction forces, respectively defined by
Next

=,

jBd2+

Td

and

Nreac=j

Pv d,

(39.9)

where 0o: 2 - S 2 is the unit outward field normal to the boundary 02. Similarly,
let Mext and Mreac denote the resultant torque associated with the prescribed system of
external loads and the resultant torque associated with the reaction forces, respectively
given by
Next=

t x Bd

t x
T

d,

NFxtfTPtXr
'Pt x~dP,
d2+J
Mreac = j ,t

x Pv

dr.

(39.10)

J.C. Simo

354

CHAPTER III

A point in the canonical phase space Z = C x Vo of the problem will be denoted


by z = (o, p) where p = poV is the generalized momenta. With this convention,
the total linear momentum and the total angular momentum about the origin of the
system are given by the standard expressions
L(z) =Jpd2

and

txpd2.

J(z) =

(39.11)

The classical Euler equations assert that any admissible motion t C11 - z(t) C Z in
the canonical phase space is to satisfy the balance laws
dL = Next + Nreac
dt

and

dt

J = Mext

Mrea.

(39.12)

Therefore, if the external forces are equilibrated, in the sense that Next = 0 and Mext =
0, then the total linear momentum and the total angular momentum are conserved for
the pure traction initial boundary value problem corresponding to Fr = 0. These
conservation laws follow immediately from (39.12) merely by noting that Nreac
Treac = 0 as a result of the condition Fr = 0.
These classical conservation laws can also be derived directly from the weak form of
the momentum equations. The explicit result is recorded below (for the case ndim = 3)
since this is the setting adopted in the numerical treatment of the initial boundary value
problem. The global conservation laws of momentum will play an important role in
the class of algorithms described in Chapter IV.
THEOREM 39.1. For the pure traction boundary value problem (,, = 0) under equilibrated loads (Next = Met = 0), the weak form of the equations yields the conser-

vation laws
d
d
d L=O and dJ=0
dt
dt

inZxE,

(39.13)

where Z = C x Vo is the canonicalphase space and L, J are defined by (39.11).


PROOF. Conservation of linear momentum trivially follows from the dynamic weak
form (39.5) by noting that for rp, = 0 arbitrary translations of the current placement
St = ot(Q2), defined by vector field x C St
r = C with
E Rndim constant,
lie in the space V, of admissible test functions. To prove conservation of angular
momentum one observes that arbitrary infinitesimal rotations of the current placement
also lie in V,. Therefore, the vector field
=

t(X) E St - r1 o(X) = C x

t(X) C V0,

(39.14)

is an admissible (material) test function for any E Rt di m, with gradient GRAD[ro0] =


CDwt where is the skew-symmetric matrix with axial vector C. A straightforward

SECTION 40

Nonlinear continuum mechanics

355

manipulation that exploits the relation r = P DotT along with the symmetry condition
r = rT then implies
P .GRAD[lo] = P

Dt
.

= PDoT *

= r.

= 0.

(39.15)

Using this result and expression (39.10)1, we conclude that the static part of the weak
form of the momentum equations, defined by (39.6), reduces to
G(P, C x

t) =

- [J

t x Bdf2+ gI

x TdFj = C - Mext,

(39.16)

which vanishes since Mext = 0 by assumption. Finally, combining (39.16) with expression (39.5) for the dynamic weak form and using definition (39.11)2 yields
Gdyn(P; C X Wt) = (

ot x po dQ = C aJ = 0,

(39.17)

which implies the result since ( is arbitrary.


The preceding argument is closely related to techniques used in the derivation of
conservation laws for systems of differential equations which are invariant relative to
a one-parameter group of transformations, see, e.g., OLVER [1986].
REMARK 39.1. The variational formulation of the equations of continuum mechanics
discussed above is formal in the sense that little is known about the appropriate mathematical structure for specific models. For instance, in elastostatics the only known
satisfactory theory is based on minimizers of the potential energy with a poly-convex
stored energy function (BALL [1977]). Existence can be proved in a suitable Sobolev
space (typically W 1 p, with p large enough as dictated by certain growth conditions),
but it is not clear in what sense the minimizers satisfy the weak form (39.6). For a
recent review of known results we refer to CIARLET [1988]. No rigorous mathematical results are currently known for finite strain plasticity, except in the context of the
infinitesimal theory where a complete existence theory does exist; see the accounts
given in Chapter I and references cited therein. The so-called deformation theory is
treated in TEMAM [1985].
40. The total and incremental weak forms of momentum balance
The numerical solution of the weak form of the momentum equations for nonlinear
problems in solid mechanics is accomplished via an iterative scheme that involves
the repeated solution of a sequence of linearized problems. These linear problems
arise from the formulation of an incremental or rate version of the weak form about
a sequence of configurations, leading to the so-called incremental method. For nonlinear elasticity, the mathematical basis of this widely used technique is examined in
BERNARDOU, CIARLET and Hu [1985]. The statement of the weak form of the equations and its associated rate version will be given below in the current configuration,

356

IC. Simo

CHAPTER III

since this is the description widely adopted in the numerical solution of plasticity
problems.
Although the weak form (39.6) can be phrased entirely in terms of fields defined
on the current placement of the body, the formulation remains Lagrangian because,
implicitly, the configurations 'pt
C are regarded as the primary variables. The key
result used in this alternative statement of the weak form arises from the following
chain-rule relation
P

GRAD[r70] =

=
where r = o0 o

GRAD[ro]

= r

GRAD[ro]F

- (V/o t)

(40.1a)

'. In components, (40.1a) is equivalent to the expression


(40. lb)

PaA(X)TOa,A(X) = [Tab()Tlab(X)] =ot(X)

Setting G,, (r; ) = G(P; ro), substitution of (40.1a) into (39.6) then produces
the result
Go,(r; ) =

r (V
'i(-

o pt) dQ -J

B (

opt)d2

t *(77 o Wt)dF.

(40.2)

To simplify the notation we shall often omit explicit indication of the composition
operation involved in the change of variables from material to spatial coordinates,
with the understanding that all the quantities involved are evaluated at appropriate
points. This will usually be clear from the context.
The rate version of the weak form of momentum balance arises in the following
context. Suppose that (pt E C is a specified configuration, with associated Kirchhoff
stress field r and spatial velocity field v, in dynamic equilibrium under prescribed
loading defined by B in
x and t on T x . Suppose further that the loading
is incremented by rates B and leading to an incremental change P in the nominal
stress. Assuming sufficient smoothness so that time differentiation under the integral
sign is permissible, the incremental version of the dynamic weak form becomes
atGdyn(P,V;o)0) =/

poat (V)

o d2 + atG(P,77) = 0.

(40.3)

The key step in the reformulation of this result in terms of objects defined on the
current placement St, involves the following identity obtained by time differentiation
of the relation P = FS
P = FS + FS = [FSFT

(PF

') (FSFT)]F T.

(40.4)

Nonlinear continuum mechanics

SECTION 41

357

Using the relations r = FSFT and Er = [FSFT] along with expression (28.22),
identity (40.4) can be written as

P = [vr + (Vv o

t)r] F- T.

(40.5)

This result together with the relation GRAD[ro] = (Vr o Wt)F arising from the chain
rule gives

a, [P

GRAD[1o]] = [vT- + (Vv o Wt)T]

(V7i o 't).

(40.6)

Assuming again sufficient smoothness so that time differentiation under the integral
sign is permissible, (40.6) and (40.2) yield the following expression for the static term
tGp (r; r7) = atG(P; 7o0) in the incremental weak form (40.3) at configuration 'pt
atG, (r; )

f
-

[vr + (Vv o pt)r] . (Vq

B.( /o

) dQ -

o W't)
(

dQ2

o t) dP,

(40.7)

for all admissible test functions rq V. Observe that the objective rate that naturally
arises in the incremental form (40.7) is the Lie derivative of the Kirchhoff stress. It
is conventional in the numerical analysis literature to use the following nomenclature
for the first two terms in (40.7):
(i) Geometric term: [(Vv o 'Pt)r] (Vr o

'Pt),

also referred to as the initial stress

term. For fixed stress field at a given configuration this term gives rise to a symmetric
bilinear form (see Chapter IV).
(ii) Material term: (7-) (V o t). For elasticity, the rate constitutive equation
f
= (Vv o Pt) gives rise to a symmetric bilinear form at fixed configuration
'Pt C (see Chapter IV).
We shall follow this nomenclature here and in subsequent chapters. The most important application of the preceding rate form of the momentum equation arises in the
numerical solution of the initial boundary value problem by the so-called incremental
method. For the quasistatic problem, the spatial velocity field v is replaced by a displacement increment Au E V,, and weak form is discretized in space via a Galerkin
finite element method. For the dynamic problem it is also necessary to perform a time
discretization. Further details will be given in Section 41 and in Chapter IV.
41. Initial boundary value problem: Dissipativity and a priori estimate
The weak form of the momentum equations discussed in the preceding section, together with the constitutive equations described in Section 34, comprise the initial
boundary value problem for elastoplasticity at finite strains. This section provides
the complete statement of this problem in the form used in our subsequent algorithmic analysis, gives the key (formal) a priori estimate responsible for the dissipative
character of the solutions.

358

J.C. Simo

CHAPTER III

To make matters as concrete as possible, attention will be restricted to a conventional


model of plasticity in which the elastic domain is specified in terms of true (Kirchhoff)
stresses defined on the current placement of the body, as summarized in Table 36.1. The
primary variables in the initial boundary value problem are the deformation qot C,
the velocity field v C V,, and the internal variables (Fe, ~) C , where
= {(Fe,o): Q

-*

GL+(ndim)

(41.1)

X Rnint}.

Observe that FP is completely determined in terms of Fe and the deformation via


the local multiplicative factorization as FP = Fe l[Dqp]. Similarly, the Kirchhoff
stress r and the hardening stresses q are determined in terms of (Fe, I,) C I via
the constitutive equations in Table 36.1. For convenience, we adopt the following
notation for the flow associated with the primary variables in the problem
t E II

1-Xt

v, F e , ,)

= (t,

E C x V,, x

= Z x .

(41.2)

The evolution of this flow is governed by the dynamic weak form of balance of
momentum, written in the form (see Section 39):

Gdyn(Xt; 7/)

fJ pot(v 0 pt)
+ G, (r;7)

(n 0 cpt) dQ

=0 Vr C71
Vt.

(41.3)

Recall that G, (7r rj) is the static part of the dynamic weak form defined by the
(virtual work) expression (40.2), i.e.,

G~,(;r) =

2r

(V

o
.*

Q.t) -

d-(ro
(q , ) dQ2

(,IWtp)dFr.

(41.4)

It is again emphasized that


is viewed here as a dependent function of primary
variables Xt Z via the local constitutive equations in Table 36.1.
The nonlinear problem for the flow t - Xt defined by (41.3) and (41.4), along
with the local equations in Table 36.1, is dissipative in the sense that there exists a
function E: Z -* I which decreases along the flow. This function is the total energy
of the system and is defined as follows.
Let Vxt: C - IRdenote the potential energy of the external loads, i.e., a functional
satisfying the condition
-dt et('-t)
dt

B VdQif'F,

Vd,

(41.5)

Nonlinear continuum mechanics

SECTION 41

359

where V = v o t is the material velocity of the motion. For dead loading the function
V,,t(ot) is obviously given by
Vext(q(t)

= -J

t d2

. t

(41.6)

t d.

The total energy function E: Z -- iR is the sum of the kinetic energy of the system,
the total internal free energy and the potential of the external loads, i.e.,
E(Xt) =

J [Pjv o Vtl2 + W(FeTFe) +7i(c)] d

+ Vext(Wt).

(41.7)

The following result gives the rate of change of the functional E(-).
THEOREM 41.1. The time rate of change of the total energy along the flow is
d E(Xt) = -DQ

in , where Do = f2dQ

is the total instantaneousdissipation at time t

(41.8)

1[and

fint

D = r. dP +

ES

c=l

is the instantaneous local dissipation.


PROOF. Assume sufficient smoothness so that time differentiation under the integral
sign is allowed. Then observe that
atw(e)

=2CeW(Ce)

Ce

= Fe [2aeW(Ce)]FeT . Fe-T [Ce]F e - 1 .

(41.9)

Since 1e = F'eFe- and d e = sym[le], by making use of the kinematic relation

= Fe-T[2e']Fe- derived in (38.3), the constitutive equation for the Kirchhoff


de
stress
in Table 36.1 implies the following alternative but completely equivalent
expression for (41.9)
at~w(Ce) =

de = r [d-dP].

(41.10)

Using this result in the computation of the time derivative of E(Xt) and recalling that
d = sym[Vv] o Spt gives
dt E(Xt) = Gdy (Xt; v)-g

dd;

dP

d+?.

(41.11)

360

J.C. Simo

CHAPTER In

The result then follows by noting that the first term on the right-hand side of (41.11)
vanishes along the flow since the spatial velocity field v E V., is an admissible test
function in the dynamic weak form (41.3) for time-independent essential boundary
condition.
Accepting the dissipation inequality Do > 0 for all t C I as an unquestioned
fundamental principle, the preceding result implies that dE(Xt)/ dt < 0 for all t eC ,
i.e., the total energy decreases along the flow as claimed. The dissipation inequality
can often be checked in examples, as illustrated in the following situation already
considered in the context of the infinitesimal theory.
EXAMPLE 41.1. Suppose that the convex functions f,(r, q)
domain are of the form
f (, q)

= (r, qx) _ ry,, < O where rye > 0, / =

that define the elastic

,...,

(41.12)

and the functions 4,(r, q0 ) are convex and homogeneous of degree one. As already
pointed out in Chapter I, this is the case of interest in applications since all the yield
criteria used in practice satisfy these conditions. Thus, for the associative flow rule in
Table 36.1 and the assumed form of the yield criteria in (41.12), the local dissipation
is given by
nint

a=l

E=~iT-l* a

(r,

qaC) +

Eq

q)

aqi(r,

(41.13)

Using Euler's theorem for homogeneous functions, it follows that


nin

at~(T, q ) +

ah (Tr q

E q

= 7 (r, qO).

(41.14)

o=1

From this relation and (41.13) one concludes that


m

D=

q) = E

rn

yf

q ) + Eylry,

mr

ince

and

as

result
o

Thus, the model obeys the local dissipation inequality.

f the Kuhn-Tucker conditions.

CHAPTER IV

The Discrete Initial Boundary


Value Problem:
Exponential Return
Mapping Algorithms

The numerical solution of the initial boundary value problem for dynamic plasticity
at finite strains involves the transformation of an infinite-dimensional dynamical system, governed by a system of quasilinear partial differential equations into a sequence
of discrete algebraic problems by means of the following two steps:
Step 1. The infinite-dimensional phase space Z = C x V is approximated by a
finite-dimensional phase space Zh C Z via a Galerkin finite element projection. This
projection induces, in turn, a discretization of the space Z h of internal variables. The
projected (finite-dimensional) dynamics on Z h XZh is governed by a system of coupled
nonlinear ordinary differential equations arising from the weak form Gdyn: Zh X h X
1[- R of the momentum equations together with the local form of the constitutive
equations for multiplicative elastoplasticity.
Step 2. The coupled system of nonlinear ordinary differential equations describes
the time evolution in the time interval ]I of interest of the nodal degrees of freedom
and the internal variables associated with the finite element Galerkin projection onto
Zh x Z h . A time discretization of this problem involves a partition
N

1[

U [t. t+l]
n=0

of the time interval . Within a typical time subinterval [t, tn+l], a time marching
scheme for the advancement of the configuration and velocity fields in Zh together
with a return mapping algorithm for the advancement of the internal variables in 2 h
results in a nonlinear algebraic problem which is solved iteratively.
361

362

J.C. Simo

CHAPTER IV

Formally, the two steps outlined above commute in the sense that the same solution
is obtained by reversing the order of application of the preceding two steps, as indicated
in the diagram below.
ELASTOPLASTIC IBVP
Time

>

Galerkin Projection

Integration

i SEMIDISCRETE IVP
Time

Integration

Galerkin Projection

|SEMIDISCRETE

BVP

ALGEBRAIC PROBLEM

The goal in this chapter is to provide a fairly complete account of the time integration schemes required to successfully complete the second step in the preceding
numerical solution scheme. The emphasis is placed on the formulation of a new class
of exponential return mapping algorithms for the integration of the plastic flow evolution equations and, to a lesser extent, on the formulation of time marching schemes for
the advancement of the solution in phase space. Issues pertaining to the construction
of the spatial Galerkin finite element projection in the first step are addressed only to
the extent needed to illustrate the main difficulties involved and the structure of the
resulting initial value problem. The approach adopted in the design of return mapping
algorithms is motivated by the following considerations.
Conventional generalizations to the finite strain regime of the return mapping algorithms described in Chapter II are based on a direct numerical integration of the
V

incremental elastic constitutive equation r = aid - dP ] to define an algorithmic


trial state. Representative examples of this approach, widely used in commercial finite element software up to the mid-eighties, are found in KEY [1974], KEY, STONE
and KRIEG [1981], NAGTEGAAL and VELDPAUS [1984], GOUDREAU and HALLQUIST
[1982], ROLPH and BATHE [1984] and HALLQUIST [1984, 1988].
Standard time-stepping algorithms, however, do not guarantee in general that the
key condition of objectivity is inherited by the resulting, numerically integrated, incremental constitutive equation. Satisfaction of the fundamental condition of frame
invariance leads to so-called incrementally objective algorithms, a notion formalized
in HUGHES and WINGET [1980], RUBINSTEIN and ATLURI [1983], PINSKY, ORTIZ and
PISTER [1983] and others. The condition of incremental objectivity precludes the generation of spurious stresses in rigid body motions. Although it would appear that such
a requirement furnishes a rather obvious criterion, a number of algorithmic treatments
of plasticity at finite strains exist which violate this condition (e.g., MCMEEKING and
RICE [1975]).
The approach described in this chapter, initiated in SIMO and ORTIZ [1985] and
SIMO [1986], completely circumvents the need for incrementally objective algorithms
and leads to objective schemes which are exact in the absence of plastic flow. The
key idea is to exploit the multiplicative decomposition of the deformation gradient
and evaluate the trial state via a mere function evaluation of hyperelastic stress-strain
relations formulated relative to a released (moving) configuration. These ideas are
further extended in SIMO [1988a,b]. Subsequent work within the context of the mul-

The discrete initial boundary value problem

SECTION 42

363

tiplicative decomposition includes WEBER and ANAND [1990] and ETEROVICH and
BATHE [1990], where an exponential approximation to the incremental flow rule is
employed, and PERIC, OWEN and HONNOR [1989]. More recently, MORAN, ORTIZ and
SHI [1990] addressed a number of computational aspects of multiplicative plasticity
and presented explicit/implicit integration algorithms. Related approaches are considered in KIM and ODEN [1990]. Methods of convex analysis, again within the context
of the multiplicative decomposition, are discussed in EVE, REDDY and ROCKEFELLAR [1991]. The preceding survey, although by no means comprehensive, conveys
the popularity gained in recent years by computational elastoplasticity based on the
multiplicative decomposition.
42. The Galerkin projection: The spatially discrete problem
The goal of this section is to outline the formal steps involved in a spatial discretization of the initial boundary value problem for finite plasticity (Step 1), leading to
a finite-dimensional, second-order initial value problem. The exposition will be restricted to the standard Galerkin finite element method since the objective here is to
provide the general algorithmic framework for the generalization to the finite strain
regime of the return mapping algorithms described in Chapter II. Details specific to
the formulation of mixed finite element methods, better suited for the problem at hand,
are of secondary interest in motivating the main issues arising in the time integration
of the problem of evolution. An account of a rather convenient class of mixed finite
element methods designed to circumvent the difficulties experienced by low-order
conventional finite element methods will be given in the following section.
Consider a spatial discretization 2 = U"' Qf2 of the reference configuration
Q c Rfndim, generically referred to as the triangularization and denoted by Th in
what follows, into a disjoint collection of nonoverlapping subsets g,. With a slight
abuse in notation we shall refer to a typical subset S2?as a finite element and denote by
h > 0 the characteristic size of an element in a given triangularization. Precise definitions of these notions, along with the technical conditions to be met by a discretization
to qualify as an admissible triangularization, are found in CARLET [1978].
Associated with the triangularization Th one introduces a finite-dimensional approximation Ch C C to the configuration manifold C, defined as

Ch

{(h

C:

i'
h E [Co(2)] d m
and oph 2e . [Pk(Qe)] ndm},

(42.1)

1. The
where Pk(S2e) denotes the space of complete polynomials of degree k
technical condition that the approximation be at least Co is dictated by the variational structure of the problem. For second-order elliptic variational problems this is a
well-known requirement (see CIARLET [1978]) which, in the present context, is merely
formal and motivated by the presence of generalized derivatives up to order one in the
(static) weak form (40.2) of the equilibrium equations. These technical considerations

J.C. Simo

364

CHAPTER IV

will be largely ignored in the exposition that follows. The finite-dimensional subspace
oh c Vo of material test functions associated with Ch is then defined as
vh = 71oh

Vo:

[cO(Q)] dim and 70hls?e C[Pk( 2 e)]ndim}.

r0lh

(42.2)

Recall that maps ph E Ch must satisfy the Dirichlet boundary condition oh|r = h,
while material vector fields on the test function space VOh satisfy the homogeneous
form of this essential boundary condition, i.e., hlr = 0.
The current configuration associated with a deformation iph
Ch is the set
Sh = h(Q). As in the continuum problem, the tangent space Vhh is spanned by
spatial test functions rh :Sh t- Rndim defined via composition with the deformation
as
= roh o pOh-l. By the chain rule one has the key relation
-h

GRAD

[h] = (ioh

[77,]

'Pi)

*)iT

ie,
i.e.,

D(V
D'Pt,

Or0

fa
aXA

_h
=

a,0

V'7a a9b
axb XA'

(42.3)

which furnishes the finite element counterpart of (39.4). The discrete canonical phase
space Zh C Z associated with the finite element discretization introduced above is
Zh = Ch V>h'. The Galerkin projection of the problem is obtained merely by restricting the weak form of the equations to Zh X Ih. Adopting a description relative
to the current placement, the projected motion t
I -* iopth C Ch, with associated
spatial velocity field vh and stress field Th, then satisfies
Gdyn (r",

h;

h) = /

A(Vh o WIh)

(h

PO

o ioh) d2

+ Gth (T; rh) = 0

(42.4)

for any (spatial) test function rh E VCh,>. As in Section 39, Gh (h ; r7h) denotes the
(static) weak form of the equilibrium equations which in a description relative to the
current configuration is given by (40.2), namely
G, (rh;rh)=

[rh
T
(,V7 *h o
'X

( h

'Ph)

- B

MOyh)
d

)]

dQ
(42.5)

The dynamic weak form (42.4) together with local constitutive equations that relate
the stress field i-h to the deformation, such as the model of multiplicative finite strain
plasticity summarized in Table 36.1, define an initial value problem on the discrete
phase space Zh for the projected motion and the projected velocity field.
REMARK 42.1. In addition to the spatial approximations for the admissible deformations and the corresponding test functions, for constitutive models of the type given
in Table 36.1 it is also necessary to specify a suitable interpolation for the internal

The discrete initial boundary value problem

SECTION 43

365

variables. For C finite element methods the approach almost universally adopted is
based on the following considerations, made with reference to the specific model in
Table 36.1 with internal variables {F e, A,} C I:
(i) A CO-approximation to the admissible deformations via polynomials of order
k > 1 defines an approximation Diph to the deformation gradient of order (k - 1)
within each element, discontinuous across elements in the triangularization Th.
(ii) Consistent with this result, the interpolated internal variables {Feh, h} E Z h
are also assumed to be discontinuous across elements and consisting of complete polynomials of order (k - 1) within each element. The construction of this local (k - 1)
interpolation for the internal variables then involves defining a local triangularization
TTh within each element DI.
The actual implementation of this approach becomes trivial if one observes that
the integrals appearing in the static weak form are always evaluated via numerical
quadrature. By choosing the vertices of the local element triangularization 5Th coincident with the quadrature points of the element, it follows that the values of the
internal variables at the quadrature points provide all the information needed for the
evaluation of the weak form (42.5).
43. The linearized (semidiscrete) initial value problem
The statement given below of the incremental initial boundary value problem is ideally
suited both for nonlinear elasticity and nonlinear elastoplasticity at finite strains, and
provides the basis for the numerical solution strategies discussed in the following chapter. For recent accounts of the relative small number of existing mathematical results,
mostly restricted to static nonlinear elasticity, see CIARLET [1988] and VALENT [1988].
The solution of the semidiscrete initial value problem, which arises from the
Galerkin finite element discretization outlined above, employs an iterative scheme
based on the solution of a sequence of incremental problems. These linear problems
are obtained from a Galerkin discretization of the rate version of the weak form
of momentum balance (see expression (40.7) for the continuum problem) and arise
in the following context. One is given a configuration qoth E Ch with velocity field
vph e ]Eo, which is not necessarily in dynamic equilibrium for prescribed loading, i.e.,
Gdym (-r, Vt; t}) 4 0. The objective, then, is to compute an incremental displacement
field
Auh(, t)

:Sth =

th(JQ)

Indi

such that Au(-, t) E 1V,h

(43.1)

by solving the linear problem obtained by linearizing the dynamic weak form about
the configuration oh G Ch. Typically, this problem is defined for t E [t,, t+l]. The
updated configuration Wot + Au o cpt then provides an improved approximation of the
solution of the dynamic weak form for prescribed incremental loading in [t,, t,+l].
Our goal is to describe the structure of this linear (incremental) initial value problem.
We do so before addressing the design of the time-stepping algorithms to provide a
motivation for the structure of the incremental constitutive equations needed in the
solution of this problem.

366

CHAPTER IV

J.C. Simo

Current Configuration
at Time to

Nearby Configuration
at Time t+

FIG. 43.1. Illustration of the configurations involved in the linearization of the semidiscrete, algorithmic

weak form. The placement of the body St =


St+

()

t+C( ) is obtained for qot+

at time t remains fixed. The nearby placement


= ot + (Au o Wt with > 0.

Consider the situation illustrated in Fig. 43.1 where the spatial velocity field vt
and the Kirchhoff stress tensor rth are given fields on the current placement of the
body Sh =
for a prescribed configuration poth E Ch. A nearby configuration
qoh+ ( Ch associated with the incremental displacement field in (43.1) is defined by
the expression
th()

for

= %olh+ AUh
hP0+h

(43.2)

> 0.

Associated with this configuration, we have a stress field rth and a spatial velocity
vt+. The linearized problem about the prescribed configuration oth is then given by
the affine approximation
dn /,Vt

-d
~~d

G
t
aGdcyn
Gdy(t+

t+, T) =
5

(43.3)
(43.3)

G V,

where, from expression (42.4) and in view of (43.2), the second term in this affine
approximation is given by
d

dGyn(r

h
,v

;)

=j

P0 82(Au

+d-j

'Pth) (n

Gusr:~;1h
T+

) d'P
)-

(43.4)

The discrete initial boundary value problem

SECTION 43

367

An explicit expression for the second term in (43.4) is derived as follows. Considering for simplicity dead loading, the only contribution to the second term in (43.4)
arises from the weak form of the stress-divergence term in (42.5). Assuming sufficient
smoothness so that the interchange of integration at differentiation is permitted, the
problem then reduces to evaluating at C = 0 the derivative with respect to , of the
term
* (Vor/h o Sth+C). In this expression V(() denotes the gradient relative to
*-h+
xh =
IIf fth+ = Dt+
Dth- 1 denotes the relative deformation gradient
= ptX)
between the placements Sh and Sh+(, from (43.2) together with the chain rule we
arrive at the relations
f-t

=1 + CV(AUh)

V/h = V,7 ft 4

and

(43-5)

where V(-) denotes the gradient relative to x = ot(X). To recast the result of the
computation in a form suggestive of the rate constitutive equation (38.16) we define
the Lie derivative of the Kirchhoff stress relative to the vector field Auh(., t) as

L-

=h

hhh

fhT

where

i rt+ t+

t+

(43.6)

From a geometric point of view, h+C is a contravariant tensor field defined on Sth
known as the pull-back under the relative deformation gradient of the stress field rt+, ,
which is defined on Sh C. Using the chain rule relation (43.5)2 along with expression
(43.6)2 for i+th,, a straightforward manipulation yields the identity:
h

_h
ri+

(,q

<.;)=

(th+

*hth+)

ft'+'ri+C). (Vr

oS

t)

(43.7)

By observing that the configuration Sth = oth (/2) remains fixed during the linearization
process, the preceding identity along with the product rule for differentiation and
definition (43.6)1 yields

d|

r/th+.* (Vc"

o t+c
(43.8)

[V(AU Who)rt + Suhrth] . (Vh o aoh).

With this result in hand, the second term in (43.8) can be written as

=d/,[V
(AUh
t(V o
j= [v(uh~ o

)
wth)tth

AUh-th]

) d.
d.

t )

(43.9)

Now suppose that a time-stepping algorithm is used to numerically integrate the constitutive equations in Table 36.1 within a time interval [t,, t] for prescribed initial data

368

J.C. Simo

CHAPTER IV

at time t. The linearization of such an algorithm at the configuration 9oth Ch, not
necessarily in dynamic equilibrium, will be shown below to result in an algorithmic
incremental constitutive equation of the form:
CP[V(Auh) o o],

A,,r-h

[CAurh]j = cPk (Au)k,I.

i.e.,

(43.10)

The incremental constitutive equation (43.10) is the algorithmic counterpart of the exact rate constitutive equation (38.16) derived in Section 38 for the continuum model
in Table 36.1. The key difference between these two incremental constitutive equations lies in the expression taken by the continuum elastoplastic moduli ctP, arising
in (38.16)1 and defined by (38.16)2, and the algorithmic elastoplastic moduli cP appearing in (43.10), which are obtained by linearization of the integration algorithm as
described below. As in the infinitesimal theory, these two expressions coincide only in
the limit as the step-size tends to zero. The algorithmic moduli will be shown below
to inherit the symmetry properties (38.15a) and reduce to the spatial elasticity tensor
c for nonlinear elasticity.
Collecting the preceding results, the linear initial value problem defined by the
affine approximation (43.4) to the Galerkin projection (42.4) of the weak form of the
momentum balance together with the rate form (43.10) of the constitutive equations
can be stated as follows.
THEOREM 43.1. Given a configuration pth Ch, with associated stress field r-t1 and
spatial velocity field vth, the linearizedproblem for an incremental displacementfield
Auh E

is given by

Vh

(pott (u

-Gdyn (h,
=

tp), (h oth)) + B

(nh, Auh)

Vt; h),

(43.11)

where (, *) denotes the L 2 (f2)-inner product and Bv h(-,o)

for all rh E VL

is the

same bilinearform arising in the incremental variationalequation (40.7), defined as


B

h
r, r/2h
h) = Bmt (OI, r2

Here B,h(-,)

+ B

(I, , r2z)

(43.12)

is the geometric term defined for fixed configuration Oh c Ch with

given Kirchhoff stress field rth by the bilinearform:


ge (r 2

B~,h (

h)

4%1:
)

=t t

*V42 ft

)dQ

\d7 , 42 E Vh:

which is clearly symmetric (but need not necessarily be coercive). B,at(,.)


material term defined in view of (43.10) and (40.7) by the bilinearform:

(43.13)

is the

The discrete initial boundary value problem

SECTION 44

369

(n , ')

Bth

=J

(v

V2 2

N)

which is symmetric (and often coercive in

d2

VJ71, r2 G

(43.14)

Vh

Vh).

PROOF. The result follows by inserting expressions (43.4), (43.9) and (43.10) into the
affine approximation (43.3).
When the initial value problem associated with the Galerkin projection is written
in matrix form, the bilinear forms Bge(-, .) and Bmat(., .) give rise to the geometric
and material tangent stiffness matrices, respectively, whereas the first term in (43.11)
gives rise to the so-called mass matrix of the Galerkin discretization.
REMARK 43.1. Suppose that configuration

Woh is in dynamic equilibrium, so that the


dynamic weak form satisfies: Gdyn(t h , vth; h) = 0 for all test functions rh
G Vh.
1
If the bilinear form Bh (, ) is coercive in Vh, then the incremental problem is
easily shown to admit the unique solution Auh = 0. If, on the other hand, fOh is not
in dynamic equilibrium, the right-hand side of the incremental problem defines the
linear functional

Lh(rh) = Gdyn(th,vth; h)

20

Vn

(43.15)

G Vch,

since the configuration Wp, the stress field r n and the spatial velocity field vh are
prescribed functions. Problem (43.11) then reduces to a standard second-order linear
system for the unknown Auh
Vh, which again admits a unique solution if the
bilinear form Bh (, ) is coercive in V,h.
44. Matrix form of the semidiscrete initial value problem
Let nelem denote the total number of elements, nnode the total number of nodes in a
triangularization Th and nnode the number of nodes in a generic element e labeled
, node}. This local numbering system is related to the
E Indim: e = 1,
as {X
global numbering system via the following fairly standard convention:
e

...

XA = X

with A = ID(e,a), e = 1,... ,nelem,

= 1,..

node,

(44.1)

where the nelem x node array ID(., ) is defined by the geometry of the triangularization.
A rather convenient implementation of the Galerkin finite element method is achieved
by writing the local polynomial basis as {Na(C)}, where
= ((... , (dim) are
normalized coordinates with domain the unit square i[ in I i dir , and introducing the
following local coordinate change known as the isoparametricmap:
node

-* Xh = ,be()

a=l

Na(() X~

Q2.

(44.2)

370

J.C. Simo

CHAPTER IV

The local polynomial basis functions N a : D - I are referred to as the local element
shape functions and satisfy the completeness condition Na((b) = ba, where a =
((la,
I,(, dim
a) are the vertices of the bi-unit square. Piecing together the local
shape functions by enforcing the CO-continuity requirement yields the global finite
element basis functions NA : Q2 -- , A = 1, . . ., node, satisfying
NA

= Na

on Q2e for A = ID(e, a).

(44.3)

The key feature that renders this basis useful is small compact support of the functions
NA(X) so defined. The isoparametric mapping can be viewed as providing a collection of charts that results in an extremely convenient parametrization of the reference
configuration.
EXAMPLE 44.1. To provide an illustration of these standard ideas in the simplest setting, assume ndim = 2, neode = 4 and k = 1 for the local (bilinear) polynomials

Pk(f2,). The bi-unit square is [I = [-1, 1] x [-1, 1] and the element shape functions
are
Na ((l,

2) =

'I(

1 a) (2

a = 1,..

2 a),

nde

(444)

defined so that the completeness condition Na(Cb) = ba is satisfied. The global


finite element basis functions are the well-known "hat functions", see STRANG and
FIx [1973], JOHNSON [1987] or any other textbook on the finite element method.
To enforce the nonhomogeneous Dirichlet boundary condition on rh C aQ2 one
typically introduces the affine decomposition
ho"(X, t) = X +- Uh(X, t)

gr (X, t)

fo

tC Q x I,

(44.5)

where U" E Voh is the unknown displacement vector field for t C I, while the function
gh :2 x - Rndm satisfies the boundary condition
gh(X, t) =

(X, t)

for (X, t)

r,

x ,

(44.6)

and is constructed as follows. Denote by


= {I E {l,. .,nnode}: XI E Pn}
the set of indices associated with the nodes on the boundary r,, with dimension
dim[g] = nboun Construct modified global shape functions N : 2 -*- R associated
with the global node numbers I C ~ such that N (XJ) = i by suitably piecing
together local element shape functions. Then define g h as the interpolant

gh (X t) =

N' X) (X (X, t).


leg

(44.7)

371

The discrete initial boundary value problem

SECTION 44

Setting
= I
{l,...,nnode}: I 5}, with dimension dim[G] = node, the
projected dynamics defined by t E
Uth G V h, along with the material test
h
functions roh E V , are written as

Uh(X,t) =

NA(X)dA(t)

and

rh(X) =

AC

NA(X)qA,

(44.8)

ACtg

where dA (t) CG
Rndim defines the current position at time t C Il of a (material) node
XA E 2 in the triangularization T h not lying in F. The coupled system of ordinary
differential equations governing the evolution of the nodal positions, obtained by
inserting the interpolations into the weak form (42.4) of the momentum equations,
can be written in standard (conservation) form as follows. Define the nnode vector
fields t ~ pA (t) of generalized momenta as

pA(t) =

MAB dB(t)

where MAR =

poNANB d2.

B=1

The nnode x nnode coefficients MA B define the ndof x ndof positive definite matrix M =
[MABI,,dJ] known as the mass matrix of the triangularization. Here nnode = nnodenboun, where ndof = node x ndim is the total number of degrees of freedom involved

in the discretization (accounting for the prescribed Dirichlet boundary conditions) and
Indim is the ndim X ndim identity matrix. Next, define the vector fields t

of

prescribed nodal forces as

Fe

=J NABdf

NAT

(44.9)

d.

Finally, define the internal nodal force vector Fint (dB, rth) by the expression
Fint(dB,r h )

r[ [VNA] df2

where VNA = DOh-TGRAD[NA].

(44.10)

With the preceding definitions in hand, it is easily verified that the projected weak
form of the momentum equations (42.4) yields the following problem of evolution:
dpA(t) = F
d
dtdA(t)=

-Fi(d(t),th)

nThode

for A E g.

(44.11)

MApB(t)
B=1

Equations (44.11) define a dissipative dynamical system on the finite-dimensional


phase space P = R ndof x ITndof for prescribed initial data, when the stress tensor

th

J.C. Simo

372

CHAPTER IV

is (implicitly) specified in terms of the dynamics via the local constitutive equations
summarized in Table 36.1. In a subsequent section where global time stepping algorithms are described, it will be shown that this finite-dimensional dynamical system
also inherits the global conservation laws (39.13). Therefore, the Galerkin projection preserves the global conservation laws of momentum as well as the dissipative
property of the infinite-dimensional dynamical system.
The matrix form of the linearized problem about a configuration 0th C h follows
immediately for the results presented in Section 43 and the matrix expressions given
above. The incremental displacement field Auh(., t) G Vh is written in the global
finite element basis as [see (44.8)]
Au th (

N A (X)AdA (t).

(X), t) = E

(44.12)

Acg

Using the linearity of the Galerkin projection, it is easily seen that the matrix form
of the linearized variational problem (43.11) coincides with the linearization of the
initial value problem (44.11) about the finite-dimensional configuration defined by
{dA(t),pA(t)} E Zh . The matrix form of (43.11) then yields the linear, finitedimensional, problem of evolution:

dtAP (t) +
dt AdA (t)

hE

(t)AdB(t)= RA(

d A ( t)

A (t )

, 'rt)

B(t)

Beg

(44.13)
ndi
where KAB (t)
RE
Indim x I
m (A, B E 5) is the nodal tangent stiffness matrix, the
ndim
matrix counterpart of the bilinear form Bh(.,-), and RA e
is the residual

vector associated with node A E 5 and defined as


RA (dA (t), pA (t),

) = Fext

Fi(dB(t), t)

_ dpA(t)/ dt

(44.14)

Obviously, RA (dA(t), pA (t), rth) = 0 if the nonlinear system (44.11) is in equilib-

rium. The decomposition (43.12) of the bilinear form Bh (, .) into a geometric part
and a material part gives, in turn, the decomposition
KAB(t) = KAeo (t) + Kma
t (t)

(44.15a)

for the tangent stiffness matrix where, in view of (43.13), the component expression
of the geometric part is given by

[K(]ABEoi=
J
ij

[geo

-JTN
NNAkij
k=l 1=1

d28ij,

(44.15b)

The discrete initial boundary value problem

SECTION 45

373

and component expression of the material part is defined in view of (43.14) as

[Kat(t)]

ij
I

j,1

NAepi E
N
i

li

d2.
I

(44.15c)

Here Nk = aNA/Oxk denotes the partial derivatives of the shape functions relative
xk = pk (XA, t).
Before closing this section it is useful to summarize the key implication of the
preceding developments in the design of return mapping algorithms. The numerical
solution of the initial value problem (44.11) for loading prescribed in terms of Fext(t)
and specified initial data exploits the following strategy. The stress tensor rh together
with the collection {Feh, S} G h of internal variables are viewed as dependent
variables which are solved for in terms of the dynamics t E [t., t.+l] -* dA(t) via
the class of return mapping algorithms described below. Since the integral that defines
the internal nodal force vector FinAt in (44.10) is evaluated by numerical quadrature,
knowledge of the stress field r is required only at the discrete points used in the
quadrature formula. Moreover, because of the interpolation scheme adopted for the
internal variables described in Remark 42.1, effectively, knowledge of the internal
variables is also required only at the quadrature points.
The specific form taken by the deformation-driven return mapping algorithm is
the subject of the following sections. The last section in this chapter addresses the
structure of the time-stepping algorithm for the advancement within the time interval
[t., t+l] of primary variables (dA, PA) in the phase space P.

to the current coordinates

45. Mixed finite element discretization: An illustration


The key difficulty involved in the construction of the Galerkin projection (Step 1) for
classical plasticity is related to the enforcement of the volume-preserving condition
on the plastic flow. This constraint arises whenever the yield condition is pressure
insensitive, a situation nearly always encountered in metal plasticity and exemplified
by classical yield criteria such as the Maxwell-Huber-von Mises condition. This problem is currently well understood and, since first pointed out in the pioneering work of
NAGTEGAAL, PARK and RICE [1974], has been extensively addressed in the literature.
In particular, use of low-order Galerkin finite element methods is well known to result
in an over-constrained pressure field that may render overly stiff (locking) numerical solutions. In the linear regime, techniques designed to circumvent this difficulty
include classical LBB-satisfying displacement/pressure mixed finite element methods, see the recent review in BREZZI and FORTIN [1991], higher-order Galerkin finite
element methods, such as the high-order triangle of SCOTT and VOGELIUS [1985],
and special three-field mixed finite element methods as in SIMO and RFAI [1990],
among others. Extensions of these methodologies to the finite strain regime have been
considered by a number of authors. In particular, generalizations of the approach in
NAGTEGAAL, PARK and RICE are presented in HUGHES [1980] and extensions of the
displacement/pressure mixed methods are considered in GLOWINSKI and LE TALLEC

374

J.C. Simo

CHAPTER iV

[1989], SIMO, TAYLOR and PISTER [1985] and SUSSMAN and BATHE [1987] among
others. Enhanced mixed finite element methods which include nonlinear versions of
the method of incompatible modes (see TAYLOR, BERESFORD and WILSON [1976] and

CIARLET [1978]) are presented in SIMO and ARMERO [1992] and have been recently
extended in SIMO, ARMERO and TAYLOR [1993].

The goal of this section is to describe a generalization to the finite deformation


regime of the assumed strain finite element methods described in Section 26. These
class of mixed finite element methods is specifically designed to circumvent the difficulties experienced by the conventional Galerkin finite element method in the nearly
incompressible regime. From a computational standpoint, this methodology offers a
key advantage over conventional displacement/pressure mixed methods. Specifically,
for boundary value problems arising in classical nonlinear elasticity and plasticity, the
convenient strain-driven format of the Galerkin (displacement) finite element method
is preserved. In particular, for classical plasticity, the structure of the local strain-driven
return mapping algorithm described in detail in the preceding sections is preserved by
assumed strain methods.
In order to present the technique in the simplest possible context, attention will be
restricted to quasistatic nonlinear elasticity following the approach described in StMo,
TAYLOR and PISTER [1985]. The generalization of the methodology to the elastoplastic
problem is entirely analogous to the situation addressed in Section 26 within the
context of the infinitesimal theory. As shown in the aforementioned reference, the only
modification needed to tackle the elastoplastic problem merely involves performing
the local return mapping algorithm with the strain field replaced by the assumed strain
field.
(A) Three-field variationalformulation of the problem. The severe locking arising
from the point-wise enforcement of a nearly isochoric deformation via a low-order
Galerkin finite element method is circumvented by introducing an independent volume
field, denoted by J, together with its dual variable r that plays the role of a Lagrange
multiplier and is interpreted as the Kirchhoff pressure. The constraint expressing the
equality between J and the point-wise value field J(qo) = det[Dqo] is then enforced
weakly via a Lagrange multiplier term. The variational setting for the problem is
completed by defining the assumed deformation gradient as

F(o, J)

J/d... F(o)

where F'(q) = [J(o)]

D~p

(45.1)

is the volume preserving part of the deformation gradient which satisfies the condition det[F(W)]
1. Observe that the assumed volume field J is subjected to the
constraint J > 0. This constraint can be automatically enforced by replacing J with
the unconstrained variable 0 = log(J), so that

J=exp[)]

and

C(q,v =)

[F(=p,0)] F(45,.),

(45.2)

375

The discrete initial boundary value problem

SECTION 45

and considering the following three-field Lagrangian functional as a point of departure


for the construction of the assumed strain mixed finite element method:

((C,
W(

L(o, , )

,)) +

r [log

(J(g)) -

] } d2 - Vext().

(45.3)

Here Vext (g) denotes the potential energy of the external forces which, for simplicity,
is assumed to correspond to dead loading. Our first objective is to compute the formal
Euler-Lagrange associated with the saddle-point problem defined by (45.3). To do so,
use is made of the relations in the result below.
THEOREM 45.1. The directionalderivatives of F(o, i9) and C(, '0) are given by the
following expressions:
(i) D F(o, 9)- r = (dev[V[r]] o o)F(o, 9),

(45.4a)

(ii) DoiF(o,0)- 0 = (1/ndim), /F(p, '),

(45.4b)

(iii) DoC((, 0)

(iv) DtC(go, 0) *.

= 2[F(,

0)]T(dev[E[r]] o o)F(, 0),


T

= (l/ndim)2[F(Wo, 0)] [/)1]F(o,29),

(45.5c)
(45.5d)

for any test spatialfunction r/ V, and any 0 E L 2 (Q), where V[r] is the spatial
gradient defined as V[rl] o = GRAD[,7 o o](Do)- ' and e[r/] = [V[r] + (V[q])T].
PROOF. (i) The directional derivative is defined by the formula
DF(,

) r

F(

Cro V,0),

(45.6)

where, in view of the definition of F(Wo, 0),

F( + (o O , )
=exp [I

0]J( +

7o o)(Do+ GRAD[r7o ]).

(45.7)

From a well-known formula for the derivative of the determinant it follows that

J(go + /o
r

g) = det [Do](D) - T GRAD[rl

= J(go) div[r/] o go.

(45.8)

Using this result together with the product rule for differentiation and the chain rule
relation GRAD[77o go] = (V[r] o p) D, expression (45.6) becomes
DqF(O,

==0)
[J()/J]-/ndim

(V[] - l/ndim div[/]) o

(45.9)
4[D5],

376

J.C. Simo

CHAPTER IV

which in view of the definition of F(qo, '0) yields result (i). Expressions (ii), (iii) and
(iv) are proved in a similar fashion. [
Making use of the preceding results, the Euler-Lagrange equations associated with
(45.3) take the following form strongly reminiscent of the results found in Section 26
for the linear theory. Upon defining the Kirchhoff stress and pressure fields
r(o, 9, r) = rl + dev[F(o, 0)2acW(C(,
p(Wo, V) =-tr[

9))F (,p

0)],

F(o,0)2acW(C(po,0)) F (o, 0)],

(45.10a)
(45.10b)

ndim

a straightforward manipulation yields the following three variational equations that


provide the stationarity conditions for (45.3):

T(Wp,,
.

, 7r)

(V[ri o o) dQ -

(
|f

log(J(

Gext()

(45.1 la)

)) - 0) dQ = 0,

(45.1 b)

(p( )t) --o7(45. ,

11c)

which are to hold for any spatial test function


GCV, and any q, b E L 2 ( Q2). Here
Gext(,q) denotes the virtual work of the external (dead) loads given by the standard
expression

Gext()

B. (o o )dQ+

) d.

(45.12)

The weak forms (45.11a,b,c) furnish the point of departure for the construction of a
mixed finite element approximation to the problem at hand.
(B) Discontinuous mixed finite element discretization. Consider a finite element triangulation Th of the reference placement 2 l2h = Ue=1 Qe, a conforming finite
element approximation Ch C C to the configuration space and the corresponding
finite-dimensional approximating subspace Vh' C Vv of test functions, as described
in Section 42. To recover the structure of the assumed strain method presented in
Section 26 within the scope of the infinitesimal theory, we introduce discontinuous
approximations both for pressure and the volume field via the finite element subspace
ph

A{sh

L 2(?):

In2

FT(X) ,b for ibe

CI

}.

(45.13)

As in Section 26, r(X) = [ (X)


...
r,(X)]T is a vector of prescribed element
local element shape functions the specific form of which is left open for the moment.
The lack of interelement continuity requirements on functions in the finite element

The discrete initial boundary value problem

SECTION 45

377

space ph is the key design condition placed on this approximation. It is precisely this
condition that enables us to obtain an explicit expression for the pressure at the volume
field at the element level in terms of the deformations eph E Ch, thus recovering the
structure of an assumed strain method. We do so by substituting the approximations
in (45.13) into the variational equations (45.1 lb,c) and solving at the element level to
obtain:
7e

(qe) = FT
=

e(

(X)H

re)

rF(X)

log(J (

(X)p(pe ,

)) df,
dQ,

(p))
e

(45.14a)
(45.14b)

where ()h = (.)hlne denotes the restriction of the finite element approximation ()h
to a typical element / 2e and H is the m x m element matrix defined as
He=

F(X)
)

F(X) dQ.

(45.15)

The collection of element functions {09(po)}=L and {r7r(qOh)}c=l defines a discontinuous global approximation to the (logarithmic) volume field and pressure field in
terms of the deformation Woh E Ch, which we shall denote in what follows by 19 h(Wph)
and 7rh(Wph), respectively. By substituting these approximations into the remaining
variational equation (45.1 la) we obtain the following variational problem:
Problem Gh: Find the deformation qoh C Ch such that

(W
(Wh),( th Ph),( Th
-

Wh))

(V [h]

qoh ) d2

Gext(77h) = 0,

for all finite element (spatial) test functions r/h C


stress field defined by expressions (45.10a,b).

(45.16)
Vhh,

with the "equivalent" Kirchhoff

The preceding variational problem merely amounts to a generalized displacement


Galerkin finite element method with "modified" constitutive equations defined in terms
of the deformation field via expressions (45.10a,b).
REMARK 45.1. Given a finite element conforming subspace, the possible choices for
the functions r(X) that define the finite element pressure/volume subspace ph are
severely restricted by the well-known inf-sup condition. While a rather complete
theory exists for linear problems, see, e.g., the account in BREZZI and FORTIN [1991],
very limited results currently exist for quasilinear problems exemplified by quasistatic
nonlinear elasticity. The conventional approach in the design of the pair of finite
element spaces Vph /ph is to adopt pairs that satisfy the inf-sup condition in the linear
theory. A well-known example is furnished by the choice Q2/P1 which renders an

378

J.C. Simo

CHAPTER IV

optimal element in the linear regime. We refer to the aforementioned reference and
the brief discussion given in Section 26 for additional details.
From an algorithmic standpoint, the solution of the "modified" variational problem G h is accomplished via an iterative technique, say Newton's method, in which
the deformation p h plays the role of the primary "driving" variable. For a given
numerical quadrature formula is given, the specific structure of which being irrelevant for present purposes, the evaluation of the assumed stress field h =
S(qth(cht),
h(,), 7rh(ph)) in (45.16) over a typical element Q, involves the following steps to be performed at each quadrature point:
Step (i). Let Nf: Q2 - 1R denote the local element shape functions associated
with the conforming approximation to the deformation field. Compute the element
deformation field via the standard interpolation formula
node

(X) = ZNa(X)r'

where r

= X'

+ d'

(45.17)

a=1

and de , a =

1,. ., nen,

denote the current displacement vectors of the node

nodal points of element 2?,, which are assumed to be prescribed in a conventional


"displacement-driven" iterative solution format.
Step (ii). Evaluate the conforming approximation to the displacement gradient and
the volume field via the standard expressions
Tnode

Dp

a
e
]
r0GRAD[_N

and

J(oh) =det [DWh].

(45.18)

al
a=1

The preceding two steps are exactly the same as in the conventional Galerkin finite
element method.
Step (iii). Evaluate the element volume field Jh = exp[nh(qehn)] by making use of
expression (45.14a), namely
e = exp

[FTrx)

F(X) log(J(e)) dQ].

(45.19)

Observe that this interpolation ensures that condition Jn > 0 holds for any X
Ce. Then evaluate the assume displacement gradient Fe and the corresponding right
Cauchy-Green tensor Ch by setting
-h

(j(

a=

h)/h)-l/hnd

/e;

and

= -h

=_

_e

(45.20)Th
(45.20)

This step represents an additional computation not present in the conventional Galerkin
finite element method.

The discrete initial boundary value problem

SECTION 45

379

Step (iv). Perform a constitutive evaluation (for each quadrature point of the element) and compute stress field Fph[2acW(Ch)]FbT. This step is exactly the same as
in the Galerkin finite element method, the only difference being that the conforming
right deformation gradient Fh = Dh is replaced by the "assumed" deformation
gradient F e defined in Step (iii). The same observation holds for models other than
elasticity. In particular, for plasticity all that is needed is to "drive" the local return
mapping algorithms with the assumed deformation gradient Fh.
Step (v). Evaluate the assumed pressure field 7rh over the element via expressions
(45.14b) and (45.10b), i.e.,
= r T (X)H

F(X)
efo

1 tr[F

2OcW(

h)FhT]

df?.

(45.21)

ndim

Then compute the assumed stress field rh over the element from expression (45.10a)
as
Teh

= rh 1 + dev[Fh2acW(Ce)FhT].

(45.22)

The pressure evaluation via (45.21) is an additional computation over the Galerkin
finite element method while the stress computation in (45.22) is formally identical.
Once the element stress field reh is computed at each quadrature point of an element S29according to the preceding steps, the evaluation of the weak form (45.16)
proceeds exactly in the same fashion as in the conventional (displacement) Galerkin
finite element method, see Section 26. In particular, the matrix formalism described in
Section 44 carries over to the present context without essential modifications. Hence,
further details will be omitted. Observe that the generalization of (45.16) to the dynamic problem is straightforward and merely involves appending the weak form of
the inertial term poo h.
From a computational standpoint, an important property of the Galerkin finite element method is the sparse structure of the Wl~-matrix that arises in the conventional
matrix expression
,node

E[r7eh = sym{V[17Jh]]
= ):'eaa,

aa
a

where a' e Rndinm.

(45.23)

a=l

for the symmetric gradient of the spatial test functions nh C Vhh. A noteworthy
feature of the mixed finite element method outline above, which becomes apparent by
inspection of the reduced weak form (45.16), is the preservation of this sparse structure.
This property holds independently of any possible coupling between volumetric and
deviatoric response induced by the specific form of the constitutive model.
REMARK 45.2. The preceding mixed finite element method can be cast into an alternative format suggested in HUGHES [1980] as an extension for linear problems of
the methodology introduced in NAGTEGAAL, PARKS and RICE [1974], without making

380

J.C. Simo

CHAPTER IV

any reference to the underlying variational structure. As noted in SIMO, TAYLOR and
PISTER [1985], such a reformulation relies on the following identity for the term in
(45.16) involving the element pressure field 7rh:

7h (div [

) d=

PJ(div [te]
,
P) dh ,

where
=
9())
CP)
eSic is defined by (45.10b), with
and div[r/h] is defined by the expression

dv[]

W=

(45.24)

rF(X)(div[r]

FT(X)H- J

( e)

given by (45.14a),

) dQ.

(45.25)

Identity (45.24) is an immediate consequence of expression (45.14b) for the element


pressure field 7rh and is verified by following the same steps used in the proof of the
theorem in Section 26. Expression (45.25) arises naturally within the present variational framework as the linearization of the element volume field Je = exp[0h( o)]
defined by formula (45.19) in Step (iii), since
DJho r7eh = Je (djiv[reh] ph).

(45.26)

This relation is easily verified and furnishes the counterpart for the assumed volume
element field of the standard result DJ(Wh) *r7h = (Wh) (div[r7h] 0o Wh). Substitution
of (45.24) into (45.16) then gives the following equivalent weak formulation:

J [Fh23cW(lh)F

hT]

7h

/h] 0a

h)

d2 - Gext( 7 h) = 0,

(45.27)

where Fh is the assumed deformation gradient defined by formula (45.20) in Step


(iii) and Vh [.] is a modified spatial gradient defined on each element 2e by

Vh[re] = dev[V[r1*]] +

1
di

iv[r/]1.

(45.28)

The implementation of the weak form (45.27) involves the same steps described above,
with the exception that the element pressure field 7rh defined by formula (45.21) in
Step (v) need not be explicitly evaluated. The drawback of this implementation over
the one outlined above lies in the structure of the matrix expression for the modified
gradient Vh[-], as defined by (45.28), which no longer retains the sparse structure
present in expression (45.23) for V[-].

SECTION 45

381

The discrete initial boundary value problem

(C) Linearization: The incremental problem. Under the assumption of dead loads,
the linearization of the weak form (45.16) yields the following bilinear form, the
structure of which was derived in Section 43:
B (rh,Auh)
=

[h] . (v[Auh]fhl)
Geometric term

e+hE
[ l]

. (aulh

h)

d2,

(45.29)

Material term

where explicit indication of the composition with the deformation is omitted to simplify the notation. We shall follow this convention throughout this section. Observe
that the stress field appearing in the geometric term is the assumed stress field defined by (45.10a,b). What remains to be done is to compute the material term aUht h
by linearizing expressions (45.10a,b). We shall do so below for the particular case
in which the volumetric and deviatoric response are uncoupled via a stored energy
function of the form
W(C) = U(J) + W(C)

with C = J-'l/ndC.

(45.30)

This situation is of special interest in metal plasticity (see Section 37) and also arises in
the numerical treatment of incompressible elastic materials via a penalty regularization
that employs elastic models of the type describe in Section 32, see, e.g., SIMO and
TAYLOR [1991]. In contrast with the general case, in this setting the deviatoric part
dev[(eh] of the element stress field is independent of the volume element Jh while
the pressure field Trh depends only on Jh. This simplification results in the following
remarkably simple result.
THEOREM 45.2. For the uncoupled stored energy function (45.30), the material term
arising in the linearization (45.29) of the reduced residual (45.16) at a configuration
Wh e Ch is given by the bilinearform
Bhat(h,

[ h]. [* 27rhI+ 6] [Au

uh)
+

] dQ

[JhU'(Jh)]'d v[h]div[Au

] dQ,

(45.31)

where Ch are the elastic moduli associated with the volume-preserving part W(Ch)
of the stored energy function, as defined by the closed-form expression (32.20), and
div[-] is the modified divergence operatordefined by (45.26).
PROOF. In view of the result obtained in Section 43, expression (45.10a) for the element
stress field th implies
LtuhTh = jAUh [

1] + CE[AUh].

(45.32)

382

J.C Simo

CHAPTER IV

A literal application of the definition for the Lie derivative together with the rule for
product differentiation then yields
Auh

= -27rE [Auh] + CE[AUh] + 1 Drh .Au.

(4533)

The first two terms in this result give rise to the first integral in expression (45.31).
All that remains to show is that the last term in (45.33) involving the linearization
of the element pressure field rh produces the second integral in (45.31). In view of
result (45.26), for the uncoupled stored energy function (45.30) the linearization of
expression (45.10b) reduces to
D7rh . Au = FT(X)H

(X) [JhU(Jh)] d- v[Auh ] dQ.

(45.34)

The final step in the proof makes use of the following identity similar to that leading
to (45.24):
|

(D7n *Au'I)

V []

dS
(45.35)

= [JhU'(Jh)]' div['h] div[Auh] dQ.


The result then follows from (45.35) and (45.33).

El

To complete the implementation of the method outlined in Step (i) to Step (v)
it remains to specify the iterative solution strategy for the solution of the nonlinear
problem (45.16). Considering Newton's method as a representative algorithm, the
results (45.29) and (45.31) show that the evaluation of the tangent stiffness matrix
associated with the bilinear form B(., ) requires the additional computation of the
modified divergence operator div[.] via expression (45.25). We remark that expression
(45.31) retains the sparse structure of the Galerkin finite element method. From a
computational standpoint, the additional term involving the modified divergence div[.]
translates into a rank-one update.
The key conclusion to be drawn from the foregoing developments is that the formulation of integration algorithms for plasticity, the objective of the following sections
in this chapter, can be addressed within the context of the conventional Galerkin finite
element method. The extension of methodology to be developed below to the mixed
finite element methods described in this section merely involves the replacement of the
standard deformation gradient Fh = Dqph with the assumed deformation gradient Fh.
REMARK 45.3. The preceding class of methods is design to circumvent the difficulties
experienced by low-order Galerkin finite element methods but is not suitable for
plasticity problems involving highly localized plastic deformations. This situation is of
interest in a number of applications and always arises in materials exhibiting softening
response, such as metals subject to very high strain rates where thermoplastic softening

SECTION 46

The discrete initial boundary value problem

383

results in adiabatic shear banding. A class of recently proposed enhanced-strain mixed


methods appear to be ideally suited for this type of problems while retaining the good
performance of mixed methods in the nearly incompressible problem. We refer to SIMO
and RIFAI [1990] and SIMO, ARMERO and TAYLOR [1993] for further background on
this subject and additional information.
46. Exponential return mapping algorithms for multiplicative plasticity
Consistent with the solution strategy outlined above, the goal of this section is to
provide a local constitutive equation for the stress field within a typical time interval [tn, t+l] in terms of the motion pt :
- ]Rd'm, which is regarded as a
prescribed field in [t, tq+l]. This algorithmic constitutive equation is derived by integration of the local constitutive equations in Table 36.1 and has the remarkable
properties of being exact for incrementally elastic processes, independent of the specific form adopted for the stored energy function and exact plastic volume preserving
for pressure insensitive plasticity models. Furthermore, it will be shown that the implementation of the resulting scheme, referred to as a local exponential return mapping algorithm in what follows, takes a form essentially identical to the standard
return maps of the infinitesimal theory for the important case of isotropic elastic
response.
In an attempt to simplify the notation involved in the exposition of the algorithm,
the following conventions are employed:
(i) The superscript h that refers to the finite element discretization is dropped
throughout. All the fields are assumed to emanate from a Galerkin projection and
interpolated as discussed in the preceding section. Accordingly, qo stands for oh E Ch,
F stands for Fh = Dph and so on.
(ii) The internal variables in h, written as Fe , ~} according to the preceding
convention, are assumed to be evaluated at a specific point X EG identified with an
arbitrary quadrature point in the triangularization Th. Thus, {F e ,
E GL(ndim)
GL}
RiRn t since the target space for fields in I is GL(ndim) x R]nRn.
(iii) A generic variable () is always assumed to be evaluated at (X,r) E
Q2 x [ta, t,+l]. The subscript r attached to that variable, as in (-),, is always understood to denote the algorithmic approximation at an arbitrary time t E [t., t,+l].
The expressions (.),, ()n+o and (),+1 denote the algorithmic approximations to the
variable () evaluated at X E 2 at times t, t,+, = 9tn + (1 - t)t,+l and t+l,
respectively, for specified E (0, 1].
Unless explicitly indicated otherwise, these notational conventions are implicitly
assumed throughout this and the subsequent sections.
The local problem of evolution to be integrated numerically is formulated as follows. Suppose that the primary variables X, = (n~, Vn, F,
nn)E Zh x 1 h that
approximate the exact solution are given at time t so that, for each X E Q2, one has
the initial data:
F n = Dpo(X) E GL(ndim)

and

(F,5,,)

E GL(ndim)

(4int.
(46.1)

384

J.C. Simo

CHAPTER IV

The problem to be addressed below is the construction of an algorithmic approximation


to the plastic flow evolution equations in Table 36.1 for the prescribed deformation
gradient

F = Dp,(X) at X CG for r E [t1 , tn+l].

(46.2)

Leaving
[t,, t,+l] as an arbitrary point within the given time interval allows an
easy generalization of the two-stage projected Runge-Kutta methods introduced in
Chapter II by specialization of the results described below. The differential algebraic
system equations to be integrated numerically is defined by the associative flow rule
in Table 36.1 within the interval [t. tn+l], namely,
= Doy?
lP:=

Tf.('r,q'

and

( = ~ aY Oq-f,(r,q'),
11=1

ii=1

(46.3a)
with 'y"> 0,

fm(Tr, q)

and

ylyf(r,q )=

0.

z=1

As it stands, however, this problem is not in standard form. The time derivative on the
left-hand side of (46.3a)i is hidden in the definition of the spatial plastic rate IP, which
is related to the time derivative PP of the plastic part of the deformation gradient by
the crucial relation (34.6) derived in Chapter III, namely,
P = Fe[LP]Fe-

with FP = LPF P .

(46.3b)

By appending to the system defined by (46.3a,b) the elastic constitutive equations for
the Kirchhoff stress r and the stress-like variables q (see Table 36.1), one arrives at a
problem in which the strain-like variables {Fe,( } become the independent variables
in the problem, while the deformation gradient (46.2) plays the role of a prescribed
forcing function. The feature to keep in mind in the design of any suitable algorithmic
approximation to (46.3a,b) is the restriction emanating from the requirement that Fe
and FP are to remain in GL(ndim) for any X E 2. Such a restriction translates into
the following constraints:
Je = det[Fe] >0 and

JP

det[FP] >0

in 2 x[t,,

t+l].

(46.4)

These two constraints are of fundamental physical significance and must, therefore,
be preserved by the algorithmic flow rule. In particular, plastic flow is often incompressible for metal plasticity, a feature that translates into the constraint JP = 1 in
Q2 x . The algorithmic procedure derived by means of the two steps described below
automatically guarantees the satisfaction of these conditions.
Step 1. Consider the following exponential approximation of the evolution of the
plastic deformation gradient defined by (46.3b)2 :
F P = exp[(-

t)LP] F p

for T E [t, t(+l] -

(46.5)

The discrete initial boundary value problems

SECTION 46

385

By well-known properties of the exponential map this approximation implies


JP = det[exp[( - tn)LP]]
for any

det[FP] = exp[(r- t)tr[L P ] J

> 0,

(46.6)

E [t,, t+l], provided that the initial data satisfies J = det[F P ] > 0.

Step 2. Use the multiplicative factorization F = FeFP of the deformation gradient


along with standard properties of the exponential map to conclude from (46.6) that
1
F, = Ffe exp [( - t-)L P] F e-FFP
= exp[(T -

t,)F L PFe-']FrePn

(46.7)

Now use definition (46.3b) 1 for the plastic rate of deformation tensor IP in the spatial
description to rewrite (46.7) as
F = exp[-(r - t)lP]Fef'

(46.8)

where Fe = FF-.

The justification for the notation Fe' in (46.8)2 should be clear. According to this
definition, Fetr is the elastic deformation gradient computed via the multiplicative
factorization under the assumption that FP = FnP; i.e., by freezing plastic flow.
Step 3. Finally, by inserting the evolution equation (46.3a)1 into (46.8), subsuming
the time step factor ( - t") into the definition of the algorithmic plastic multipliers
A-y > 0 and using a conventional backward Euler approximation for (46.3a)2 one
arrives at the following result:

Fr
exp_
=f

ttyM
tr,

Fre

ll=

m
a

r = ( n+

A- y aq f (r,

(46.9a)

q),
/=1

Ayl >0,

flu (r, qr ) < 0

and

Ay

f,

(r,q)=O.

'U=J

In this algorithmic version of the flow rule, the generalized stresses (-r, qT) are defined
via the elastic stress-strain relations evaluated at (F, ( T), namely,
r- = Fe [2ae(FeTFe)]FeT and q7 = -aH(T).

(46.9b)

A geometric illustration of the algorithm defined by (46.9a,b) is given in Fig. 46.1.


The derivation of the algorithmic flow rule (46.9) is a generalization of a scheme
proposed in SIMO [1992] for the case of isotropic elasticity, improving on earlier
schemes introduced in SIMO [1988b] and SIMO and MIEHE [1992] which is based

386

J.C. Simo

CHAPTER IV

Fixed Reference Configuration

""\

fl
ben+1

Fixed Intermediate Configuration

Updated Current Configuration

FIG. 46.1. Interpretation of the update procedure defined by the trial elastic state. The (local) reference
and intermediate configurations remain fixed, while the current configuration is updated by the prescribed
relative deformation gradient.

on conventional backward Euler approximations. A noteworthy feature of the present


approach is provided by the following result.
THEOREM 46.1. For models of plasticity possessing a pressure insensitive yield criterion, defined by the condition tr[0,-fp(r, q`)] = 0, the algorithm (46.9) preserves
plastic volume in the sense that JP = 1.

PROOF. Set Jr = det[FT]. Taking the determinant on both sides of (46.9a)1 and using
a well-known property of the exponential map gives
det F] = exp [

Ay4 tr(a,-fT
(r,,

q-))

det[Fetr]

(46.10)

The condition that the yield criterion is pressure insensitive then implies det[Fe] =
det[Fetr]. Moreover, from the definition of Fetr in (46.8)2 it follows that det[F
det[Fr] det[F P - l] = J/J p. Consequently,
JP = det[Fee-lFr] = J/det[F

e tr

] = JT(JP/JT) = JP.

e tr ]

(46.11)

By induction it follows that JP = 1 since J = JP = Je = 1 at the reference state.

REMARK 46.1. In the present context, the generalization of the classical backward
Euler return mapping algorithms of the infinitesimal theory is obtained by setting
T = t+l
so that t,+l - t = At is the full step size. This scheme is first-order
accurate. The generalization within the present approach of the second-order accurate return mapping algorithms based on the use of either backward differentiation
formulae or projected Runge-Kutta schemes is described below.

The discrete initial boundary value problem

SECTION 47

387

47. Exponential return mappings for isotropic elastic response


The actual implementation of the general algorithm described above involves the solution of the highly nonlinear system of equations (46.9a,b) by an iterative solution
scheme, typically Newton's method. Remarkably, for the important case of isotropic
elastic response a closed-form solution is possible, as described below. This situation
covers most of the plasticity models currently used in large-scale metal forming applications including the classical J2 -flow theory at finite strains. The results that render
the implementation of the scheme closed-form are summarized below.
THEOREM 47.1. For isotropic elastic response the following properties hold:

(I1) The stored energyfunction takes thefunctionalform


=

w(ce)

W(b

e )

(,...,e

(47.1)

where be = FeFeT, eeA = log(AA) are the elastic logarithmic stretches, and (Ae) 2
are the eigenvalues of be (A = 1,..., ndim).
(12) The Kirchhoff stress tensor r and the elastic left Cauchy-Green tensor be are
coaxial, i.e., their principaldirections coincide. In addition, we have
ndim

r=2be[abeW(be)] orequivalently r =E

rAn(A) n(A)

(47.2)

A=l

where TA = aw/laeA and be = Edml( eA) 2 n(A) n ( A) (spectral decomposition).


(13) Denoting by e = log(be) the elastic logarithmic strain tensor, one has the
constitutive equations
r = a,e(E

e)

where w (Ee) = w(Er,..., Edi)

(47.3)

and Ee is defined via the spectral decomposition Ee = EAdiml

Ee

(A)

n(A).

PROOF. (11) Elastic isotropy is equivalent to the assertion: W( e) = W(QieQT)


for any Q G S0(3). Setting Q = R e , where F e = ReUe is the polar decomposition

of F e , and noting that be = ReCeReT gives the first part of (I1). The second part
of (I1) follows from the representation theorem for isotropic functions [see (31.23)],
since the principal invariants are a function of the principal stretches and hence of
their logarithms.
(12) That r and be are coaxial follows immediately from the representation theorem
for isotropic tensor functions [simply replace b by be and II, I2, 13 } by {If, I2, I3e}
in expression (31.27)]. For a direct proof of (47.2)1 observe that
be

a, (FeFeT) =

lebe + beleT

where

l e = FeFe-1

(47.4)

Time differentiation of the stored energy function and use of this relation gives
W(be) =

[2abeW(be)be]

le =

[2dabW(be)

e]

*d ,

(47.5)

388

J.C. Simo

CHAPTER IV

since abeW(be) commutes with be by isotropy. The result then follows by noting that
W(be) = r . de [see (41.10)]. Expression (47.2)2 is a rewrite of result (33.7) since,
by the chain rule, A'aiv/8aA = aw/a log(A).
(13) Result (47.3) is merely a restatement of expression (47.2)2.
Since the elastic response is completely determined from the elastic left CauchyGreen tensor, consistent with definition (46.8)2 for the trial elastic deformation gradient, we define the trial elastic left Cauchy-Green tensor as
ndim

(Ae 2nr()

bet= Fetr (Fetr)T =

n(

).

(47.6)

A=]

With this definition, the algorithmic flow rule (46.9) can be written as
bet = exp [Z>AY"a.f/ b exp [-Eaiaf

1 1j

(47.7)

By result (12) in the preceding theorem both be and e = 1 log(b) commute with r
and, therefore, they also commute with a,f, since fS(, qu) is an isotropic function.
In particular, we have the spectral decomposition
ndim

T An

Tr = E

(A

) n(A)

(47.8)

A=l

By observing that the two sides of (47.8) must have the same spectral decomposition,
the preceding considerations yield the following remarkable results:
(i) The principal directions nrA) of the (unknown) stress Tr and the (unknown)
left Cauchy-Green tensor coincideA with the principal directions nr(A) of the trial left
(A)
t,( )
elr
Cauchy-Green tensor b., i.e., nr = n,t)
forA 1
dim
(ii) By taking the logarithm on both sides of (47.7), using results (47.3) in item
(13) and noting that all the tensors on the right-hand side of (47.7) commute, the
algorithmic flow rule (46.9) reduces to
e=

an ftb(T~,Xq

etr _-

)v

/~=1

5A
'y' )

fAT,)q-f
S=
(r,

a),
m

(47.9a)

389

The discrete initial boundary value problem

SECTION 48

(iii) The stresses r and qa are defined from (47.3) via the hyperelastic constitutive
equations
r

and

= eew(e. )

q'

(47.9b)

7-().

= -O~

The algorithmic equations (47.9a,b) define a return mapping algorithm in logarithmic


strain space with functional form identical to that arising in the infinitesimal theory.
Observe that result (i) implies that the return mapping takes place at fixed principal
axes defined by the trial state b. tr
The advantage of the preceding scheme is clear: It allows the extension to the finite
strain regime of the classical return mapping algorithms of the infinitesimal theory
without modification but with the following additional simplification. The return map
is now formulated in the principal (Eulerian) axes defined by the trial state, which can
be computed in closed form directly from ber, as described below.
REMARK 47.1. From the point of view of implementation, the optimal coordinate
expression of (47.9a) is in the principal axes n() = ntr ( A). Setting f, (ri, T2, r3, qO) =
fz(r, q) one arrives at the reduced system:
m
gerA

A=
t'-ErA
=

jAY]aTA f (

A,qc),

-7;tar~fil~rA~/r

/b=1

,ar = ~n +

ATlaq

l (rA, qT ),

(47.10)

P/=1

Ay

0,

f(TTA, q)

<0,

YAy"

(TrTA,qT)

0.

/1=1

It should be noted that no added computations are needed since the principal directions
n()
(which coincide with nA) must be computed in order to evaluate the trial
logarithmic strain tensor e'trial - lg(b).
REMARK 47.2. Observe that in the isotropic case only the stretching part Ve in the
polar decomposition F e = VeRe is defined. The orientation of the intermediate configuration remains completely arbitrary. The preceding algorithm is consistent with
this observation. In fact, by making use of the kinematic relation
be = FeFe T = FCP-FT where CP = FpTFp ,

(47.11)

the trial elastic left Cauchy-Green tensor defined by (47.6) can be written as
b

= FC- 1FT

with CP

F'PTFP.

(47.12)

Therefore, for given F,, the algorithm can be completed without any knowledge
of the plastic rotation tensor R which defines the orientation of the intermediate
configuration. Thus, only CP is required.

390

J.C. Simo

CHAPTER

48. Implementation of exponential return mapping algorithms


This section gives a detailed implementation of the exponential return mapping algorithms described above. Throughout this presentation, attention is restricted to the
case of isotropic elastic response. Consistent with the observation made in the last
remark, in the implementation of the algorithm described below the set of internal
variables {be, n}
, is adopted in the data structure. Accordingly, denoting by f the
relative deformation gradient defined by
fr

O (,,

= FFn

-1

'

f(Xn)

= I + V7uT(r),

xn

S,

(48.1)

the trial elastic left Cauchy-Green tensor is computed from (47.12) via the identity
be=

LFF-]Fh

IFnTF

= f,be f.

(48.2)

Obviously, the choice {C P, a n} as a set of internal variables is equally admissible.


The step-by-step procedure outlined below summarizes the actual implementation of
the updating scheme defined by the preceding algorithm within the interval [tn, t,+l].
In a finite element context, this update procedure takes place at each quadrature point
of a typical element.
Step 1. Trial state. Given {b, ,,n} [at a specific quadrature point x, E (p,,)]
compute the relative deformation gradient fT and the trial elastic left Cauchy-Green
tensor be tr for prescribed deformation ur:

, ()

IRndim

as

be tr = fbef.t

f-(x,): = I + V(x))

(48.3)

Step 2. Spectral decomposition ofb tr. Find the principal stretches {)etA`} by solving

the characteristic (cubic) equation in closed-form via Cardano's formulae. Compute


the principal directions {n (A)} via the closed-form formula in (33.2). In the case of
three different roots:
r=

/ _ ,etrB) I I - betr
( e tB )2_( XetA)21 2
(2 trI )

etr 2
C

(48.4)

for A = 1,2, 3, with B = 1 + mod(3, A) and C = 1 + mod(3, B).


Step 3. Return mapping in principal (elastic trial state) axes. Compute the loga-

rithmic stretches and the principal trial stresses:


etA = log (eT A)

1,..

dim,

(48.5)
TrA = aEA (

If the trial stress (in principal axes) lies outside of the elastic domain expressed in
term of principal stresses, then perform the return mapping algorithm (see remark

The discrete initial boundary value problem

SECTION 48

391

below) to compute the principal logarithmic strains {ET A, : T }. Then reconstruct the
Kirchhoff stress tensor via the explicit formula
n'dim

?~ = Z

=Cr~nrL~(")e:
a;A nflT (A)
n~'"':
tr (A)

(48.6)
(48.6)

A=l

where {r, A, q' } are the principal stresses, also computed as part of the return mapping
algorithm, defined by (48.5).
Step 4. Update of the intermediate configuration. Reconstruct the updated left
Cauchy-Green tensor via the spectral decomposition
7ndim

b_ = E exp [2 ETA] n

) (

(48.7)

A=i

where ntr (A) and e_A are defined in Step 2 and Step 3, respectively.
For an arbitrary stored energy function (Ae, A', AS), the return mapping algorithm
required in Step 3 of the preceding implementation, must be formulated in strain space.
The classical return mapping algorithms in stress space of the infinitesimal theory described in Chapter I are exactly recovered in the present finite strain context, for a free
energy function quadratic in the principal logarithmic stretches, of the form (ndim = 3)

(EA)

(Ee)2 + (e)e]

+ / _,]
[(le)2

2 A [El + 2 +

(48.8)

where A > 0 and pt > 0 are the Lamd constants, and E := log(AeA), for A = 1,2,3.
This stored energy function (48.8) is often referred to as the Henky model. From the
definition of the elastic logarithmic stretches, for the case ndim = 3 it follows that
3

4e=
A=l

lIog(e) 10g(2A3)

= log(J )

(48.9)

A=l

Now let A = Je-11/3e denote the volume-preserving principal stretches, introduced


in FLORY [1961], which satisfy the condition 1/2)3 = 1. From (48.9) one concludes
that
3

EA = E~ - 3 iEE

= log(i;) =

B=I

E A =

(48.10)

A=1

By analogy with the infinitesimal theory, one refers to e as the deviatoric (logarithmic) stretches. The Henky stored energy function (48.8) can then be written as
w(E ) = [log(J)]J+

where

"

[(,)+2

= A + 3~
2 isi the
I bulk
UI modulus.
IVUL~

(2)2+

(?e)],

(48.11)

392

J.C. Simo

CHAPTER IV

REMARK 48.1. From expression (48.11) one concludes that wr has the correct behav-

ior for large strains in the sense that


oo as Je
0 and, likewise, tD
oo
as Je - o. Unfortunately, Cwis not a convex function of Je and hence fw cannot
be a poly-convex function of the deformation gradient; see, e.g., CIARLET [1988] for
an explanation of this terminology. Therefore, the stored energy function defined by
(48.11) cannot be accepted as a correct model of elasticity for extreme elastic strains.
Despite this shortcoming, the model provides an excellent approximation for moderately large elastic strains, vastly superior to the usual Saint Venant-Kirchhoff model
of finite elasticity. Furthermore, this limitation has negligible practical implications in
realistic models of classical plasticity, which are typically restricted to small elastic
strains, and is more than offset by the simplicity of the return mapping algorithm in
stress space described below.
Below, a brief discussion is given of the actual implementation of the return mapping
algorithm in principal stress space, with attention restricted to an elastic response
governed by the Henky stored energy function defined in (48.11). The generalization
to arbitrary elastic models is straightforward, see SIMO [1992]. The goal here is to
demonstrate that the algorithms described in Chapter II within the context of the
infinitesimal theory apply literally to this model without any modification. To do
so, we introduce the following vector notation for vectors and matrices containing
components relative to the principal axes:

ndim

=
Enint

ndim

and

(48.12)

qnint

The ndim x ndim matrix


of elastic moduli in the principal axes n(A)
associated with the Henky model (48.11) is given by

ntr(A)

ndim

and :=,,N DIAG1,..., 1] is the ndim x ndi identity matrix. Here,=


] r:
( -A
is the bulk-modulus and f > 0 is the shear modulus. We shall use the notation
h to designate the nint x nint matrix of generalized plastic moduli, with entries
h = 2(
)/aa assumed to be constant for simplicity. Following the same
convention used in Chapter II, we define the following (ndim + nint) x 1 vectors of

The discrete initial boundary value problem

SECTION 48

393

generalized stresses and generalized strains along with the corresponding matrix of
generalized moduli:

Ee= {

{~}

and

G=DIAG[C,h].

(48.14)

We use the convention E to label the elastic domain in stress space, defined as:
E

{Z E ]Indim+ni-t' f ()

= f (,

q)

0,

= 1,2,..., m}.

(48.15)

The gradient Vf,,(L) is interpreted throughout as an (ndim + nit ) x 1 vector. With


this notation in hand the return mapping algorithm (46.9) in principal axes takes the
form
m

tr _

A-yVf, (),

with

r = GEtr,

(48.16)

fv(-r) < 0,

Ay> O

and

Af'f

,(E ) = 0.

which is identical to that discussed in detail in Chapter II. The analysis presented
there shows that this constrained algebraic problem arises as the optimality conditions
of the dissipation inequality
[T - X] G - [

- X,],

VTE .

(48.17)

Accordingly, the final state E, in principal axes is the closest-point projection onto
the elastic domain E (specified in terms of principal stresses) in the metric defined
by the generalized elasticity matrix G (see Fig. 48.1). The two general strategies
described in Chapter II, i.e., the general closest-point projection algorithm and the
cutting plane algorithm, can be used to compute -. No modifications in these two
algorithms are required.
To summarize, from a practical standpoint, the preceding developments translate
into the following prescription for the implementation in the finite strain regime of
any infinitesimal model of classical plasticity.
(i) Express the model in the principal axes defined by the trial state. The classical
linear infinitesimal elastic stress-strain relations then translate into a linear relation
between principal Kirchhoff stresses and principal logarithmic strains.
(ii) Construct the algorithmic counterpart of the model (in principal axes) by applying a standard return mapping algorithm in the principal axes defined by the trial
elastic state.
This defines Step 3 in the step-by-step outline of the algorithm given above. The
other steps in this algorithm are independent of the specific elastoplastic constitutive
model and remain, therefore, unchanged. To complete the formulation of the algorithm

394

J.C. Simo

CHAPTER IV

tr1

tn+ 1

T3=U
FIG. 48.1. Geometric interpretation of the return mapping algorithm for multisurface plasticity in principal
stress space for two possible trial elastic states.

it only remains to compute the algorithmic elastoplastic moduli in the incremental


constitutive equation (43.10). Below it is shown that results derived in the context of
the infinitesimal theory remain essentially unchanged.
49. Linearization: The exact algorithmic tangent moduli
The step-by-step procedure outlined in the preceding section defines the Kirchhoff
stress tensor r7 at time r C [t, t+l], along with the updated internal variables, in
terms of the motion pT. By inserting this algorithmic expression into the Galerkin
projection (44.2) of the dynamic weak form, with matrix expression defined by (44.11),
one obtains a nonlinear dynamical system in P =lRdo x Ridof for the motion ~co
and the velocity field. The iterative solution of this nonlinear initial value problem is
accomplished by solving a sequence of linearized problems, each of them defined by
(43.4), with matrix expression given by (44.13). The linearized problem is completely
specified once an explicit expression is given for the consistent elastoplastic moduli
ceP in the algorithmic incremental constitutive equation (43.10), i.e.,
AUTr-:

= CrP[V(Au) o r],
L

equivalently,

[AT],,ij

jPkl(AU)k,1.

(49.1)

Recall that Au E V, is the incremental displacement, a vector field on S, = qp(D)


which is the primary unknown in the linearized problem, and ar,- is the Lie derivative defined by (43.6). The goal of this section is to provide an explicit expression
for the algorithmic moduli CeP for the case of elastic isotropy formulated in principal
stretches.

The discrete initial boundary value problem

SECTION 49

395

f =FxFnl
currentConfiguration
at Time tn

FLIT1]
Fixed Inten
on.

a:

-1 ll
nd11

L,,-

Ic X C

S.I

= 1 + V(A)
Nearby Configuration
at Time tn+5

Fi-, -X<
FIG. 49.1. Local configurations involved in the linearization of the algorithmic constitutive equations.

Consider a nearby configuration ,+ E C and let f+C denote the relative deformation gradient between the placements S, and S,+, respectively defined by
(,+

+ (Au

and

f-+C = 1 + (V(Au),

(49.2)

where V(-) denotes the spatial gradient relative to the coordinates xk =

'pk(XA,

r).

The strategy used below in the derivation of the algorithmic moduli mimics the procedure used in Chapter III to obtain the elastic moduli for models of elasticity formulated in principal stretches. Briefly, the algorithmic constitutive equations (49.1)
are first reformulated in a fixed reference configuration, chosen to be the intermediate
configuration at time t. The algorithmic moduli are then computed in this fixed configuration using ordinary differentiation and the result is finally push-forward to the
current configuration. Recall that the trial elastic deformation gradient Fet,: a linear
map between the local intermediate configuration at time t and the current (local)
configuration, is defined as (see Fig. 49.1)
F'e,

= F+(Fjp-t = fT+FF

= f+Fe

r.

(49.3)

To implement this strategy we introduce the second Piola-Kirchhoff and the right
Cauchy-Green tensors, relative to the intermediate configuration at time t, respectively defined as the pull-backs of the Kirchhoff stress Tr,+ and the unit tensors on
the placement S,+(, i.e.,

s,,+=

F'e,)

1Tr,+(F'er(

S =and
and

r
,+

=(F+)
(Fe()T

(FIr
lt+(]

(
(49.4)

396

J.C. Simo

CHAPTER V

Using (49.3), (49.4)1 and definition (43.6) for the Lie derivative, it is straightforward
to verify the following push-forward relation:
ar

= F rt [dS,+t/

(49.5a)

dl=o] (Fr )T.

An identical manipulation using (49.4)2 and (49.3) yields the analogous result for the
right Cauchy-Green tensor, namely,
(

2sym[V(Au)] = (Fetr)-T[djC/etd

(49.5b)

Relations (49.5a,b) are the key results used in the derivation of the consistent
elastoplastic moduli Cep appearing in the incremental form (49.1) of the algorithmic
constitutive equations. The closed-form expression is given in the following.
THEOREM 49.1. Let [cerPAB] denote the ndim x ndim matrix obtained by linearization
of the algorithm (47.9a,b) formulated in principal axes, i.e.,
CrA B

a-rA/e',tB, for A, B = 1,..., ndim.

:=

(49.6)

The tensor of algorithmic elastoplastic moduli in the constitutive equation (49.1) is


given by the closed-form expression
ndim

ep =

dim

[CtpAB n

nt(A)
(A)

tr(B) 3 ntr(B)] + gt,

(49.7a)

A=1 B=I

where the nonzero components gtIJKL of the tensor gt

relative to the basis {nt(A)

are defined by the following expressions (I, J = 1,..., ndim):


gHlIII =-2%,

(49.7b)
-tr

tr-tr

-t(

T, I (re tr )r)2
(Ae t r 2

TT
J(,\e I
t
2
(Ae r

for I

J.

The tensor gtr thus depends only on the form of the stored energy function in principal
stretches; only the matrix [CeAB] depends on the specific model of plasticity under
consideration via the algorithm (47.9a,b).
PROOF. Since the principal directions of both r, and b coincide with those of be tr
defined in the trial state phase, the spectral decomposition of the Kirchhoff stress r,+
and the second Piola-Kirchhoff stress tensor Sr+( take the form
ndim

~l'r =

Antr(A)

Tr+~ATr+
.

ntr(A)

0 '~*+~

A=I

(49.8a)
ndim

S
A=I

=ES+Ttr(A

) O N-r(A)

The discrete initial boundary value problem

SECTION 49

397

where
nt(A)Aetr
Ar=

Fe

Ntr(A)

and

S,+CA = TT+(A/(

+4CA)
.

(49.8b)

Obviously, the unit vectors {Ntr(A) } are the eigenvectors of the right Cauchy-Green

tensor

The unit norm condition INtrA)I = 1 then implies the relation

d Ntr(A)

m
di

x N7t(A )=

=W

W AN(9)

=
_

AfB=l

where WT AB = -W

BA. As in the derivation of the tangent elastic moduli given in

Section 33, the argument now proceeds in two steps.


Step (i). Computation of the Lie derivative fAr, using expression (49.5a). The

spectral decomposition (49.8a) 2 for S,+ together with (49.9) imply

d(

(=O

A=l B=1

dSZZ
d(

ndim

ndim

)
(A
[S B - ST-A]W ABNrt
Nr(

B)

(49.10)

A=1 BOA=I

Using the chain rule, from expression (49.8b)2 one easily concludes that

C =0
-

+A
1

,s
=

)2

[je}r

=ld

I 7Ar B

26ABTB

d.Er+CB
-e

(49.11)

t;

Inserting relations (49.11) and (49.10) into expression (49.5a) for the Lie derivative
and using (49.8a,b) yields the result
ndim

dim
fdE

FEt
ahT,

A=I/3=1 I
ndim

ndim

q-

Ar

B '"

2
2S
B nl d
I d
e

tr

T- (T A) -rA()X
t
etr
A=I B3A=I

er

tr(A)

tr(A)

=O
2_

IWABtrA)

(49.12)
B)

(49.12)

rB

Step (ii). Computation of the Lie derivative Aul using expression (49.5b). From
the spectral decomposition of the right Cauchy-Green tensor C er,
,r+(, relation (49.10)

398

J.C. Simo

CHAPTER IV

and the chain rule, it follows that


d
2

fl1rtdim

dC

nZdi.

etr

de

dim

A)

I= E
E
(Aet
r
(=O
A=l B=I

N(A)

d_

4A
A

ndim

2lf[(Aetr)
2
B[(,\

1(A)

22r

JU2
WrABNT

'

(49.13)

Inserting this result into (49.5b) and noting that [Fetr]-TNtr(A) =

tr(A) /.e t, yields

(AectA)

Nr
-

A=1 B#A=I

the result
ndim
I

n dim
i

E5 (etr

A)

_E

rB

tr(A)

ntr(A)

A=l B=
ndim

fndim

(etr

E"

E
2
A= B7A=~1i
E

T /A -

,etr

(etr

)2

e tr

WT_

ABtr) One(B).

Wrgnr(A)

3 nr)

(49.14)

A direct comparison of formulae (49.12) and (49.14) yields the closed-form result
(49.7a,b) thus completing the proof of the theorem.
The ndim x ndim matrix [CTPAB] of algorithmic moduli appearing in expression
(49.7) contains the derivatives of the principal stresses Tr A relative to the elastic trial
logarithmic strains e trB. This matrix has an expression identical to the algorithmic
elastoplastic moduli of the infinitesimal theory derived in Chapter II. In terms of the
notation introduced in Section 3, the general result (17.6) now yields

-Gp-PG Vf(Xr) PGTVf(ZT)

-eBp

IcAB] =PG

where

Vf(-)

G, = [G- l + AyV 2f(,)]


t

PGTPVf(E)

(49.15)

l and P denotes the orthogonal projection from

R(ndi+nLin ) onto R"ndim, thus defined by P[Z] =

In summary, the computation of the algorithmic moduli reduces to evaluating the


ndim x ndim matrix [ce AB] using expression (49.15) since the components in principal
axes of the tensor t, defined by (49.7b), can be computed once and for all with total
independence of the specific model of plasticity.
50. A generalization of J2 -flow theory to finite strains
The preceding theory and numerical implementation will be illustrated within the
specific setting of J 2-flow theory. The generalization of this classical model of metal
plasticity considered below follows the lines outlined in Example 37.1 and incorporates
finite elastic strains within the framework of a multiplicative decomposition of the
deformation gradient. A main goal is to demonstrate that the radial return algorithm
is exactly recovered within the algorithmic framework described in the preceding

The discrete initial boundary value problem

SECTION 50

399

sections merely by setting = t,+l so that At = t, - t+l, i.e., for the backward
Euler form of the preceding class of exponential return mapping algorithms.
(A) The case of nonlinear isotropic hardening. Recall that for J2 -flow theory the
elastic domain E in principal (true) Kirchhoff stresses is the cylinder defined by the
von Mises yield criterion

f(,

) := [dev[t] | -

[y + K'()] < 0,

(50.1)

where oy > 0 is the flow stress, is the equivalent plastic strain and dev[r] is the
vector containing the deviatoric principal Kirchhoff stresses, i.e.,
dev[t]

tr[]i

and

Idev[r][ = v/dev[e]

dev[],

(50.2)

ndim

where tr[e] =
i1 is the sum of the principal Kirchhoff stresses, i.e., the true
hydrostatic pressure. The unit normal to the von Mises cylinder is v = dev[]/Idev[r] d
and we have v

*1

= 0. By setting ;r,+

= de"t,, the specialization of the general

scheme (48.16) to the present example yields the radial return algorithm described in
Chapter II, now formulated in principal axes as
Tn+l =

-2LA-yn+l
2+l

and

,+i =

Ay,

(50.3)

supplemented by the standard Kuhn-Tucker loading/unloading conditions. The constraint i'+l *i = 0 yields n+ 1 1= T+ 1, which implies J++ = Jtr = J+
As in the infinitesimal theory, Eq. (50.3)1 determines Pv+l solely in terms of the trial
state. The resulting expression along with the yield condition evaluated at the trial
state take the form

dev[-

and

f t+ = dev[[ml]

- V/[y

+ K(n)].
(50.4)

Finally, again as in the infinitesimal theory, one observes that the algorithmic loading
condition f +l > 0 implies that Ay > 0. The consistency condition fn+l = 0 then
yields the following scalar equation which defines A-y during plastic loading:
fn-+I: =

[K-

3-2/iA-'
)
-

()]

= .

(50.5)

Due to the convexity of K(.), Eq. (50.5) is ideally suited for an iterative solution by
Newton's method for fixed f4+4 . Of course, the resulting algorithm is identical to the

400

J.C. Simo

CHAPTER IV

radial return method with nonlinear isotropic hardening, now formulated in principal
stresses.
REMARK 50.1. An extension of the preceding algorithm to accommodate viscoplastic
response is constructed exactly in the same fashion as described in Chapter II. One
considers the standard viscoplastic regularization in which the Kuhn-Tucker relations
are replaced by the constitutive equation y =]g(f(r,q)[ / )> 0, where ]x[:= (x +
H(x))/2 with H(.) denoting the Heaviside function. To circumvent the characteristic
ill-conditioning exhibited by standard viscoplastic algorithms in the rate-independent
limit as the viscosity q1- 0, the viscoplastic regularization is rewritten in terms of
the inverse function g-l() as

f+

= g-'

A)

for ft+l > 0 A:

Ay > 0.

(50.6)

Observe that g' l(.) is given in closed-form for practically all rate-dependent models
of interest, which typically characterize g(-) via exponential functions or power laws.
Combining (50.6) with the consistency condition (50.5) results in the modified scalar
equation

-9 (t A-y) + nl-2/jAy-[k
1
(tin +
K'(n)]

J23

= 0,

(50.7)

which defines Ay > 0 and can be easily solved by Newton's method. Observe that
(50.7) is well-conditioned for any value of the viscosity coefficient
[0, oo). In
particular, it is possible to set = 0 in (50.7) to recover exactly the inviscid limit
defined by Eq. (50.5).
(B) A unified scheme for kinematic hardening and viscoplasticity. The extension
of the preceding model of J2-flow theory at finite strains to incorporate kinematic
hardening relies on the following observation. By frame invariance, a general yield
criterion of the form f (r, q ) < 0 must be an isotropic function of its arguments and
can only depend on the principal values of r and q. Accordingly, in the presence of
kinematic hardening, the von Mises yield criterion (50.1) in principal axes takes the
form

f (dev[I]-

) = dev[t]

[yu + K'()]

(50.8)

where E R'nd m denotes the vector of principal values of the back-stress tensor in
principal deviatoric (Kirchhoff) stress space; i.e., I = 0. Equations (50.3) defining
the return mapping algorithm for J2-flow theory in principal stresses are then replaced

401

The discrete initial boundary value problem

SECTION 50

by the following relations


+1 =+

2pAyvi',

~n+l =Cn + >A2HV+l,


n+ =
/

(50.9)

Ay,

]g(f(dev[] - L,))[ 3 0,

where H is the kinematic hardening modulus, r+


to von Mises yield surface now defined by
,+

n + 1/13/n+li,

with

= a e

and v'+l is the normal

n+: = dev[in+] - n+,.

(50.10)

Relations (50.9) define a return mapping algorithm in principal stress space which
incorporates a form of kinematic hardening consistent with the Prager-Ziegler evolution law of the infinitesimal theory. Proceeding exactly as in the derivation leading to
(50.7), one concludes that the trial state completely determines the unit normal to the
von Mises yield surface by expression (50.10). Similarly, one sees that plastic loading
occurs whenever fr+l > 0. The multiplier A-y > 0 is then determined by solving the
scalar equation

-9
-

(+
(A

[K '(

A-y[2y +

2H]

-A)-k/((

)) = 0.

(50.11)

Since g(x) = 0 X x < 0, Eqs. (50.4)-(50.11) collapse to the standard radial return
method described in Chapter II for combined linear isotropic/kinematic hardening by
setting q = 0 and assuming that K'(.) is linear (see Fig. 50.1).
Observe that the form of (50.11) precludes ill-conditioning for any value of the
viscosity parameter Trl [0, oo). To complete the algorithmic treatment of J2-flow
theory it only remains to compute the algorithmic elastoplastic moduli P+, associated
with the return map (50.4)-(50.11), and defined by the general expression (49.14) in
the rate-independent case. A direct calculation identical to that described in Chapter II
yields the result

[e'np+IAB]

= rd; 1i

+ 2[sn,+(I-

311

1) -

+,(Fn+, (n+l)],\

(50.12)
(50.12)

JC. Simo

402

CHAPTER IV
A-

FWitr I

12
U

L nJ

FIG. 50.1. Geometric illustration of the standard radial return method for combined isotropic-kinematic
hardening in principal stretches obtained here for zero viscosity ( = 0). In the present approach, the
algorithm takes place in principal Kirchhoff stress space and is the same as in the infinitesimal theory.

where D~+l is defined by (50.4), while the scaling factor s,l


6,+, are given by the explicit formulae
2pAy
sn+ :=

and
1)3

Observe that the (dim

tr

1
l

and the coefficient

2/A-y

K"+H q_+ t

A+1:=
'3t

(50.13)

3Observe
algorithmic
that
the
(nx)atrx
[f elastoplastic moduli in

X fldim) matrix [c + IAB] of algorithmic elastoplastic moduli in

principal axes defined by (50.12) is symmetric. For r7 = 0 one recovers from (50.12)
the algorithmic moduli of the rate-independent infinitesimal theory.
51. Two-stage projected exponential return mapping algorithms
The goal of this section is to design the counterpart of the two-stage projected implicit Runge-Kutta scheme described in Section 15 of Chapter II and summarized in
Table 15.1 within the framework of the exponential return mapping algorithms introduced in this chapter. The main difference between this scheme and its generalization
presented below lies in the form taken by the linear extrapolation step that defines
the trial stress field in the second stage of the product algorithm. For the finite strain
theory, a literal transcription of Step 3 in Table 15.1, for the purpose of defining the
trial stress rt+l in Stage II of the algorithm, would result in open violation of the
fundamental principle of frame invariance, thus leading to completely meaningless
numerical results. The underlying reason is that the stress fields 7-,+, and r,+l involved in the extrapolation update are defined in different configurations and cannot

The discrete initial boundary value problem

SECTION 51

403

be naively added. Substantiation of this claim requires a detailed examination of the


basic kinematics involved in the generalization of the two-stage algorithm to the finite strain regime. Before doing so, and to motivate the subsequent developments, we
record below the form of the linear extrapolation update in Stage II which is amenable
to generalization when finite deformations are involved.
First, one operates with elastic strains in place of stresses. This distinction is immaterial if the constitutive equations are linear but results in considerable simplification
in the large deformation regime. Second, the linear extrapolation update is carried out
in two steps. Recall that Au is the total incremental displacement field in the time
step [t,,t,+l] and set
where AE,+ti = 9sym[V(Au)].

En+g = En + Ae,+

(51.1)

The first step in the extrapolation update produces an intermediate strain field E *
defined by the linear extrapolation formula, while the second step defines the final
trial state t, in Stage II of the algorithm according to
1

e+

an

+1 and

etr

e*

zn+ = n+ +-

1I

9- AEn+,d

(51.2)

By combining the two relations in (51.2), using (51.1) and multiplying the result
by the elasticity matrix C, one recovers the linear extrapolation formula given in
terms of stresses in Table 36.1. The key observation to keep in mind, however, is
that the extrapolation formula (51.2)1 remains meaningful in the finite deformation
theory since the finite strains involved are all defined on the same configuration. It will
be shown below that the extrapolation formula (51.2)1 together with an appropriate
generalization of the update formulae (51.1) and (51.2)2 arise naturally by applying
the methodology presented in Section 46.
To give a concise formulation of the algorithm it proves convenient to rewrite the
evolution equations that define plastic flow as follows. Define the forcing functions
fndim

ndm

/P =

ylaTf,(r,qa)

and

ha = yaqf,(r,q),

JL=1

(51.3)

I=1

where the Lagrange multipliers y > 0 either obey the usual Kuhn-Tucker conditions
for the rate-independent problem or, for single surface rate-independent models, are
defined via the viscoplastic constitutive equation
a'y

= ]g(f,

(r,qA))[

with

= 1.

(51.4)

Here ]x[= (x + H(x))/2 is the ramp function, 7rE [0+, oo) is a viscosity parameter
and g(-) is a function satisfying g(x) = 0 X x = 0. For multisurface plasticity
(/ > 1) models of this type are generally not well-defined and must be replaced by

404

J.C. Simo

CHAPTER

the Duvaut-Lions viscoplastic regularization described in Chapter II. The problem of


evolution is then defined by the rate equations
FP = LPFP
where LP = F eFPl t=t, = F P

and

ds = ha

in [tn,tI+l],

(51.5a)

PFe, together with the initial conditions

and

(alt=t,

= Sn.

(51.5b)

The forcing functions IP and ha in this problem, defined by relations (51.3), depend
on the generalized stresses (r, qa) which, in turn, are functions of the internal variables
{Fe , Iai}
via the constitutive equations
' = Fe[CueW(FeTFe)]FeT and

qa = -Oa7().

(51.6)

The primary variables in the problem are, therefore, the internal variables {Fe, a},
although the evolution equations (51.5) are cast in terms of FP to facilitate the formulation of the algorithm described below. Before doing so, we introduce some kinematic
preliminaries.
(A) Kinematic preliminaries: The generalized mid-point configuration. Recall that
in the infinitesimal theory the displacement field in the first stage of the algorithm,
corresponding to time t,+ = t,+l + (1 - )t,, is a convex combination of the
displacement fields at time t, and time t+, respectively. To generalize this set-up
to the finite deformation theory, consider the one-parameter family of generalized
mid-point configurations po+ C defined by the convex combination
Wn+f: =

qOn+l + (1 -

0 e (0 1],
dE)(,,

(51.7)

and illustrated in Fig. 51.1. Setting F,:= Dp, and F,+l := DP,,+l the deformation gradient associated with the configuration p,+o is also given by the convex
combination
F,,+,: =

F,,++ (1 - 2O)F..

(51.8)

In the finite element implementation of the algorithm it is convenient to work with the
relative displacement field between the configuration S,+., = (p+ (Q) and S, =
p, (), parameterized by points in S,+; i.e., the vector field u,+ :'Sn+
,+
Rndim
defined as
un+9(xn+z9) = n+aO - W.n qn+,(a(X+z)

for Xz+,

E Sn,+.

(51.9)

1
As before, let f,+o = Fn+,gF
denote the relative deformation gradients between
the configurations S,+s and S,. Similarly, we shall denote by f,+ = Fn,+1F-'
the relative deformation gradient between the configurations S,+ and S+,y (see

The discrete initial boundary value problem

SECTION 51

405

Referenc
Configu:

iguration
e tn

FIG. 51.1. Illustration of the three configurations involved in the formulation of the modified return mapping
algorithm and the corresponding relative deformation gradients. Attached to each configuration are shown
the corresponding deformation tensors.

Fig. 51.1). The following expressions in terms of u,,+ are an immediate consequence
of the chain rule and (51.8):
fn+,7

[1 - Vu+]

and

f+l =

1+

0- Vun+J.

(51.10)

Observe that (51.10)1 differs from (48.1) since the vector field u is assumed parameterized in this latter expression by points x, in the configuration S,. Finally, let
{n,, be}, ~{,+o, bn+,} and {T+l , ben+} denote the (true) Kirchhoff stress field and
the left Cauchy-Green tensor at the configurations Sn, Sn+ and S,,+, respectively.
These tensor fields are also illustrated pictorially in Fig. 51.1. Consistent with the
remarks made at the beginning of this section, two crucial facts should be kept in
mind when a Lagrangian description of the motion is employed:
(i) Fields defined on a given configuration, say for example Tr+d defined on Sn+ ,
obey the usual frame invariance requirement only relative to isometrics superposed
on that particular configuration; for the example at hand this configuration is S,+e.
Superposed isometries on any other configuration leave these fields unchanged. For
the example at hand, isometries superposed on either S, or S,+1 do not affect r,+g.
(ii) A linear combination of frame invariant tensor fields defined on different configurations renders tensor fields which are no longer frame invariant. For instance,
ar,+o + 3r,+1 is not frame invariant for a, 3 E IR.
It is precisely because of these two facts that a literal transcription of the linear
extrapolation formula in Step 3 of the two-stage algorithm summarized in Table 51.1

406

J.C. Simo

CHAPTER IV

TABLE 51.1
Return mapping algorithm in principal axes.
For prescribed input data {ber, ,

} perform the following steps:


t

(i) Solve the eigenvalue problem: betrn (A) =


A = ,..., ndim.

ernttr(A), for all the eigen-pairs {Ae,

nt()},

REMARK. Although closed-form solutions are possible the most robust technique is Jacobi iteration
given the small problem size (ndim < 3).
(ii) Compute the trial logarithmic strain components setr = log(AAt), for A = 1,..., ndim to evaluate
the trial stresses. For the Henky model we have
rAt

= Alog(J.") +21E
Ar

while

q r = -a(-

/a..

REMARK. Note that log(Je t r) = Adm EeAr


(iii) Compute the principal stresses {TA ,, q } and the principal logarithmic strains {EeA r,
T } from
the trial state by performing a closest-point projection return mapping algorithm in principal axes.
REMARK. For the Henky model the return map is the classical stress-space return mapping algorithm.
(iv) Reconstruct the Kirchhoff stress and the elastic left Cauchy-Green tensor via the spectral decompositions as
ndim

ndim

VTE

tr(A)

tr (A)

b=

A=L

(Ae

)2t()

ntr(A)

A=i

where XeA7 = exp[EA T] for A = 1,.

, ndim.

is completely meaningless in the finite deformation regime. The two-stage algorithm


described below carefully avoids updates of this type.
51.1. The practical implications of an integration algorithm which involves
objects lacking frame invariance are hardly ignorable. For instance, numerical results
computed with such an algorithm can be altered merely by reorienting the coordinate
axes for the initial finite element mesh. It is clear that a method which produces results
that depend on the mesh orientation can hardly be regarded as useful.
REMARK

(B) The general two-stage projected return mapping algorithm. The general set-up
for the construction of the two stages involved in the single step projected return
mapping algorithm is the same as presented in Section 46. To provide a concise
formulation of the algorithm that accommodates both the rate-independent and ratedependent problems it proves convenient to introduce the following notation:
m

Al

= E AY

,ifi

(r,

E),

=1

(51.11)

The discrete initial boundary value problem

SECTION 51

407

As in the continuum problem, the Lagrange multipliers Ayl > 0 obey either the
discrete version of the Kuhn-Tucker conditions for the rate-independent problem or,
for single surface rate-dependent models, are defined via the algorithmic constitutive
equation
(51.12)

[ with /t = 1.

r,T))

A7y = -]g(fp(

With this notation in hand, the update of the given initial data (Fen, c, n) G I at
time t, for a prescribed deformation gradient obeying relation (51.8) proceeds in two
stages as follows.
STAGE I. The algorithmic solution {F+o(X), cu,n+(X)} C ]GL(rndim) x ] nint that
approximates the actual solution at time t,+g is computed by the same algorithm
described in Section 46. Specializing the results given there to t = tn+s gives
F
-tL
1 ] nP

FP+, = exp [

and

n+o

+ t9 Aha+,9.

(51.13)

Using the multiplicative factorization F,+, = F +,FP+,9,the same arguments presented in Section 46 now yield the following update formula for the elastic part of
the deformation gradient:

F,+, = exp[-6i9Pn+,9F,,%

where Fn+4" = F,f+FP-.

(51.14)

Equations (51.13)2 and (51.14), along with (51.11) and the elastic constitutive equations (46.9b) define Stage I.
STAGE II. This stage constructs an algorithmic approximation {Fe+lE, n+l} E
GL(ndim) x RniR to the exact solution for the internal variables at time t+l via a
product formula algorithm. Consistent with a mid-point rule approximation, a trial
state is defined in Stage II by the forward Euler step
exp [(1-

FP,=

)AtL P+,]
P F-+
(51.15)

a n+

+ (1-

r)Ah n+ .

This explicit update defines internal variables for which the yield condition will, in
general, be violated. Accordingly, they are regarded as trial values which are subsequently corrected via a return mapping algorithm. To recast the preceding formulae in
terms of the elastic part of the deformation gradient, which is regarded as the primary
internal variable, the trial elastic deformation gradient is defined via the multiplicative
factorization as Fn+l = F FP Using this definition, the update formula (51.15)
is expressed in terms of the elastic deformation gradient as

Fn r=exp
exp[(1F+=F."
[(I -

)Af Pn+l,] FP+,+

Fe' IF,,-' exp [(I t

e)Fee+29
)+oA Alp+19Fp61-+-1 ]eF+O p, o

F, r exp[( - )AlrP,,+]F+.
=

(51.16)

J.C. Simo

408

CHAPTER IV

Using expression f,+l = F,,+lF-,+9 for the relative deformation gradient, a rearrangement of (51.16) gives the following explicit formulae
F.+ = fn+l exp[-(1 - )Al1P+]Fn+,9,
(51.17)
i n+I =

ao+9

(1

Sho

n+9

which define the trial state for Stage II of the algorithm. The final state {Fn+ l, n.}
is then computed via an exponential return mapping algorithm essentially identical
to that used in Stage I. It will be shown below that the resulting two-stage scheme
becomes closed-form under the assumption of elastic isotropy.
52. Closed-form of the two-stage projected IRK for elastic isotropy
Under the assumption of elastic isotropy, the constitutive response is completely characterized in terms of the internal variables be, }. The results in Section 48 then
apply literally to Stage I, without any modification, leading to the algorithmic flow
rule
n
En+ - 9AIP
and I, +r = {o,+, + 0Ah,+oa,
(52.1)
where
E' +19
log~b' + and e tr
where En+o9 = 2 log(bn+og) and %En = log(bt +) are the final and trial logarithmic
elastic strain tensors at the configurations Sn+o, respectively. Using these results,
Stage II can be reformulated in a form ideally suited for implementation as follows.
Upon defining the strial elastic left Cauchy-Green tensor via the standard expression
b+= (Fe )T, the update formula (51.17)1 implies the sequential update
F

b;

= exp[-(l -

)AtlP+

P
] be+l exp[-(l - 9)AI
+

(52.2)
bt + = f+bn+

tfT+1 .

Now observe that the right-hand side of (52.2)1 involves symmetric tensors defined
on the same configuration Sr,+, all of which commute as a result of the restriction
to isotropy. Consequently, taking logarithms on both sides of this expression and
eliminating both AlP,+d and Ah, +,o
0 using (52.1) gives
ns

+1O9 -

wC*7+9 =

n+

n+9

and
(52.3)

L+

where Ee,+ = 2 log(ben+). These formulae involve tensor fields defined in the same
configuration Sn,+g, thus furnishing the correct generalization of the linear extrapolation update (51.2)1. In addition, from (51.17b)l we conclude that
bn+l =

fn,+l
exp[2e+
n~~~~~~~~~~+
l

] fT+

and

trn+1
=52n+
a
=+ cn+0'

(52.4)

SECTION 52

The discrete initial boundary value problem

409

We show below that the actual implementation of this two stage scheme can be
performed in a rather efficient and cost-effective fashion.
A concise implementation is achieved by grouping together the steps common to
each of the two stages of the algorithm, as summarized in Table 51.1, and making a
systematic use of the spectral decomposition in principal Eulerian axes. For prescribed
trial values {bt, i}, the steps in Table 51.1 define a exponential return mapping
algorithm in principal axes, thus producing the following results: (1) The spectral
decomposition of b, (2) The principal logarithmic strains, { eA, T}, (3) The
principal Kirchhoff stresses {rA,,q' }, and (4) The reconstructed Kirchhoff stress
tensor Tr and elastic left Cauchy-Green tensor be.
From the point of view of implementation, it is convenient to view Table 51.1 as a
subroutine that is invoked with different values of the input parameters {betr, tr} at
appropriate points during the computation, according to the following strategy.
STAGE I. The first stage of the algorithm is identical to the scheme already described
in Section 48. Explicitly, the following steps are performed for prescribed initial data
{b, e n} and a given displacement increment un+o:
Step 1. Define the trial state according to
be+tr

=f.+b

and

'n+

n,

(52.5)

where fn+g is the relative deformation gradient computed via (51.10)1.


Step 2. Perform the return mapping algorithm in Table 51.1, with input parameters
be tr

and

tr=+

At this point, the following converged results are available: the principal logarithmic
stretches eA n+,
, nh} and the principal stresses {rA , qT} together with the Kirchhoff stress tensor r,+, and trial principal directions n+ ), which coincide with final
principal directions at the configuration Sn+og. Observe that the left Cauchy-Green
tensor b+g,also computed in Table 51.1, is not necessary in Stage I.
STAGE IIA. Since the trial values Ee tr
and the converged principal values A
are available from Stage I, the computation of the tensor fields e= 1 and e entering the linear extrapolation formula (52.3) is trivially accomplished via the spectral
decompositions
ndim

etr _E
n+d-d

etr
tr(A)
HA .+,9 n +

tr(A)
n+9X

Fe n
ntr (A)n
An+ n+
(

tr (A)
n+ O
9

A=1

(52.6)

ndim

ce
:

'

A=1

The computation of the linear extrapolation formula (52.3) and the push-forward
relation (52.4) is then carried out according to the following steps:

410

J.C. Simo

CHAPTER IV

Step 1. Compute the linearly extrapolated components -An+


etr*
- 0

1e* e
An+O

19

1
(oe n+,9

,-

etr
An+O
no

nd

1 -W

1-

tr

A = I,..

and

C,

, ndim-

as

(52.7)

oE~
n+9

This defines
ndim

e
= 2i log(be*)
n+79
'Y~iiAn

n+ e*
An +,r(Atr
A=l

Step 2. Compute the tensor field ben*,

(A)
n+oo
n+7.

on S

by exponentiation as

ndi

be*+

-n+ L~"~
9

exp [2EeA *n +o]

tr(A)
T129

n+

n+9

(52.8)

A=I

Note that ben~_ = exp[2en+*] is a (contravariant) tensor field on Sn -.


Stage IIA can be viewed as defining the initial data {ben*+9o (n+o} for the second
stage in a product formula algorithm. Observe that these formulae have the remarkable
property of preserving the constraint det[be * 9] > 0. In other words, the initial data
be'_*o lies in S+ C GL(ndim) for all x.,+ E Sn+ , as required.
STAGE IIB. The second part of Stage II is identical to Stage I. The prescribed initial
data now becomes be * ,*+,} and the steps involved in the update are:
Step 1. Define the trial state according to
betr = f+IbefT+1 and

Ean+l =

(52.9)

where f+l is the relative deformation gradient computed via (51.10)2.


Step 2. Perform the return mapping algorithm in Table 51.1, with input parameters
and tr = ,1

betr = be t

Stage IIB thus defines the final values {be+, > n+l} which become the initial data in
a subsequent time step. These updated variables satisfy the constraint det[bet+ ] > 0 in
S+ I thus remaining within the admissible set 2i. Observe that the full Kirchhoff r+ I
also computed in Table 51.1 is admissible (lies within the elastic domain) but is not
necessary in this stage and its computation can be omitted. Furthermore, because of
the structure of the conserving time stepping algorithms described in the next section,
it will be shown that Stage II of the algorithm need be performed only once within a
typical time step.

The discrete initial boundary value problem

SECTION 53

411

53. Global time-stepping algorithms for dynamic plasticity


This section is concerned with the last step in the time-discretization of the initial
boundary values problem for dynamic plasticity at finite strains: The formulation of
global time stepping algorithms that relate the velocity field to the motion in configuration space. This global time stepping algorithm, together with the local exponential
return mapping algorithm, produces a nonlinear algebraic system after the Galerkin
discretization in space described in Section 42 is introduced (see the commutative
diagram in the introduction of this chapter). Before describing alternative global time
stepping algorithms we prove an important property of the Galerkin projection already
mentioned in Section 42: The finite dynamical system (44.11) inherits the global conservation of the infinite-dimensional system. One of the goals in this section is to design global time-stepping algorithms that preserve this unique feature of the Galerkin
projection.
Let rA = XA + dA be the position vector of a node XA in the triangularization.
Define the Galerkin approximation to the resultant force and the resultant torque of
the prescribed system of loads by
Nxt =

Fx

and

,,xt
=

rA

Fxt

(53.1)

AE9

AEg

From (44.9) it is easily concluded that if body force B and prescribed traction T are
constant, expression (53.1) reproduces the exact results (39.9)1 and (39.10)1 for the
infinite-dimensional dynamical system. Similarly, the Galerkin approximation to the
total linear momentum and the total angular momentum are defined by
Lh =

EpA

and

AEg

jh = E rA x pA.
Ac9

(53.2)

The following result provides the counterpart for the finite-dimensional dynamical
system (44.11) of Theorem 34.1.
THEOREM 53.1. The Galerkin projectionpreserves the conservationlaws for the pure
traction initial boundary value problem (r, = 0) under equilibratedloads (Nhxt =
Mehx = 0), in the sense that
dLh=O
dt
where P =

Rnd f

and
x

dJh=0
dt

]tRdof

inlIxP,

(53.3)

is the phase space for the system (44.11).

PROOF. The result follows from the proof of Theorem 39.1 since Voh C Vo. A direct
verification of (53.3)2 starts by noting that
(C ErA
AEg

xFin = (A)t
oN
Ac9

(53.4)

412

J.C. Simo

CHAPTER IV

for any constant


E I di., as a result of (44.10). Using expression (44.10)2, a
straightforward manipulation yields

( x r4A) ThVNA = Th
AEg

(rA X GRAD[NA]) Dh-i

(535)

Ag
O

Upon noting that the expression DSDh = ZAEg rA GRAD[NA] is the Galerkin approximation to the deformation gradient, relation (53.5) collapses to h . which
vanishes since rh is symmetric and C is skew-symmetric. Consequently, (53.4) vanishes for any C RIndm
i which implies the result ZAEg rA x Fi
= 0. This result
together with the identity
d jh
dt

d
dt

AeQ

rA x p
MAIpB X PA] -

=[
ACg BEg

rA X Ft

(53.6)

AE9

which follows from (44.11) and the fact that MA is symmetric while PB x PA is
skew-symmetric, completes the proof of relation (53.3)2.
To describe a number of widely used global time-stepping algorithms it proves
convenient to adopt a conventional matrix notation and rewrite the finite-dimensional
dynamical system (44.11) in the standard form
Pt = Ftext - Fint(dt, rt)
ddt
dt = M-pt = vt
dt

for t EG,

(53.7)

where the global vector dt of nodal displacements is subject to the initial condition
(dt,pt)lt=o = (do,po). Here M is the mass matrix and t G 1I H- (dt,pt) E P

is the flow in the finite-dimensional phase space P = IRdo


x RLdof, with ndof =
.
3 x n.ode In addition, t E I
Fext denotes the vector containing the prescribed nodal
forces F At(t) and Fint (dt, t) is the vector containing the internal nodal forces Fjnat,
respectively defined by (44.9) and (44.10) for A c g.
The following time-stepping algorithms provide a representative sample of alternative global discretization schemes for the initial value problem (53.7) which are
widely used in large-scale simulations. Only one of this schemes exactly preserves
the conservation laws of total linear and total angular momentum inherited by the
Galerkin discretization. For a proof of the results quoted below see SIMO, TARNOW
and WONG [1992] and SIMO and TARNOW [1992].

The discrete initial boundary value problem

SECTION 53

413

(A) The classical Newmark family of algorithms. This time stepping algorithm is
defined by the semidiscrete momentum equation (53.7)1 enforced at t+l along with
the standard Newmark formulae; i.e.,
=

lMa+la

'next - Fmn(Tn+l,dn+ ),

dn+l = dn + Atvn + At 2 [(2-3)an + 3a+],

(53.8)

v,+ = v + At [(1 - y)an + yan+l].


Two noteworthy properties of this class of schemes are
(i) The only member of the classical Newmark family that preserves exactly conservation of total angular momentum (for equilibrated loading) is the explicit central
difference method obtained for /3 = 0 and y =
(ii) The Newmark-trapezoidalrule, corresponding to the values -y= and / = 1,
does not inherit the conservation property of total angular momentum.
Observe that the Newmark-trapezoidal rule defines an acceleration-dependent time
stepping algorithm. Furthermore, in the nonlinear regime, the trapezoidal rule does
not define an A-contractive algorithm (see the counterexample of WANNER [1976]).
Since the implementation of the preceding class of algorithms is fairly standard further
detailswill be omitted.
(B T e-stepping algorithms in conservation form. As an illustration of this class
of methods, consider the algorithm defined by the semidiscrete momentum equation
(53.7) enforced in conservation form at t+,g along with the standard Newmark formulae; i.e.,
M[vn+l-

v.]

= FX,

- Fint(rn+,9 d.+

dn+ = ds + At[(1 -/3/y)v


a.+l

[v.+

),

+ /3yvn+l] + At2 (' - /3/')an,

(53.9)

- vn]/'At + (1 - 1/?)an,

where d,+a: = dn,+l + (1 -)d, and Fext. = F'+ + ( - )Fxt. In the preceding
expressions r+o denotes the Kirchhoff stress field evaluated at the generalized midpoint configuration p,+ : B -- IR3, with nodal position vector dn,+. Noteworthy
properties of this algorithm are:
(i) Exact conservation of the total angular momentum (for equilibrated loading)
is achieved for the values
= l3/-y= 2 corresponding to the conservation form
of the mid-point rule. These values define an acceleration-independenttime-stepping
algorithm.
(ii) In general, formulae (53.9) define a one-leg multistep method. Second-order
accuracy is obtained if and only if = . For the parameter values = /-y = a
linear analysis shows that the spurious root at zero sampling frequency vanishes if and
only if y = 1. This value yields the post-processing formula a,+l
(v,+l - v,)/At.
In sharp contrast with the trapezoidal rule, the implicit mid-point rule does define
an A-contractive algorithm in the nonlinear regime. In addition to the property of

414

J.C. Simo

CHAPTER IV

exact momentum conservation, both the explicit central difference method and midpoint rule possess the strong property of defining symplectic schemes for nonlinear
elastodynamics.
(C) Algorithms incorporating high frequency dissipation. Extensions to the nonlinear regime of a number of algorithms exhibiting numerical dissipation in the high
frequency range often require the evaluation of the stress field at a generalized midpoint configuration. As an example, consider the algorithm defined by a modified
momentum equation and the standard Newmark formulae as
Man+ = Fext - Fint(Tn+, d+ ),
dn+l = dn + Atvn + At2 [(

- 3)a

+n aa,+I],

(53.10)

Vn+ l = Vn + At [(1 - y)an + 'yan+l1.


For linear elastodynamics, this scheme reduces to the 9i-method of HILBER, HUGHES
and TAYLOR [1977]. However, in the nonlinear regime the preceding scheme uses a
generalized mid-point rule evaluation of the stress divergence term in place of the
commonly used trapezoidal rule evaluation. The former preserves A-contractivity and
leads to nonlinearly stable schemes for infinitesimal plasticity (SIMO [1991]) whereas
the latter does not. Noteworthy properties of this algorithm are
(i) The values 9 = , = , y = 1 reduce the algorithm to (53.9) and yields the
acceleration-independent, exact momentum-preserving, mid-point rule in conservation
form. As pointed out above, this algorithm possesses zero spurious roots (compare
with HILBER [1976]).
(ii) The values 9 = 1, = 4 and y = give the acceleration-dependent Newmarktrapezoidal rule in (53.8). The acceleration field predicted by this algorithm is notoriously noisy.
For 0 E (, 1) the algorithm exhibits numerical dissipation in the high frequency
range but the property of exact conservation of total angular momentum no longer
holds.
54. Remarks on the implementation of return mapping algorithms
In the classical Newmark family of algorithms, the discrete version of the momentum
balance equation is enforced at the end of each time step. However, in the last two
global time-stepping algorithms outlined above, enforcement of the discrete version of
the momentum balance involves a single evaluation of the external force vector at an
intermediate time t+, a task easily accomplished via interpolation, as well as a single
evaluation of the internal nodal force vector Fint (rn+O, dn+o) also at time t,,+. This
at the generalized mid-point
evaluation requires knowledge of the stress field Tr,+
configuration. It is precisely for this class of conserving time-stepping algorithms that
the two-stage projected return mapping algorithms described in Section 51 are ideally
suited.

SECTION 54

The discrete initial boundary value problem

415

The reason lies in the observation that the stress field ,n+o is computed in Stage I
which is uncoupled from Stage II. The only purpose of the second stage is to compute
the stress field r,+1 at time t,+l once r,+o is determined. Consequently, the entire
iterative solution procedure of the discretized weak form of momentum balance can
be performed by repeated evaluation of Stage I. Stage II of the algorithm need be
performed only once per time step, once convergence of the iterative solution scheme
is attained, and can be viewed merely as a means of updating the initial data from
time t, to time t+l from the solution at time t,+,g. The nodal configuration and
velocity updates are trivially accomplished either via formulae (53-9)2,3 or through
formulae (53.10)2,3. Stage II thus defines the mapping
(b+,c+o}

e S+ X IR+

} E S+ x 1R+
x {be+,,~n+l

(54.1)

which provides the initial conditions for the internal variables in the return mapping
algorithm at the subsequent time step.
From the preceding observation it also follows that only the algorithmic moduli
associated with Stage I, with a closed form expression given by (49.7) for t, = t+o,
enter into the iterative solution of the algorithmic version of the projected momentum
equation (44.10). In fact, the entire iterative solution of the nonlinear algebraic system arising in each time step is independent of Stage II and remain unaffected even
if a scheme other than the one described above is adopted. From a computational
standpoint the preceding two-stage scheme is particularly attractive since the implementation of Stages I and II is essentially identical, the added computational cost of
Stage II is negligible since the update is purely local, while the overall accuracy of
the resulting scheme is increased to order two.
As in the infinitesimal theory, Stage II of the algorithm furnishes one among the
many possibilities for specifying the map (54.1). Two alternative definitions of this
map considered in SIMO [1992] are constructed as follows. Define the trial elastic left
Cauchy-Green tensor by the following formulae:
fn+l (+~)
::

I+ I

Vun+9(x+)],(
(54.2)

It follows that b+ 1 defined in this fashion is the push-forward of the converged


solution b+ to the current configuration S,+ 1 with the relative deformation gradient
f,+l between the configurations S,+0 and Sn+1. The final values {b',+, n+l} can
be defined by the following two alternative schemes that replace Stage II:
(i) Shifted backward Euler scheme (SBE). This update is defined merely by setting
{b+l), Gn+l} = {btr+l, n+q} It can be shown that the resulting scheme yields
a backward Euler return mapping algorithm but within time steps shifted by 9; i.e.,
[tn+,9 tn+l+], for n = 0, 1,....
(ii) Productformula algorithm (PFA). This update procedure defines the internal
variables {be,+l, n+l } via a return mapping algorithm identical to the one em-

416

J.C. Simo

CHAPTER IV

ployed in Stage I1 of the projected IRK method, but with the trial state now given by
{bet , >+o}, as defined by (54.2).
In the first of these two alternative strategies, labeled as (SBE), only the stress
field at the generalized mid-point configuration Sn,,+ remains in the elastic domain,
i.e, (+,g
q,, ) E E. In general, however, the stresses (n+l, qn+l) will not lie in
the elastic domain E. This feature is often viewed as a disadvantage from a practical
standpoint. In the second of these two strategies, labeled as (PFA), consistency is
enforced at the end of the time step via an additional return mapping algorithm exactly
as in Stage II of the projected IRK method. The only difference lies in the alternative
definition used for the trial state. For the linear theory, the accuracy analysis described
in Section 20 shows that definition (52.7) results in a second-order accurate scheme.
The same analysis also shows that the alternative definition (54.2) yields a scheme
which is only second-order accurate.
REMARK 54.1. At the time of writing this article, we have had no occasion to conduct extensive numerical experiments with the projected return mapping algorithm
described in Section 53. On the other hand, extensive numerical experiments conducted with the two alternative strategies outlined above, some of which are reported
below, support their excellent performance in spite of being only first-order accurate.
Preliminary numerical experiments also suggest that the second-order accurate projected scheme described in Section 53 does not appear to exhibit a performance nearly
as robust as these two, nominally less accurate, update schemes. For this reason, the
numerical simulations described below will be conducted only with the well-tested
alternative schemes (SBE) and (PFA).
55. Representative numerical simulations
The formulation presented in the preceding sections is illustrated below in a number of
full three-dimensional numerical simulations taken from SIMo [1992]. The goals are
to provide a practical accuracy assessment of the different return mapping algorithms
in actual calculations and to demonstrate the robustness of the overall finite element
formulation in both static and dynamic analyses. The calculations are performed with
an enhanced version of the finite element program FEAP developed by R.L. Taylor
and the author from the version documented in ZiENKIEWICZ and TAYLOR [1989].
(A) Three-dimensional necking f a circular bar. This simulation is the threedimensional version presented in SIMO and ARMERO [1992] of a well-known twodimensional problem considered by a number of authors. A bar possessing a circular
cross section with radius Ro = 6.413 and total length L = 53.334 is subjected to a
displacement-controlled pure tension test with simply supported boundary conditions
corresponding to perfectly lubricated end grips. For a perfect specimen these boundary conditions lead to a bifurcation problem from an initially homogeneous uniaxial
state of stress. The bifurcation problem is transformed into a limit point problem via a
geometric imperfection induced by a linear reduction in the radius along the length of
the bar to a maximum value Rsym = 0.982 x Ro at the cross section in the symmetry

SECTION 55

The discrete initial boundary value problem

417

plane. By obvious symmetry considerations only one eighth of the specimen is discretized in the analysis. The constitutive response of the material is characterized by
the model of J2 -flow theory described above, together with the Henky stored energy
function (48.11) and the von Mises yield criterion (50.1). The material properties are
= 164.21,

j/ = 80.1938

and

0y = 0.45.

(55.1)

Hardening in the material is characterized by an isotropic hardening function of the


saturation type, with functional form
K(,) := HJ + [y
where a-y >
ar

0y

ay] (1 - exp[-6]),

(55.2a)

> 0, H > 0 and 6 > 0 are material constants chosen here as

= 0.715,

6 = 16.93,

H = 0.12924.

(55.2b)

The specimen is subjected to a total increase in length of AL/L x 100 = 26.25%.


The finite element discretization used consists of three-dimensional Qi/PO mixed finite
elements, as described in SMO, TAYLOR and PISTER [1985]. This simulation is used
in SIMO [1992] as a practical accuracy assessment of a number of alternative return
mapping algorithms in different finite element meshes. The results presented there
(see Fig. 55.1 for a representative sample) support the remarkably good performance
exhibited by the exponential return mapping algorithms described in the preceding
sections. In particular, the lack of significant degradation in accuracy exhibited by these
algorithms suggests that the choice of time step should be dictated by the convergence
of the global solution scheme.
The simulations described above are performed with a global Newton iterative solution procedure together with a linear line-search algorithm which renders the scheme
globally convergent. Remarkably, successful computations are completed in only ten
load increments; a feature which demonstrates the robustness of the formulation. In
spite of the increase in the condition number of the Hessian with finer meshes, a
successful solution for the 960 finite element mesh can also be accomplished in 10
load steps scheme employing 6 BFGS updates, a line search with factor 0.6 and periodic refactorizations. Simulations with 20, 50 and 100 load steps were successfully
completed with only two refactorizations per load step.
(B) Three-dimensionaldynamic impact of a circularbar The goal of this simulation
is to illustrate the performance of the time-stepping algorithms for dynamic plasticity
described in the preceding section; in particular, the conservation form of the midpoint rule. These results are compared with with those obtained via the standard form
of the Newmark time-stepping algorithm. The model problem selected is the dynamic
impact on a rigid frictionless wall of a three-dimensional bar, with length L = 32.4 mm
and circular cross section with radius Ro = 3.2mm, presented in HALLQUIST and
BENSON [1987]. The initial velocity is v 0o = 0. 2 2 7 mm//t s. The material response is

J.C. Simo

418

CHAPTER IV

0
-0.5
Q)

-1.0
-1.5

g2

-2.0
-2.5
-3.0
-3.5
-4.0
0

Total Elongation (7 = nsteps

Al)

FIG. 55.1. Displacement at the section of extreme necking versus total elongation. Exponential return

mapping algorithm. 960 finite element mesh.

po$P

S TRESS15
< 2.133E-01

sTRESS 2

< -3,845E-01

STRESS 3
< 3775E-03

FI-. 55.2. Fine finite element mesh consisting of 960 QI/PO elements. Contours of the equivalent plastic
strain and stress distribution for the three-dimensional bar. Final solution obtained in 20 time steps with
the product formula (PF) algorithm.

419

The discrete initial boundary value problem

SECTION 55

STRESS 3
< -6.659E-01

Adinhhh,

I HL

< -4.238E-01

STRESS 3
< -3.21 E-01

FIG. 55.3. Three-dimensional impact of a circular bar. Small mesh consisting of 144 QI/P0 mixed finite
elements. Sequence of deformed shapes corresponding to times t = 20, t = 40 and t = 80, obtained with
the conservation form of the mid-point rule in 16 time steps.

again characterized by the same model of J2 -flow theory as in the preceding example,
with the following values of the material constants
K = 0.4444 GPa,

wy = ao = 0.40 GPa,

H = 0.1 GPa.

(55.3)

The density in the reference configuration is taken as p = 8930.00kg/m 3 .


The sequence of deformed configurations corresponding to times t = 20 s, t = 40 s
and t = 80 s, computed with the conservation form of the mid-point rule, is shown in
Fig. 55.3 (for the 144-finite element mesh) and Fig. 55.4 (for the 972-element mesh).
These simulations are successfully completed in 16 and 32 time steps, respectively.
Despite the extremely large time steps used in the simulation, the final shape of the
specimen at time t = 80 s agrees well with the results reported in HALLQUIST and
BENSON [1987]. The time histories of the maximum lateral displacement at the impact
section, computed with different time steps, are shown in Fig. 55.5 (144-element
mesh) and in the following figure for the 972-element mesh. The same simulations
were attempted for the same two meshes with the standard Newmark algorithm and
the one-step exponential backward Euler return map. Successful computations were
completed only when the time steps where reduced to 2.5 and 1.25, respectively. In

420

J.C. Simo

CHAPTER IV

STRESS 3
< -7.830E-01

FIG. 55.4. Three-dimensional impact of a circular bar. Large mesh consisting of 972 Ql/PO mixed finite

elements. Sequence of deformed shapes and axial stress distribution corresponding to times t = 20, t = 40
and t = 80, obtained with the conservation form of the mid-point rule in 32 time steps.

all these simulations a quadratic rate of asymptotic convergence was attained in a


Newton iterative solution strategy.
(C) Simple shear of a block: Pure kinematic hardening law. This final simulation is
concerned with the simple shear of a rectangular three-dimensional block, with constitutive response characterized by J2 -flow theory, in the ideal case of pure kinematic
hardening. As first noted in NAGTEGAAL and DE JONG [1981], this situation gives rise
to spurious oscillations in the stress-strain response when the kinematic hardening law
is described by the Jaumann derivative. The goal of this simulation is to assess the
response exhibited by the extension to the finite strain regime of the classical PragerZiegler kinematic hardening law described above. The finite element mesh consists
of 27 three-dimensional Ql/PO elements as shown in Fig. 55.8. The elastic properties
are the same as in Example 47.1 while the model of J2-flow is now characterized
by a flow stress y = 0.45 and a kinematic hardening modulus H = 0.1. The final
configuration is also shown in Fig. 55.8 and the plot of the Kirchhoff shear stress rl2
versus the amount of shear is recorded in the last figure. It is apparent from these
results that the stress response is monotonically increasing; in fact, essentially linear,
and exhibits no spurious oscillations in the entire range of deformations.

421

The discrete initial boundary value problem

SECTION 55

5
q2

4
3

a)

2
1
0
0

20

40

60

80

Total Simulation Time t


FIG. 55.5. Impact of a circular bar: 144-element mesh (Ql/PO elements). Time history of the maximum

lateral displacement at the impact section for different time steps, computed with two schemes: Newmarktrapezoidal rule with exponential backward Euler (NT/BE) and mid-point with exponential product formula
(MP/PF).

5
a)
Pi

a
3
._l
,--q

.2

,-C~
X

-4
Cd

1
0
0

20

40

60

80

Total Simulation Time t


FIG. 55.6. Impact of a circular bar: 972-element mesh (Ql/PO elements). Time history of the maximum
lateral displacement at the impact section for different time steps, computed with two schemes: Newmarktrapezoidal rule with exponential backward Euler (NT/BE) and mid-point with exponential product formula
(MP/PF).

422

J.C. Simo

CHAPTER IV

K1

DISPLACEMENT 2
< 4.286E+00

FIG. 55.7. Simple shear of a three-dimensional block. J2-flow theory with pure kinematic hardening. Initial
mesh and final meshes consisting of 27 Q1/PO mixed finite elements.
_

a~ An
U.4U

0.38

U)

0.36

0.34

0.32

0.30
-

0.28

0.26
0

0.5

1.0

1.5

2.0

2.5

3.0

Amount of Shear Strain


FIG. 55.8. Simple shear of a three-dimensional block. J2-flow theory with pure kinematic hardening. Plot
of the Kirchhoff shear stress versus the amount of shear deformation.

(D) Plane strain localization problem. This example is concerned with the plane
strain tensile test of a rectangular bar, and has been studied by a number of authors;
in particular, TVERGAARD, NEEDLEMAN and Lo [1981]. The goal is to demonstrate
the inability of conventional mixed methods to accurately resolve the sharp gradients
arising in localization problems. This unsatisfactory performance is contrasted with a
recently proposed technique (SIMO and ARMERO [1992]) ideally suited for this class
of problems. A strong localization in the specimen is induced by considering a steep

The discrete initial boundary value problem

SECTION 55

423

,,

U.Y

0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1

Previous law (satur.)


,

0.5

, I

1.0

Equivalent plastic strain

----.

1.5

2.0

FIG. 55.9. Plane strain localization problem. Steep softening law, defined by a double-parabola, used in the
simulation.

softening law, the double-parabola depicted in Fig. 55.9. Steep softening laws of this
type are often used in the numerical simulation of localization problems.
The specimen considered in this study possesses a width of 12.826 and a length
of 53.334 and is subject to ideal plain strain loading conditions; hence, the third
dimension is assumed to be infinite. To trigger the localization, an initial geometric
imperfection is considered in the form of a linear reduction of the width from its initial
value at the top to 0.982% of this value at the center of the specimen. A quarter of the
specimen is discretized imposing symmetry boundary conditions along the boundaries.
The finite element mesh consists of 200 quadrilateral elements and the simulation is
performed under displacement control of the top surface.
Figure 55.10 shows the load/displacement curves obtained for the double-parabola
hardening/softening law with a conventional mixed method (Q1/PO) and the recently
proposed enhanced-assumed-strain method (SIMO and ARMERO [1992]) labeled as
Q1/E4. The conventional Q1/P0 mixed method leads to an overly diffuse response
that provides a very poor resolution of the sharp strain localization in the specimen. In
sharp contrast with these results, the load displacement curve computed with Q1/E4
exhibits the sudden drop which is typically found in localization problems of this
type. These results are confirmed by the deformation patterns shown in Fig. 55.11(a).
The Q1/E4 element method exhibits a very well defined line of discontinuity in the
specimen at 450 with the load direction while the Q1/P0 element shows a very diffuse
localization pattern. Figure 55.11(b) depicts the distribution of the equivalent plastic
strain for both formulations at the same imposed top displacement. Once more, the

424

J.C. Simo

CHAPTER IV

1u

m0

Top displacement
FIG. 55.10. Plane strain localization problem with double-parabola hardening/softening law. Computed
load/displacement curve.

clear shear band exhibited by the Q1/E4 formulation should be contrasted with the
diffuse shear pattern exhibited by the mixed QI/PO element.
(E) Sheet metal forming. This three-dimensional example, very similar to one considered in REBELO, NAGTEGAAL and HIBBITT [1990], is taken from LAURSEN and
SIMO [1993] and proved an illustration of an industrial application of the preceding
theory. A flat elastoplastic sheet is formed into a pan, by forcing it to conform to the
shape of a rigid die pressed against it (see Fig. 55.12). A J 2-model of multiplicative
plasticity is assumed, with material properties E = 70 GPa, v = 0.3, ay = 140 MPa
and isotropic hardening with hardening modulus H = 100 MPa. A Coulomb friction
model for the contact interface between the die and the sheet is assumed with coefficient of friction of 0.25. A detailed account of the algorithm issues involved in
the treatment of the contact interface is given in LAURSEN and SIMO [19931. This
reference should be consulted for additional information on this aspect of the simulation. The sheet initially measured 600 mm long, 560 mm wide, and 5 mm thick. As
may be noted in Fig. 55.12, the die consists of a lower flat region with an inclined
section leading to it; the simulation was continued until the lower region had been
moved through a distance of 100 mm. For symmetry reasons only half the geometry
was modeled, with 800 continuum elements being utilized for the discretization of the
sheet. The loading was achieved in 100 load steps, through displacement control of
the die (the table was fixed).
The mixed finite element method described in Section 45 is employed in this simulation, with pice-wise constant interpolations for both pressure and volume along with

425

The discrete initial boundary value problem

SECTION 55

Q1/PO

Q1/E4

II
I HHH
IIIF
I

0.15

F HI]H t

0.60

Hill
IIH

I:
-11
F
b

F
F

i
I4:I.:4

.F-h4

;:;i i [ 5

FIG. 55.11. Plane strain localization problem (double-parabola softening law). Results at an imposed top
displacement of ui = 4.24: (a) Deformed configuration. (b) Distribution of the equivalent plastic strain.

426

J.C. Simo

CHAPTER IV

FIG. 55.12. Initial geometry for the pan forming problem.

FIG. 55.13. Deformed configurations for the pan forming problem.

a bilinear interpolation for the displacement field. This choice of continuum elements
is not optimal for the problem at hand, particularly with regard to capturing bending behavior. It nonetheless demonstrates the capability of the model of plasticity to
handle large deformations of the type typically encountered in metal forming applications. The outer edges of the plate were considered to be fixed, but even in light
of this constraint some redistribution of material occurs between plate and table and
(obviously) between plate and die. Figure 55.13 shows deformed configurations for
the plate at four equally spaced intervals during the calculation.
One notices immediately that the most severe straining of the material occurs at the
corners of the deepest part of the forming. This assertion is borne out by Fig. 55.14,
which depicts a contour of the effective plastic strains at the final state. The computed
strain in this corner (over 100 percent) is partly an artifact of the coarse zoning in
this region, but also indicates the likelihood of tearing of the pan in this corner for
the geometry considered.
(F) Post buckling of a cylinder This two-dimensional problem, taken from LAURSEN
and SIMO [1993], combines the phenomena of inelastic buckling, frictional contact,
and post buckling response. This reference should be consulted for a detailed account of the algorithmic treatment of the contact problem not discussed here. We

SECTION 55

427

The discrete initial boundary value problem


Plastic Strain
< 2.304E-01
I

> 1.382E+00
FIG. 55.14. Contour of effective plastic strains for the pan forming problem, at the final state.

remark that problems of this type are encountered in crash-worthiness research, where
structural elements may buckle and subsequently crumple, creating regions of self
contact. Figure 55.15 depicts the physical situation considered. In this axisymmetric problem, a steel cylinder (inner diameter = 27 mm, outer diameter = 31.75 mm,
length = 180 mm) is forced downward (via displacement control of its top surface) into
a rigid fixture. The J2 -flow theory described in the preceding sections is employed,
with material properties defined as K = 175 GPa, G = 80.8 GPa, y = 700 MPa, and
H (isotropic hardening modulus) = G/100. The spatial discretization of the cylinder
employs the same mixed finite element methods used in the preceding example, with
finite element mesh consisting of a total of 177 elements. As the calculation proceeds, the cylinder initially moves down into the die, but when a critical axial load is
reached, it begins a series of buckles, as is shown in Fig. 55.15. Importantly, this buckling occurs without any initial geometric imperfections. As indicated in Fig. 55.15,
calculations are performed both with and without friction between the cylinder and
die, with frictionless response assumed in the self-contacting buckle regions.
Initial examination of Fig. 55.16 may not reveal significant differences between the
frictionless and frictional responses. Careful review of the figure will show, however,
that in the frictionless case the cylinder moves a little farther into the fixture before
buckling is induced. This result is as one would expect, since in the frictional case
the axial force builds more rapidly and buckling is induced sooner. This assertion is
substantiated by Fig. 55.17, which shows the total axial load induced in the cylinder
as a function of the top displacement. As expected, the frictional fixture induces the
buckling (indicated by the drop in load) sooner than does the frictionless fixture.
Interestingly, this offset in the curves due to the friction only persists through the first
buckle cycle, after which time the curves rejoin each other for the last two buckle
cycles. This suggests that once the initial buckling has been induced and completed
it is the conditions existing above the first buckle which dominate the response (and
thus the presence of friction below the first buckle becomes of secondary importance).

428

J.C. Simo

DIlt

CHAPTER TV

C-1

FIG. 55.15. Initial and final configurations for the post buckling cylinder problem, shown for the case of
frictionless response between cylinder and fixture.

Also apparent in Fig. 55.17 is the contribution of the self-contact occurring between
buckles. It is this contact, manifested in the figure as a reversal of load at the bottom
of each buckle cycle, which provides the stiffening mechanism necessary to trigger
the next buckle. The regions in which self-contact occurs are also readily apparent in

Figs. 55.15 and 55.16. Finally, it is important to emphasize that the deformation induces large plastic strains in the cylinder, rendering the problem extremely nonlinear.
Figure 55.18 shows the plastic strains at the final state for the frictionless case. As can
be seen, virtually the entire specimen is yielded, with particularly large strains (over

The discrete initial boundary value problem

SECTION 55

ImL

429

:Ulil~~~~~~~~~~~ci~~1
(a)

rVLI

L
LJ JL1

liii

LALLJT

LLj L

LL

(b)
FIG. 55.16. Buckling sequence in the steel cylinder problem, shown for both the frictionless (a) and frictional
(b) cases.

100%) occurring in the buckles. The fact that this problem was executed in a moderate number of load steps (approximately 260 for the frictionless case and 350 for the
frictional case) even in the presence of the complicated phenomena present demonstrates the efficiency and utility of the algorithmic treatment of both the multiplicative
plasticity and the contact techniques.
(G) Cylinder impact on deformable rails. In this final example, also taken from
LAURSEN and SIMo [1993], a steel cylinder (ID = 9cm, OD = 10cm, length =

80 cm) is dropped, with initial velocity 80 m/s, onto a pair of steel rails (5 cm deep,
4.67 cm wide, and 80 cm long), as shown in Fig. 55.19 The rails are parallel to each
other and 58.67 cm apart. Because of the symmetry of the problem only one-quarter of
the geometry was modeled. Both the cylinder and the rails were given the properties
K = 175 GPa, G = 80.77 GPa, ay = 700 MPa, H = 400 MPa, and p = 7850 kg/m 3 .

Roller boundary conditions were assumed on the bottom of the rails, and the ends

CHAPTER IV

J.C. Simo

430
250

200

z
Z

150

'

100

50

fi

20

40

60

80

100

120

Top Surface Displacement (m)


FIG. 55.17. Computed load-displacement curve for the steel cylinder post buckling problem.

Plastic Strain
< 1747E-01
_I_..........

> 1.048E-00

FIG. 55.18. Computed plastic strains at the final state for the steel cylinder post buckling problem (frictionless
case).

SECTION 55

431

The discrete initial boundary value problem

FIG. 55.19. Initial geometry for the cylinder-rail impact problem.


Eff. Plas. Strain
< 2.259E-02

1_

> 1.356E-01

FIG. 55.20. Contour of plasticity in the cylinder at time t = 0.625 ms.

of each were fixed. Coulomb friction, with


= 0.1, was assumed to prevail in the
contact between cylinder and rails. A conventional Newmark time integration (with
y = 0.9 and = 0.49) was used to integrate the global equations.
Figure 55.21 shows deformed configurations at various stages of the calculation,
which was achieved in 100 time steps. By the final state shown (2.5 ms), the cylinder
has obviously rebounded. The contact region, which is quite localized in this problem,
is manifested by large dents in the cylinder, as well as small yielded regions near

432

J.C. Simo

CHAPTER IV

(a)

(b)

(c)

(d)

(e)

FIG. 55.21. Deformed configurations for the cylinder-rail impact problem at (a) t = O.Oms, (b) t =
0.625 ms, (c) t = 1.25 ms, (d) t = 1.875 ms, and (e) t = 2.5 ms.

the contact points on the rails. Figure 55.20 shows the contours of plastic strain
in the cylinder at the time t = 0.625 ms. As the figure indicates, this is yet another
example where significant plasticity occurs. Since this problem also involves dynamics
and the frictional contact between deformable bodies, it provides a particularly good
illustration of the applicability of the proposed techniques to a wide range of problems.

CHAPTER V

The Coupled
Thermomechanical Problem:
Product Formula Algorithms
A significant portion of the energy dissipated during plastic deformation in metals is transformed into heat. Heat conduction in the solid affects, in turn, characteristic properties in the solid that govern plastic flow, such as the flow stress. The
modeling and simulation of this combined process, which accounts for the coupling
effects between mechanical deformation and heat conduction in a solid undergoing
large plastic deformations in the cold-working regime, leads to a coupled thermomechanical problem. The numerical analysis of this problem constitutes the central
topic of this chapter. The exposition below emphasizes the formulation of the coupled
initial boundary value problem and, in particular, the description of numerical techniques currently used in scientific computing and large-scale simulation of coupled
problems. No attempt is made to provide a comprehensive treatment of the physical
mechanisms involved in the deformation of metals in the cold work range. Background material on the physical foundations of the subject, initiated in the fundamental work of G.I. Taylor and his associates (see, e.g., TAYLOR and QUINNEY [1933]),
can be found in the books of REED [1964], COTTRELL [1967] and the monograph
of BEVER, HOLT and TITCHENER [1973]. Expositions of thermoplasticity that emphasize the connection between continuum theories, underlying physical mechanisms and
available experimental results are given in DILLON [1963], PHILLIPS [1974], ZDEBEL
and LEHMANN [1987] and references therein. An outline of the topics covered below
is as follows.
In the first part of this chapter, a brief account is given of the basic balance laws
governing the coupled thermomechanical problem, together with an extension of the
multiplicative model of plasticity described in Chapter III that incorporates thermal
effects. This completes the formulation of the coupled initial boundary value problem. Next, a formal a priori stability estimate for this initial boundary value problem
introduced in ARMERO and SIMO [1992] is described which generalizes to coupled
nonlinear thermoplasticity the canonicalfree energy of DUHEM [1911]. This functional was shown in ERICKSEN [1966] to define a Lyapunov function for coupled
433

434

J.C. Simo

CHAPTER V

thermoelasticity and plays a central role in the mathematical treatment of the quasistatic thermoelastic problem (BALL and KNOWLES [1986]). For thermoplasticity, the
canonical free energy defines a functional of the dynamics which does not increase
along the flow. This property provides a natural notion of nonlinear numerical stability
which does not preclude interesting physical phenomena such as formation of shear
bands in the presence of thermoplastic softening. Alternative notions of nonlinear stability, B-stability in particular, are tailored to contractive problems of evolution and,
therefore, not applicable to the problem at hand.
The second part of this chapter is devoted to numerical analysis aspects involved in
the simulation of the thermoplastic, coupled, initial boundary value problem. Coupled
thermomechanical problems typically involve different time scales associated with
the thermal and mechanical fields. It is widely accepted that an effective numerical
integration scheme for the full coupled thermomechanical problem should take advantage of these different time scales. It is considerations of this type that motivate
the so-called staggered algorithms, whereby the problem is partitioned into several
smaller sub-problems which are solved sequentially. This technique is specially attractive from a computational standpoint since the large and generally nonsymmetric
system that results from a simultaneous solution scheme is replaced by much smaller,
typically symmetric, sub-systems. A sound numerical analysis is possible by interpreting staggered schemes as product formula algorithms arising from an operator
split of the coupled problem, in the spirit of classical fractional step methods (see,
e.g., YANENKO [1971]). Within this framework, the following two specific staggered
schemes are examined.
First, the conventional approach which, as noted in SIMO and MIEHE [1992], arises
from a formal split of the problem into a mechanical phase with the temperature held
constant, followed by a thermal phase at fixed configuration. Staggered schemes of this
type are widely used in practice and go back to the work of ARGYRIS and DOLTSlNIS
[1981] and others, see, e.g., LEMONDS and NEEDLEMAN [1986] and WRIGGERS, MIEHE,
KLEIBER and SIMO [1992] along with the review articles of PARK and FELIPPA [1983]
and DOLTSNIS [1990] for an overview. The well-known restriction to conditional
stability is the crucial limitation of this approach, which often becomes critical for
strongly coupled problems. Stabilization techniques designed to remove the restriction
to conditional stability have been devised by a number of authors with mixed success
and, typically, within the context of the linearized problem. The early augmentation
schemes of PARK, FELIPPA and DERUNTZ [1977], the iterative scheme of ARGYRIS
and DOLTSINIS [1980] and the more recent augmentation strategy of FARHAT, PARK

and DUBOIS-PELERIN [1991] furnish sample representative approaches.


Second, a staggered scheme was recently proposed in ARMERO and SIMO [1992,
1993], which completely circumvents the stability restrictions inherent to the conventional approach while retaining its computational approaches. The idea is to partition the coupled problem into an adiabatic mechanical phase, in which the total entropy of the system is held constant, followed by a thermal phase at fixed
configuration. The replacement of the isothermal phase in the conventional staggered scheme with an adiabatic phase results in an operator split which, remarkably, preserves the a priori stability estimate for the coupled problem. An uncon-

The coupled thermomechanicalproblem

SECTION 56

435

ditionally stable staggered algorithm is then constructed as the product formula of


two unconditionally stable fractional steps. The chapter concludes with an account
of several representatives of numerical simulations of thermoplasticity. The numerical analysis results described in this chapter rely on joint work of this author with
F. Armero.
56. Integral, local and weak forms of the general conservation laws
The first step towards the statement of the initial boundary value problem for coupled thermoplasticity is the formulation of the balance laws of momentum, energy
and entropy for a continuum body undergoing finite strains, in the presence of heat
conduction. The statement of these basic laws of mechanics is standard, see, e.g.,
TRUESDELL and TOUPIN [1960] for a complete account including the effects of jump
discontinuities. This section provides a summary of these classical results starting with
the integral form of the balance laws and concluding with their weak formulation,
which is the form ideally suited for numerical approximations employing the finite
element method. The background material and the notation on nonlinear continuum
mechanics introduced in Chapter IV is assumed throughout this exposition.
The simplest statement of the conservation laws is obtained by adopting a Lagrangian description of the motion t E I -- cpt C of the continuum body in terms
of the nominal stress tensor P, the material velocity field V = Cbt and the reference
density po. Recall that B denotes the body force per unit of reference volume and
T = Pv is the nominal traction vector field specified on the part of the boundary
FT C a.
(A) Integral and local form of the balance laws. Consider an arbitrary open set
13 c Q in the reference placement of a continuum body that undergoes a motion pot.
In terms of the preceding notations, the integral form of the balance law of linear
momentum takes the form
dt

/poVd

J= BdQ+

Pm dr,

(56.1a)

while the integral form of the law of balance of angular momentum requires

-t

1Jt x poVd

gJt x Bd2++J

Pvod.

(56.lb)

Here vo:
- S 2 is the outward unit normal to the boundary aB of the open
set B, which is assumed to be smooth. Equations (56.1a,b) furnish the statement for
a continuum body of Euler's laws of motion. In a thermodynamical context, these
classical laws are supplement by the balance of energy equation and the entropy
production inequality.
Let E denote the internal energy in the solid per unit of reference volume, Q the
nominal heat flux and R the heat source per unit of reference volume. The integral

CHAPTER V

J.C. Simo

436

form of the balance of energy in an arbitrary open set B C 2Qis the statement that
[ppoIV 2 + E] d Q

d/
dt

=Ij[B. V+ R]d+

[V Pvo-Q-vo]dF,

(56.2)

while the integral form of the Clausius-Duhem inequality requires that the net entropy
production in B be nonnegative in time. The entropy production per unit of volume,
denoted by a, is an extensive quantity defined in terms of the time rate of change of
the (nonequilibrium) entropy in the solid, denoted by H, and the absolute temperature
O > 0 by the following expression:
d
fs

=-

dt B

Hd

R/
B

Q v/ /Od

> 0.

(56.3)

los

Within the scope of rational thermodynamics, it is assumed that the internal energy E,
the entropy function H, the nominal stress P and the nominal heat flux Q are constitutive functions defined in terms of the deformation o and the absolute temperature
O. Inequality (56.3) is then viewed as a restriction to be satisfied by these constitutive
functions for prescribed body force B and specified heat source R in 2. This point
of view is advocated in COLEMAN and NOLL [1963], COLEMAN and GURTIN [1967]
and is also adopted in this exposition.
Assume next that the Lagrangian fields involved in the statement of the preceding
balance laws are smooth. A standard result known as the localization theorem, see,
e.g., TRUESDELL and TouPIN [1960] or GURTIN [1981], then yields the following local
form of the balance of linear momentum equation (56.1a) introduced in Chapter III:
po V = DIV[P] + B

in

x .

(56.4)

The local form of the balance of angular momentum equation (56.1b) results in the
symmetry condition PFT = FPT on the nominal stress where, as usual, F = Dt
denotes the deformation gradient. The local balance law (56.4) together with the
localization theorem leads to the following local form of the energy balance equation
(56.2):
E

= P

GRAD[V] +- R - DIV[Q]

in 2 x 11,

(56.5)

which is known as the reduced energy equation. Finally, the localization theorem
applied to the entropy production inequality (56.3) yields the result
eO

= OH - R + DIV[Q]h5 GRAD[9] . Q >

in

(56.6)

known as the Clausius-Duhem form of the second law. It is conventional to recast


inequality (56.6) in the following alternative form which can be found in TRUESDELL

The coupled thermomechanicalproblem

SECTION 56

437

and NOLL [1965]. Define the total dissipation D, the dissipation arising from heat
conduction D,,, and the internal dissipation Dint in the solid by the expressions
D = OE,

Dcon
=--

GRAD[O]

and

D= Dint +Dcon.

(56.7)

Introducing the additional hypothesis that the constitutive equation for the nominal
heat flux Q is such that D,,,on 0 in F2 x , the Clausius-Duhem inequality D > 0 in
$2 x implies (but is not implied by) the reduced inequality
Dint = (9/H- R + DIV[Q] > 0

in Q x II,

(56.8)

often referred to as the Clausius-Plankform of the second law.


REMARK 56.1. The assumption that D,,on > 0 leading to inequality (56.8) is quite
reasonable and always satisfied for models of heat conduction such as Fourier's law.
According to this classical model, the Kirchhoff heat flux q = F-TQ is proportional
to minus the spatial gradient V[9] = F-TGRAD[19] of the temperature field. In other
words, the following constitutive equation is assumed to hold:
q = -k(9)V[]

Q = -k(O)C-GRAD[9],

(56.9)

where k(O) is the conductivity and C = FTF is the right Cauchy-Green tensor.
Assuming that the conductivity k() > 0, this model of isotropic heat conduction
clearly satisfies the condition q V[(9]
0 or, since C is positive definite, the
equivalent condition Q GRAD[9] < 0 which implies Dcon > 0. The conductivity
k(9) is replaced by a positive definite, symmetric, conductivity tensor k(O) in the
conventional extension of the classical Fourier's law to anisotropic heat conduction.
(B) Local governing equations for the initial boundary value problem. To summarize the preceding discussion, the local momentum equation (56.4) together with the
reduced dissipation inequality (56.8) provide the two basic balance laws used in the
statement of the initial boundary value problem, i.e.,
poV = DIV[P] + B

in Q x ]l.

(56.10)

/H = -DIV[Q] + R + Dint
In this equations, the motion pt and the absolute temperature field 9 are regarded as
the primary variables in the problem while the body force B and the heat source R
are prescribed data and the nominal heat flux Q is defined via constitutive equations,
say Fourier's law, subject to the restriction Dcon > O.In addition, both the entropy H
as well as the nominal stress P are also defined via constitutive relations, typically
formulated in terms of the internal energy E, as described below. These constitutive

438

J.C. Simo

CHAPTER V

equations are subject to the following restriction on the internal dissipation Dint in the
solid:
Dint = O9H + P

GRAD[V] -E > 0

(56.11)

x 1[,

in

which is obtained by combining the energy equation (56.5) with inequality (56.8).
Equation (56.11) asserts that the internal dissipation Dint appearing in (56.10)2 equals
the thermal power 9H plus the mechanicalpower P GRAD[V] minus the time rate of
change of the internal energy E, and must be nonnegative. Effectively, this equation
provides a constitutive relation for D)int > 0 in terms of the constitutive equations
for P, H and E. Local balance of angular momentum is automatically enforced by
considering constitutive equations that satisfy the symmetry condition PFT = FPT.
The basic governing equations (56.10) and the constitutive restriction (56.11) are
supplemented by the standard boundary conditions for the mechanical field

Sot = t on , x[

and

Pv 0 = T

on T x ,

(56.12a)

together with the analogous essential and natural boundary conditions for the thermal
field, namely,
6=0

onFe x I

and

(56.12b)

Qvo=-Q on Q x [.

Here (6 and Q are specified functions on Fr C as2 and Q C aQ, respectively. In


addition to the usual restriction placed on Fr, and and T, described in Chapter III,
we also require that Fo n r = 0 and Fe U rQ = Oa2. Finally, we assume that the
following initial data is specified for the mechanical and thermal fields
(PtIt=o =

so0,
ott=o = Vo and Olt=o = 9o in Q.

(56.13)

The preceding relations, together with the constitutive equations introduced in the
next section, comprise the local formulation of the initial boundary value problem for
coupled thermoplasticity.
(C) Weak formulation of the governing equations. As repeatedly pointed out in this
monograph, the weak form of the governing equations provides the most convenient setting for the design of numerical approximations to the initial boundary value
problem. For the mechanical field, the weak form of Eq. (56.10)1 has already been
described in Chapter III and is summarized below for the convenience of the reader.
Assuming for simplicity time-independent essential boundary conditions, the admissible deformations comprise the infinite-dimensional configuration manifold denoted
here by
Cmech = J{g

E W'P()ndim: det[DSo] > 0 in

2 and

lu =

6c.}.
(56.14)

The coupled thermomechanical problem

SECTION 57

439

Associated with the configuration manifold, we have the tangent space of material test
functions defined by
V = {r0 C WI'P(Q)nd

: r0o = 0 on

rF

(56.15)

Recall that for fixed time t E Il the material velocity field V(., t) is a vector field on
Q that lies in Vo. Similarly, the admissible temperature fields comprise the infinitedimensional manifold
Cther =

{O

Wlq(SQ)ndi:

> 0 in 2 and

lr, = 9},

(56.16)

while the admissible temperature test functions are contained in the associated linear
space
i im
Go= Oon ro}.
0 = {o0 e Wl'q() nd:

(56.17)

As pointed out in Chapter III, the choice of exponents p and q in the preceding
definitions are dictated by growth conditions on the response functions. These technical
considerations, however, will play no role in our subsequent developments. With these
definitions in hand, the weak form of the local governing equations (56.10) take the
following form:
(poV, 'r0)+ (P, GRAD[ro]) = (B, 70 ) + ( T, 76.)r1-,

(O,

) - (Q,

RAD[])
] = (R,
(
) + ) , Go

(56.18)

(
(o,o)pr,

for all material test functions r0o E Vo and 0 E To. In the preceding statement,
the symbol (., ) denotes the standard of the L 2(Q2)-inner product, whereas (., )r
and (., *)r stand for the inner product of vector fields and functions defined on FT
and FQ, respectively. The weak form of the equations can be derived directly from
the integral form of the balance laws (see ANTMAN and OSBORNE [1979]) and, as
this latter statement, is considerably more general than the local form of the balance
laws. We remark that in the weak formulation of the problem defined by (56.18), the
restriction on the constitutive equations placed by the second law is still enforced via
the local inequality (56.11). The role played by this local condition becomes apparent
in the developments described in the next section.
57. Constitutive equations for multiplicative thermoplasticity
The goal of this section is to generalize the model of multiplicative plasticity described
in Chapter III to incorporate thermomechanical (coupling) effects. The methodology
employed in this extension is patterned after the construction presented in Section 35
and Section 36 within the context of the purely mechanical theory. In short, motivated
by underlying physics of the problem, we consider a specific functional form for the

440

J.C. Simo

CHAPTER V

internal energy function E and exploit the dissipation inequality to arrive at constitutive equations for the fields involved in the coupled problem. The model is completed
by introducing evolution equations for the internal variables that describe plastic flow
in the solid.
In classical thermodynamics, it is conventional to regard the internal energy E as a
function of the entropy H in the system. For crystalline plasticity, two basic physical
mechanisms contribute to the total entropy in the solid, see, e.g., COTTRELL [1967]
and BEVER, HOLT and TITCHENER [1973]. The entropy associated with the lattice is
herein denoted by H e, and the configurational entropy associated with the motion of
dislocations and defects through the crystal lattice is denoted by HP . Noting that the
entropy of a system is an extensive quantity, the total entropy in the solid is written
as H = H e + HP. This decomposition, along with the multiplicative factorization of
the deformation gradient motivated by the micromechanics of single-crystal plasticity
(see Section 34), yield the two basic relations
F= FeFP and

(57.1)

H = He + HP

Consistent with the physical interpretation assigned to the part H e of the total entropy,
the internal energy associated with the lattice is assumed to depend both on the lattice
distortion, characterized by F e , and the lattice entropy H e . The frame-invariance
argument used in Section 35 then yields an internal energy function of the form
E(Ce, He,f ) = E(Ce, He) + (

),

(57.2)

where we have used the notation C e = FeTFe introduced in Section 34. This functional form for the internal energy in the elastoplastic solid is motivated by micromechanical considerations (justifiable only for crystalline plasticity) and will be regarded
as the central constitutive assumption on the basis of which phenomenological constitutive relations are developed. We remark that, in a phenomenological context,
the assumption implicit in expression (57.2) that thermal effects are uncoupled from
hardening mechanisms is usually made in formulations of metal plasticity, see, e.g.,
GREEN and NAGHDI [1966], GERMAIN, NGUYEN and SUQUET [1983] and LEHMANN
[1982]. From a micromechanical point of view, such an assumption is motivated by
the observation that the lattice structure, which determines the thermoelastic response,
remains essentially unaffected by the plastic deformation.
(A)
the
for
the

Constitutive equationsfor stress, temperatureand internal variables. To identify


restrictions placed by the dissipation inequality (56.11) on the assumed form (57.2)
the internal energy in the solid, we compute the time rate of change of E using
identity
Ce = sym[FTGRAD[V]FP-I - FeTIPFe],

(57.3)

The coupled thermomechanicalproblem

SECTION 57

441

also derived in Section 34, with 1p = FeLPFe-l and LP = FPF p - 1. The time rate

of change of the internal energy then becomes


nint

[Fe(2ace,)F

P-

T] . GRAD[V] + [agHE]H

[a,,,.

(57.4)

Inserting this expression into the constitutive restriction (56.11), a standard argument
in irreversible thermodynamics yields the following potential relations:
TP = Fe(2aceE)FP
,

9 = aHeE and

qa = aW.

(57.5)

Since the second Piola-Kirchhoff stress tensor S relative to the intermediate configuration, the nominal stress tensor P and the Kirchhoff stress tensor r are related via
the transformation rules
P = FeSFP T

FT
F=

(57.6a)

the constitutive equation (57.5)1 for the nominal stress tensor P is equivalent to the
stress-strain relations
S= 2 cE

and

r = Fe(2aceE)FT .

(57.6b)

Expression (57.4), together with the preceding constitutive equations and the decomposition (57.1)2, also yields the following reduced form for the dissipation inequality
(56.11):
nint

Dint = T dP + 9HP +

5q
a=1

q > 0

in

x [,

(57.7)

where dP = sym[lP]. Using the kinematic relations derived in Section 34, the plastic
work term in (57.7) can be written in a number of equivalent forms depending of
which stress measured is preferred, in particular, r dP = S DP. This equation
defines the internal dissipation Dint in the solid once the rate equations that specify
the evolution of dP and ai in terms of r and qa are given.
REMARK 57.1. As indicated in the preceding section, the absolute temperature field
(9 along with the motion pt are the primary variables in the formulation of the initial
boundary value problem for thermoplasticity. Accordingly, it is customary to cast
the constitutive equations in terms of an alternative thermodynamical potential, the
free energy function T(F, 9, ~), obtained by performing the change of variables
He
(9 in the internal energy function for the solid via the Legendre transformation
T1(Co, e,

~,, = E(Fe, He, ) -

HOO.

(57.8)

442

J.C. Simo

CHAPTER V

Setting f(Ce, O) = E(C e H e ) - He H, the uncoupled form (57.2) for the internal
energy along with (57.8) results in the following functional form for the free energy
function in the solid:
T( eI,

,) (

, +T(c9
z,

(57.9)

o)
0)

It is readily seen that the preceding constitutive equations in terms of the internal
energy can be written in the equivalent form
He = -a(

7 = Fe[2IC,(Ce,)]FeT and

(e,).

(57.10)

As before, the corresponding stress-strain relations for either the nominal stress P or
the Piola-Kirchhoff stress tensor S follow immediately from the transformation rules
(57.6a). In the subsequent developments, the free energy function will be regarded as
the basic thermodynamic potential.
(B) The evolution of the total entropy. For the uncoupled free energy function (57.9),
the local evolution equation for the total entropy entropy H in the solid is computed
as follows. Time differentiation of the constitutive relation He = _-ao (C e, 9) and
use of the additive decomposition (57.1)2 yields, after changing the order of partial
differentiation,
OH = OH + coq-

(57. 11)

7- heat,

where c is the specific heat at constant deformation (and internal variables) per unit of
the reference volume and h eat is the structuralelastic heating, respectively, defined
by the relations
c= -O2e0 2o9(Oe,e)
'Heat = -oa

and

[2a=fi(Oe,H)]

ce.

(57.12)

Because of the form (57.9) adopted for the free energy function in the solid, the
thermoelastic heating -heat contains no latent plastic terms. Therefore, as in elastic
solids, the thermoelastic heating is associated in this model with nondissipative (latent) elastic structural changes and leads to the so-called Gough-Joule effect, see,
e.g., CARLSON [1972]. Using the stress relations along with the kinematic relation
= FeTdeFe, which is implied by the multiplicative decomposition (57.1)1 and

was derived in Section 34, the structural elastic heating can be expressed in terms of
the temperature derivative of the stress field as
heat =

-ea

9 s. -eD=

-Oar

- de.

(57.13)

The coupled thermomechanicalproblem

SECTION 57

443

The additive decomposition H = He + H P implies that the internal dissipation in the


solid can be written as the additive contribution
Din t = Dmech + Dther

in Q2 x

]I,

(57.14a)

where Dmech is the mechanical part of the internal dissipation and Dther is the part
of the internal dissipation induced by the temperature field in the solid, respectively
defined by
hint

Dther = OHP

and

Dmech =

d + E

q'

(57.14b)

a=1

The local evolution equation for the absolute temperature field in the solid is obtained by inserting into the local balance law (56.10)2 the constitutive equation for
the nominal heat flux and Eq. (57.11) for the rate of entropy in the solid. In view of
the preceding relations we have
c/

= -DIv[Q]

+ R +

[Dmecb -

YhFeat]

in Q x

(57.15)

It should be noted that the plastic entropy H P does not appear explicitly in the temperature evolution equation. The second bracket in (57.15) is the result of heat conduction
in the solid and, therefore, vanishes in an adiabaticprocess.
EXAMPLE 57.1. To provide a specific illustration of the preceding ideas, consider one
possible thermomechanical extension of a purely mechanical model characterized by
a stored energy function W(C e) and a hardening potential 7-I( ). Such an extension
relies on the rather strong assumption that the heat capacity c of the solid is a constant.
Integration of the defining relation (57.12)1 then yields the following expression for
the free energy function
kr(Ce,

, ~,) = T(O)- (

- Oo)M(C

e)

+ W(Ue) +
t(),

(57.16)

where 90 is interpreted as a reference temperature and the function T(O) is given by


the explicit expression
T(O) = co[(9 - 0o) - log(/0o)]

with c = co (constant).

(57.17)

The term -(6 - Oo)M(C e) describes the thermomechanical coupling due to thermal
expansion and furnishes the potential for the associated elastic structural entropy,
whereas the function T(O) is the potential for the purely thermal entropy. The part
H e of the total entropy then becomes
He = co log(9/90) + V(ee).

(57.18)

J.C. Simo

444

CHAPTER V

This relation can be solved in closed form for the absolute temperature field 6 to
obtain the inverse constitutive relation
9 =

(ce, He)

90 exp{ [He - M(e)]/co}

(57.19)

Using this result, the closed-form expression for the internal energy in the solid follows
from (57.8) as
E(Ce,He, ) =

(C e, (Ce,He),

) + He9(Ce, He).

(57.20)

It is immediate to verify that constitutive equation (57.19) satisfies the potential relation O(C e, H e) = aHE(Ce, He, ).
Consider next the specialization of this result to the two specific models of interest in
metal plasticity examined in Example 37.1, labeled as Model I and Model II, respectively. These two models are isotropic and exhibit uncoupled volumetric/deviatoric
response. Consistent with these two restrictions, it is reasonable to assume that the
function M(Ce) specifying the coupling term in (57.16) depends only on the volumetric part Je = J of Ce. Setting M(Ce) = G(J) and using the notation in Section 37,
the free energy function becomes
(C e , , a) = T(O)- (

- Oo)G(J) + U(J) iF(6e) + 7(),

(57.21)

where Ce = FeTFe and Fe is the volume preserving part of the elastic deformation
gradient F e . The functions W(Ce) and U(J) are those defined in Example 37.1.
Specifically, setting be = FeFeT we have
Model I. 1[J
Model II.

1 - log(J)] and W(e)

U(J) = 1j[log(J)]2

and (C

e)

(tr [be] _ dim)

= 4 log (e)

The constitutive equation (57.18) for H e , together with the inverse relation (57.19)
for the absolute temperature field 9, then take the following forms:
H e = colog(9/0o) + G(J)

and

O = 90 exp{ [He - G(J)]/co}.

(57.22)

A possible choice for the function G(J) inspired by the linear theory is G(J) =
ndimoU(J) where a > 0 is the linear coefficient of thermal expansion.
(C) Evolution equations for thermoplastic flow. To generalize the results described
in Section 36 to a temperature-dependent yield criterion, it suffices to consider the
case in which the elastic domain E is specified in the current configuration in terms
of the Kirchhoff stress tensor. An identical extension holds if E is specified in terms

The coupled thermomechanicalproblem

SECTION 57

445

of the second Piola-Kirchhoff tensor S. To account for a temperature-dependent yield


criterion it is assumed that E is specified as

S x Rni t X IR+:

E = (r, q{ O)

= 1,2, ... , m.

fu (r, q, 69) < O for

(57.23)

As usual, it is further assumed that the m-constraints, now depending also on the absolute temperature, are qualified and defined via convex functions. To gain insight into
the effect of the temperature field, consider once more the case of associative plasticity. Under this hypothesis, the evolution equations that define the rate-independent
model are those which render a maximum for the internal dissipation Dint in the solid,
subject to the constraints specified by (57.23). In view of expression (56.11) for Dint,
this assumption translates into the inequality
ninth

IT-

d + [-

6] H + E

[qa

o
0]

oa1

V(t,

E,

(57.24)

which, given the functional form (57.23) adopted for the elastic domain E, implies
the following evolution equations:
m.
=

h a,

z=l

(57.25)

Eo yl q fp (r, q , ),
=A
m=1

iHP =

aeOof4(r, q=, O).

For the rate-independent theory, the multipliers in these evolution equations obey the
standard Kuhn-Tucker conditions
m

,yl > 0

and

75f,(r,q,

) = 0,

(57.26)

=1

which enforce the constraints f(-r, q, 6) < 0, for t = 1,2,..., m. Interestingly,


the evolution equation (57.25)3 arising from the principle of maximum dissipation
provides a phenomenological interpretation for the evolution of the configurational
entropy HP in terms of the temperature change of the yield criterion. This interpretation first appears in StMO and MIEHE [1992]. The extension of the associative flow

446

J.C. Simo

CHAPTER V

rule (57.25) to the rate-dependent theory is accomplished via the same regularization
technique described in Chapter II within the context of the infinitesimal theory.
EXAMPLE 57.2. To illustrate the simplest form of the rate-dependent model, consider
for simplicity single-surface plasticity obtained by setting m = 1 in (57.23). The internal dissipation Din t defined by (56.11) is then replaced by the regularized dissipation
function

Di =
= Diint

I
-j+(f,

) >0

in Q2 x ,

(57.27)

where the regularization parameter 7rC (0, oo) has the dimensions of a viscosity coefR is a differentiable function which satisfies,
ficient. In this expression j+:RIx JR+
for fixed but arbitrary temperature 9 E 1t+, the two conditions: (i) j+ (x, 6) > 0, for
all x E IR, and (ii) j+(x, 6) = 0 if and only if x < 0. Setting g(f, (9) = afj+(f, 6),
the rate-dependent flow rule for single-surface associative thermoelasticity is obtained
by replacing the Kuhn-Tucker conditions (57.26) with the constitutive equation
= g(f(, q, )), 6)

if f(r, q, 9) > 0,

(57.28)

otherwise.

Recall that the simplest choice of the regularization function in the purely mechanical theory is obtained by setting g(x) = 2[H(x) + x], where H(-) is the Heaviside
function, and corresponds to the linear viscous model (see Chapter II). For the thermomechanical theory, the temperature-dependent part g(-, (9) of the regularization
function is often assumed to correspond to an Arrhenius term with functional form
exp[-A/k6)], where A is the activation energy and k is the Boltzmann constant, see
MCCLINTOCK and ARGON [1968] for additional details.
It remains to show that the evolution equations (57.25) are consistent with the
constitutive restriction (56.11) implied by the Clausius-Plank form of the second law.
It is shown below that this restriction implies thermoplastic softening in the solid for
a wide class of material models arising in metal plasticity. In other words, the flow
stress must decrease with increasing temperature field. This result is consistent with
experimental observations on plastic flow in metals within temperature ranges below
phase transitions, see PHILLIPS [1974] and ZDEBEL and LEHMANN [1987], and will be
demonstrated within the setting afforded by a representative model problem.
EXAMPLE 57.3. Assume that the yield criteria in definition (57.23) for the thermoelastic domain IE have the functional form
f (r, a,(9) = p (,q

) -

y(9) < 0

for i = 1,2,..., m,

(57.29)

where the functions 4, (-, -) are convex and homogeneous of degree one and
uy,((9) > 0 are interpreted as suitable flow stresses. As pointed out in Chapter
II, this is the conventional setting adopted in the description of plastic flow in metal

The coupled thermomechanical problem

SECTION 58

447

plasticity. Implicit in (57.29) is the assumption that the effect of the temperature field
on the yield criterion arises only via the temperature dependence on the flow stress,
since the hardening potential 7-( ) is assumed temperature-independent. This simplifying hypothesis agrees, at least to first-order effects, with a number of experimental
observations, see the references quoted above and LEHMANN and BLIX [1985].
The issue to be addressed here is the sign of the derivatives ayL, (9), the case
yus, () < 0 corresponding to thermoplastic softening. By recalling that Euler's theorem for homogeneous functions implies the property
q')

q)
q ")a~r(~
.rqaqp,rIL'7Q

) qY,

(57.30)

the associative flow rule (57.25) together with (57.29) yield the following expression
for the internal dissipation Dint in inequality (56.11):
m

Dint=

yylf, (r, q, O) +

E -'

[aiy(e) -ar

(y,)]

0.

(57.31)

/z=l
=1

To fix ideas, consider the rate-independent theory. The first term in (57.31) then
vanishes as a result of (57.26)2 while the second term is positive since yM" > 0 and
ory,() > 0. The last term in (57.31) is the dissipation Dther in the solid arising from
the dependence of the flow condition on the temperature field. In view of (57.14b)
and (57.31), it follows that
m
Diher

=-

9P

= -E

Zyr

y((9)

in /2

x 1[.

(57.32)

From (57.31) and (57.32) one concludes that the requirement Dther > 0, which ensures
0, automatically holds under the assumption
satisfaction of the inequality Dint
d (9) < 0, i.e., for thermoplastic softening. The same conclusion is easily shown
to hold for the rate-dependent theory.
58. Formal a priori stability estimate and conservation laws
The central issue in a nonlinear stability analysis of coupled problems concerns the
appropriate notion of nonlinear stability. In the linear regime, there is a well notion of
stability going back to the fundamental work of Lax, see RICHTMYER and MORTON
([1967], Chapter 2), which exploits the underlying semigroup structure of the problem of evolution. A brief exposition of this notion is given in the following section.
For nonlinear dissipative problems of evolution, nonlinear stability is often phrased
in terms of an a priori estimate on the dynamics. Typical examples include the incompressible Navier-Stokes equations in fluid mechanics, see TEMAM [1984], and the
system of coupled nonlinear thermoelasticity in solid mechanics, where the a priori

448

J.C. Simo

CHAPTER V

estimate arises from the second law, see ERICKSEN [1966], GURTIN [1975] and BALL
and KNOWLES [1986].
The goal of this section is to present a (formal) stability result (ARMERO and
SIMO [1993]) for the nonlinear initial boundary value problem of thermoplasticity.
Existing stability results are either restricted to the mechanical theory, as described in
Chapter II and Chapter III, or limited to a linear perturbation analysis for the ratedependent coupled problem of the type initiated in CLIFTON [1980], see MOLINARI
and CLIFTON [1983], ANAND, KIM and SHAWKI [1987] and the review in SHAWKI and
CLIFTON [1989]. The nonlinear stability estimate described below will be shown to
arise as a direct consequence of the Clausius-Duhem form of the second law.
To state the result, recall that the deformation poE Cmech, the material velocity V =
q of the motion, and the absolute temperature field (9 E Cther comprise the primary
variables of the initial boundary value problem. The plastic part FP of the deformation
gradient and the hardening variables A, comprise the internal variables, with evolution
defined in terms of the primary variables via the flow rule. For convenience, the
following compact notation will be adopted
Z = {., V,

and

F =

(58.1)

HP}.

{FP,,

We shall further assume that the mechanical loads derive from a potential Vext(pO)
according to the standard potential relation
DVext(p) -*0= -Gext(n)

- (T, no)r

= -(B,,)

Vn0o

Vo.

(58.2)

The form Gext(r0o) defined by the preceding expression is linear in r70 E Vo and gives
the virtual work of the external loads. For a given system of conservative loads with
potential Vet(qo) and a general free energy function !(Ce, O,(), not necessarily
of the uncoupled form (57.9), in the absence of heat sources (R O0)we define the
Lyapunov-like function
r(Z, r)
=

J [ POlVI

(C'e .

,)

- O)He] dQ2 + Vext(q),

(58.3)

where He = -a,9fo(Ce, 0, a). We emphasize that in this expression Ce = FeTFe


is regarded as a dependent variable defined in terms of (Z, F) via the multiplicative
factorization Fe = DqoF p- 1. With these conventions we have the following result.
THEOREM 58.1. Assume that thefollowing conditions on the datafor the initial boundary value problem of dynamic thermoplasticity hold:
(i) No heat sources, i.e., R = 0 in Q x I.
(ii) Conservative mechanical loading with potential Vext(cp).
(iii) Boundary condition for the temperaturefield of the Dirichlet type, with prescribed constant temperature 9o > O, i.e., 0 = 0o on a2Q x I and FQ = 0.

The coupled thermomechanicalproblem

SECTION 58

449

Assume further that the constitutive equationfor the heat flux Q and the evolution
equationsfor the internal variables F satisfy the dissipation inequalities
'int

Dcon =

--

GRAD[]

O0 and

DP +

)mech= S

qa

0,

c=l

(58.4)
where DP = sym[CeLP] is the plastic rate of deformation tensor; with LP =

FPFp -1, while the generalized stresses (, qu) satisfy the constitutive equations
S= 20ueT(fe

, 5,)

and q' = -al

P(ze,o,

).

(58.5)

Under these hypotheses, the weak form of the balance laws (56.10) in the statement
of the initialboundary value problem imply the a priori estimate
d

(Z,

0 [Dcon +

Dmech ] d2

< 0 in Il,

where (Z, F) is defined by (58.3) with He = -oeP(Ce, O,

(58.6)

).

PROOF. Since the domain of integration 2 in the integrals involved in definition (58.3)

is fixed, assuming sufficient smoothness it is permissible to interchange integration


and time differentiation. The key identity used in the computation below is
nint

l(Ve 6?,

) =P GRAD[V] -

P -

qc',

Ca=I

= P

GRAD[V] - Dmech,

(58.7)

which follows from (57.3), the relation P = FeSFP-T, constitutive equations (58.5)

and definition (58.4)2 by exactly the same argument as leading to (57.4). Using this
result, the time derivative of (58.3) reduces to
d-L(Z, F)= [(poV, V) + (P, GRAD[V]) - Get(V)]
+

X2[He

+ J

+ a

[-- mech

( ),]e 0,

+ (9

d2

o0 )fe] d.

(58.8)

The first term within brackets in (58.8) is precisely the weak form (56.10)1 of the
balance of momentum evaluated with o = V chosen as test function which, therefore,
vanishes. The second term within brackets in (58.8) also vanishes as a result of the
constitutive equation for the elastic part H e of the total entropy in the solid. Finally,

450

JC. Simo

CHAPTER V

since Dint = D1)mech + Dther with Dther = OHtP, using the additive decomposition
H = H + HP the remaining term in (58.8) can be written as

= Lr)Dint +

dt (Z

,,Dther+

90) Ot dQ.

(58.9)

The proof is concluded by noting, in view of definition (56.17) for the space To
of admissible temperature test functions, that assumption (iii) on the boundary data
renders the function ( - Oo)/e an admissible test function, i.e.,
~o = (O -

o)/O E Vo

since 60 = 0 on a2 x E1.

(58.10)

For this specific choice of test function, the weak form (56.10)2 reduces to

etH (

eo)
0)
Q- GRD [] + Din.)

'Din d.

(58.11)

Combining (58.11) with (58.9) and using definition (58.4)1 yields the result.
It is again emphasized that the formal a priori estimate (58.6) does not preclude
physical phenomena to be expected in the presence of thermoplastic strain softening, such as the possible formation of shear bands and strongly localized and plastic
deformations. Therefore, estimate (58.6) provides a meaningful notion of nonlinear
stability suitable for general models of multiplicative plasticity. This estimate plays a
central role in the analysis of nonlinear numerical stability described below.
REMARK 58.1. The case of coupled nonlinear thermoelasticity is recovered within the

present context merely by setting F P = 1 and (>. = 0 in 2 x . In this situation, F e =


Dcp and the preceding a priori estimate reduces to the classical result in DUHEM [1911].
For thermoelasticity the functional L(Z, F) is a Lyapunov function, a fact first proved
in ERICKSEN [1966] and subsequently exploited in the analysis in BALL and KNOWLES
[1986] of the quasistatic problem. For linear thermoelasticity in the absence of external
loads, the linearization of (Z, F) about the reference configuration reduces to a
quadratic functional shown in DAFERMOS [1976] to define a norm in the Hilbert space
Z = H(f)nd im x [L2 (f] n dim x L 2 (Q) of admissible displacements, velocities and
(linearized) temperature fields. The a priori estimate (58.6) then renders linearized
thermoelasticity as a semigroup of contractions in Z.
In addition to the preceding a priori stability estimate, for traction boundary conditions the initial boundary value problem retains the two global conservation laws of
momentum described in Section 39 and summarized in the following.

SECTION 59

The coupled thermomechanicalproblem

451

58.2. For the pure traction boundary value problem ( = 0) under equilibrated loads (Next = Mext = 0), the weak form (56.10) of the governing equations
yields the conservation laws
THEOREM

-L(Z,r)=O0
dt

and

d
dt

J(Z,r)= 0

in ,

(58.12)

where L(Z,r) = f 2 poV d2 and J(Z,r) = f cp x poVdf2 are the total linear
momentum and the total angularmomentum, respectively. These definitions agree with
those given in (39.11).
PROOF. The proof of these results are identical to that given in Theorem 39.1.

The remaining of this chapter is concerned with the design of numerical schemes
for the solution of the initial boundary value problem of dynamic thermoplasticity
which preserve the a priori estimate (58.6) and inherit exactly the conservation laws
of momentum (58.12).
59. Time integration of the coupled problem: General considerations
Given a partition II = UNO=[t,, t,+I] of the time interval II of interest, algorithms for
the numerical integration in time of the initial boundary value problem of dynamic
thermoplasticity are typically designed by rewriting this system as an abstract firstorder problem of evolution for the primary variables Z and r defined by (58.1). The
main numerical analysis issues involved can be addressed with reference to a general
first-order system with the following form:
dtZ(.,'t)=A[z(-,t)]+ f(.)
z(, to) = zo(.)

in

x ,

(59.1)

in 2,

where z(o, t) lies in a suitable function space, typically a Banach space of the Sobolev
class with norm denoted by lI zI for z e Z, A[.] is a nonlinear elliptic operator with
domain dense in Z, f is a prescribed forcing term and zo C Z is some specified
initial data. Under suitable technical assumptions; see, e.g., MARSDEN [1973, 1974],
the homogeneous version (f = 0) of the abstract problem (59.1) defines a local
semiflow, denoted by
Ft: Z x [t t+l] -- Z,

(59.2)

which advances the initial data z(., to) E Z to the solution of problem (59.1) at time
t according to z(., t) = Ft[z(-, to)], and satisfies Ft+, = FPto F for t > s. In the
jargon of dynamical systems, one refers to A[-] as the vector field associated with the
flow Ft. In what follows, we shall assume that the technical conditions which ensure
the existence (at least locally in time) of the flow hold, and concern ourselves solely
with its algorithmic approximation.

452

JC. Simo

CHAPTER V

Let Kat: Z x R - Z be a one-parameter family of maps, referred to as the algorithm


in what follows, which depends continuously on the parameter At > 0 herein referred
to as the time step. Recall that the algorithm IMAt is said to be consistent with the
flow (59.2) if the following two conditions hold for any z Z (see RICHTMYER and
MORTON [1967]):

im IAt[z] = z

At-,0

and

lim

At-

At

[KAt[z] -

A[z].

(59.3)

Stability considerations on the algorithmic approximation are typically made for the
homogeneous problem obtained from (59.1) by setting f = 0. For the linear problem,
A[] is the infinitesimal generator of a semigroup for which a well-defined notion of
stability due to Lax exists. This concept of stability essentially amounts to requiring
that the iterated algorithm KI = KAt o... o Kat, with At = (t - t,)/k, be uniformly
bounded in bounded t-intervals.
For quasicontractive semigroups, the appropriate notion of stability is furnished by
a stronger condition known as B-stability. Recall that a semigroup IFt is said to be
quasicontractive if the condition lFt[z] - Ft[] liz < liz - llz holds for all t E E1
and any z,z E Z, relative to some norm I Ilz, equivalent to the standard II on
Z and called the natural norm for the problem. An algorithm IKAt consistent with
the semigroup, in the sense of (59.3), is called B-stable if it inherits the contractivity
property relative to the natural norm II z, i.e.,
|IKAt[Zn] -

I;At[n] 11Z

lizn - Zni

for At > 0.

(59.4)

and any two initial conditions z,, ,, E Z. This definition was exploited in Chapter
III in the stability analysis of the initial boundary value problem for infinitesimal
elastoplasticity formulated in stress space. Unfortunately, contractivity of the flow
is a very restrictive assumption which does not hold for many dissipative dynamical
systems of interest and, in particular, for thermoplasticity at finite strains. The situation
for general nonlinear problems of the type (59.1) is, therefore, significantly more
complex and several alternative notions of nonlinear stability are possible, see the
introductory exposition given in Section 13 and references therein. One possibility is
to assume that the algorithm IKAt is consistent, in the sense of (59.3) and linearized
stable (or locally Z-stable, see CHORIN, HUGHES, MARSDEN and MCCRACKEN [1978],
pp. 237, 239). These two assumptions plus technical hypotheses on the flow yield
convergence of the algorithm in the sense that
lim

k-oo

AKt[z] = Ft[Zn],

for At = (t- t)/k and t G [tn,t+],

(59.5)

thus providing a nonlinear version of the classical Lax equivalence theorem.


Alternatively, for dissipative problems of evolution, nonlinear stability can be formally phrased in terms of a (formal) a priori estimate on the dynamics of the form
[Ft(zo)] < 0

for t e

I9,

(59.6)

SECTION 59

The coupled thermnnomechanicalproblem

453

where : Z -- IRis a nonnegative function. Under suitable technical assumptions, it


is the existence of such a (Lyapunov-like) function that provides the natural notion
of nonlinear stability for a dissipative problem of evolution of the form (59.1). In
particular, for nonlinear thermoplasticity, the function is the canonical free energy
functional defined by (58.3) and the a priori estimate is precisely inequality (58.6)
with zero forcing. An algorithm KAt for problem (59.1) will be called dissipative
stable (or simply stable for short) if it inherits the estimate (59.6), in the sense that
(Kt[z]) -

(z) ~ 0 for z C Z and some At < At,,i,,

(59.7)

for zero forcing, i.e., f = 0 in Q x I and T = 0 on FT x I. If no restrictions are placed


on the critical time step Atcit > 0 or, equivalently, if Atcit = o00, the algorithm is said
to be unconditionally stable; otherwise, the algorithm is said to be only conditionally
stable. This is the notion of stability adopted in the remainder of this chapter.
EXAMPLE 59.1. The simplest example of a linear dissipative problem of evolution is
provided by heat conduction in a rigid isotropic solid with linearized heat capacity co
and constant conductivity k > 0. The standard governing equations are

az

cat

= kA[z] + f

z=0
zlt=o = z0

on a

in

x [,

x ,

(59.8)

in Q,

which is obtained from (59.1) by specifying A[.] as the constant k/co times the
Laplacian with zero boundary conditions, and letting Z = L 2 ( Q). In this situation
A[.] is an unbounded closed operator with domain the Sobolev space Ho'() dense
in the space Z = L 2 (9). Thus 11-11 is the standard L 2 (2)-norm and the associated
inner-product (-, ) is the standard L2 (9)-pairing of square integrable functions. The
natural inner product (, )z and the associated natural norm 1I z for the problem
at hand are respectively defined by
(z, Z2)z = (Po zi, z2)

and

lizliz = Vz,/)z.

(59.9)

Obviously, the standard L 2-norm 11 11and natural norm 11- 1iz are equivalent for
co > 0. Under the assumption of no forcing term (i.e., f = 0 in Q x II), the a priori
estimate (59.6) for the flow map follows via integration by parts as
dt [lZ

= (z,

kA[z])- = -klVzlj

-Ckll Il 2 < 0,

(59.10)

where C > 0 is the positive constant arising in Poincare's inequality. Thus, the positive
function (z) = I Z11 satisfies (59.6) and, therefore, decreases along the flow. It is
well known that problem (59.8) is also contractive relative to the natural (and hence
relative to the standard L 2-norm). As an example of a (generally implicit) definition of

454

J.C. Simo

the algorithm z, + l
mid-point rule

= KAt [zn],

CHAPTER V

consider the approximation of (59.8) by the generalized

co(Zn+ - z,) = AtkA[zn+,]

where z,n+ = Oz,+X + (1 -

)zn,.

(59.11)

Multiplying this expression by (zn + z,+ ) and integrating over Q, a straightforward


manipulation that exploits the identity z,+o = 2(Zn + zn+L) + ( - 2[Z+yields
21z+Ilz -

11

-(

--tktIlVznol

z.

(59.12)

Since L(z) = [z112, this result implies that (z,,+) - (zn) < 0 for any At > 0 if
0 >
and k > 0. Therefore, the concept of stability embodied by inequality (59.7)
reproduces the well-known unconditional stability result for the generalized mid-point
rule when applied to the simple linear problem (59.8).
Considerations on the stability of algorithms for the first-order system (59.1) are
generally made without reference to the forcing term f (). In the linear regime,
for instance, the classical concept of A-stability for an algorithm is a test on the
homogeneous problem z' = z, with A e C, known as Dahlquist's model equation
(see Chapter II). For dissipative systems, considerations on the effect of the forcing
term lead to questions concerning the long-term behavior of both the dynamical system
and the algorithm, as the following representative example illustrates.
59.2. Consider again the classical problem of heat conduction, now without
introducing the assumption of a vanishing forcing term, i.e., for f f 0. The objective
is to examine the effect of the forcing term on the dynamical system under the simplest
hypothesis that f is time-independent. Following the same steps that were leading to
(59.10) one arrives at
EXAMPLE

d(
dt

z)= (z, ) z

-k1

IVz2

+(f,z)z

-CkIlzl

+ f,z)zl.

(59.13)

The term in (59.13) involving the forcing term f can be estimated using Young's
inequality as

for any

Z) Z I< IfIIZ IzI


> 0. Setting

2<,

(59.14)

Ck and inserting the result into (59.13) yields

d L(z) + Ck L(z) < fllk


dt
2Ck

(59.15)

SECTION 59

The coupled thermomechanical problem

455

By application of the classical Gronwall lemma, see, e.g., TEMAM [1979], one arrives
the result
L(z) < exp[-Ckt](Zo) + 2

(1 - exp[-Ckt]).

(59.16)

This inequality implies, in particular, the following estimate governing the long-term
behavior of the dynamical system for a given (time-independent) forcing function:
lim sup[C(z)] <

[fllz/(Ck)]2.

(59.17)

A similar estimate holds for nonlinear dissipative dynamical systems far more general than the classical model of heat conduction considered here, the incompressible
Navier-Stokes equations being an specific example, see TEMAM [1979] and references
therein. A natural question to be asked is whether a given algorithm IKt [-] inherits the
long-term dissipative property embodied in the long-term estimate (59.17). To answer
this question consider again as a model algorithm the generalized mid-point rule, with
stability properties derived in the preceding example. Following the same steps as
leading to (59.12) one arrives at the identity
( +,)

- r(z)

-- (oL -)lz,+

ll

-AtklJVz+o1

+ At(f, zn+)z.

(59.18)

Assume that the algorithm is unconditionally stable, so that 9


I Then the first
term in (59.18) is negative while the second term in (59.18) can be estimated using
Poincare's inequality as -lVzn+ll
< -Cllz+2Z. The last term in (59.18) is
estimated using Young's inequality as
(f Zn+9)3

IZ((fzn+)zI <

(59.19)

for any > 0. Setting e = Ck and inserting the result into (59.18) yields the following
algorithmic counterpart of inequality (59.15):
(Zn+l)-

L(zn) + AtCkL(zn+o)

AtlIlk
t 2Ck

(59.20)

To proceed further, one uses an analog of the classical Gronwall's lemma, which is
obtained by estimating the term (zn+,) using Young's inequality (with = 2) as
follows. First observe that
(Z,+9) ) 092 C(Z+, 1) + (1 - ) 2L(Zn) -0(1
>

2 (Zn+l) + (1 -

= 0(20 - l)L(zn+l) - (1 -

-)2(z,)

(1 -

O)(Z Z,n+l)zl
) [(Zn+1) + ,C(zn)]

)(29 - 1)Z(zn).

(59.21)

456

J.C. Simo

CHAPTER V

Next, by inserting this result into (59.20) one obtains the recurrence relation
At
L(zn+,) 4 A(At)L(z.) + 1 +

IfI

AXt 2Ck

(59.22a)

where A(At) is a is a norm-amplification factor defined as


A(At) =

1 + 2At

1 + ( 1At

with

= 1 + AtCk(2

- 1)0,

w2 = 1 + AtCk(21 - 1)(1 - ).

(59.22b)

Recursive application of (59.22a) and use of the well-known formula for the sum
of the terms in an arithmetic progression then yields the result
(z)

<

[A(At)

(z)

At
IlI 1 - [A(At)](9
+ 1 + At 2Ck 1 - A(At)

(59.23)

which furnishes the algorithmic counterpart of relation (59.16) implied by Gronwall's


lemma. If (59.23) is to have a limit for fixed At as n -- o,

the norm amplification

factor A(At) must satisfy the condition IA(At) I < 1 which implies that [A(At)]"
as n
o. A straightforward computation gives the At-independent estimate
limn sup[L(zn)]

2C2k2(20
22
2C

(29

1)

if
if

(t)

< 1,

(59.24)

which completely characterizes the long-term behavior of the algorithm for a fixed
(time-independent) forcing function f and all possible initial data z0o. The condition
> , which is more reil(At)l < 1 is easily seen to imply the strict inequality
strictive than the condition 9 > 2 arising solely from stability considerations. In fact,
for the value
= 2 corresponding to the second-order accurate Crank-Nicholson
scheme, the long-term estimate (59.24) blows up. Remarkably, by comparing (59.24)
with (59.10) we conclude that the algorithm reproduces the long-term a priori estimate of the dynamics for the value L9= 1 which corresponds to the backward Euler
method. The preceding considerations can be shown to generalize to arbitrary dissipative dynamical systems and are motivated by the analysis for the incompressible
Navier-Stokes equations presented in SIMO and ARMERO [1993]. Issues related to
the long-term behavior of the algorithmic dynamics will not be pursued further here
and the analysis of coupled problems described below will be restricted to standard
stability and accuracy considerations.
60. Monolithic and staggered schemes: Product formula algorithms
With the preceding background in hand we are in the position of addressing the main
issues arising in time integration of the initial boundary-value problem for dynamic
thermoplasticity. The first step is to cast the coupled problem into a first-order system
of the form (59.1). Once this is accomplished, the time integration of the coupled

SECTION 60

The coupled thermomechanicalproblem

457

problem is typically performed via two alternative class of algorithms referred to


as monolithic and staggered schemes, respectively. The main characteristic of these
two methods are summarized below. The reader is referred to LEWIS and SCHREFLER
[1987] and WOOD [1990] for an overview of some representative schemes currently
used in coupled problems arising in a wide range of engineering applications.
(A) Monolithic schemes. These methods are obtained by applying a suitable time integrator to the full first-order system (59.1), such as the the generalized mid-point rule
algorithm considered in the preceding example. For coupled thermoplasticity, the treatment of algebraic constraints introduced by the Kuhn-Tucker conditions necessitates
an additional projection scheme leading to the return mapping algorithms described in
Chapter III. This approach lends itself to a relatively easy development of algorithms
possessing unconditional stability, by adaptation of classical schemes for systems of ordinary differential equations. However, for coupled problems such as thermoplasticity,
monolithic schemes do not take advantage of the different time scales often involved
in this class of problems. Moreover, the linearization of the algebraic problem resulting
from a combined spatial and temporal discretizations typically leads to large nonsymmetric matrices that add considerably to the total computational cost. It is for this
reason that monolithic schemes are not widely used in the numerical solution of coupled problems, in spite of their good stability properties. A fairly complete treatment of
monolithic schemes for coupled thermoplasticity, including the design of mixed finite
element methods and the closed-form expressions arising from the exact linearization
of the nonlinear algebraic problem, can be found in SIMO and MIEHE [1992].
(B) Staggered schemes. This class of algorithms is designed to exploit the different
time scales often involved in a coupled problem via partitioning of the associated
first-order system into two or more sub-problems. For coupled thermoplasticity this
partitioning of the problem is accomplished by considering a mechanical phase, typically at fixed temperature, and taking the result as initial conditions of a subsequent
thermal phase, performed at a fixed configuration, that defines the temperature field.
Strategies of this type will be examined in detail in the next section and are strongly
reminiscent of fractional step methods designed on the basis of a global operator split
of the first-order coupled problem of evolution. To set the stage for the discussion of
these methods, we sketch below the basic strategy for the abstract problem of evolution (59.1). For a detailed exposition of fractional step methods the reader is referred
to the review article of MARCHUK [1990] and references therein.
The key idea is to introduce the additive operator split A[-] = A(')[.] + A( 2) [.]
of the vector field A[-] that defines the first-order problem of evolution (59.1), where
A(') [] and A( 2) [] are two (hopefully simpler) vector fields, and consider following
two sub-problems:
Problem 1:

-z(., t) =A() [z(, t)].

Problem 2:

dz(., t) = A(2) [z(, t).


dt

(60.1)

J.C. Simo

458

CHAPTER V

The critical restriction on the design of the operator split is that each of the subproblems must preserve the underlying dissipative structure of the original problem. In
other words, if IFta) [] denotes the semiflow associated with the vector field A(a) [.],
(c = 1,2), then the following estimates must hold:
dL(IFt )[z]) < 0

for t C
I[ and a = 1,2.

(60.2)

Now consider algorithms IK() [] consistent with the flows F? ) [-], a = 1,2, and stable
in the sense that each algorithm obeys the a priori stability estimate (59.7). Then, the
algorithm KAt[.] defined by the product formula
Kat[-] =

(2)

,)[A],

in Z x [tn, t+l],

(60.3)

is also consistent and dissipative stable. Formally, the stability of the product formula
(60.3) in the sense of the a priori estimate (59.7) is easily concluded from the stability
of each algorithm, merely by noting that

{(KAt [At [z]])

(KAIt[Z]) - L(z)

(KAt [])

+- 2{ (K(it)[z]) -

(Z)}< 0,

(60.4)

since neither of the two terms within braces in this identity is nonnegative as a result
of the assumed stability on each of the algorithms. Under appropriate technical conditions, consistency and stability of the algorithm defined by (60.3) implies convergence
of the iterated algorithm as the time step At -- O0;i.e.,
lim [K(2) o K[]k
ltZtoo
A t A
[Zn]

k- -

= Ft[z,], for At = (t - t,,)/k, with t e [ta, t,+l].

(60.5)

Expression (60.5) is a classical result going back to Lie, see KATO [1976]. Nonlinear
versions of this formula have been given by a number of authors, in particular BREZIS
[1973] and MARSDEN [1974] within the context of a product formula for the numerical
solution of the incompressible Navier-Stokes equations due to CHORIN [1969].
REMARK 60.1. For quasicontractive semigroups, if each of the algorithms is unconditionally B-stable in the sense of (59.4), then the product algorithm (60.3) is also
unconditionally B-stable in the sense of (59.4). This property follows immediately
from the inequalities
n

||

KAt[n

at

<_||K(1 [z,,]

[] At

Bat1 [.]lZ

6At

1Zn lz

which are implied by the assumed B-stability of the algorithms.

il

(60.6)

The coupled thermomechanical problem

SECTION 60

459

It should be noted that the product formula algorithm defined by (60.3) is only
first-order accurate, even if each of the underlying algorithms is second- or higherorder accurate. A second-order accurate product formula of two (at least) second-order
accurate algorithms is given by the modified expression
ast =

(t)
At/22

( 1)
K(2)
At oK At/2

(60.7)

which goes back to STRANG [1969] and is known as the double pass technique (or
"Strang's trick"). A simple example illustrates this result.
EXAMPLE 60.1. Let A E S+ (n) be a symmetric, positive definite matrix and consider

the standard initial value problem

dz(t) = Az(t)
dt

with z(0) = zo

In.

(60.8)

For this elementary example it follows that Z = REn. The exact solution to problem
(60.8) is given in closed-form via the the matrix exponential as
z(t) = Ft[zo] = exp[-At]zo.

(60.9)

The problem is obviously dissipative, with (z) = zTAz and natural norm the
matrix norm IlzI2I = 2(z). Since Z is finite-dimensional, the equivalence between
this norm and any other matrix norm follows by standard results in linear algebra. Let
A = Al + A 2,

with A 1,A 2 cE $+(n),

(60.10)

be any decomposition of the matrix A into two positive definite, symmetric matrices.
Such a decomposition defines an operator split of problem (60.8) since the dissipative
property relative to the natural norm is preserved by the additive partition (60.10). The
exact solution of each sub-problem is again given by the matrix exponential formula
(60.9). Accordingly, the algorithms defined by
K(?)[z] = exp[-A~At]z,

for z E Rn

and c = 1,2,

(60.11)

furnish the exact solution to each of the sub-problems in (60.1). Let Tat denote the
local truncation error associated with a specified product formula algorithm IKat[[],
defined by
Tat [z,] = exp[-(Al + A 2 )At] zn - KAt [z,(],

(60.12)

J.C. Simo

460

CHAPTER V

where z, E R is the prescribed initial data in a typical time step [t, tn + At].
Assume that the matrices Al and A 2 do not commute, so that the matrix commutator
[Al, A 2] = A 1A 2 - A 2A 1 7 0, and consider the following two product algorithms:
(i) One-pass product formula. This algorithm is defined by the conventional expression KAt[zn]
=K K(I o IK( ) [zn]. Making use of the standard series expansion
for the matrix exponential, an elementary computation yields
Tt[Zn]

= At[A 1, A2][z,] + O(At3 ).

(60.13)

This result demonstrates that the standard product formula is only first-order accurate,
even if each of the sub-algorithms is exact, as in the present elementary example.
(ii) Two-pass productformula. This algorithm is defined by expression (60.7). Using again the series expansion for the matrix exponential, a similar computation reveals
that the modified formula (60.7) is designed precisely as to cancel the second-order
term involving the matrix commutator in the conventional formula (60.13). Since
the calculation involves only terms up to order O(At 2 ) in each of the algorithms
K(t) [ ], a = 1, 2, second-order accuracy of both algorithms suffices for this conclusion to remain valid. For instance, the product formula (60.7) with algorithms now
defined by the Crank-Nicholson approximation
K(t)[Z] = (I -

tA)

(I + AtAa)z,

ac= 1,2,

(60.14)

is also second-order accurate, while the conventional formula is not.


The preceding considerations generalize immediately to multiple operator splits of
the form A[.] = rN=I A() [.]. From a computational standpoint, the success of the
technique relies critically on the choice of the specific operator split on the basis of
which the fractional step method is constructed. This choice is problem-dependent
and nearly always motivated by physical considerations, as illustrated below within
the context of a classical model problem.
61. Model problem: The coupled system of linearized thermoelasticity
This section provides an illustration of the ideas introduced above within the context of a concrete model problem, the coupled system of linearized thermoelasticity,
for which a complete mathematical theory currently exists going back to DAFERMOS
[1976]. Specifically, the goals are the construction of an operator split and the design of a product formula algorithm which render an unconditionally stable staggered
algorithm for this coupled problem of evolution. Conceptually, the results described
below can be formally generalized to the significantly more complex system of coupled thermoplasticity at finite strains. We shall do so in the next section by exploiting
the insight gained in the analysis of the present model problem.
The linearized system of thermoelasticity is obtained by coupling the hyperbolic
system of linear elastodynamics with the parabolic system of transient heat conduction,
via two terms: The linearized structural elastic heating and the thermally induced

The coupled thermomechanical problem

SECTION 61

461

stress field arising from thermal expansion. The resulting first-order system can be
obtained by linearization of the nonlinear theory described in the preceding sections,
with internal variables set to FP
1 and , _ 0, a = 1,..., nint, i.e., from the
quasilinear system of nonlinear thermoelasticity. The linearized conservation laws
(56.10), together with the elastic constitutive equations and Fourier's law of heat
conduction, yield the following first-order system (see, e.g., CARLSON [1972]):
i =v,

poi = div[CVu - AOm] + b in 2 x l,


co

= div[kV] -m

(61.1a)

Vv + r,

where u(., t) E (Ho (2))ndm is the linearized displacement field and 9(., t) L 2 (Q)
is the relative temperature field. In this statement of the problem po and co are the
reference density and the linearized heat capacity of the solid, respectively, C denotes
the linearized elasticity tensor and k is the conductivity tensor, which is assumed to
be symmetric and positive definite. In addition, m is a rank-two symmetric tensor
that couples the displacement and temperature fields via two terms: The linearized
structural elastic heating 'heat = -m Vv and the thermal stress field Other = -m.
For the isotropic theory, the coupling tensor m is given by the familiar expression
m = ndimal, where c > 0 is the coefficient of thermal expansion. Finally, b and
r denote the body force and the heat source term, respectively, both assumed to be
constant in Q x I. For simplicity the following Dirichlet boundary conditions will be
assumed throughout the analysis:
'0=0 onOa2x:1

and

u=O onaf2x:.

(61.lb)

The formulation of the coupled initial boundary value problem is completed by specifying the initial data
91t=o = 90,

ult= = u0

and

vlt=o = v0o

in 2.

(61.1c)

In keeping with the notation introduced in the preceding section, we let z =


{u, v, V} denote the primary variables and, according to the well-established functional setting for this problem (see DAFERMOS [1976]), adopt the function space
Z = Hl(Q)ndi m x L2 (Q)ndim x L 2(Q2).

(61.2)

The standard Sobolev norm of any z


Z will be denoted by lzilz. This norm
is induced by an inner product denoted by (., .). Recall that Ilzll is defined by the
standard expression
Ilzll 2

=J

[vl

Vu 2 +

Pl2]

df?,

(61.3)

462

J.C. Simo

CHAPTER V

46
CHPTERim
easily shown that for the homogeneous
where 2Vu 2 = E1i"'I i,jui,j. It canJ. beSimo
problem, obtained from (61.1a) by setting b = 0 and r = 0, a quadratic approximation
to the canonical free energy function introduced in Section 58 gives the functional
L: Z R defined as
(z) = 1 J

[POIV12 + Vu. CVu + co02] dQ.

(61.4)

The first two terms in (61.4) give the kinetic energy and the elastic stored energy in
the solid, respectively, while the last term in (61.4) agrees with the function used
in the analysis of the heat conduction problem in Example 59.1. The natural norm for
the problem is defined as lztlIz = 2(z, for any z E Z. As before, the associated
natural inner product on Z is denoted by (, *)z. Using Korn's inequality, see, e.g.,
CIARLET [1988], it is easily shown that 11 Hlz does in fact define a norm equivalent
to the standard Sobolev norm 11-IIon Z. This well-known fact implies, in particular,
that (.) > 0 in Z x 11.The analysis below rests on the following result.
THEOREM 61.1. The canonical free energy function defined by (61.4) satisfies the
following properties:
(i) For the homogeneous problem, L(.) satisfies the estimate
dt (z) = (z, A[z])z = -

VV * kV

df <

in .

(61.5)

It follows that (.) is a positive function that decreases along the dynamics, thus
qualifying as a Lyapunov function for the problem.
(ii) Let t e l - z() C Z, a = 1,2, denote the solutions of (61.1a) corresponding
to two initial data z~t ) Z. Then
Z(2) z <
1z( ) -Z(2)I1z
<

(1)>- z
14ZO

z(2 in

1,

so that the problem is contractive relative to the norm


(iii) The continuous problem has a unique solution t
data zo e Z.

(61.6)

z
-

2/I.)=
z for prescribed initial

PROOF. (i) Time differentiation of (61.4) and use of integration by parts together with
Eq. (61.1a)l and boundary condition (61.1b)1 yields
dt ()

v * div[mO9] + 9cor] dQ.

(61.7)

Inserting Eq. (61.1a) 2 into this expression gives


dt(z) = f

(-v.

div[m]

im

Vv + 9div[kVt9]) dQ.

(61.8)

The coupled thermomechanicalproblem

SECTION 61

463

Using integration by parts along with boundary condition (61.1b)l results in the cancellation of the first two terms in (61.8) and produces the a priori estimate (61.5).
(ii) The difference z = z() _ Z( 2 ) of the two solutions satisfies the homogeneous
version of the initial boundary value problem (61.1a), with initial data to = z ) -o2)
It follows that the flow t E II - z E Z satisfies the a priori estimate (61.5). Integration
in time yields the contractivity result (61.6).
(iii) Suppose that C - z('), ca = 1,2, are two solutions associated with the initial
data z0 G Z. The contractivity property (61.6) then yields 0 < Ilz() - z(2 ) lIz < 0,
which implies that z(1) = z(2 ) thus proving uniqueness. El
Consider next the numerical approximation in time of the thermoelastic problem
(61.1 a) via algorithms which inherit the contractivity property proved in the preceding
theorem. It is a fairly simple matter to construct a monolithic scheme that preserves
the contractivity property. In fact, in view of the results summarized in Chapter II, any
member of the sub-class of the Runge-Kutta methods characterized by the condition
that the matrix M be positive semidefinite (see Section 13) is a B-stable time-stepping
algorithm which, therefore, inherits the contractivity property. A typical example is
the generalized mid-point rule considered in Example 59.1, with 9 > . Our objective
is to construct a staggered scheme, to be interpreted as a product formula algorithm,
designed so that the contractivity property (61.6) is preserved.
(A) The conventional isothermal split. The obvious strategy in the design of a staggered scheme for the solution of (61.1la) is to partition the problem into a mechanical
phase, solved at constant temperature, followed by a purely thermal phase solved for
fixed displacement and velocity fields. As noted in SIMO and MIEHE [1992], this approach can be formally interpreted as a product formula algorithm based on the split
A[-] = A(s [.] + A2 [], where
V

As)O [z] =

div [CVu - m ]
(61.9)

A[Z]
div[kV]

m * Vv

A first observation is that neither Al) [] nor As2o) [] generates a quasicontractive


semigroup in Z. For instance, if F()[.] is the semigroup associated with AsW) [], a
calculation analogous to that used in the proof of Theorem 61.1 reveals that

dt
dt (

[z] ) =(A
' ([z]
ISO

z)

=-div[md],v)

~CIll112

VZ(EZ
2

(61.10)

J.C. Simo

464

CHAPTER V

for some constant C, since the inequality 11V9 11L2 Ml|10L2 does not hold for some
constant M > 0 and V0 C H l ( 2) . The Lumer-Phillips theorem (see, e.g., PAZY
[1983]) then implies that Ai(s [] cannot generate a quasicontractive semigroup, thus
breaking the contractive structure of the original problem. In fact, further analysis
reveals that each operator does not even generate a semigroup in Z implying that
each problem, when considered separately, is improperly posed.
In spite of this negative result, for the discrete problem a mesh-dependent constant
Mh = M(h-l) does exist, such that llV9hllL 2 < Mhll 09hlL2 as a consequence of a
standard inverse estimate for finite element spaces. This last result implies, therefore,
that time-stepping algorithms based on the isothermal split will work in practice for
a given mesh-size h (maximum element size), although the nonuniformity of the
estimate in h renders the method conditionally stable at best.
The practical implications of this stability restriction will be illustrated within the
context of a specific model problem. Consider a fractional step method in which each
phase in the product formula is approximated by the (unconditionally stable) CrankNicholson scheme, together with a conforming finite element approximation in space
Z h C Z defined by piece-wise linear polynomials. The resulting product formula
algorithm is clearly consistent and first-order accurate in time. Sufficient conditions
for stability are provided by the following result (ARMERO and SIMO [1992]).
THEOREM 61.2. Consider a generalized Crank-Nicholson algorithm associatedwith
(61.1 a), defined by formulae
n+l

Un = AVn+l/27

po (Vn+ -Vnh)

= At div [CVU+ 1 /2 -_19m],

CO(On+ - Vn) = At(div[kVOn+l/2]

-m

(61.11)

. Vvh+,),

where f E [0, 1] is an algorithmic parameter and the superscript h refers to the


discretized field via a conforming Co-finite element approximation defined by piecewise linear polynomials. The scheme is stable provided that
(i) The algorithmicparameter is restricted to the value = 1.
(ii) The following restrictions on the time step hold
At < 2-pc0

ljmllM

CFL = aisohAt

2
-- m

(61.12)

where ai,, is the largest wave speed for the isothermal problem, CFL denotes the
Courant number 6 > 0 is a normalized parameter that measures the strength
of the coupling and M is the constant in the inverse estimate 11V9h L2 (Q) <

(M/h)ll hllL2().

For the isotropic case, aiso = /(A + 2 l)/po and = m 2 /(2/ co), where A, / > 0
are the Lame constants and m = ml with lImllo, = Iml.

The coupled thermomechanicalproblem

SECTION 61

465

PROOF. Sufficient conditions for stability can be derived via the energy method,
whereby an algorithmic norm of the numerical solution to the problem is constructed
which does not increase along the algorithmic flow. Multiplying Eqs. (61.11)2,3 by
n+l/2 and 9n+l/, 2 respectively, and integrating by parts yields the identity
Ch (znh+l )

h (zh)

-(1
=

3)At[(div [d+1 ,] v n)-

2(div [9h+ / 2],

Vh+l)]

(61.13)

- (kV n+l/2, V6n+/ 2),


with the function h: Zh - iR defined by
Lh(zh) = (zh) - At(div[mth],vh) Vzh E Zh .

(61.14)

For l = 1, the first term on the right-hand side of identity (61.13) vanishes leading
to the estimate h(zh+l) - (zn) < 0 implied by the assumption of positive definiteness on k. The property that Lh (.) decreases along the algorithm flow for /3 = 1
implies stability of the algorithm, provided that the function h(-) defines a norm,
see RICHTMYER and MORTON [1967]. The stability analysis then reduces to estimating
a lower-bound that ensures positive definiteness of the function h(-). To do so, we
first estimate an upper bound for the additional term in (61.13) using integration parts
and standard inequalities as
(div [hm], vh) <j (m. V0 h, vh) < mlVIh{L2lvhL2.

(61.15)

Inserting this estimate into (61.14) and making use of Young's inequality yields
Ch(zh) > (zh)

Atlmll.
2 cop

(zh) _ At llll

[(, pllv
01

L) (C<oI

I hL)]

[PIIv [hI + o0h 2 /],

(61.16)

for any constant y > 0. The crucial result used in the proof is the standard inverse
estimate IlV9h llL2 < (M/h)[dhllL2 , for some constant M > 0, valid for conforming
linear finite element spaces see CIARLET [1978]. Substituting this inverse estimate into
inequality (61.16) gives
h (zh)> I
x (

hVU
. CVuh + [1

I2mlI[MAth

PoIvh12 + M Coh12) dQ.

466

J.C. Simo

CHAPTER V

Setting y = M/h > 0, a sufficient condition for positive definiteness of h(-) is


obtained by requiring that the bracket in this expression be positive, which gives
condition (ii) in the theorem. D
Observe that the algorithmic norm h () is not uniformly equivalent to the natural
norm in Z, even for a fixed value of the Courant number satisfying the stability condition (61.12). Thus, strictly speaking, we cannot bound (uniformly) the algorithmic
norm of the solution at some t in terms of the natural norm of the initial conditions.
As a result, it is not possible to infer conditional B-stability of the algorithm. The
constant M can be estimated first for a single element and then maximizing the result
over all the elements. This technique is also used in the derivation of the inverse
estimates (JOHNSON [1987]). For the 1D case, the optimal value is /12 (> 1).
REMARK 61.1. Necessary conditions for stability are provided by a classical von Neu-

mann analysis of the finite difference stencil associated with the one-dimensional problem, see RICHTMYER and MORTON [1967]. Denoting by E the Young modulus and
considering, for simplicity, lumped approximations to the mass and capacity matrices,
a von Neumann stability analysis yields the following stability condition (ARMERO
and SIMO [1992]):
At

A
h

2 /poo

P2
jml

CFL = aiso

At

<-

'

(61.17)

where ais = V/Epo is the uncoupled or isothermal wave speed and 6 = m 2/(Eco)
is the nondimensional parameter that measures the strength of the coupling. In the
multidimensional isotropic case, ai,, is to be understood as the isothermal longitudinal
wave speed, i.e., ai,, = /(A + 2 p)/po. Note that the adiabatic wave speed aad is given
by aad = a 0(1 + 3). The necessary condition (61.17) should be compared with the
sufficient condition (61.12).
REMARK 61.2. Similar stability analyses can be performed for the quasistatic problem,
formally obtained by setting po = 0 in (61.1) while keeping co > 0. We assume

that k
0 since, otherwise, the problem becomes rate-independent and reduces to
elastostatics with adiabatic elasticities. To give an indication of the restrictions imposed
by stability considerations, consider the simplest situation afforded by a single onedimensional finite element discretization with piece-wise linear polynomials. In this
situation, the stability requirements are easily shown to imply
At
m 2 - 2Eco
jZ
2Ek
h ~> 2Ek

2k At
C

(S (2

),

(61.18)

2. Thus, for strongly coupled


a condition which is automatically satisfied if 6
problems (corresponding to 6 > 2), condition (61.18) is seen to place a restriction of
At/h 2 reminiscent of the conditional consistency condition arising in certain explicit
unconditionally stable methods. An alternative interpretation of this stability condition
arises when the product algorithm in the quasistatic regime is viewed, following the

The coupled thermomechanical problem

SECION 61

467

terminology of ORTEGA and RHEINBOLDT [1970], as a one-pass block Gauss-Seidel


method for the solution of a (coupled) linear system of equations. It is easily shown
that the condition for convergence of this scheme is precisely given by the result in
(61.18). We refer to ARMERO and SIMO [1992] for further details.
The preceding analysis demonstrates that algorithms based on the isothermal split
(61.9) are not suitable for strongly coupled problems, since the stability restriction
phrased in terms of the Courant number becomes increasingly restrictive the higher
the strength of the coupling, as measured by the nondimensional parameter . On the
other hand, these algorithms become rather useful in the solution of weakly coupled
problems for which the stability conditions derived above do not place severe restrictions on the time step. Nevertheless, an implicit algorithm possessing only conditional
stability, as in the present case, is always difficult to justify from a computational
standpoint. We show below that these difficulties can be completely circumvented at,
essentially, no additional cost.
(B) The adiabatic operator split. The proof of a priori stability estimate (61.5) in
Theorem 61.1 relies on the cancellation of the coupling terms t9m. Vv and v. div[9m]
arising as a result of integration by parts, see Eq. (61.8). Preservation of this property,
suggests considering an operator split of problem (61.1) in which the coupling terms
-div[nm] and -m * Vv appear in the same phase. These considerations lead to the
alternative split A[.] = A()[.] + Ad)[], where

A( [z]:

odiv [Q [u] - n],

* Vv
(61.19)
Ad [z]

1div[k V]

Thus, the treatment of the structural elastic heating term Heeat = -m


Vv is the
key difference between the conventional split (61.9) and the operator split defined
by (61.19). As a result, the temperature field in the mechanical phase of the split
(61.19) now changes in such a way that the (linearized) entropy remains constant in
this phase, hence the denomination adiabaticsplit. From a mathematical standpoint,
this seemingly unimportant modification has profound implications. First, A 2) [-] now
generates a semigroup (in fact, the analytic semigroup of classical heat conduction)
and, as a consequence, A( )[] also generates a semigroup. Second, the semigroups
generated by Al )[-] and Ad2)[.] are contractive relative to the same natural norm
induced by (.). In fact, denoting by t - F(a) [-], a = 1,2, the flows generated in

468

J.C. Simo

CHAPTER V

each of the phases, the same calculation used in the proof of Theorem 61.1 now gives
the two estimates

dtlF()[z]) =0 and

+d(F(2)[z])
=

dt

V09 kV,9dQ2 < 0.

(61.20)

From result (61.20)1 it follows that the infinitesimal generator A(J) [-] defining the
mechanical phase produces a semigroup of isometries relative to the natural norm
induced by C(). Consequently, no dissipation takes place in the mechanical phase,
an obvious conclusion since the total entropy is held constant, while the dissipation
in the subsequent thermal phase is precisely the total dissipation Doon generated by
heat conduction.
From the results described in the preceding section one concludes that a stable
algorithm KI[at for problem (61.1) can be obtained merely as the product formula 1Kat =
KI(2)
At o K(
At of two stable algorithms consistent with each of the phases in (61.19). The
product algorithm is B-stable if each of the algorithms is B-stable. Convergence of
the product formula (in the linear case) is an immediate result of the Lax equivalence
theorem, see RICHTMYER and MORTON [1967].
REMARK 61.3. In principle, the quasistatic case is not encompassed by the estimates
in (61.20). A separate stability analysis confirms the unconditional stability of the
product formula for the quasistatic case. The reader is referred to ARMERO and SIMO
[1992] for the details involved in the proof of this result.
Although both phases in the operator split (61.19) involve the evolution of the temperature field, it will be shown below that the actual implementation of the product formula reduces the first phase to a purely mechanical problem with modified (adiabatic)
elasticities, thus retaining the computational advantages inherent to the isothermal split.
For a typical time step [t,, t+l] with prescribed initial data z,, = {Ur, vn, n} e Z,
the key idea is to enforce the strong form of the temperature evolution equation in
the mechanical phase, even though the evolution equation for the mechanical field is
typically enforced weakly in a standard finite element discretization of the problem.
Doing so, the temperature field in the mechanical phase can be solved for explicitly
in terms of the displacement field as
9=

hn

1
- -m
CO

V[u - u,,],

for (x, t)

Q2 x [tn, t + At],

(61.21)

since it = v. Inserting this result into the second evolution equation in the mechanical phase, which defines the evolution of the mechanical field, yields the reduced
momentum equation
pot = div

(C +-m

m VU

+ b- div9nm+

(m m)Vu],
1CO
YIv 11

(61.22)

SECTION 62

The coupled thermomechanical problem

469

which involves only the displacement field. Equations (61.21) and (61.22) comprise
the reduced mechanicalphase in the operator split (61.19). The solution of this problem is then taken as initial conditions for the thermal phase in (61.19), which remains
unchanged. Observe that the reduced momentum equation (61.22) is essentially identical to that arising in the mechanical phase of the conventional isothermal split (61.9),
with the elasticity tensor C replaced by the adiabatic elasticity tensor C+(1/co)m m
and the body force b replaced by the effective body force in (61.22). The same elimination process can be performed if the momentum equation in the mechanical phase is
written in weak form. A scheme directly formulated on the discretized system arising
in soil consolidations problems, related to the product formula algorithm based on the
split (61.19), is suggested in ZIENKIEWICZ, PAUL and CHAN [1988].
62. Generalization: A staggered scheme for nonlinear thermoplasticity
The goal of this section is to generalize the product formula algorithm based on an
adiabatic operator split of the linearized thermoelastic problem, as described in the
preceding section, to the full nonlinear system of multiplicative plasticity. An entirely
analogous extension of the product algorithm based on a formal isothermal operator
split of the problem of evolution is also possible. However, in view of the inherent stability restrictions alluded to above, and since no additional computational advantages
are to be gained from this class methods, further details on the construction of such
an extension are omitted. The interested reader is referred to SIMO and MIEHE [1992]
for a full account of staggered schemes for thermoplasticity based on an isothermal
split of the problem.
(A) Multiplicativeplasticity as a first-ordersystem. The first step in the construction
of the staggered scheme is to cast the initial boundary value problem for multiplicative
plasticity as a first-order problem of evolution. Since the developments in Section 61
show that the constant entropy condition plays a key role in the design of the operator
split, it is convenient to replace the absolute temperature field 09 by the entropy
function H of the solid as the primary variable. In view of the decomposition H =
He + H P , we replace the set Z of primary variables defined by (58.1) with the set of
conservation/entropy variables Z defined as
Z = {o, p, He, HP}

along with r = {FP,,},

(62.1)

where p = poV denotes the material momentum. All the remaining variables in the
problem are defined in terms of Z and r by the kinematic and constitutive relations
derived in Section 57. In particular,
(i) The elastic kinematic variables Fe = DqpFP- I and Ce = FeTFe.
(ii) The stress fields P = Fe[23ceE(Ce,He)]F p- T and r = PDp-T.
(iii) The temperature
= aHeE(Ce,He) and the nominal heat flux Q =
Q(Ce, o).

Whenever any of the variables listed above appears, it will be implicitly understood
that the appropriate relation linking the specific variable to the primary variables

J.C. Simo

470

CHAPTER V

(Z, F) in the problem is employed. With this convention, the evolution equations
for the internal variables r and the conservation/entropy variables Z can be written
as follows. Define the nonlinear operators A[.] and G,[.], t = 1,..., m, via the
expressions

DIV[P]
-A[ZDIV[Q] +

Dmech

A1
Dther
G
[2, r

(62.2)

Fe-I [arf,] D
aqlit A,

Using the kinematic relation FP = Fe-dDDo, which follows from the result in
Theorem 45.1 under the assumption of zero plastic spin, the associative flow rule
(57.25) gives the local evolution equations
m

r=

in

yGZ,F]

(62.3)

x ,

with the multiplier 7y > 0 defined either by the Kuhn-Tucker conditions (57.26) in
the rate-independent theory, or by constitutive equations analogous to (57.28) in the
rate-dependent theory. The mechanical and thermal dissipations in definition (62.2)1
are understood to be given by expressions (57.14b) or, equivalently, as
m

Dmech = Z

G,[Z,r]T
h=Ll

and

Dthle

= 9eyZ"af

0,

j=1

(62.4)
where we have set T = [pTqi .. qfn,]. With this notation in hand, the first-order
problem of evolution for multiplicative plasticity takes the form
d=A[Z,r]
jjZA[dtAZ[FZ.

r ]

ins2xa,
in
x(62.5)

Zt=0 = ZO in Q,
with the internal variables r defined via the local evolution equations (62.3) and subject to the initial conditions Flt=o = {1, O} in Q. In the formulation of the staggered
method described below, it is essential to regard the internal variables r as implicitly
defined in terms of the conservation/entropy variables Z via the rate equations (62.3).

The coupled thermomechanical problem

SECTION 62

471

Recall that an identical point of view is adopted in the formulation of the return mapping algorithms described in Chapter IV. Effectively, therefore, the only independent
variables in the problem are the conservation/entropy variables Z.
The stability considerations in the preceding section can be formally extended to
the problem at hand by considering the canonical free energy function for the homogeneous problem (i.e., B = 0 and R = 0 in S2 x 1[together with T = 0 on ET x II
and Q on EQ x Il), driven only by initial data now expressed in terms of the conservation/entropy variables Z. This change in variables is accomplished via the Legendre
transformation (57.8) and yields the following expression for C() defined by (58.3):

(Z

)=

[I p2/Po +

E(CeHe) +

9l(o,) dO.
OHe]

(62.6)

A straightforward computation again verifies that time rate of change of (o) obeys
the a priori estimate (58.6) along the flow generated by the homogeneous problem.
(B) The adiabaticoperatorsplit. Motivated by the strategy used in the design of the
adiabatic split for linear thermoelasticity, we consider an operator split with mechanical
phase at fixed total entropy and a thermal phase at fixed configuration. To implement
this idea, we define the vector fields

A(d[,r =

DIV[P]

(62.7)
0

AdI)[[ Z.T-

-DIV[Q]

Dmech

D)ther

and consider the following two problems of evolution:


Problem 1:

A= )[ :G],

El
/1=1

(62.8)
Problem 2:

-Z=Ad) [Z,
dt

=Ea

G,[

,Y]

j/=1

Observe that local evolution equations governing plastic flow are appended to both
problems as a means of defining the internal variables F in terms of the primary
variable Z in each phase of the split. The mechanical phase defined by Problem 1
takes place at constant total entropy since H = He + HP = 0, hence the name
adiabatic split attached to the partition of the problem specified by (62.8). The design

J.C. Simo

472

CHAPTER V

condition Hp = 0 placed on the mechanical phase requires further elaboration. As in


the thermoelastic problem, a change in the temperature field occurs in the mechanical
phase in order to enforce the condition He = 0 of constant structural entropy. If
plastic flow also takes place, then there exists at least one /lo e {1,.. ,m} such
that y"0 > 0. As a result, we would have y1 0aef 0 > 0 and, therefore, a nonzero
evolution fHP 0 of the configurational entropy, unless the yield criterion is held at
constant temperature. It follows that the design condition
ftP = 0 A= aDf, =

( = 1,..., m) for Problem 1.

(62.9)

In other words, the plastic flow takes place at fixed temperature in the mechanical
phase. This condition becomes critical for preservation of symmetry in the linearized
(incremental) problem.
By construction, it is clear the A[] = Al)[
A []). Therefore, according to the
general results in Section 59, all that remains is to show that each of the two problems
defined by (62.7) generate flows that obey the a priori estimate L[Z, F] < 0 for the
split to be (formally) well-posed. The verification of this property is given in the
following result.
THEOREM 62.1. Let t - (Z(a),F(r)), a = 1,2, denote the flows generated by each
of the problems in (62.9). Then
dt
dt

] =-j

)
12 smechdQ
< 0,

(62.10)
el
( 2)

d L [ (2 ), (2 )] =

[D)ch

D(] dQ h

0,

where Dnmech
0 (a = 1, 2) is the mechanical dissipation in Problem a, computed
using expression (62.4)1, and Dn = _Q( 2). GRAD[0( 2 )]/9(2) > 0 is the dissipation
due to the heat flux Q(2) in the second problem.
PROOF. Since the structural part H e is held constant in the mechanical phase, the
canonical free energy (.) coincides, modulo the constant term O90 He, with the total
energy in the functional defined by (41.7) for the purely mechanical problem. Result
(62.10)1 is therefore a restatement of the energy decay estimate in Theorem 41.1 for
the mechanical problem, since the yield criterion is held at constant temperature as a
result of condition (62.9). Result (62.10)2 is proved by making use of the relation
d [(Ce

,He) + -i()

= -Dmech

+ (

- 9

)He,

oHe] V=constant
(62.11)

together with the evolution equation for He in Problem 2 and integration by parts.
The computations involved are essentially identical to those in the last part of the
proof of Theorem 58.1 and hence are omitted.

The coupled thermomechanicalproblem

SECTION 62

473

An unconditionally stable algorithm KAt for coupled thermoplasticity at finite strains


is therefore formally constructed as the product formula IKAt = (2) o K(1) of two
unconditionally stable algorithms IK(), a = 1,2, consistent with the two problems in
(62.7). A brief indication on how this product algorithm is implemented in practice is
given below.
(C) Implementation of the product formula as a staggered scheme. Concerning algorithm KI1 ) for the mechanical phase (Problem 1), the implementation relies on the
observation that the resulting algorithmic problem is essentially identical to that arising
in the treatment of the purely mechanical theory, as described in Chapter IV. The only
difference results from the replacement of the stored energy function W(Ce) in the
purely mechanical problem with the internal energy function E(C e, H e ) at constant
structural entropy H e for the problem at hand. Assume that an explicit expression for
E(Ce, H e ) is available. The temperature change in the mechanical phase within a
typical time step [t,, t, + At] is then computed via the local constitutive equation
O = aHeE(Ce

H e) He=He,

(62.12)

where Hn is the structural entropy at the beginning of the time step. Similarly, the
nominal stress P in the mechanical phase is evaluated via the local constitutive equation

= FSFP -T

where S = [aeE(Ce, He)] He=H

(62.13)

The internal variable FP in this expression is expressed in terms of the deformation in


the mechanical phase via the return mapping algorithms described in Chapter IV. The
Kirchhoff stress field r appearing in the yield criterion is computed via the standard
formula r = FeSFeT. Inserting the resulting expression for the stress field into the
algorithmic version of the weak form (56.18)1 of the momentum equation produces
a nonlinear algebraic problem for the deformation field. The iterative solution of this
problem involves the computation of the linearized weak form of the momentum
equations. All the expressions given in Chapter IV remain valid, with the elasticity
tensor C = 4C2
C e))W(now replaced by the rank-four tensor
Cad =

[4ae eE(Ce, He) He:He,

(62.14)

which defines the adiabaticelasticities of the solid relative to the intermediate configuration. The corresponding expression for the adiabatic elasticity tensor cad in the
current configuration is obtained via the standard push-forward transformation with
the elastic deformation gradient Fe.
REMARK 62.1. In practice, one is often given the free energy function (Ce , 6) in
place of the internal energy function E(C , He). In such a situation, application of

474

.C. Simo

CHAPTER V

the preceding strategy requires the local solution for the absolute temperature field of
the (generally nonlinear) constitutive equation

9 (Cle) - H, = 0 to obtain

O(Ce,He)IH=He.

(62.15)

With this expression in hand, defined implicitly for the general case via a local iterative solution scheme, the internal energy function is computed via the Legendre
transformation as
E(e,

He)

(l

e, (e

He)) ++=
a(ee

He) e

(62.16)

l , ) the rank-four tensor of isothermal elasticities,


Denoting by Ciso = 4 a2_c(ce
a straightforward application of the chain rule gives the expression
Cad =

Ciso + - (2)

( (2e9)l

(62.17)

is the heat capacity of the


for the adiabatic elasticity tensor, where c = -ae9o
solid. As noted in Example 57.1, all the preceding expressions become closed-form
under the assumption of constant heat capacity. In particular, the solution of (62.15)
for the temperature field is given by (57.19), while expression (59.15) for the internal
energy is given by (57.20). Finally, expression (62.17) for the adiabatic elasticities is
trivially evaluated by noting that 2 ae,,t = -2u,,M(Ce).
The thermal phase of the product formula algorithm (Problem 2) is performed by
applying the algorithm KI(2) with initial data defined by the solution of Problem 1. In
view of (62.7) both the deformation and the velocity field remain fixed at the solution
computed in the mechanical phase, while the temperature field is updated by solving
the heat conduction problem in conservation form. To do so, it is necessary to evaluate
both the mechanical dissipation Dmech and the thermal dissipation t)int induced by the
evolution of plastic flow. This results in a further update of the internal variables r in
the second phase, performed by applying the return mapping algorithms described in
Chapter IV. Observe that plastic flow will, in general, take place during the thermal
phase of the product formula as a result of thermoplastic softening. A detailed stepby-step description of the implementation of the product formula algorithm is given
in the appendix in ARMERO and SIMO [1993]. The reader is directed to this reference
for additional information.
REMARK 62.2. A noteworthy feature of the product formula algorithm outlined above

is the possibility of exact preservation of the global conservation laws of momentum


for the pure traction problem (see Theorem 58.2). This is accomplished merely by
employing in the mechanical phase a momentum conserving algorithm of the type
described in Chapter IV. Concerning the design of spatial finite element discretizations, all the results described in Chapter IV carry over to the present setting without
modification.

The coupled thermomechanicalproblem

SECTION 63

475

63. Representative numerical simulations


The performance of staggered algorithms based on either the classical isothermal split
or on the proposed adiabatic split is assessed below in three representative simulations
which closely replicate the stability estimates derived in the preceding sections while
confirming the excellent performance anticipated for the adiabatic split. The threedimensional simulations described below are performed with a slightly more general
version of the model of coupled thermoplasticity labeled as Model II in Section 57.
Specifically, the internal energy function is assumed to take the following form:
E(

be

, 6, He) =

K[log J]2 + 4
- co

'l|log be

+ ndimt9eo0 log J

( - exp [(He - ncdimna log J + Ke ()) /co])

(63.la)

+ [K(, 0o) - OoKe(()],


with the temperature hardening potential K((, 09) assumed to be given by
K(I, ) =

h(e) 2 + [yo(0) - y ()] H(),

(63.1b)

where H(~) = - (1 - exp[-6])/8, and the function Ko(() defined by the closedfrom expression:
Ke(6) = h(Oo)Wh~2 + [YO(90)wo - Y (O0)Wh] H().

(63.lc)

Observe that this expression for the free energy incorporates the more general
situation in which the hardening potential is also temperature-dependent. The model
is completed by considering the standard von Mises yield criterion with isotropic
hardening, written as
f(-ir, q, 9) =

3ldev[r] + q-

(63.2a)

(),

with the stress-like hardening variable q defined by the potential relation


=

-{h(O)J + [yo(O) - y.()]

(1 - exp[-])}.

(63.2b)

For simplicity, attention is restricted to material properties that experience a linear


thermal softening, as defined by the relations
yo(O) = yo(Oo) [1 - wo( -

o)],

h(O) = h(6o) [1 - wh(O - 0)],


yoo()

= y.

(Oo) [

- Wh(o -

(63.2c)

0)],

where wo and Wh are material constants. In addition, to make the model agree with
widely used formulations of thermoplasticity, the term Dint in the heat conduction

476

J. C. Simo

CHAPTER V

equation has been replaced by a fraction of the total plastic power, as defined by the
term X-r . dP. Here X is a numerical factor estimated in TAYLOR and QUINNEY [1933]
as X = 0.9.
(A) D linear thermoelastic vibration problem. The objective of this simulation,
taken from ARMERO and SMO [1992], is to demonstrate the unconditional stability
property of the adiabatic split and the conditional stability estimates for the isothermal split by means of a D linear thermoelastic vibration problem. Consider a onedimensional problem governed by the equations of linear thermoelasticity written in
the following nondimensional form:
a2u a a2f
02 -- 2

a
f

a
a-

2
aT
and

a2w
at 2-

ar

a2u

(63.3a)

The dimensionless variables are defined as


-=-x,

cais

T=

ca 2

cEai,

= -u
cmdo

and

=-.
90

(63.3b)

Recall that the dimensionless parameter 6 measures the strength of coupling present
in the problem, and that the restriction on the Courant number are defined in terms
of this parameter. The problem is posed in the interval [O, L], with the boundary
conditions:
tU(0, T) =

'u(L, ) = 0 and 09(0, T) = (L, ) = 0,

(63.4)

supplemented by the initial data


u((, 0) = 0,

v(f, 0) = sin

7t6

and

(f,0) =0.

(63.5)

The uncoupled or isothermal solution to this problem corresponds to free vibration


in the first mode, with a phase speed of di, = 1 or, equivalently, with a period
Tis = 2L. By contrast, the coupled solution exhibits amplitude damping as well as
dispersion. It is easily concluded that low wave numbers propagate adiabatically, i.e.,
at a phase speed near d2 = 1 + 6, while high wave numbers propagate effectively
in an isothermal manner. This result is well known (see, e.g., ACHENBACH ([1987],
p. 394) or PARKUS ([1976], p. 100)), and implies that the damping induced by thermal
effects has a significant effect only on the high wave numbers. Since the problem
under consideration involves low wave numbers, one expects a decrease of the period
of the order T Tis//I1 + , with a moderate amplitude decay.
The results reported below correspond to simulations performed with the values L =
100 and 6 = 1. This value of 6, although leading to a relatively strong coupling for this
class of problems, is often considered as a suitable choice in the standard literature on
the subject assessing the performance of algorithms. The numerical solution involves

The coupled thermomechanicalproblem

SECTION 63

477

CFL = 1

Id

CFL = 3

60

60

40

40

20

20

-20
-40
-60

-20

AD ---

Z2

SIM - -uncoupl. ---I, 1,


1

50

-40
, ,

100 150 200 250 300

-60

50

100 150 200 250 300


Time,

Time, r

1.0

1.0

0.5

0.5

I1

111111111111
111

II

'
li IlIllllll
I ll I
0

-0.5

-0.5

-1.0

-1.0
0

50

100 150 200 250 300


Time, r

SAD -

... uncoup].
_ iiiiiII
I
IuncoupI.I
0

50 100

'1'"00250 300

Time, 7

FIG. 63.1. Coupled D vibration problem. Evolution of the displacement at L/2 and the temperature at
L/4 obtained with different algorithms for CFL = 1 and CFL = 3 (CFLcrit = 2 for the isothermal split).
Coupling parameter 6 = 1.

a spatial discretization consisting of 100 C linear finite elements (i.e., h = 1.0) and
three time-stepping strategies:
(i) The adiabaticsplit, with a Crank-Nicholson scheme for the adiabatic mechanical phase and a backward Euler algorithm for the heat conduction phase. Recall that
the resulting single pass staggered scheme is unconditionally stable.
(ii) The isothermal split, with the algorithm defined by Eqs. (61.11). Recall that
the resulting staggered algorithm is only conditionally stable.
(iii) A simultaneous solution (or monolithic) scheme obtained by applying a CrankNicholson scheme to the full coupled system written in conservation form. This
scheme is unconditionally stable.

478

J.C. Simo

CHAPTER V

20

10
Ar

1'

-10

-20
0

50

100

150

Time,

200

250

300

FIG. 63.2. Coupled D vibration problem. Accuracy of the adiabatic split (AD) with respect to the monolithic
scheme based on a simultaneous solution (SIM). Coupling parameter 6 = 1.

Lumped mass and capacity matrices are used throughout the simulation. Two numerical simulations are considered, corresponding to the values CFL = 1 and CFL = 3
respectively, which are below and above the critical value CFLcrit = 2 for stability
of the isothermal split. Figure 63.1 shows the displacement i at the center of the
bar = L/2 and the temperature at = L/4 computed in these two cases. We
observe a good agreement of the three solutions in both displacements and temperature for CFL = 1. On the other hand, while the adiabatic and simultaneous solutions
agree perfectly for CFL = 3, the isothermal split shows disastrous oscillations leading
eventually to overflow. This unstable behavior confirms the conditional stability of
this split and replicates closely the estimate (61.12). It is interesting to observe that
the temperature appears to start oscillating first in this unstable computation, then
followed by oscillations in the displacements. Note also that the damping and period
decrease of the stable solutions agree with the comments above.
A more detailed comparison between the solutions obtained with the adiabatic staggered algorithm and the monolithic scheme is contained in Fig. 63.2. These results
demonstrate the good accuracy exhibited by the adiabatic split. Figure 63.3 shows the
spatial distribution of the displacements and temperature at T = Tis/4 = 50 and r =

3Ti,/2 = 300 for the three strategies with CFL = 1. The physical dispersion present in
the problem is apparent. Again, a good agreement is found between the three solutions.
We remark that results not reported here obtained in a series of numerical tests involving the high wave number range (i.e., wave propagation problems) confirmed also the
same good accuracy and unconditional stability properties of the adiabatic split.

SECTION 63

The coupled thermomechanical problem

479

r = Tji/4 = 50

40

0.6

'

0.4
'; 30
co+=

20

0.2
0

11

ADIS -SIM - - uncoupl. ---.

-0.2

i 10
-0.4
0

20

40

60

80

-0.6

100

[,

20

40

60

80

100

20

40

60

80

100

r = 3Tis/2 = 300

0.2
'1

'-

0.1

co

-1
-2
0

20

40

60

80

100

-0.1
-0.2

FIG. 63.3. Coupled ID vibration problem: Spatial distribution of the displacement U and temperature i at
T = 50 and T = 300 (CFL = 1). Coupling parameter 6 = 1.

The conditional stability property of the isothermal split is of special concern in


situations for which the response is dominated by low wave numbers, as in the present
example. The reason is that, in the uncoupled case, problems of this type (i.e., structural dynamics) are often solved in practice with implicit algorithms and large values
of CFL. If the same approach is used in the coupled problem, condition (61.12) may
be violated leading to unstable computations.
(B) Dynamic impact of a thermoplastic cylindrical rod. This example corresponds
to the dynamic impact of a three-dimensional bar on a rigid frictionless hot wall.
The goal of this problem is to check the performance of the algorithms based on the
adiabatic split in this dynamic setting.

480

J.C. Simo

CHAPTER V

FIG. 63.4. Dynamic impact of a cylindrical rod. Discretization of the reference configuration (972 3D mixed
finite element brick).

TABLE 63.1

Dynamic impact of a cylindrical bar: Material properties.


Bulk modulus

130.

GPa

Shear modulus
Flow stress

/p
Yo

43.3333
0.40

Linear hardening
Density
Expansion coefficient
Conductivity
Specific capacity
Flow stress softening
Hardening softening

h
Po
c

0.10
8930.
1.0 x 10 5

GPa
GPa
GPa
kg/m 3
K-

k
c,
wo
wh

45.
460.
2.0 x 10 3
0.

N/sK
m2 /s2 K
KK-'

The bar considered in the simulation has a length of lo = 32.4 mm and a circular
cross section of ro = 3.2 mm. Figure 63.4 shows the reference mesh. A quarter of
the bar is discretized, with 972 isoparametric 8-node tri-linear bricks with piece-wise
constant pressure and volume along with a tri-linear interpolation for the displacement
field, as described in Section 45. The model of J2 -flow theory summarized at the
beginning of this section is used in the simulation, with material properties recorded
in Table 63.1. The initial velocity of the bar is vo = 0.227 mm//t s along the axis of
the cylinder. The temperature at the free face is assumed fixed at the reference value
090 = 293.15 K, while the wall temperature is = 00
i + 100 K. The lateral face of the
bar is assumed to be thermally insulated.

The coupled thermomechanicalproblem

SECTION 63

481

t = 40 is

t = 20 s

TEMPERATURE[K]
1 50E*Cl

2.90E-C1
4.33E01
7.IE-0 1
8 5E-01

t = 60

us

t = 80

FIG. 63.5. Dynamic impact of a cylindrical rod. Relative temperature distribution at t = 20, 40, 60, 80 / s.

All the deformed configurations are at the same scale. Observe that the wall is hot, at a temperature
100 + o0K.

Simulations are performed with both the adiabatic and isothermal splits. In both
cases, the dynamical mechanical phase is integrated with the standard Newmark
= , Y =algorithm, i.e., trapezoidal rule (see SIMO [1991] for a detailed discussion of other alternatives) with a lumped mass matrix. The thermal phase is integrated
by a backward Euler scheme in both simulations, together with a lumped capacity
matrix. The simulation is carried out for t [0, 80] u s, with equal time increments
of At = 1.25 ts.

482

J.C. Simo

CHAPTER V

t = 20 I/s

t = 40

EQ P
z_

-__

STN
OE-01
4.8OE-01

1 24Et00
1.62E00
200E+00

t = 60 s

t = 80

FIG. 63.6. Dynamic impact of a cylindrical rod. Equivalent plastic strain distribution at t
20, 40, 60, 80 p s. All the deformed configurations are at the same scale.

Figure 63.6 shows the temperature distribution as well as the deformed configurations at t = 40/ s and t = 80 s, obtained with the adiabatic split. Figure 63.7 shows
the maximum radial displacement versus time for both simulations. We observe a
perfect agreement between both simulations, which permits us to conclude that the
presented adiabatic split results in a good numerical accuracy in this dynamic context
as well. We note that the same remarks pointed out in the previous section hold in this

483

The coupled thermomechanical problem

SECTION 63

Isothermal

Hot wall

*r:L

Cd
6r

20

40

60

80

Time [ps]
FIG. 63.7. Dynamic impact of a cylindrical rod. Equivalent plastic strain distribution at t = 80 s obtained
for a hot wall and an isothermal simulation (same scale of temperatures as in Fig. 63.6). The lower plot
depicts the maximal radial displacement versus time for these two solutions.

J.C. Simo

484

=
11~~~~~
fr

CHAPTER V

f U~~~~i

R
K
R
R

q,=0
s

F-

SE

K-

fr
Sr

q.= 0
E

tU,

I I

banding. Reference configuration and boundary conditions.


qFIG.= 0--shear
FIG. 63.8. Plane strain nearly adiabatic shear banding. Reference configuration and boundary conditions.

case. The isothermal split presents conditional stability in these dynamic simulations;
but, as shown in ARMERO and SIMO [1992], the stability condition is inversely proportional to the strength of coupling. For these weakly coupled problems, the isothermal
and adiabatic splits will then perform very similarly. The superior stability properties
of the adiabatic split become critical, however, in the presence of a stronger coupling.
(C) Plane strain, nearly adiabatic shear banding. This final example, taken also
from ARMERO and SIMO [1993], furnishes the extension to the full thermomechanical
regime of the localization in plane strain of a rectangular bar under uniaxial tension
considered in Section 55. In this more general setting, thermoplastic softening is the
physical mechanism that induces response in the material leading to the formation of
shear bands. For high strain rates, this process is nearly adiabatic since heat conduction
in the material is precluded by the extremely small characteristic time involved in the
loading. A localized generation of heat then takes place at the center of the bar as
a consequence of plastic dissipation, resulting in a high temperature rise and a sharp
decrease in the effective value of the yield stress due to the thermal softening. It is
this softening response that triggers the localization of the deformation leading to the
formation of shear bands which, for the problem at hand, are oriented at 450 with the
axial direction of loading. This phenomenon is commonly referred to as "adiabatic"
shear banding.

SECTION 63

The coupled thermomechanicalproblem

485

TABLE 63.2

Nearly adiabatic shear banding: Material properties.


Bulk modulus

Kt

164.206

GPa

Shear modulus
Flow stress
Linear hardening
Saturation hardening
Hardening exponent
Density
Expansion coefficient

A
Yo
h
yo
6
Po
c

80.1938
0.450
0.12924
0.715
16.93
7800.
1 x 10- 5

GPa
GPa
GPa
GPa

Conductivity

45.

Specific capacity
Flow stress softening

c,
wo

460.
2.0 x 103

N/sK
m2 /s2 K
K- 1

Hardening softening

wh

2.0 x 10 - 3

K- 1

kg/m 3
K- 1

The specimen considered in the numerical simulations has a width of wo =


12.826mm, a length of lo = 53.334mm and is subjected to uniaxial loading under plane strain conditions. Figure 63.8 shows the mesh of the initial configuration
with the assumed boundary conditions. The bar is assumed insulated along its lateral
face, while the temperature is kept constant to the reference value (90 = 293.15K
on the upper and lower faces. Because of the symmetry in the problem, a quarter
of the specimen is discretized by imposing the corresponding symmetry boundary
conditions. The finite element mesh consists of 200 4-node isoparametric quadrilaterals based on the enhanced formulation presented in SIMo and ARMERO [1992].
The enhanced strain interpolations are chosen so that, in the linear regime, the onepoint quadrature element is recovered. The additional (nonlinear) terms arising in the
enhanced formulation provide a stabilization mechanism for the spurious modes associated with the one-point quadrature element. The model of J2 -flow theory outlined
at the beginning of this section is employed, with material properties summarized in
Table 63.2 and chosen to replicate those associated with steel. It should be noted that
the effect of work hardening is eventually overcome at high strain rates by the (small)
thermal softening. As in the preceding simulation, the source term Dint in the energy
equation is taken to be a fraction of the total plastic work. Neither geometric nor
material imperfections are introduced in the simulation; the final localized pattern of
the deformation is triggered solely by the thermal field.
Figure 63.9 shows the final configuration at an imposed top displacement i =
5.0 mm. The simulations are performed under quasistatic conditions with a staggered
algorithm based on the adiabatic split. Two different nominal strain rates are considered:
(a) i/lo = 4 x 102 s-l

and

(b) I/lo = 4 x 10 4 s.

The first strain rate leads to a diffuse necking mode while the second strain rate
produces sharp shear bands which appear to be well-resolved in the simulation. Fig-

486

J.C. Simo

CHAPTER V

I
i
l1.

TEMPERATURE [K

EC, PL ST

< 1.000E+01

< 2.30DE-1

, 6.200E+01

- ------ 11
TEMPERATURE

EQ. PL ST.

6 OOOE+01

< 6000E-9

=-~~

-11
- ----- 11
I'

> 1.500E+02

, '.200E+30

b
FIG. 63.9. Plane strain nearly adiabatic shear banding. Distribution of the relative temperature ((9 - (90) and
the equivalent plastic strain at u = 5.0 mm for two different nominal strain rates: (a) /lo = 4 x 10- 2 s - 1,
(b) i/lo = 4 x 104 - 1 . (Notice the different scales between (a) and (b).)

The coupled thermomechanicalproblem

SECTION 63

487

I rs

lU

6
o

4
2
n

Top displacement

[mm]

FIG. 63.10. Plane strain nearly adiabatic shear banding. Plots of the load/displacement curve for two different
nominal strain rates.
TABLE 63.3

Residual norm for a typical time increment


(i/lo = 4 x 104 s-l).
Mechanical phase
1.96287 x
1.13417 x
1.08594 x
5.09266 x

10+ 01
10-02
10- 05
10-10

Thermal phase
2.41033
2.00488
1.48953
6.13528

x
x
x
x

10+5
10+02
10 - 04
10- 09

ure 63.10 shows the load/displacement curves obtained for these two strain rates.
Table 63.3 summarizes the values of the Euclidean norm of the residual, obtained
within typical time increment, in an iterative solution procedure employing Newton's
method. The quadratic rate of convergence exhibited by the iteration is the result of an
exact linearization of the two symmetric sub-problems leading to an exact expression
for the algorithmic tangent moduli.

References
ACHENBACH, J.D. (1987), Wave Propagationin Elastic Solids, 5th print (North-Holland, Amsterdam).
ANAND, L. (1979), On H. Henky's approximate strain-energy function for moderate deformations, J. Appl.
Mech. 46, 78-82.
ANAND, L. (1985), Constitutive equations for hot working of metals, Internat. J. Plasticity 1, 213-231.
ANAND, L., K.H. KIM and T.G. SHAWKI (1987), Onset of shear localization in viscoplastic solids, J. Mech.
Phys. Solids 35, 407-429.
ANTMAN, S.S. and J.E. OSBORN (1979), The principle of the virtual work and integral laws of motion,
Arch. Rational Mech. Anal. 69, 231-262.
ARGYRIS, J.H. (1965), Elasto-plastic matrix analysis of three dimensional continua, J. Roy. Aeronaut. Soc.
69, 633-635.
ARGYRIS, J.H. and J.ST. DOLTSINIS (1979), On the large strain inelastic analysis in natural formulation. Part
I: Quasi-static problems, Comput. Methods Appl. Mech. Engrg. 20, 213-251.
ARGYRIS, J.H. and J.ST. DOLTSINIS (1980), On the large strain inelastic analysis in natural formulation. Part
II: Dynamic problems, Comput. Methods Appl. Mech. Engrg. 21, 213-251.
ARGYRIS, J.H., J.ST. DOLTSINIS, P.M. PMENTA and H. WUSTENBERG (1982), Thermomechanical response
of solids at high strains-natural approach, Comput. Methods Appl. Mech. Engrg. 32, 3-57.
ARMERO, E and J.C. SMO (1992), A new unconditionally stable fractional step method for coupled thermomechanical problems, Internat. J. Numer Methods Engrg. 35, 737-766.
ARMERO, F. and J.C. SIMO (1993), A priori stability estimates and unconditionally stable product algorithms
for nonlinear coupled thermoplasticity, Internat. J. Plasticity 9, 749-782.
ARNOLD, V.I. (1989), Mathematical Methods of ClassicalMechanics, 2nd edition (Springer, Berlin).
ASARO, R.J. (1979), Geometrical effects in the inhomogeneous deformation of ductile single crystals, Acta
Metal. 27, 445-453.
ASARO, R.J. (1983), Micromechanics of crystals and polycrystals, in: Advances in Applied Mechanics 23.
ASARO, R.J. and J.R. RICE (1977), Strain localization in ductile single crystals, J. Mech. Phys. Solids 25,
309-338.
BALL, J.M. (1977), Convexity conditions and existence theorems in nonlinear elasticity, Arch. Rational
Mech. Anal. 63, 337-403.
BALL, J.M. and G. KNOWLES (1986), Lyapunov functions for thermomechanics with spatially varying
boundary temperatures, Arch. Rat. Mech. Anal. 92, 193-204.
BERNARDOU, M., P.G. CIARLET and J. Hu (1984), On the convergence of the semi-discrete incremental
method in nonlinear three-dimensional elasticity, J. Elasticity 14, 425-440.
BERTSEKAS, D.P. (1982), Constrained Optimization and Lagrange Multiplier Methods (Academic Press,
New York).
BEVER, M.B., D.L. HOLT and A.L. TITCHENER (1973), The Stored Energy of Cold Work (Pergamon, Oxford).
BOYCE, M.C., G.G. WEBER and D.M. PARKS (1989), On the kinematics of finite strain plasticity, J. Mech.
Phys. Solids 37(5), 647-666.
BRENAN, K.E., S.L. CAMPBELL and L.R. PETZOLD (1989), Numerical Solution of Initial-Value Problems in
Differential-Algebraic Equations (North-Holland, New York).
BREZIS, H. (1973), Operateurs Maximaux Monotones et Semi-Groupes des Contractions (North-Holland,
Amsterdam).
BREZZI, F and M. FORTIN (1991), Mixed and Hybrid Finite Element Methods (Springer, Berlin).

489

490

J.C. Simo

BURRAGE, K. and J.C. BUTCHER (1979), Stability criteria for implicit Runge-Kutta methods, SIAM J.
Numer Math. 16(1), 46-57.
BURRAGE, K. and J.C. BUTCHER (1980), Nonlinear stability of a general class of differential equation
methods, BIT 20, 185-203.
BUTCHER, J.C. (1975), A stability property of implicit Runge-Kutta methods, BIT 15, 358-361.
CARLSON, D.E. (1972), Linear thermoelasticity, in: S. Fluegge, ed., Handbuch der Physik VI/2a (Springer,
Berlin) 297-346.
CASEY, J. and P.M. NAGHDI (1981), On the characterization of strain in plasticity, J. Appl. Mech. 48,
285-295.
CASEY, J. and P.M. NAGHDI (1983a), On the nonequivalence of the stress space and strain space formulations
of plasticity theory, J. Appl. Mech. 50, 350-354.
CASEY, J. and P.M. NAGHDI (1983b), On the use of invariance requirements for intermediate configurations
associated with the polar decomposition of a deformation gradient, Quart. Appl. Math. 41, 339-342.
CHADWICK, P. (1976), Continuum Mechanics (Wiley, New York).
CHORIN, A.E. (1969), On the convergence of discrete approximations to the Navier-Stokes equations, Math.
Comp. 23, 341-353.
CHORIN, A.J., T.J.R. HUGHES, J.E. MARSDEN and M. MCCRACKEN (1978), Product formulas and numerical

algorithms, Comm. Pure Appl. Math. 31, 205-256.


CHRISTOFFERSEN, J. and J.W. HUTCHINSON (1979), A class of phenomenological theories of plasticity,
J. Mech. Phys. Solids 27, 465-487.
CIARLET, P.G. (1978), The Finite Element Method for Elliptic Problems (North-Holland, Amsterdam).
CIARLET, P.G. (1988), Mathematical Elasticity I: Three-DimensionalElasticity (North-Holland, Amsterdam).
CIARLET, P.G. (1989), Introduction to Numerical Linear Algebra and Optimization, English translation of
1982 edition in Cambridge Texts in Applied Mathematics (Cambridge University Press, Cambridge).

CLIFTON, R.J. (1980), Adiabatic shear banding, in: Materials Response to Ultra-HighStrain Rates, Chapter 8
(National Materials Advisory Board (NRC), Washington, DC).

COLEMAN, B.D. and M. GURTIN (1967), Thermodynamics with internal variables, J. Chem. Phys. 47,
597-613.
COLEMAN, B.D. and W. NOLL (1963), The thermodynamics of elastic materials with heat conduction and
viscosity, Arch. Rational Mech. Anal. 13, 167-178.
COMMI, C. and G. MAIER (1989), On the convergence of a backward difference iterative procedure in elastoplasticity with nonlinear kinematic and isotropic hardening, in: D.R.J. Owen, E. Hinton and E. Onate,
eds., Computational Plasticity, Models, Software and Applications (Pineridge, Swansea) 323-334.
CORIGLIANO, A. and U. PEREGO (1991), Convergent and unconditionally stable finite-step dynamic analysis of elastoplastic structures, in: Proc. Europ. Conf on New Advances in Computational Structural
Mechanics, 577-584.
CORMEAU, I.C. (1975), Numerical stability in quasi-static elasto/visco-plasticity, Internat. J. Numer Methods
Engrg. 9, 109-127.
COTTRELL, A.H. (1967), Dislocations and Plastic Flow in Crystals (Oxford University Press, London).
CROUZEIX, M. (1979), Sur la B-stabilit6 des mrthodes de Runge-Kutta, Numer Math. 32, 75-82.
CURTISS, C.F. and J.C. HIRSCHFELDER (1952), Integration of stiff equations, Proc. Nat. Acad. Sci. USA 38,

235-243.
DAFALIAS, YF. (1984), A missing link in the formulation and numerical implementation of finite transformation elastoplasticity, in: K.J. Willam, ed., Constitutive Equations: Macro and Computational Aspects

(ASME, New York).


DAFERMOS, C.M. (1976), Contraction semigroups and trend to equilibrium in continuum mechanics, in:

Lecture Notes in Math. 503 (Springer, Berlin) 295-306.


DAHLQUIST, G. (1963), A special stability problem for linear multistep methods, BIT 3, 27-43.
DAHLQUIST, G. (1975), Error analysis of a class of methods of stiff non-linear initial value problems, in:

Numerical Analysis, Dundee 1975, Lecture Notes in Math. 506 (Springer, Berlin) 60-74.
DAHLQUIST, G. (1978), G-stability is equivalent to A-stability, BIT 18, 384-401.
DEMENGEL, F. (1989), Compactness theorems for spaces of functions with bounded derivatives and applications to limit analysis problems in plasticity, Arch. Rational Mech. Anal. 105(2), 123-161.

References

491

DENNIS, J.E. and R.B. SCHNABEL (1983), Numerical Methods for Unconstrained Optimization (PrenticeHall, Englewood Cliffs, N.J.).
DESAI, C.S. and H.J. SIRIWARDANE (1984), Constitutive Laws for Engineering Materials (Prentice-Hall,
Englewood Cliffs, NJ).
DIENES, J.K. (1979), On the analysis of rotation and stress rate in deforming bodies, Acta Mech. 32,
217-232.
DILLON, O.W. (1963), Coupled thermoplasticity, J. Mech. Phys. Solids 11, 21-33.
DIMAGGlo, EL. and I.S. SANDLER (1971), Material models for granular soils, J. Engrg. Mech., 935-950.
DOHERTY, W.P., E.L. WILSON and R.L. TAYLOR (1969), Stress analysis of axisymmetric solids utilizing
higher order quadrilateral finite elements, SESM Report No. 69-3, Department of Civil Engineering,
University of California, Berkeley.
DOLTSIMS, I.ST. (1990), Aspects of modeling and computation in the analysis of metal forming, Engrg.
Comput. 7, 2-20.
DUHEM, P. (1911), Traite d'Energetique ou de Thermodynamique Gendrale (Gauthier-Villars, Paris).
DUVAuT, G. and J.L. LIONS (1972), Les Inequations en Mecanique et en Physique (Dunod, Paris). English
translation: (1976) (Springer, Berlin).
ERICKSEN, J.L. (1966), Thermoelastic stability, in: Proc. 5th National Cong. Appl. Mech., 187-193.
ETEROVICH, A.L. and K.J. BATHE (1990), A hyperelastic-based large strain elasto-plastic constitutive formulation with combined isotropic-kinematic hardening using logarithmic stresses and strain measures,
Internat. J. Numer Methods Engrg. 30(6), 1099-1115.
EVE, R.A., B.D. REDDY and R.T. ROCKEFELLAR (1990), An internal variable theory of elastoplasticity
based on the maximum work inequality, Quart. Appl. Math. 48, 59-83.
FARHAT, C., K.C. PARK and Y. DUBOIS-PELERIN (1991), An unconditionally stable staggered algorithm
for transients finite element analysis of coupled thermoelastic problems, Comput. Methods Appl. Mech.
Engrg. 85, 349-365.
FLORY, R.J. (1961), Thermodynamic relations for highly elastic materials, Trans. FaradaySoc. 57, 829-838.
FoIAS, C., M.S. JOLLY, I.G. KEVREKIDIS and E.S. TITI (1991), Dissipativity of numerical schemes, Nonlinearity 4, 591-613.
GEAR, C.W. (1971), Numerical Initial Value Problems in OrdinaryDifferential Equations (Prentice-Hall,
Englewood Cliffs, NJ).
GERMAIN, P., Q.S. NGUYEN and P. SQUET (1983), Continuum thermodynamics, Trans. ASME 50, 1010-

1020.
GIRAULT, V. and P.A. RAVIART (1986), Finite Element Methods for Navier-Stokes Equations. Theory and

Algorithms (Springer, Berlin).


GLOWINSKI, R. and P. LE TALLEC (1984), Finite element analysis in nonlinear incompressible elasticity, in:
J.T. Oden and G.E Carey, eds., Finite Elements V: Special Problems in Solid Mechanics (Prentice-Hall,
Englewood Cliffs, NJ).
GLOWINSKI, R. and P. LE TALLEC (1989), Numerical Simulation of Continuous Media by Augmented Lagrangian and OperatorSplitting Methods, SIAM Studies in Applied Mathematics (SIAM, Philadelphia).
GOUDREAU, G.L. and J.O. HALLQUIST (1982), Recent developments in large-scale finite element Lagrangian
hydrocode technology, Comput. Methods Appl. Mech. Engrg. 33, 725-757.
GREEN, A.E. and P.M. NAGHDI (1964), A general theory of an elastic-plastic continuum, Arch. Rational
Mech. Anal. 18, 251-281.
GREEN, A.E. and P.M. NAGHDI (1966), A thermodynamic development of elastic-plastic continua, in:
Proceedings of IUTAM Symposia, Vienna, June 22-28.
GREEN, A.E. and W. ZERNA (1960), Theoretical Elasticity, 2nd edition (Clarendon Press, Oxford).
GURTIN, M.E. (1972), The linear theory of elasticity, in: C. Truesdell, eds., Handbuch der Physik VIa/2,
Mechanics of Solids II (Springer, Berlin).
GURTIN, M.E. (1975), Thermodynamics and stability, Arch. Rational Mech. Anal. 59, 53-96.
GURTIN, M.E. (1981), An Introduction to Continuum Mechanics (Academic Press, Orlando, FL).
HAIRER, E., S.P. NORSETT and G. WANNER (1993), Solving Ordinary Differential Equations I, Nonstiff
Problems, 2nd edition (Springer, Berlin).
HAIRER, E. and G. WANNER (1991), Solving Ordinary Differential Equations II, Stiff and DifferentialAlgebraic Problems (Springer, Berlin).

492

J.C. Simo

HALLQUIST, J.O. (1984), NIKE 2D: An implicit, finite deformation, finite element code for analyzing the
static and dynamic response of two-dimensional solids, Lawrence Livermore National Laboratory, Report
UCRL-52678, University of California, Livermore.
HALLQUIST, J.O. (1988), DYNA 3D: An explicit, finite element code for dynamic analysis in three dimensions, Lawrence Livermore National Laboratory University of California, Livermore.
HALLQUIST, J.O. and D.J. BENSON (1987), DYNA3D User's Manual, Report No. UCID-19592, Rev. 3,
Lawrence Livermore National Laboratory.
HALPHEN, B. and Q.S. NGUYEN (1975), Sur les materiaux standards generalises, J. Mecanique 14, 39-63.
HARREN, S.V. (1991), The finite deformation of rate-dependent polycrystals I. A self-consistent framework,
J. Mech. Phys. Solids 39(3), 345-360.
HAVNER, K.S. (1971), A discrete model for the prediction of subsequent yield surfaces in polycrystalline
elasticity, Int. J. Solids Structures 7, 719-730.
HAVNER, K.S. (1992), Finite Plastic Deformation of Crystalline Solids, Cambridge Monographs on Mechanics and Applied Mathematics (Cambridge University Press, London).
HERRMANN, L.R. (1965), Elasticity equations for incompressible and nearly incompressible materials by a
variational theorem, AIAA J. 3(10).
HILBER, H.M. (1976), Analysis and design of numerical integration methods in structural dynamics, Report
No. E.E.R.C. 76-29. Earthquake Eng. Research Center, University of California, Berkeley, CA.
HILBER, H.M., T.J.R. HUGHES and R.L. TAYLOR (1977), Improved numerical dissipation for time integration
algorithms in structural dynamics, Earthquake Engrg. &Struct. Dyn. 5, 283-292.
HILL, R. (1950), The Mathematical Theory of Plasticity, Last edition (Oxford University Press, Oxford).
HILL, R. (1958), A general theory of uniqueness and stability in elastic-plastic solids, J. Mech. Phys. Solids
6, 236-249.
HILL, R. (1966), Generalized constitutive relations for incremental deformation of metal crystals by multislip, J. Mech. Phys. Solids 14, 95-102.
HILL, R. (1978), Aspects of invariance in solid mechanics, Adv. in Appl. Mech. 18, 1-75.
HILL, R. and K.S. HAVNER (1982), Perspectives in the mechanics of elastoplastic crystals, J. Mech. Phys.
Solids 23, 5-22.
HILL, R. and J.R. RICE (1972), Constitutive analysis of elastic-plastic crystals at arbitrary strain, J. Mech.
Phys. Solids 20, 401-413.
HINTON, E. and D.R.J. OWEN (1980), Finite Elements in Plasticity: Theory and Practice (Pineridge Press,
Swansea, Wales).
HOGER, A. and D.E. CARLSON (1984a), Determination of the stretch and rotation in the polar decomposition
of the deformation gradient, Quart. Appl. Math. 42, 113-117.
HOGER, A. and D.E. CARLSON (1984b), On the derivative of the square root of a tensor and Guo's rate
theorems, J. Elasticity 14, 329-336.
HUGHES, T.J.R. (1980), Generalization of selective integration procedures to anisotropic and nonlinear
media, Internat. J. Numer. Methods Engrg. 15(9), 1413-1418.
HUGHES, T.J.R. (1983), Analysis of transient algorithms with particular emphasis in stability behavior, in:
T. Belytschko and T.J.R. Hughes, eds., Computational Methods for Transient Analysis (North-Holland,
Amsterdam).
HUGHES, T.J.R. (1984), Numerical implementation of constitutive models: rate-independent deviatoric plasticity, in: S. Nemat-Nasser, R. Asaro and G. Hegemier, eds., Theoretical Foundationsfor Large Scale
Computations of Nonlinear Material Behavior (Martinus Nijhoff Publishers, The Netherlands).
HUGHES, T.J.R. (1987), The Finite Element Method (Prentice-Hall, Englewood Cliffs, NJ).
HUGHES, T.J.R. and R.L. TAYLOR (1978), Unconditionally stable algorithms for quasi-static elasto viscoplastic finite element analysis, Comput. & Structures 8, 169-173.
HUGHES, T.J.R. and J. WINGET (1980), Finite rotation effects in numerical integration of rate constitutive
equations arising in large-deformation analysis, Internat. J. Numer Methods Engrg. 15(9), 1413-1418.
HUNDSDORFER, W.H. (1985), The Numerical Solution of Nonlinear Stiff Initial Value Problems: An Analysis
of One Step Methods, CWI Tract, Centrum voor Wiskunde en Informatica (Center for Mathematics and
Computer Science, Amsterdam).
IWAN, D.W. and P.J. YODER (1983), Computational aspects of strain-space plasticity, J. Engrg. Mech. ASCE
109, 231-243.

References

493

JOHNSON, C. (1976a), Existence theorems for plasticity problems, J. Math. Pures Appl. 55, 431-444.
JOHNSON, C. (1976b), On finite element methods for plasticity problems, Numer. Math. 26, 79-84.
JOHNSON, C. (1977), A mixed finite element for plasticity, SIAM J. Numer. Anal. 14, 575-583.
JOHNSON, C. (1978), On plasticity with hardening, J. Appl. Math. Anal. 62, 325-336.
JOHNSON, C. (1987), Numerical Solutions of PartialDifferential Equations by the Finite Element Method
(Cambridge University Press, Cambridge, MA).
JOHNSON, C. and D.J. BAMMANN (1984), A discussion of stress rates in finite deformation problems,
Internat. J. Solids Structures 20, 725-737.
KACHANOV, L.M. (1974), Fundamentals of the Theory of Plasticity (MIR Publishers, Moscow).
KARMANOV, V. (1977), ProgrammationMathematique (MIR Publishers, Moscow).
KATO, T. (1976), PerturbationTheory for Linear Operators, 2nd edition (Springer, Berlin).
KEY, S.W. (1969), A variational principle for incompressible and nearly incompressible anisotropic elasticity, Internat. J. Solids Structures 5, 951-964.
KEY, S.W. (1974), HONDO - A finite element computer program for the large deformation response of
axisymmetric solids, Sandia National Laboratories, Albuquerque, NM, Report 74-0039.
KEY, S.W. and R.D. KRIEG (1982), On the numerical implementation of inelastic time dependent and time
independent, finite strain constitutive equations in structural mechanics, Comput. Methods Appl. Mech.
Engrg. 33, 439-452.
KEY, S.W., C.M. STONE and R.D. KRIEG (1981), Dynamic relaxation applied to the quasi-static, large
deformation, inelastic response of axisymmetric solids, in: Wunderlich et al., eds., Nonlinear Finite
Element Analysis in Structural Mechanics (Springer, Berlin).
KIM, S.J. and J.T. ODEN (1990), Finite element analysis of a class of problems in finite strain elastoplasticity
based on the thermodynamical theory of materials of type N, Comput. Methods Appl. Mech. Engrg. 53,
277-302.
KOITER, W.T. (1960), General theorems for elastic-plastic solids, Chapter IV in: Progressin Solid Mechanics
1, 165-220.
KRATOCHVIL, J. (1973), On a finite strain theory of elastic-inelastic materials, Acta Mech. 16, 127-142.
KRIEG, R.D. and S.W. KEY (1976), Implementation of a time dependent plasticity theory into structural
computer programs, in: J.A. Stricklin and K.J. Saczlski, eds., Constitutive Equations in Viscoplasticity:
Computational and Engineering Aspects, AMD-20 (ASME, New York).
KRIEG, R.D. and D.B. KRIEG (1977), Accuracies of numerical solution methods for the elastic-perfectly
plastic model, J. Pressure Vessel Tech. 99.
KRONER, E. and C. TEODOSIU (1972), Lattice defect approach to plasticity and viscoplasticity, in:
A. Sawczuk, ed., Problems of Plasticity (Noordhoff, Leiden).
LAURSEN, T.A. and J.C. SMO (1993), A continuum based finite element formulation for the implicit
finite element solution of multi-body, large deformation, frictional contact problems, Internat. J. Numer
Methods Engrg. 36, 3451-3485.
LEE, E.H. (1969), Elastic-plastic deformations at finite strains, J. Appl. Mech. 36, 1-6.
LEE, E.H. and D.T. Llu (1967), Finite strain elastic-plastic theory particularly for plane wave analysis,
J. Appl. Phys. 38.
LEE, R.L., P.M. GRESHO and R.L. SANI (1979), Smoothing techniques for certain primitive variable solutions
of the Navier-Stokes equations, Internat. J. Numer Methods Engrg. 14(12), 1785-1804.
LEWIS, R.W. and B.A. SCHREFLER (1987), The Finite Element Method in the Deformation and Consolidation
of Porus Media (Wiley, New York).
LEHMANN, T. (1982), General frame for the definition of constitutive laws for large non-isothermal elasticplastic and elastic-viscoplastic deformations, in: T. Lehmann, ed., The Constitutive Law in Thermoplasticity, CISM Courses and Lectures No. 281 (Springer, Berlin).
LEHMANN, T. and U. BLIX (1985), On the coupled thermo-mechanical process in the necking problem,
Internat. J. Plasticity 1, 175-188.
LEMONDS, J. and A. NEEDLEMAN (1986), Finite element analysis of shear localization in rate and temperature
dependent solids, Mech. Mater 5, 339-361.
LORET, B. and J.H. PREVOST (1986), Accurate numerical solutions of Drucker-Prager elastic-plastic models,
Comput. Methods Appl. Mech. Engrg. 54, 259-278.

494

J.C. Simo

LUBLINER, J. (1972), On the thermodynamic foundations of non-linear solid mechanics, Internat. J. Nonlinear Mech. 7, 237-254.
LUBLINER, J. (1973), On the structure of the rate equations of materials with internal variables, Acta Mech.
17, 109-119.
LUBLINER, J. (1986), Normality rules in large-deformation plasticity, Mech. Mater 5, 29-34.
LUBLINER, J. (1990), Plasticity Theory (MacMillan, London).
LUENBERGER, D.G. (1972), Optimization by Vector Space Methods (Wiley, New York).
LUENBERGER, D.G. (1984), Linear and Nonlinear Programming (Addison-Wesley, Menlo Park, CA).
MAENCHEN, G. and S. SACKS (1964), The tensor code, in: B. Alder, S. Fernback and M. Rotenberg, eds.,
Methods of Computational Physics 3 (Academic Press, New York) 181-210.
MAIER, G. (1970), A matrix structural theory of piecewise linear elastoplasticity with interacting yield
planes, Meccanica, 54-66.
MAIER, G. and D. GRIERSON (1979), Engineering Plasticity by Mathematical Programming (Pergamon
Press, New York).
MALVERN, L. (1969), An Introduction to the Mechanics of a Continuous Media (Prentice-Hall, Englewood
Cliffs, NJ).
MANDEL, J. (1964), Contribution theorique a l'etude de l'ecrouissage et des lois de l'ecoulement plastique,
in: Proc. 11th Internat. Congress on Applied Mechanics, 502-509.
MANDEL, J. (1965), Generalisation de la theorie de la plasticite de W.T. Koiter, Internat. J. Solids Structures
1, 273-295.
MANDEL, J. (1972), Plasticite Classique et Viscoplasticite', Course held at the Department of Mechanics of
Solids, Sept.-Oct. 1971 (Springer, New York).
MANDEL, J. (1974), Thermodynamics and plasticity, in: J.J. Delgado Domingers, N.R. Nina and
J.H. Whitelaw, eds., Foundations of Continuum Thermodynamics (Macmillan, London) 283-304.
MARCAL, P.V. and I.P. KING (1967), Elastoplastic analysis of two-dimensional stress systems by the finite
element method, Internat. J. Mech. Sci. 9, 143-155.
MARCHUK, G.I. (1990), Splitting and alternating direction methods, in: P.G. Ciarlet and J.L. Lions, eds.,
Handbook for Numerical Analysis I (North-Holland, Amsterdam) 197-462.
MARSDEN, J. (1973), On product formulas for nonlinear semigroups, J. Funct. Anal. 13, 51-72.
MARSDEN, J. (1974), A formula for the solution of the Navier-Stokes equations based on a method of
Chorin, Bull. Amer Math. Soc. 80, 154-158.
MARSDEN, J.E. and T.J.R HUGHES (1983), Mathematical Foundations of Elasticity (Prentice-Hall, Englewood Cliffs, NJ).
MARTIN, J.B. (1988), Convergence and shakedown for discrete load steps in statically loaded elastic-plastic
bodies, Mech. Structures Mach. 16(1), 1-16.
MATTHIES, H. (1978), Problems in plasticity and their finite element approximation, Ph.D. Thesis, Department of Mathematics, Massachusetts Institute of Technology, Cambridge, MA.
MATTHIES, H. (1979), Existence theorems in thermo-plasticity, J. Mech. 18(4), 695-711.
MATTHIES, H. and G. STRANG (1979), The solution of nonlinear finite element equations, Internat. J. Numer
Methods Engrg. 14(11), 1613-1626.
MATTHIES, H., G. STRANG and E. CHRISTIANSEN (1979), The saddle point of a differential, in: Glowinski,
Rodin and Zienkiewicz, eds., Energy Methods in Finite Element Analysis (Wiley, New York).
MCCLINTOCK, FA. and A.S. ARGON (1968), Mechanical Behavior of Materials (Addison-Wesley, Reading,
MA).
MCMEEKING, R.M. and J.R. RICE (1975), Finite-element formulations for problems of large elastoplastic
deformations, Internat. J. Solids Structures 11, 601-616.
MEHRABADI, M.M. and S. NEMAT-NASSER (1987), Some basic kinematical relations for finite deformations
of continua, Mech. Mater. 6, 127-138.
MOLINARI, A. and R.J. CLIFTON (1983), Localisation de la deformation viscoplastique encisaillement simple:
Rbsultats exacts en thkorie non lindaire, C. R. Acad. Sci. Sr 11 296, 1-4.
MORAN, B., M. ORTIZ and E SHI (1990), Formulation of implicit finite element methods for multiplicative
plasticity, Internat. J. Numer. Methods Engrg. 29, 483-514.
MORAN, B., M. ORTIZ and CF. SHIH (1992), An analysis of cracks in ductile single crystals I. Anti-plane
shear, J. Mech. Phys. Solids 40(2), 291-314.

References

495

MOREAU, J.J. (1976), Application of convex analysis to the treatment of elastoplastic systems, in: P. Germain and B. Nayroles, eds., Applications of Methods of FunctionalAnalysis to Problems in Mechanics
(Springer, Berlin).
MOREAU, J.J. (1977), Evolution problem associated with a moving convex set in a Hilbert space, J. Differential Equations 26, 347-374.
MORMAN, K.N. (1987), The generalized strain measure with application to non-homogeneous deformations
in rubber-like solids, J. Appl. Mech. 53, 726-728.
NAGHDI, P.M. (1960), Stress-strain relations in plasticity and thermoplasticity, in: Proc. 2nd Symposium on
Naval Structural Mechanics (Pergamon Press, London).
NAGHDI, P.M. and J.A. TRAPP (1975), The significance of formulating plasticity theory with reference to
loading surfaces in strain space, Internat. J. Engrg. Sci. 13, 785-797.
NAGTEGAAL, J.C. (1982), On the implementation of inelastic constitutive equations with special reference
to large deformation problems, Comput. Methods Appl. Mech. Engrg. 33, 469-484.
NAGTEGAAL, J.C. and J.E. DE JONG (1981), Some computational aspects of elastic-plastic large strain
analysis, Internat. J. Numer Methods Engrg. 17, 15-41.
NAGTEGAAL, J.C., D.M. PARKS and J.R. RICE (1974), On numerically accurate finite element solutions in

the fully plastic range, Comput. Methods Appl. Mech. Engrg. 4, 153-177.
NAGTEGAAL, J.C. and F.E. VELDPAUS (1984), On the implementation of finite strain plasticity equations in
a numerical model, in: J.F.T. Pittman, O.C. Zienkiewicz, R.D. Wood and J.M. Alexander, eds., Numerical
Analysis of Forming Processes (Wiley, New York).
NAYAK, G.C. and O.C. ZtENKIEWICZ (1972), Elastoplastic stress analysis. Generalization of various constitutive equations including stress softening, Internat.J. Numer Methods Engrg. 5, 113-135.
NEEDLEMAN, A. (1982), Finite elements for finite strain plasticity problems, in: E.H. Lee and R.L. Mallet,
eds., Plasticity of Metals at Finite Strains: Theory, Computation, and Experiment, 387-436.
NEEDLEMAN, A. and V. TVERGAARD (1984), Finite element analysis of localization plasticity, in: J.T. Oden
and G.F Carey, eds., Finite Elements V: Special Problems in Solid Mechanics (Prentice-Hall, Englewood
Cliffs, NJ).
NEMAT-NASSER, S. (1983), On finite plastic flow of crystalline solids and geomaterials, J. Appl. Mech. 50,
1114-1126.
NGUYEN, Q.S. (1977), On the elastic-plastic initial boundary value problem and its numerical integration,
Internat.J. Numer Methods Engrg. 11, 817.
OGDEN, R.W. (1982), Elastic deformations in rubberlike solids, in: H.G. Hopkins and M.J. Sewell, eds.,
Mechanics of Solids, Rodney Hill 60th Anniversary Volume (Pergamon, Oxford), 319-323.
OGDEN, R.W. (1984), Nonlinear Elastic Deformations (Ellis Horwood Ltd., West Sussex, England).
OLVER, P.J. (1986), Applications of Lie Groups to Differential Equations (Springer, Berlin).
ORTEGA, J.M. and W.C. RHEINBOLDT (1970), Iterative Solution of Nonlinear Equations in Several Variables
(Academic Press, San Diego, CA).
ORTIZ, M. and J.E. MARTIN (1989), Symmetry preserving return mapping algorithms and minimum work
paths: A unification of concepts, Internat. J. Numer. Methods Engrg. 28, 1839-1853.
ORTIZ, M. and E.P. PoPov (1985), Accuracy and stability of integration algorithms for elastoplastic constitutive equations, Internat. J. Numer Methods Engrg. 21, 1561-1576.
ORTIZ, M. and J.C. SIMO (1986), An analysis of a new class of integration algorithms for elastoplastic
constitutive relations, Internat.J. Numer Methods Engrg. 23, 353-366.
PARK, K.C. and C.A. FELIPPA (1983), Partitioned analysis of coupled problems, in: T. Belytschko and T.J.R.
Hughes, eds., ComputationalMethods in Transient Analysis (North-Holland, Amsterdam).
PARK, K.C., C.A. FELIPPA and L.A. DERUNTz (1977), Stabilization of staggered solution procedures for
fluid-structure interaction analysis, in: T. Belytschko and T.L. Geers, eds., Computational Methods for
Fluid-Structure Interaction Problems, ASME Applied Mechanics Symposia Series, AMD 26, 94-124.
PARKUS, H. (1976), Thermoelasticity, 2nd edition (Springer, Wien).
PAZY, A. (1983), Semigroups of Linear Operators and Applications to Partial Differential Equations
(Springer, New York).

PERIC, F., D.R.J. OWEN and M.E. HONNOR (1989), A model for finite strain elastoplasticity, in: R. Owen,
E. Hinton and E. Onate, eds., Proc. 2nd Internat. Conf on ComputationalPlasticity (Pineridge, Swansea)
111-126.

496

J.C. Simo

PERZYNA, P. (1971), Thermodynamic theory of viscoplasticity, in: Advances in Applied Mechanics 11


(Academic Press, New York).
PHILLIPS, A. (1974), The foundation of thermoplasticity-Experiments and theory, in: J.L. Zeman and
F. Ziegler, eds., Topics in Applied Continuum Mechanics (Springer, Wien).
PrNSKY, P., M. ORTIZ and K.S. PISTER (1983), Numerical integration of rate constitutive equations in finite
deformation analysis, Comput. Methods Appl. Mech. Engrg. 40, 137-158.
PITKARANTA, J. and R. STERNBERG (1984), Error bounds for the approximation of the Stokes problem using

bilinear/constant elements on irregular quadrilateral meshes, Report MAT-A222, Helsinki University of


Technology, Institute of Mathematics, Finland.
PRAGER, W. (1956), A new method of analyzing stress and strains in work-hardening plastic solids, J. Appl.
Mech. 23, 493-496.
REBELO, N., J.C. NAGTEGAAL and H.D. HiBBITT (1990), Finite element analysis of sheet forming process,
Internal. J. Numer Methods Engrg. 30, 1739-1758.
REED, R.E. (1964), Physical Metallurgy Principles (Van Nostrand, New York).
RESENDE, L. and J.B. MARTIN (1986), Formulation of Drucker-Prager cap model, J. Engrg. Mech. 111(7),
855-865.
RICE, J.R. and D.M. TRACEY (1973), Computational fracture mechanics, in: S.J. Fenves, ed., Proc. Symposium on Numerical Methods in Structural Mechanics, Urbana, Illinois (Academic Press, New York).
RICHTMYER, R.D. and K.W. MORTON (1967), Difference Methods for Initial Value Problems, 2nd edition
(Interscience, New York).
ROLPH III, W.D. and K.J. BATHE (1984), On a large strain finite-element formulation for elasto-plastic
analysis, in: K.J. Willam, ed., Constitutive Equations, Macro and ComputationalAspects (Winter Annual
Meeting, ASME).
RUBINSTEIN, R. and S.N. ATLURI (1983), Objectivity of incremental constitutive relations over finite time
steps in computational finite deformation analyses, Comput. Methods Appl. Mech. Engrg. 36.
SANDLER, I.S., F.L. DIMAGGIO and G.Y. BALADI (1976), Generalized CAP model for geological materials,
J. Geotech. Engineering Division 102(GT7), 683-699.
SANDLER, I.S. and D. RUBIN (1979), An algorithm and a modular subroutine for the cap model, Internat.
J. Numer. Anal. Methods Geomech. 3, 173-186.
SCHREYER, H.L., R.L. KULAK and J.M. KRAMER (1979), Accurate numerical solutions for elastic-plastic
models, J. Pressure Vessel Tech. 101, 226-334.
SCOTT, L.R. and M. VOGELIUS (1985), Conforming finite element methods for incompressible and nearly
incompressible continua, Lectures Appl. Math. 22, 221-244.
SCOVEL, C. (1991), Symplectic numerical integration of Hamiltonian systems, in: T. Ratiu, ed., The Geometry of Hamiltonian Systems, Proc. Workshop June 5-15, 1989 (Springer, Berlin) 463-496.
SHAWKI, T.G and R.J. CLIFTON (1989), Shear band formation in thermal viscoplastic materials, Mech.
Mater. 8, 13-43.
SIMO, J.C. (1986), On the computational significance of the intermediate configuration and hyperelastic
relations in finite deformation elastoplasticity, Mech. Mater. 4, 439-451.
SIMO, J.C. (1987), A J2-flow theory exhibiting a comer-like effect and suitable for large-scale computation,
Comput. Methods Appl. Mech. Engrg. 62, 169-194.
SIMO, J.C. (1988a), A framework for finite strain elastoplasticity based on maximum plastic dissipation and
the multiplicative decomposition: Part I. Continuum formulation, Comput. Methods Appl. Mech. Engrg.
66, 199-219.
SIMO, J.C. (1988b), A framework for finite strain elastoplasticity based on maximum plastic dissipation and
the multiplicative decomposition: Part II. Computational aspects, Comput. Methods Appl. Mech. Engrg.
68, 1-31.
SIMO, J.C. (1991), Nonlinear stability of the time discrete variational problem in nonlinear heat conduction
and elastoplasticity, Comput. Methods Appl. Mech. Engrg. 88, 111-121.
SIMO, J.C. (1992), Algorithms for multiplicative plasticity that preserve the form of the return mappings
of the infinitesimal theory, Comput. Methods Appl. Mech. Engrg. 99, 61-112.
SLMO, J.C. and F. ARMERO (1992), Geometrically nonlinear enhanced strain mixed methods and the method
of incompatible modes, Internat. J. Numer. Methods Engrg. 33, 1413-1449.

References

497

SIMO, J.C. and E ARMERO (1993), Unconditional stability and long term behavior of transient algorithms
for the incompressible Navier-Stokes and Euler equations, Comput. Methods Appl. Mech. Engrg. 111,
111-154.
SIMO, J.C., F ARMERO and R.L. TAYLOR (1993), Improved versions of assumed enhanced strain tri-linear
elements for 3D deformation problems, Comput. Methods Appl. Mech. Engrg. 110, 359-386.
SIMO, J.C. and S. GOVINDJEE (1988), Exact closed-form solution of the return mapping algorithm in plane
stress elasto-viscoplasticity, Engrg. Comput. 4(3), 254-258.
SIMO, J.C. and T.J.R. HUGHES (1986), On the variational foundations of assumed strain methods, J. Appl.
Mech. 53(1), 51-54.
SIMO, J.C. and T.J.R. HUGHES (1987), General return mapping algorithms for rate independent plasticity,
in: C.S. Desai, ed., Constitutive Equations for Engineering Materials, 221-231.
SIMO, J.C., J.W. Ju, K.S. PISTER and R.L. TAYLOR (1988), An assessment of the cap model: Consistent
return algorithms and rate-dependent extension, ASCE, J. Engrg. Mech. 144(2), 191-218.
SIMO, J.C., J.G. KENNEDY and S. GOVINDJEE (1988), Non-smooth multisurface plasticity and viscoplasticity.
Loading/unloading conditions and numerical algorithms, Internat. J. Numer Methods Engrg. 26, 21612185.
SIMO, J.C., J.G. KENNEDY and R.L. TAYLOR (1988), Complementary mixed finite element formulations of
elastoplasticity, Comput. Methods Appl. Mech. Engrg. 74, 177-206.
SIMO, J.C. and J.E. MARSDEN (1984), On the rotated stress tensor and the material version of the DoyleEricksen formula, Arch. Rational Mech. Anal. 86, 213-231.
SIMO, J.C., J.E. MARSDEN and P.S. KRISHNAPRASAD (1988), The Hamiltonian structure of elasticity. The
convective representation of solids, rods and plates, Arch. Rational Mech. Anal. 104(2), 125-183.
SIMO, J.C. and C. MIEHE (1992), Coupled associative thermoplasticity at finite strains. Formulation, numerical analysis and implementation, Comput. Methods Appl. Mech. Engrg. 98, 41-104.
SIMO, J.C., X. OLIVER and F ARMERO (1993), Analysis of strong discontinuities in rate independent
softening materials, Internat. J. Comput. Mech. in press.
SMO, J.C. and M. ORTIZ (1985), A unified approach to finite deformation elastoplasticity based on the use
of hyperelastic constitutive equations, Comput. Methods Appl. Mech. Engrg. 49, 221-245.
SIMO, J.C. and K.S. PISTER (1984), Remarks on rate constitutive equations for finite deformation problems:
Computational implications, Comput. Methods Appl. Mech. Engrg. 46, 201-215.
SIMO, J.C., T.A. POSBERGH and J.E. MARSDEN (1991), Stability of relative equilibria II: Three-dimensional
elasticity, Arch. Rational Mech. Anal. 115, 61-100.
SIMO, J.C. and M.S. RIFAt (1990), A class of mixed assumed strain methods and the method of incompatible
nodes, Internat. J. Numer Methods Engrg. 29, 1595-1638.
SIMO, J.C. and N. TARNOW (1992), Conserving algorithms for nonlinear elastodynamics: The energymomentum method, Z. Angew. Math. Mech. 43, 757-792.
SIMO, J.C., N. TARNOW and K. WONG (1992), Exact energy-momentum conserving algorithms and symplectic schemes for nonlinear dynamics, Comput. Methods Appl. Mech. Engrg. 100, 63-116.
SIMO, J.C. and R.L. TAYLOR (1985), Consistent tangent operators for rate independent elasto-plasticity,
Comput. Methods Appl. Mech. Engrg. 48, 101-118.
SIMO, J.C and R.L. TAYLOR (1986), A return mapping algorithm for plane stress elastoplasticity, Internat.
J Numer Methods Engrg. 22(3), 649-670.
SIMO, J.C. and R.L. TAYLOR (1991), Quasi-incompressible finite elasticity in principal stretches. Continuum
basis and numerical algorithms, Comput. Methods Appl. Mech. Engrg. 85, 273-310.
SIMO, J.C., R.L. TAYLOR and K.S. PISTER (1985), Variational and projection methods for the volume
constraint in finite deformation elastoplasticity, Comput. Methods Appl. Mech. Engrg. 51, 177-208.
SOKOLNIKOFF, I.S. (1956), Mathematical Theory ofElasticity, 2nd edition (McGraw-Hill, New York).
STRANG, G. (1969), Approximating semigroups and the consistency of difference schemes, Proc. Amer
Math. Soc. 20, 1-7.
STRANG, G. (1986), Introduction to Applied Mathematics (Wellesley-Cambridge Press, Wellesley, MA).
STRANG, G. and G.J. Fix (1973), Analysis of Finite Element Methods (Prentice-Hall, Englewood Cliffs,
NJ).
STRANG, G., H. MATTHIES and R. TEMAM (1980), Mathematical and computational methods in plasticity,
in: S. Nemat-Nasser, ed., VariationalMethods in the Mechanics of Solids (Pergamon, Oxford).

498

J.C. Simo

STUART, A. and A.R. HUMPHRIES (1993), Model problems in numerical stability theory of initial boundary
value problems, SIAM Rev. 36, 222-257.
SUQUET, P. (1979), Sur les equations de la plasticit, Ann. Fac. Sci. Toulouse 1, 77-87.
SUQUET, P.M. (1981), Sur les equations de la plasticite: Existence et regularite des solutions, J. Mecanique
3, 39.
SUSSMAN, T. and K.J. BATHE (1987), A finite element formulation for nonlinear incompressible elastic and
inelastic analysis, Comput. &Structures 26(1/2), 357-109.
TAYLOR, G.I. (1938), Analysis of plastic strain in a cubic crystal, in: J.M. Lessels, ed., Stephen Timoshenko
60th Anniversary Volume (Macmillan, New York).
TAYLOR, G.I. and CF. ELAM (1923), Bakerian lecture: The distortion of an aluminum crystal during a
tensile test, Proc. Roy. Soc. London A102, 643-667.
TAYLOR, G.I. and C.F. ELAM (1925), The plastic extension and fracture of aluminum crystals, Proc. Roy.
Soc. London A108, 28-51.
TAYLOR, G.I. and M.A. QUINNEY (1933), The latent energy remaining in a metal after cold working, Proc.
Roy. Soc. London A143, 307.
TAYLOR, L.M. and E.B. BECKER (1983), Some computational aspects of large deformation, rate-dependent
plasticity problems, Comput. Methods Appl. Mech. Engrg. 41(3) 251-278.
TAYLOR, R.L., P.J. BERESFORD and E.L. WILSON (1976), A non-conforming element for stress analysis,
Internat. J. Numer Methods Engrg. 10(6), 1211-1219.
TAYLOR, R.L., K.S. PISTER and L.R. HERRMANN (1968), A variational principle for incompressible and
nearly-incompressible elasticity, Internat. J. Solids Structures 4, 875-883.
TAYLOR, R.L., J.C. SIMO, O.C. ZIENKIEWICZ and A.C. CHAN (1985), The patch test: A condition for
assessing finite element convergence, Internat. J. Numer Methods Engrg. 22, 39-62.
TEMAM, R. (1985), Mathematical Problems in Plasticity (Gauthier-Villars, Paris) (Translation of 1983
French original edition).
TEMAM, R. (1984), Navier-Stokes Equations. Theory and Numerical Analysis, 3rd edition (North-Holland,
Amsterdam).

TEMAM, R. (1988), Infinite-Dimensional Dynamical Systems in Mechanics and Physics, Applied Mathematical Sciences, 68 (Springer, Berlin).
TEMAM, R. and G. STRANG (1980), Functions of bounded deformation, Arch. Rat. Mech. Anal. 75, 7-21.
THORPE, J.A. (1979), Elementary Topics in Differential Geometry (Springer, Berlin).
TING, T.T. (1985), Determination of C1/2, C- 1/ 2 and more general isotropic tensor functions of C,
J. Elasticity 15, 319-323.
TRUESDELL, C. and W. NOLL (1965), The nonlinear field theories, in: Handbuch der Physik 111/3 (Springer,
Berlin).
TRUESDELL, C. and R.A. TOUPIN (1960), The classical field theories, in: Handbbuch der Physik lI/1
(Springer, Berlin).
TSAI, S.W. and E.M. Wu (1971), A general theory of strength for anisotropic materials, J. Composite Mater
5, 58.
TVERGAARD, V., A. NEEDLEMAN and K.K. LO (1981), Flow localization in the plane strain tensile test,
J. Mech. Phys. Solids 29, 115-142.
VALENT, T. (1988), Boundary Value Problems in Finite Elasticity, Springer Tracts in Natural Philosophy
31 (Springer, New York).
VOCE, E. (1955), Metalurgica 51, 219.
VON MISES, R. (1925), Z. Angew. Math. Mech. 5, 147.
WANNER, G. (1976), A short proof of nonlinear A-stability, BIT 16, 226-227.
WEBER, G. and L. ANAND (1990), Finite deformation constitutive equations and time integration procedure
for isotropic hyperelastic-viscoelastic solids, Comput. Methods Appl. Mech. Engrg. 79, 173-202.
WILKINS, M.L. (1964), Calculation of elastic-plastic flow, in: B. Alder et al., eds., Methods of Computational
Physics 3 (Academic Press, New York).
WOOD, W.L. (1990), Practical Time Stepping Schemes (Oxford University Press, Clarendon).
WRIGGERS, P., C. MIEHE, M. KLEIBER and J.C. SIMO (1992), On the coupled thermomechanical treatment
of necking problems via finite element methods, Internat. J. Numer. Methods Engrg. 33, 869-883.
YANENKO, N.N. (1971), The Method of FractionalSteps (Springer, New York).

References

499

YODER, P.J. and R.L. WHIRLEY (1984), On the numerical implementation of elastoplastic models, J. Appl.
Mech. 51(2), 283-287.
ZAREMBA, S. (1903), Sur une forme perfectione de la thorie de la relaxation, Bul. Int. Acad. Sci. Cracovie
8, 594-616.
ZDEBEL, U. and T. LEHMANN (1987), Some theoretical considerations and experimental investigations on
a constitutive law in thermoplasticity, Internat. J. Plasticity 3, 369-389.
ZIEGLER, H. (1959), A modification of Prager's hardening rule, Quart. Appl. Math. 17, 55-65.
ZIENKIEWICZ, O.C. (1977), The Finite Element Method, 3rd edition (McGraw-Hill, London).
ZIENKIEWICZ, O.C. and I.C. CORMEAU (1974), Viscoplasticity-plasticity and creep in elastic solids-a
unified numerical solution approach, Internat. J. Numer. Methods Engrg. 8, 821-845.
ZIENKIEWICZ, O.C., D.K. PAUL and A.H. CHAN (1988), Unconditionally stable staggered solution procedure
for soil-pore fluid interaction problems, Internat. J. Numer Methods Engrg. 26, 1039-1055.
ZIENKIEWICZ, O.C. and R.L. TAYLOR (1989), The Finite Element Method I (McGraw-Hill, London).
ZIENKIEWICZ, O.C., R.L. TAYLOR and J.M. Too (1971), Reduced integration technique in general analysis
of plates and shells, Internat. J. Numer Methods Engrg. 3, 275-290.

You might also like