Nothing Special   »   [go: up one dir, main page]

A New Class of Algorithms For Classical Plasticity

Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/226529717

A New Class of Algorithms for Classical Plasticity Extended to Finite


Strains. Application to Geomaterials

Article  in  Computational Mechanics · July 1993


DOI: 10.1007/BF00371865

CITATIONS READS

85 210

2 authors, including:

Günther Meschke
Ruhr-Universität Bochum
259 PUBLICATIONS   2,549 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Collaborative Research Center SFB 837 “Interaction modeling in mechanized tunneling” View project

All content following this page was uploaded by Günther Meschke on 05 June 2014.

The user has requested enhancement of the downloaded file.


Computational Mechamcs (1993) 11,253 278
Computational
Mechanics
@ Springer-Verlag 1993

A new class of algorithms for classical plasticity extended


to finite strains. Application to geomaterials*
J. C. Simo and G. Meschke
Division of Applied Mechanics, Department of Mechanical Engineering, Stanford University, Stanford, CA 94304, USA

Abstract. A recently proposed methodology for computational plasticity at finite strains is re-examined within the context of
geomechanical applications and cast in the general format of multi-surface plasticity. This approach provides an extension to
finite strains of any infinitesimal model of classical plasticity that retains both the form of the yield criterion and the hyperelastic
character of the stress strain relations. Remarkably, the actual algorithmic implementation reduces to a reformulation of the
standard return maps in principal axis with algorithmic elastoplastic moduli identical to those of the infinitesimaltheory. New
results in the area of geomechanics include a fully implicit return map for the modified Cam-Clay model, extended here to
the finite deformation regime, and a new semi-explicit scheme that restores symmetry of the algorithmic moduli while retaining
the unconditionalstability property. In addition, a new phenomenological plasticity model for soils is presented which includes
a number of widely used models as special cases. The general applicability of the proposed methodology is illustrated in several
geomechanical examples that exhibit localization and finite deformations.

1 Introduction

Currently, there is a well-developedframework for the algorithmic treatment of classical infinitesimal


plasticity. Two key ingredients are (i) The so-called return mapping algorithms which, for associative
plasticity reduce the incremental integration of the plastic flow evolution to a convex minimization
problem, and (ii) The so-called algorithmic elastoplastic moduli, obtained by consistent linearization
of the return mapping algorithm. Return mapping algorithms were simultaneously introduced by
Wilkins (1964) and Maenchen and Shack (1964) in the context of J/-flow theory, and subsequently
generalized by a number of authors in a different context; see Simo and Hughes (1990) for a com-
prehensive review. The key interpretation as a convex optimization problem goes back at least to
Moreau (1976). The notion of algorithmic (as opposed to continuum) tangent elastoplastic moduli
was first introduced in Simo and Taylor (1984), leading to a significant improvement in performance
of iterative solution schemes employing Newton and quasi-Newton methods.
Extensions to the finite strain regime of classical models of plasticity have typically relied in
the past on hypoelastic rate formulations of the stress response. Fundamental questions aside,
such extensions make the task of generalizing the standard return maps of the infinitesimal theory
cumbersome, while rendering practically impossible the derivation ~f the algorithmic moduli. In
sharp contrast with this state of affairs, the extension to finite strains of classical infinitesimal
plasticity models advocated herein possesses the following properties:
(i) The stress response is hyperelastic and, consistent with a multiplicative decomposition of
the deformation gradient, characterized by means of a stored energy involving only the elastic part
of the deformation gradient.
(ii) The definition of the elastic domain is identical to that of the infinitesimal theory, with the
stresses now interpreted as true stresses acting on the current configuration of the body. We remark
that the specification of the elastic domain in terms true stresses necessarily restricts the theory to
isotropic yield criteria.

* Partial support provided by the Max Kade Foundation under Grant No. 2-DJA-616 and with Stanford University, and
the Naval Civil Engineering Laboratory at Port Huaneme
254 Computational Mechanics 11 (1993)

(iii) The associative form of the flow rule compatible with a multiplicative decomposition of
the deformation gradient is uniquely defined by resorting to the principle of maximum plastic
work.
From a computational perspective the central issue concerns the numerical integration of the
evolution of plastic flow. Remarkably, by applying an exponential approximation to the flow rule
in the form recently proposed in Simo (1991), one exactly recovers the standard return mapping
algorithms of the infinitesimal theory, now formulated in principal Eulerian axes. Within this
approximation, the algorithmic elastoplastic moduli in principal axes coincide with the algo-
rithmic moduli of the infinitesimal model. In summary, the actual implementation in the finite
strain regime of a classical model of infinitesimal plasticity involves two steps:
(iv) A trial elastic state computed by an approach identical to that recently described in Simo
and Taylor (1991) for finite hyperelasticity formulated in principal stretches.
(v) A corrector step defined by a return mapping identical to that of the infinitesimal theory,
now formulated in fixed Eulerian principal axes defined by the trial state.
We remark that the tangent operator (the Hessian) in this formulation is symmetric and possesses
a structure identical to that of finite elasticity formulated in principal stresses, with material tangent
in principal axis identical to the algorithmic elastoplastic moduli of the infinitesimal theory.
A main goal of this work is to demonstrate the effectiveness and general applicability of the
preceding computational framework to the complex models of phenomenological plasticity
that arise in geomechanical applications. Two specific applications are considered. In Sect. 4 a
formulation and algorithmic treatment of the modified Cam-Clay model of Roscoe and Burland
(1968) is presented within the realm of true finite strains. In addition to (necessarily non-symmetric)
fully implicit schemes, an alternative class of new semi-implicit return maps is described which
restores the symmetry of the algorithmic moduli while retaining the property of unconditional
stability. In Sect. 5 we undertake the formulation of a fairly general class of phenomenological
models of plasticity which include as special cases widely used models in soil mechanics. Since the
algorithmic treatment of this class of models is identical to that of the infinitesimal theory, now
formulated in principal stretches, further discussion on computational aspects is omitted. The
reader is directed to Simo et al. (1988) for a comprehensive discussion. Section 6 describes a number
of representative simulations in geomechanics, exhibiting localization and finite deformations,
which demonstrate the effectiveness of the proposed methodology. Conclusions are drawn in
Sect. 7.

2 Multiplicative elastoplasticity at finite strains

We give a concise outline of a formulation of plasticity at finite strains in which the elastic domain
IE is specified by any classical model of infinitesimal plasticity, with the stress tensor now inter-
preted as the true (Kirchhoff) stress tensor acting on the current configuration of the body. As in
the infinitesimal theory, in the present formulation the associative form of the evolution equations
is uniquely defined via the principle of maximum plastic work (Hill 1950). Throughout this presenta-
tion attention is restricted to the purely mechanical theory. For a complete thermodynamical
treatment we refer to the recent work of Simo and Miehe (1991).

2.I Motivation: The infinitesimal theory

To motivate our subsequent developments we summarize first the key assumptions underlying a
classical model of infinitesimal elastoplasticity. In fact, the formulation of finite strain plasticity
described below will be constructed by specifying a direct counterpart of these assumptions. In
particular, the standard evolution equations describing associative plastic flow can be uniquely
extended to finite strains via their equivalent statement in terms of maximum plastic work.
i. Elastic domain. A convex elastic domain IE defined by a m-convex yield functions defined in
J. C. Simo and G. Meschke: A new class of algorithms for classical plasticity 255

stress space and intersecting (possibly) in a non-smooth fashion; i.e.,


IE = {(~, q)~S x lR'~=~:q~u(~,q) < 0, for # = 1, 2,..., m}, (2.1)
where S is the vector space of symmetric rank-two tensors and q is a suitable vector of m~,t > 1
(stress-like) internal variables characterizing the hardening response of the material.
ii. Additive decomposition and free energy. The total infinitesimal strain tensor is decomposed
additively into elastic and plastic parts as
~.(X, t) = ~.e(x, t) + gP(X, t). (2.2)
The elastic response of the material is assumed to depend only on the elastic part e~ and is character-
ized by means of a free energy function ~b(e~, or). Here a~is the vector of (strain-like) internal variables
conjugate to q; i.e., q:= - c3~.
iii. Associative evolution of plastic flow. Restricted to the purely mechanical theory the dissipation
function ~ is the difference between the stress power and the rate of change of the internal energy.
The second law then implies
~@ = 'I"~ - - ~(,~e, 0D ) 0. (2.3)
Clearly ~ -= 0 for an elastic material. The simplest model that describes the evolution of the internal
variables {Z, ~} corresponds to the assumption of maximum dissipation. According to this postulate,
among all possible admissible states ('~,~)eIE the actual state (~, q)slE is that leading to maximum
dissipation for prescribed rates of deformation. Maximum plastic dissipation is the key property
of associative plastic flow amenable to a unique and straight-forward generalization to the finite
strain regime. As a motivation, we show below how this assumption yields the classical evolution
equations describing associative plastic flow.

2.1.1 Constitutive equations. Dissipation inequality. The three assumptions above completely
define the evolution equations of classical plasticity, as follows. First, using the chain rule expres-
sion (2.3) for the dissipation function becomes
= [~ - c~o~b].~ + ~e~b"E~ -/;el + [ _ c~,~b]"d~> 0. (2.4)
Since Eq. (2.4) is to hold for all admissible processes, a standard argument produces the stress-strain
relations and the reduced dissipation inequality:
= ~?~b(~e, ~) and ~ = t" [/~ - U] + q" & > 0, (2.5)
where q = - c3a@(~e, 00. Second, with this expression in hand the assumption of maximum plastic
dissipation translates into the inequality:
[~-~].[~-U]+[q-~].&>0, V(~,q)~IE. (2.6)
Consequently, if the elastic domain IE is defined by the convex set (2.1), a standard result in convex
analysis shows that inequality (2.6) holds if and only if the following conditions hold:

U=k- ~ 7,~?,qSU(~,q), &= ~ 7u~q(aU(~,q), 7 , > 0 , q~u(T,q)<0, ~ 7,qSU(~,q) -- 0, (2.7)


,u=l #=1 #=1

which furnish the standard local form of the evolution equations for associative plasticity. We
remark that the maximum dissipation inequality (2.6) leading to (2.7) plays a central role both in
the mathematical theory and in the numerical analysis of classical plasticity (see e.g., Simo (1991a)
and references therein).

2.2 The nonlinear theory. Basic assumptions

In our formulation of classical models of infinitesimal elastoplasticity in finite strain regime we


retain the form of elastic domain IE in the infinitesimal theory with v now interpreted as the
256 Computational Mechamcs 11 (1993)

Kirchhoffstress tensor in the current configuration. This key design condition defines the structure
of the nonlinear theory as follows.
i. Elastic domain. For simplicity, attention is restricted in this section to a single scalar hardening
variable; thus
IE:= {(~, q)eS x IR: q~"(~,q) __<0 for/~ = 1,2,...,m}. (2.8)
We remark that the principle of objectivity restricts the possible forms of q5in Eq. (2.8) to isotropic
functions. Consequently, a classical formulation of the yield criterion in terms of true stresses
necessarily imposes the restriction to isotropy of the resulting theory.
ii. Multiplicative decomposition and free energy. The generalization to finite strains of the additive
decomposition used in the infinitesimal theory is motivated by the structure of the single crystal
models of metal plasticity, and takes the form of the local multiplicative decomposition
F(X, t) = Fe(x, t)FP(X, t) (2.9)
where F(X, t):= Do(X, t) denotes the deformation gradient of the motion q~(.,t) at X ~ . From a
phenomenological point of view F p is an internal variable related to the amount of plastic flow
while F e- 1 defines the local, stress-free, unloaded configuration. Consistent with the restriction to
isotropy it is further assumed that ~ is independent of the orientation of the intermediate configura-
tion. Frame invariance then implies the functional form
0=tp(b~,~), where b~:=FeF ~r. (2.10)
The measure of elastic deformation b e defined above is the elastic left Cauchy-Green tensor.
iii. Associative evolution of plastic flow. The local dissipation function ~ per unit of reference
volume associated with a material point Xe:~ is defined as the difference between the local stress
power and the local rate of change of the free energy; i.e.,
~ : = x ' d - ~(be, c~)> 0. (2.11)
Here d = sym [1] denotes the rate of deformation tensor and 1:= ~'F- 1 is the spatial velocity gradient.
As in the infinitesimal theory, the simplest model describing the evolution of plastic flow is obtained
by postulating maximum dissipation of the flow.
Following an argument entirely analogous to that used for the infinitesimal theory, we show
below that the preceding assumptions uniquely define a model of associative plasticity at finite
strains.

2.3 Constitutive equations. Dissipation inequality

Our first objective is to obtain an explicit expression for the dissipation. The rate of change ~ is
computed by exploiting the relation
be = FCP-IF r with CP:= FprF p, (2.12)
which follows from the multiplicative decomposition (2.9), and identifies C p as the plastic right
Cauchy-Green tensor. Time differentiation of this expression yields the identity

[je:lbe+belr+~.~vbe with ~q~b~:=F~[Cp-1]F r, (2.13)


dt
where ~vb e is referred to as the Lie derivative of b e. Time differentiation of the free energy function
and use of the relation (2.13)1 yields the result
d/,4
~-~ = aboO'l~ e + cg,0~ = [28boObe] 9El + 89176 e- 1] + 8 ~ . (2.14)
dt
Since be commutes with c?b~p as a result of the restriction to isotropy, combining Eq. (2.14) and
J. C. Simo and G. Meschke: A new class of algorithms for classical plasticity 257

Eq. (2.11) yields the following local form of the dissipation inequality:
= [~ - 2C3be0be]-d + [2~bo~tbe].[-89 ] + [-- 0a~]~ =>0, (2.15)
which furnishes the finite strain counterpart of (2.4). Since (2.15) must hold for all admissible
processes, the same standard argument used in the infinitesimal theory now gives the following
constitutive equations and the reduced form of the dissipation inequality
~ = 2c3b~0be, ~ =,c.[ 15(~vb e ) b e - 1] + qo~ >0,~___ (2.16)
with q:= -~?=~p. These relations are the finite strain counterpart of relations (2.5).
2.3.1 Associative evolution equations. Maximum dissipation. According to the postulate of maximum
dissipation the actual state (~, q)elE in the plastically deformed body at a given (fixed) configuration,
with prescribed intermediate configuration and prescribed rates {&t'~,be, 0i}, renders a maximum of
the dissipation function 9. Equivalently, the following inequality holds:
Ix - ~3.[ - l(~q~~ 13 + [q - ~]0i > 0, (~, ~/)6IE. (2.17)
This result is the counterpart of inequality (2.6) in the finite strain regime. As in the infinitesimal
theory, inequality (2.17) holds if and only if the coefficients {--l(~.~vbe)be-l,~ } lie in the cone
normal to c?IEat the point (x, q). In particular, if ~?IEis defined by Eq. (2.8), the evolution equations
read
7u0~qSr z, q)] be
#=i

7u~qqSU(~, q)]
,u=l

7,>0, q~U(~,q)__<0 and ~ 7u~bU(~,q)=0. (2.18)


,u=l
Despite its unconventional appearance (compare with the corresponding expression (2.7) in the
infinitesimal theory) the flow rule (2.18) possesses a number of interesting properties. In particular,
it gives the correct evolution of plastic volume changes as the following observations show.
(i) The total and elastic volume changes are given by J : = det IF] > 0 and je: = (det [b e])1/2 > 0,
respectively.
(ii) Let JP:= det [FP]. Setting e"'=
/)" log [JP] the rate of plastic volume change predicted by the
flow rule (2.18)1 is given by the evolution equation

i p= ~ 7,tr[~,b"], (2.19)
#=1
which implies exact conservation of plastic volume for pressure insensitive yield conditions; i.e. if
tr[0~4~"] = 0.
The proof of the result in Eq. (2.19) follows easily from Eq. (2.13) along with Eq. (2.18)1. The
foregoing theory will be applied to a number of classical models in soil mechanics in the following
sections.
2.3.2 General non-associative evolution equations. Expression (2.18) provides a crucial insight in
the extension to finite strains of non-associative models of infinitesimal plasticity. For instance,
for single-surface plasticity (i.e.,/t = 1) the non-associative flow rule takes the form
k e = ~ - - 7(~9('~, q), 0i = 70qg(~, q), (2.20)
where g:S x IR ~ IR is the plastic flow potential; in general g # q5. The extension of a model of this
type to the finite strain regime then takes the form
- - ~1 ' v b e = [7~?~g(~,q)]b e, ~ = 7~qg(~, q), (2.21)
where g(~, q) has identical form as in the infinitesimal theory.
258 C o m p u t a t i o n a l M e c h a m c s 11 (1993)

3 Exponential return mapping algorithms

Following Simo (199 lb) we consider an algorithmic approximation consistent with the associative
flow rule derived above which is 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. Remarkably, the implementation of this algorithm takes a
form essentially identical to the standard return maps of the infinitesimal theory.

3.1 The local problem of evolution. The algorithmic flow rule

Consider a typical time sub-interval [t,, t,+ 1], an arbitrary material point X e ~ and assume that
the deformation gradient F, along with {be, c%}ES+ x R+ which define the constitutive response
of the material, are prescribed initial data at XsN. The objective is to construct an algorithmic
approximation to the plastic flow evolution equations for incremental displacement field u t pres-
cribed on tp,(~) x It., t,+ 1], where %(-) is a given configuration at time t, defined on ~. Denoting
by ft(x,):= I + Vut(x,) the relative deformation gradient at x, = %(X), the local evolution of plastic
flow is governed by the first order constrained system

b" te _- [ltb t + b te Itr ] __ 2 bt

7u >0, 0 ( t, qt) < 0 and ~ 7u4 ( t, qt) O, (3.22)

where i, = ~,ff 1 is the spatial velocity gradient. The solution of (3.22) defines {b7, c~t} at the point
X within the interval of time [t,, t, + 13 for configuration prescribed by the relation tPt = tp, + uto ~p,.
The key idea in the design of an integration algorithm for Eq. (3.22) is to recast this system as
a standard problem of evolution, of the form j? = H(y, t)y, and then apply the exponential approxima-
tion y, = exp [AtH(yt, t)]yr This idea is carried out in two steps:
Step 1. Recast Eq. (3.22) in the form of a nonlinear problem of evolution of the form j~ = H(y, t)y.
In the present context this is accomplished by pull-back of Eq. (3.22) to the fixed initial configuration
tp,(N), as follows. Define the tensor fields
b;*:=ftlbff/r, ct:=ftrft and n~:=fr~(~u('c,,qt)f,. (3.23)
In geometric terms nt," # = 1, 2 , . . , m are interpreted as the pull-back of the normal d~q~Uto the
configuration (p,(N), with the tensor c t = ffft viewed as a Riemannian metric on ~p,(N). A straight-
forward manipulation shows that the system (3.22)1 is equivalent to

~ge* =- -- 2 -1 ~ b e*
t, with h~*lt=t,, = ben' (3.24)
~t t =1 ~/uct nt

which is now written in standard form.


Step 2. Using an exponential approximation to Eq. (3.24) within the interval It,, t ] c [t,, t,+ 1]
yields the first order accurate formula b~* ~ exp I - 2 ~ AT.et- ln~l be' where ATu >
= 0 is a Lagrange
iz=l
multiplier to be determined by enforcing consistency as described below. Using Eq. (3.23)1 this
relation becomes

b:=(f'explh2 tr c n lf l)b:l where b:tr=fb:f 325,


J. C. Simo and G. Meschke: A new class of algorithms for classical plasticity 259

Fixed ReferenceConfiguration

Fn @ b~ctn

"<, Fn+l ~tr


b n + l o~n
Fig 3.1. Interpretation of the trial
state: The (local)referenceand inter-
mediate configurationsremain fixed,
while the current configuration is
updated for prescribedrelativedefor-
Fixed IntermediateConfiguration Updated CurrentConfiguration mation gradient

where the symbol ~ has been replaced by the equal sign. The algorithmic approximation is
completed by exploiting the well-known property F exp[A]F-1 = exp[FAF-1], which yields

bt=ex p [-2 .=*~ ATu~qSU(rt, qt)]bt tr

at = ~. + ~ ATu0q~u(~t, qt)
#=1

AT,.>0,
= q~'%,q3 <0, ~ A , ~,~b".c
( t, q0 = 0, (3.26)
#=1

where (% qt) are defined by the elastic stress-strain relations evaluated at (bt, at). A standard
argument using a Taylor series expansion confirms that the product formula algorithm (3.26) is
first order accurate. Unconditional stability follows from the properties of the backward Euler
scheme and the exponential map. A geometric illustration of Eq. (3.26) is given in Fig. 3.1.
The derivation of the algorithmic flow rule (3.26) is identical to the algorithmic treatment in
Simo (1988), with the backward-Euler scheme replaced here by an exponential approximation.
Although both algorithms possess identical accuracy properties, the backward Euler method does
not inherit the following conservation properties which are identically satisfied by Eq. (3.26).
(i) Preservation of the constraint det[b[] > 0; a property which follows from the identity
det [exp [A] ] = exp [tr [A] ].
(ii) Exact conservation of the plastic volume. For pressure insensitive yield criteria the plastic
volume preserving property of the exact flow rule is preserved by the algorithmic approximation,
since Eqs. (3.26) implies tr [~?,q~"(~t,q,)] - 0,~JtP = JP,.
A modification of the backward Euler method which does inherit property (ii) above is given
in Simo and Miehe (1991). We remark that both the backward Euler and the exponential approxima-
tions become exact for incrementally elastic steps since ATu = 0.

3.2 Spectral decomposition: Formulation in principal axes

By isotropy the principal directions {nlA)}A= 1,2,3 of the Kirchhoff stress tensor ~t coincide with
the principal directions of the elastic left Cauchy-Green tensor b~. Let {O-A}and {2A2} denote the
respective set of principal values. The key to the implementation of the algorithm lies in the follow-
ing observations:
t and '~A
(i) Let n tr(A) etr
> 0 denote the principal directions and the square root of the associated
principal values of b~ tr, respectively. Then, from (3.26)~ one immediately concludes that the final
principal directions coincide with the principal directions; i.e., n~A) = nlr(A) for A = 1, 2, 3.
260 Computational Mechanics l 1 (1993)

(ii) Let I~tA 9. = log [),ta ] a n d etA


etr.. = log t-"tAI-
]etrq.j denote the logarithmic stretches. Setting ~b(b~, ~) =
q~(e], ~), the constitutive equations become
ff A = ~ . e . 4 ~ ( e Ae' O~t)'
qt = - - (3 1if(8;, ~t) ' (3.27)
which possess a functional form identical to that of the infinitesimal theory.
(iii) Set ~b"(,, q) - q~U(al,0"2, 0"3, q)- Then the algorithmic flow rule collapses to

,e etr ~ A~#~aA ~#(O.t, qt )


~tA ~ ~'tA - -
#=1

#=1

aTu>0, 6~'(at,qt)<O, ~ aT,6~'(o',,qt)=0, (3.28)


/t=l
which along with Eq. (3.27) define a return mapping algorithm at fixed principal axis (defined by
the trial state) identical to that of the infinitesimal theory.
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) axis defined by the trial state, which can be computed in closed form directly from b~t~,
as explained next.

3.3 Closed-form implementation

The step-by-step procedure outlined below summarizes the actual implementation of the updating
scheme defined by preceding algorithm within the interval [t,, t,+ 1]. In a finite element context,
this update procedure takes place at each quadrature point of a typical element.
Step 1. Trial state. Given {b~,c~,} (at a specific quadrature point x,~q~,(~)) compute the trial
elastic left Cauchy-Green tensor and the relative deformation gradient f, + 1 for prescribed deforma-
tion u, + 1: q~,(~) ~ jR3 as
f , + l ( x , ) : = I + Vu,+l(x,), b n~t~
+l
=f,+ 1 bef T
n n+l"
(3.29)
Step 2. Spectral decomposition of~beta.+1. Compute the principal stretches t~2 ,etr+ 1A } by solving the
cubic characteristic equation in closed form via Cardano's formulae. Compute the principal direc-
tions ,fn . + l J via the closed form formula
t. tr(A)/,
[- betr (,~etr ~21 7[- h etr (~Letr "12"1- -]
Htr(A) (~ ntr(A) = | - n + 1 ~ '.--n + 1 B ~ | [ --n+ 1 ~ V*n + I_C.' - | (3.30)
n+l'~ n+l |r "12__(~etr ~2]|ggetr "~2 (~etr ~2|'
t_,. n + l B / p'~n+lA ? -II-\"~n+IC] --~ n+lA ! -J

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 logarithmic stretches
and the principal trial stresses:
etr = [-lo_Z~etr ~Letr etr T
~ L g~ n + l l ) , l o g ( n+12),log()~n+13)]
o'tr (~ *l~t(getr ~n),
tr
qn+l =--~ I~/(Ee tr ~n)"
~t't" ~ n + l
(3.31)
n+l ~ "r', n + l ~
Evaluate q~t~+1 = q~(tr~§ ~,q~+ ~). If q~tr+1 > 0 perform the return mapping algorithm (see remarks
below) to compute {e~+ t, ~, + 1}. Compute the Kirchhoff stress tensor via the explicit formula
3
~n+l = E ffn+lAS|
~tr(A) ~ B,tr(A) (3.32)
A=I
where 0-,+ 1A = ~,~tp(ee + 1, ~,+ 1) are the principal Kirchhoff stresses.
J. C. Simo and G. Meschke: A new class of algorithms for classical plasticity 261

o ,/' \ , \ Ck=O

Fig. 3.2. Geometric interpretation of the return mapping


................... . 1 / algorithm for multlsurface plasticity in principal stress space
for two possible trial elastic states
r

Step 4. Update of the intermediate configuration. Compute the updated left Cauchy-Green tensor
via the spectral decomposition
3
b~,+2 = ~ exp[2~+IA-j•tr(A), + l ~~ - -ntr(A)
,+l, (3.33)
A=I

n + l (A = 1, 2, 3) are defined in Step 2.


w h e r e n tr{A)
It can be shown that the linearization of the preceding algorithm takes the following closed-
form expression:
3 3 3
Ct= Z Z he" mtr(A)(~mttr(B)-t- 2 ~, '- otr(A) (3.34)
~tAB'"t UtA~.t
A=I B=I A=I
The moduli c~mA) in the last term of Eq. (3.34) are independent of the specific model of plasticity
under consideration and depend only on the form of the stored energy function in principal stretches.
A closed-form expression is given in Simo and Taylor (1991). Similarly, the tensor products
mma)
t = ,tr(A)
"'n + 1 ~
"r ,tr(A)
"'n + 1 are independent of the specific model of plasticity and, therefore, the algorithmic
moduli defined by Eq. (3.34) are completely specified once the coefficients a~]~ are determined.
These coefficients depend solely on the specific model of plasticity and, remarkably, possess identical
structure as the algorithmic elastoplastic moduli of the infinitesimal theory.

Remarks 3.1 1. The explicit formula for the principal directions given in Eq. (3.30) above assumes
three different principal stretches. Although similar expressions can be derived for repeated
principal stretches, this latter situation is most effectively handled numerically by introducing a
perturbation of the repeated roots and applying (3.30).
2. The classical return mapping algorithms in stress space of the infinitesimal theory are
exactly recovered in the present finite strain context if the free energy function is quadratic in
the principal logarithmic stretches; i.e.:
~(,~A,~) = 1 e + + ;12 + + (3.35)
where 2 > 0 and/~ > 0 are the Lam6 constants,/((.) is a function which characterizes isotropic
hardening in the material and Ca:= log(2]), for A = 1, 2, 3. Using vector notation, the linear elastic
stress-logarithmic strain relations emanating from Eq. (3.35) read
o'=ag e, a : = t c l | +2//[I 3 - 89174 (3.36)
where 13 denotes the 3 x 3 identity matrix, ~c:= 2 +-~// is the bulk-modulus, a is the (3 x 3)
matrix of constant elastic moduli in principal stress space, and 1 = [1 1 1] r.
3. The geometric interpretation of the stress-space return mapping algorithms of the infinitesimal
theory, now phrased in principal stress space, is illustrated in Fig. 3.2 for the general case of multi-
surface plasticity. Assuming for simplicity linear isotropic hardening, as in the infinitesimal theory
262 Computational Mechanics 11 (1993)

one can show that (Simo 1991b)


[fl -- o.+13.a-1 [o,+
_tr 1-a.+l]+[~/-q.+~]H-l[q.-q.+l]<0 V(II,q)~IE. (3.37)
Standard results in convex analysis yield the following interpretation of inequality (3.37): The final
state ( ~, + 1, q, +1)~IE in principal stress space is the closest point projection of the trial state (at,r+1, q,)
onto the elastic domain in the metric defined by the inverse moduli (a -a, H-1). This result reduces
the algorithmic problem in principal Kirchhoff stress space to the standard problem in convex
optimization: Find the distance of a point (the trial state) to a convex set (the elastic domain) in
a given metric (the inverse elastic moduli). The standard algorithm for this problem is a variant
of the gradient-projection method proposed in Simo et al. (1988).
Form a practical standpoint, the preceding developments translate into the following prescrip-
tion for the implementation in the finite strain regime of any infinitesimal model of classical
plasticity.
(i) Express the model in principal axes. Note that the classical linear infinitesimal elastic stress-
strain relations 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 the
standard return mapping algorithms, along with the algorithmic elastoplastic moduli a ep (in
principal axes).
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.

4 Finite deformation Cam-Clay

The methodology presented in the preceding sections is exploited here in the extension to the
finite strain regime of a classical model of soil mechanics: the modified Cam-Clay model of Roscoe
and Burland (1968). This model remains one of the most widely used plasticity models in geo-
technical applications due to its capability of predicting (at least qualitatively) the most relevant
characteristics of the mechanical behavior of cohesive soils; see e.g. Wood (1990) for an elaborated
discussion. The modified Cam-Clay Model is characterized by ellipsoidal-shaped yield surfaces
in the stress space, an isotropic hardening and softening mechanism and by a nonlinear elastic
constitutive law.
Algorithms for the Cam-Clay model restricted to the infinitesimal theory have been considered
by Ortiz and Simo (1986), within the framework of their (explicit) cutting plane algorithms, and
Borja and Lee (1990) using the standard return maps of classical plasticity in conjunction with an
explicit treatment of the nonlinear elastic law, among others. An implicit treatment of the elastic
law is proposed in Borja (1991), outside the scope of the standard return maps, with the evolution
of the void ratio treated explicitly and non-symmetric algorithmic moduli. Here, in addition to a
formulation of the Cam-Clay model in the regime of finite strains, we address a number of novel
algorithmic aspects; in particular:
(i) An implicit algorithmic treatment of the nonlinear elastic constitutive law along with an
implicit integration of the flow rule and the hardening laws which conforms to the now standard
return mapping algorithm of classical plasticity.
(ii) An alternative semi-explicit integration scheme which retains the implicit character of the
return mapping algorithm while adopting an explicit integration of the hardening law. The key
advantage of this approach is the restored symmetry of the algorithmic tangent matrix, achieved
without loss of unconditional stability.

4.1 Nonlinear elasticity

A characteristic feature of Cam-Clay is the nonlinear elastic relationship between the (Kirchhoff)
pressure p and the elastic volumetric strain ee. Here we incorporate this effect via a stored energy
J. C. Simo and G. Meschke: A new class of algorithms for classical plasticity 263

function of the form:


1 3
co(ee, e~):= U(ee) + CYO(eA)= U(ee) + 2G-izA=~I(eeA)2' A = 1,2,3, (4.1)

where e~, e 2, e 3 are the principal deviatoric elastic (logarithmic) strains. In what follows, it is more
convenient to use the dual form of Eq. (4.1) obtained via the Legendre transformation as
3
2 A = 1,2,3,
Z(P, SA):= )?(p) + 2(SA) = ~(p) + 4 AZ=1 sA, (4.2)

where st, s2, ss are the principal deviatoric (Kirchhoff) stresses and G denotes the (constant) shear
modulus. The specific form of)~(p) is derived below from the following rate form of the volumetric
stress-strain specific to the Cam-Clay model
~e:= )7,,(p)/~ = f~e(p)/v(p). (4.3)
Here v is the specific volume in loading processes while v~ is the specific volume in unloading/
reloading processes (Roscoe and Burland 1968). These specific volumes are related to the pressure
via the relations
v(p):=N-21og(-p), ve(p):=N--Klog(--p) for pN-1, (4.4a)

where N is a material parameter defined as the specific volume at unit compressive pressure, 2
denotes the consolidation index and ~cis the swelling/recompression index. (Note that compression
is associated with a negative sign). Since positive values ofp may occur in practical boundary value
problems, the following linear extension of Eq. (4.4a) is considered in what follows:
v(p):=N+2(l+p), ve(p):=N+K(l+p) for p>-l. (4.4b)

Substituting Eq. (4.4) into Eq. (4.3) gives


K
-p(N_21og(_p))~ if p _ - < - l ,
V (4.5)
K
N + 2(1 + p ) / ~ otherwise.

1
The constitutive equation for the elastic deviatoric (logarithmic) shear strains is eA = ~ SA, while
the volumetric part is computed by integrating Eq. (4.5) in time to obtain

61~ 1 - r~-r-rl
;/ l7-7S.
~ ] if p < - l ,
~e = )~,(p) = (4.6)
6log ( l + l + t ~/
l p) otherwise,

2
where 6 = K and q = --. Finally, the inverse relation gives the pressure as the gradient of the volu-
2 N
metric part of the stored energy function via the following explicit expression:

--expU!---(~-~)exp(-~)lLr/ if e:<61og(1-~q),
p = U'(e~) = 1 + t1V ['e~'~ 7 (4.7)
-~- L exp ~ ~ ) -- 1 J otherwise.
264 Computational Mechanics 11 (19931

3~z Critical State Line

subcritical

p-

Fig. 4.1. Cam-clay: Meridional sections and critical state hne

4.2 Structure of the model

The yield function may be formulated in the following (generalized) convex format:

f(I1,J2,q(oO):=~fJ2-~-(I+M 2 - +3
1 ~q)
2
2M
, / ~ q ----0,
~ g
(4.8)

where T1 = 11 - 3t, t denotes the hydrostatic strength of the soil, M is the slope of the critical state
line in the ~//~z - p-plane and the hardening parameter q(~) represents the consolidation pressure
(Fig. 4.1).
The evolution of the plastic strains and of the hardening variable ~ is governed by an associative
flow rule
Of ~P= Of
e~=?~s~'~, v Y~pp, A = 1 , 2 , 3 , (4.9)

along with the hardening rule

v(/3) q, d ~ ~p___Y~pp'
Of /3:= p -- t. (4.10)
(l=-H(p,q)d~, H:= 2--~

of of
Note, that the hardening law is nonassoeiative in the sense that 0i = 7 ~pp ~ Y~qq.

Remarks 4.1 1. The yield function (4.8) and the hardening law (4.10) differ from the original
formulation of the modified Cam-Clay model (see Burland and Roscoe 1968) in that the soil
cohesion represented by t is taken into account. Consideration of t may also be useful in the simula-
tion of hydrostatically prestressed conditions arising, for example, in standard triaxial tests. If t
is to be considered, the elastic constitutive laws (4.6), (4.7) need to be modified by substituting p
and e~by/~: = p - t and g~ = ~ + s_~, respectively, where e~,t is obtained from Eq. (4.6) as s~(p = - t).
2. Since c)changes sign with --,~ stress states on the subcritical (supercritical) side of the critical
0p
state line (Fig. 4.1) are associated with hardening (softening).

4.3 Fully implicit integration algorithm

The numerical integration scheme for the model falls within the general scheme described in Sect. 2
and will not be further discussed here. The only observation to be made concerns the specific
(non-associative) form of the hardening law (4.10). A convenient approach which reduces the
J. C. Simo and G. Meschke: A new class of algorithms for classical plasticity 265

problem to the standard situation discussed in Sect. 2 and involves only replacing ~e_ by ~- adding
cq cp
2 terms in the algorithmic tangent, proceeds as follows. We integrate (4.10)~ over a finite time step
It,, t, + ~] and use the fact that c~= B~. In the k-th iteration within It,, t, + ~], the elastic volumetric
strain is given as
B;(k)= /~;,tr 0~(k)..~~n, (4.11)
where B~'tr = 1og(je). Hence, the pressure p(e~) may be expressed as a function of the plastic harden-
ing variable c~. Implicit integration of Eq. (4.10)1 yields the following expression for q(k)(ee(k)(c~(k))):

q(k)(~tk))=q, exp p exp exp~--))], (4.12)

where
+
ge = /~e + Be ge.tr = Be,tr _~_ Be and p -
V I) v~t' U V V~t'
)~-- tr
The complete algorithm is summarized in Box 4.1. The resulting algorithmic tangent matrix a ep
is obtained from the moduli Atj, i,j = 1,..., 4 as
a~a~= ~An--(nA'mB)/(h/A~,'h,), A,B= 1,2,3, r,s = 1,...,4, (4.13)
with
3
~An = Aan + AA4 + A4B + A44 -- 2 (AAc + Agc)/3 (4.14a)
C=I
and
na = (Aar + A4~)'h~, m a -= (Ara + Ar4)'~lr. (4.14b)
As a result of the nonassociative hardening law, a ep is nonsymmetric.

Box 4.1. Fully implicit return mapping algorithm for Cam-Clay

1. Initialize: k = O, eP(~+ x = ee., gev,(O) = Bv,etr ..~ Bey,t,~ n +~(O1) = O~n'In +~(O1) = O.
2. Evaluate the pressure, the hardening parameter, the elastic bulk modulus and the hardening
modulus:
p~k) U,(ge~k~)
[- // //~e(k) ~ //~e,tr"~ '~ q
q,k)= q. exp L p ~exp ~ ~ ) -- exp ~ - ) ) J,

Klk = l(plk _ t),


_ v(p - t) q(k .

3. Evaluate residual and check convergence:


f(k) = f(s~k), p(k), q(k)), sT = [-St , $2 ' $3],

/,(k) ~ -~(k)-{-0~tr )~_~f/(k), hT=[~fs,~fpl"

if f ( k ) < T O L and Ihrl < T O L then EXIT. (Continued)


266 Computational Mechanics 11 (1993)

4. Linearize the above equations and compute the algorithmic moduli:

r (k) + A (k)-' i{ As \/(k)+ A? ~(k)= 0,


\Ap/ Lgs'~qqJ'

!(.~• 3) 1 + ~__
FA(k) ] - t _ ~s2
? k TsOp + OsDq k--S/
K(,
t-" ~(4 x 4)~ O2f

Yt?p t?s + \ Op
5. Solve the linearized problem to obtain A7(k), A e p'(k) and A~(k):

As (k) = - 2G Ae p(k)
6. Update s, 0~and compute ~"~9"
s(k+ 1) = s(k) -I- A s (k), ~(k+ 1) __~ (z(k) _~ h a ( k ) g;,(k + 1) ~-. g;.tr __ g(k+ 1) ~_ (~n"

Set k ~ k + 1 and goto 2.

4.4 Explicit treatment of the hardening law


The issue of a (hardening-induced) nonsymmetric tangent may be entirely by-passed by retaining
the implicit form of the return map in the flow while introducing an explicit treatment of the
hardening law. Accordingly, integration of the evolution laws (4.9), (4.10) over a finite time-interval
It., t.+ 1] is now achieved by means of the formulae

ePA,. + 1 = epA,. + ~ ~SA(SA,. + 1' Pn + 1, q . ) , A -= 1, 2, 3

~P
V,n q- 1
s
~,n
+ ~p (s~,. + 1, P. + 1, q.)

q"+ l = q" -- TH'(p"' q") ~P n (sa,., P., q,), (4.15)

U is evaluated at t.. The consistency parameter 7(k) in the k-th iteration is


where the gradient ~pp.
obtained from linearization of Eq. (4.15) altogether with
f(k)
- I n + 1 = jtsa--~(k)._~.1' ~'.+
(k)
n pq.) = 0, A = 1,2,3, (4.16)
as

f(k)--h(k)A(k)r(k) ]'!T [~T2f ~?f l (k) (4.17)


AT(k)- c3f(k) ~fl ' ='C u~' ~P J "
h(k)rA(k)h Ik) + ~ H, ~p ,

As pointed out above, the resulting tangent matrix a ep has now a symmetric form which is obtained
from Eq. (4.13) along with A Ck)in Box 4.1 merely by setting H (k) = 0 and h = h.
J. C. Simo and G. Meschke: A new class of algorithms for classical plasticity 267

= 25
30
= 1.33 ~0 -- ~- ~ ' ~ ' t ." ,,_%~= 2.0 % --4x--
2 20 ~- r 'At = 0.33 % - - 15 [- .~" - At=0..5%--
;~ ~ implici[:~: = 1.33 % --m-,- 10 l-J~ im~lici+:A+ = 2.0 ~, --+~--
+ Io ,+,+ = + ++ =

0 ~ , I , I , ~ , I , I , 0c~' ' ~ ' ~ ' ~ ' + ' + '

0 2 4 6 8 10 12 0 2 4 6 8 10 12

a displacement (cm) b displacement (cm)


Fig. 4.2a and 5. Triaxml test: Pressure-displacement curves from the fully-implicit and the semi-explicit algorithm for different
time steps and different b o u n d a r y conditions ( C a m - C l a y model): a rigid platens, 5 frictionless platens

V O L . PL. S T
9 ; ..'~': ...... :- : i-' '"
~?,,t.,' . ; i " : , '-., <-5 900E-02
!--,,..,.. i ......... f +-- .' ; - + " , .~&,~,
:~;';~"!i !2. L" ;-;-;-.'-.2.
r - - ~ - - ' - . , i ' ~ " , ' + - , f .-..-. ~ ' o - + - ; ~ . . ; . .
]
-.. r.~...-.z:...~:...~_=
I "~ ,4.
.....
9 'r-'~ .... '-2.- .... Z, . . , . , . . . . . . .

"'- ..... 1"-: +. . . . . . . . . J .,,I--+i . . . . . .


i't ....... a_. i ...... ~- ........ i ......
~-. ..... ~,., ............. ,. _ .
:.L.... + ~"]T.2 . . . . . . ,. . . . . .
, ..... ~.,,1 & ...,,..--..~ . . . . . . ~ . . . . .
9
,. u- .........
~...-~....~ .~._t
~". ./._. ." '.'-~ + - .'.-
t . . ~ . ,..~ . ; -
,..~ ~.-.b--T-"E. "T . ' ..,,..] -f" ....... ~....+.,'.
9: . . . . . . . . ",' ~ ' " ; " ; " . . . . . . . r " " ~ ,.r. . . , . . . . :: . , , ?, 9 , . , ,
~-~"+"+,2,..";.' .'.'-. . . . ~. ~;-" " ' + .......

t';"..,-.,. "-'"l ~ --: :''-~"'"


~:~-:~-}~'-~" r.H :-:::.:.:
':","T, - .... L_ :Y~ ."~":".+ Fig. 4.3. Triaxial test with rigid platens: Distribution of plastic
~k " ~ . . . . . . ~.~,ff >-2 900E-02 volumetric strains at u = 8.8 cm ( C a m - C l a y Model)

4.5 Numerical assessment of the algorithms


In this subsection, a comparative evaluation of the time-step sensitivity exhibited by the two
proposed algorithms (fully implicit and semi-explicit) is presented in the context of axisymmetric
analyses of triaxial compression tests9 Due to axisymmetry, only one quarter of the specimen
(diameter d = 21.5 cm, height h - - 5 3 . 3 4 c m ) is discretized by 120 axisymmetric 4-node elements.
Loading is incrementally applied by prescribing a constant vertical top displacement Au.
Simulation (a) assumes rigid loading platens (fixed horizontal boundary conditions), thus
inducing a " b a r r e l i n g ' - - t y p e of deformation, as shown in Fig. 4.3. Simulation (b), on the other
hand, assumes perfectly frictionless platens leading to a h o m o g e n e o u s stress state. The material
parameter used in the analyses belong to normally consolidated Weald Clay, taken from W o o d
(1990) as: 2 = 0.088, ~c= 0.031, N = 2.097, M = 0.882, G = 3000 kPa, q0 - t = 138 kPa.
Figure 4.2 contains the pressure-displacement curves obtained from the (symmetrized) fully-
implicit and the semi-explicit algorithm for different time steps (Au = 0.0133h/2 and Au = 0.0033h/2
for example (a) and Au = 0.02h/2 and Au = 0.005h/2 for example (b)). As expected, considerable
influence of the time-step length is observed for the semi-explicit integration scheme as long as the
material is hardening. For moderately large time steps ( < 1~o), however, the accuracy of the semi-
explicit scheme is within the range of the accuracy of the fully-implicit algorithm.
268 Computational Mechanics 11 (1993)

Figure 4.3 shows the distribution of the volumetric plastic strains in the deformed configuration
of the cylindrical specimen, corresponding to a vertical displacement u = 8.8 cm (33~o of hi2).

5 General plasticity model for granular materials

As far as classical plasticity models in geomechanical applications are concerned, a main effort
since the development of the Cam-Clay concept has been focused on improving the predictive
capability of the models on the supercritical side of the critical state line. Various models in the
context of multisurface plasticity (Cap-Models) have been proposed, see, e.g. Chen and Mizuno
(1990) for a review of these formulations, and Simo et al. (1988) for the algorithmic treatment of
this class of models.
Recent work has been devoted to the use of single surface models. Representative for these
developments are the works by Lade and Kim (1988) and of Desai et al. (1986), respectively. The
functional format of these models, however, does not allow the straightforward application of the
standard return mapping algorithms. In particular, the yield function proposed by Lade and Kim
is not continuous while the yield function proposed by Desai et al. (1986) is not convex with respect
to the hardening parameter. Therefore, stability of the algorithm cannot be guaranteed (see Simo
1991) and was not observed in actual calculations.
Motivated by the preceding observations, we formulate a fairly general class of single surface
phenomenological plasticity models for granular materials, possessing the following two features:
(i) General applicability to different geologic materials, including sands, cohesive soils and
rocks.
(ii) Most of the standard phenomenological plasticity models can be obtained as special cases;
in particular, J2-flow theory, Drucker-Prager and modified Cam-Clay models.
The proposed class of phenomenological plasticity models use an associativeflow and hardening
rule which formulated in principal axes reads

~P=7 ~faa =7(t?f+t?f~ A=1,2,3, 0=-H76 ~ q- f f H - ~ ? q (5.1)


\ •S A ~P)' ' 0~"
We remark, however, that alternative (nonassociative) expressions may be used without changing
the general structure of our finite strain algorithm.

5.1 Formulation of the model


The elastic response is characterized by the quadratic model (3.35) formulated in terms of
principal logarithmic stretches. We propose an isotropic single-surface yield criterion f(SA, P, ~) =
f(I1, J2, J3, q(~)), with the following specific form:
f(I1,J2,Ja,q(oO):=[f1(J2,Ja).Jz +cllql"(T1+q)"]l12-c2rl-~/cl[qL"+m- f2(q)<O, (5.2)
where T1:= 3 t - 11 and the symbol t denotes the hydrostatic tensile strength of the material. The
physical significance of the functions and the calibration of the parameters used in Eq. (5.2) are
summarized below:
1. The function fl(J2, J3) determines the shape of the yield function in the deviatoric planes.
Following Desai et al. (1986), we chose

fl(J2,J3): = (1 -- ~ x / ~ J 3
)"~1/2,
where/7 is determined from different shear tests.
2. The function fz(q) provides an extension of the model to include cohesive hardening character-
istics. Here, to account for the possibility of matching the model to materials other than soil, metals
in particular, we simply set fz(q) = c3q. For sands one sets c3 = 0.
J. C. Simo and G. Meschke: A new class of algorithms for classical plasticity 269

25
20
15 60 ~ - M e r i d i a n ~_*'f---'-~=-0.5 /
10
5
0
-5
-t0
-15
-20
10 20 30 40 50 Fig. 5.1. Plots of the yield function in the ~ -
Ii-Plane (O = 60, O = 0 ~ (calibration for Munich
Ii (psi) sand)

3. The hardening function q(c0 is assumed to take a form proposed by Desai et al.:
q(~):= _ t/l(C~+ %),2 < 0, (5.4)
where c~o controls the initial elastic domain. The parameter t/1 and q2 have to be calibrated for
specific materials.
4. The exponents n and m determine the shape of the yield function in the meridional planes.
For sands we use n = 4 and m = - 1. Note that the elliptic shape of the C a m - C l a y models is
recovered by setting n = 2 and m - 0. Figure 5.1 shows plots of the yield function in the meridional
planes characterized by lode angles O = 0 ~ and O = 60 ~ obtained from the calibration of the
model to Munich Sand.
5. The ultimate state is determined by the parameters c2, fl and t, thus representing a
J3-dependent D r u c k e r - P r a g e r model (Fig. 5.1). Hence, assuming that c3 = 0 and for given fl, the
parameter c 2 and t may be determined from the known friction angle ~b and the cohesion c by
matching it in a particular meridional plane to the M o h r - C o u l o m b model.
6. For given values for c2, fl and t, the parameter cl is determined numerically from the condi-
tion that the variation dM(q)/dq of the slope M of the critical state line relative to the hardening
parameter q in a particular meridional plane does not exceed a given tolerance within an interval
of q of interest. The critical state line is characterized by ~ ( q ) = MI](q), where J~(q) and I] (q)
are determined from # ~ - f ( l ] , ~ 2 ) = 0 and f ( / ] , x / ~ 2 ) = 0.
op
7. The model includes as particular cases: J2-Theory: c 1 = 0, c 2 = 0, fl = 0, c 3 # 0, D r u c k e r -
Prager: c 1 = 0, c 3 = 0, fl = 0, t # 0, c 2 # 0 and C a m - C l a y : n = 2, m = 0, c 2 = 0, c 3 = 0, fi = 0, c 1 # 0.

Remark 5.I. The algorithmic treatment of the proposed plasticity models conforms to the standard
return mapping algorithms but with the simplification that the scheme is formulated in principal
stretches, as discussed in Sect. 3. In the solution of practical b o u n d a r y value problems, in particular
in the presence of large tensile trial stress states, it is essential to include a line search procedure
in the solution strategy. In all the simulations presented below the line search scheme incorporates
a quadratic fit procedure with selective three point pattern (Luenberger 1984, p. 206).

5.2 Numerical examples

We demonstrate the performance of the phenomenological model outlined above in the context
of reanalyses of typical drained triaxial tests from a particular type of sand (Munich Sand). Since
the emphasis of this paper is primarily on algorithmic aspects of the proposed approach to finite
strain plasticity, no attempt is made to achieve optimal fittings of the model to the various types
270 ComputaUonal Mechamcs 11 11993)

go = 1 3 p s i ~o = 1 3 p s i
I [ I I 25 I I [ k [
16
2O
14 /
12 [] "or'-,, /~ 15
rnl 13
TE - P a t h /
'.~ I0 I 1 lO experiment a / /
TC - Path - / /
-
~~
8
6
experiment a
I

~4(
t
5
0~
experiment
V /
//

4
T C : Path -- ~J S" []
2 ex'periment [] ? -5
E-3
0 I , [ , I~ ' r -1( { i

-0.02 -0.01 0 0.01 0 3 6 9 12 15 18


E-3 (0.1%)
a ~1,2 (%) b E1

ao = 6.5 p s i ao = 6.5 p s i
40 35
35
L I L I I I

30 CTC'-P ' h
experiment zx
'- ' '
/
'/
25
30 model bv Desai . . . . . /
20 (~ss~
25 ,e.-t 15
b 2O 10
j~
5
I0 exp eriinent a 0
5 -5
E-3
0 r , I , ~, , I , I -10 , I ~ I L I ,

-0.04 -0.02 0 0.02 0.04 0 0.01 0.02 0.03 0.04

c el,2 (%) d q (%)


Fig. 5.2a-d. Comparison of model predictions with experimental results from Munich sand: a z o -gl,2-curves from TE- and
TC-path, % = 13 psi; b el - e v -curves from TE- and TC-path, ao = 13 psi; e Z o - el,z-curves from CTC path, ao = 6.5 psi; d
el - ev-curves from CTC-path, ao = 6.5 psi (general plasticity model)

of soils. Consequently, the specific comparison with experimental results described below merely
represents a typical application of the model. Specific details on the actual determination of the
parameter in the model are omitted.
The test results as well as some of the model parameters are taken from Frantziskonis et al.
(1986): E = 9200 psi, v = 0.21, f l = 0.747 and c 2 = 0.35. c 1 is determined as 10 -4, stemming from
the requirement that dM/dq < T O L , with T O L taken as 10 -3 in the interval - 3 < q < - 0 . 0 1 .
The parameters concerning the hardening law are obtained by an iterative fitting procedure from
experimental results reported in Frantziskonis et al. as ~/1 = 0.82, rh = 2.06 and % - 0.016.
Figures 5.2a and c show comparisons between experimental results and model predictions for
tests following a triaxial extension (TE) and a triaxial compression (TC) path at % = 13psi
(Fig. 5.2a) and a conventional triaxial compression path (CTC) at a 0 = 6.5 psi (Fig. 5.2.c), respectively.
The respective a 1 - av-curves are contained in Figs. 5.2b and d. Figure 5.2a illustrates the relatively
strong dependency of the soil strength on the lode angle. Generally, good agreement between test
results and model predictions is observed. With regard to the dilatational behavior, however,
the model overestimates the a m o u n t of dilatation in all cases (Figs. 5.2.c, d), which is typical for
associative models. A comparison with the predictions from an associative version of a single
surface model developed by Desai et al. (Fig. 5.2.d) shows, however, the relative superior perfor-
J. C. Simo and G. Meschke: A new class of algorithms for classical plasticity 271

mance of the proposed associative model. Conceptually, it would be fairly straight forward to
incorporate modified (non-associative) flow rules as a device for reducing the amount of predicted
dilatancy.

6 Representative numerical examples


We demonstrate the performance of the proposed algorithm for isotropic finite strain plasticity
in the context of finite element analyses of three representative geomechanical problems. A main
goal of this section is to illustrate the general applicability of the methodology presented in the
preceding sections to a wide class of problems treated with an ample spectrum of finite element
techniques. Specifically,
(i) three fundamentally different simulations are considered: i.1) An analysis of shear band
formation in soil specimen, i.2) a rigid footing problem and i.3) the simulation of pile penetration
in soil.
(ii) two types of fairly general plasticity models at finite strains are employed: ill) The modified
Cam-Clay model and ii.2) the proposed general plasticity model, which is applied in one case in
a degenerated from (Drucker-Prager model).
(iii) two different finite element methods are used: iii.1) The standard Galerkin formulation
and iii.2) a recently proposed mixed finite element method based on a three field variational principle
and suitable for strain localization.
We remark that both quadrilaterals (bilinear 4-node QUAD) and triangles (CST and cross
triangles) are used in the simulations presented below which include plain strain as well as axi-
symmetric problems.

6.I Shear band formation in soil specimen


The goal here is the simulation of the (unsymmetric) shear bands formation typically observed in
conventional drained triaxial compression tests on overconsolidated soil specimen. We employ
the Cam-Clay model and the general plasticity models discussed in Sects. 4 and 5, extended to the
finite strain regime by the methodology described in Sects. 2 and 3. The soil specimen (b = 10.35 cm,
h = 26.67 cm) is discretized by 450 plain strain elements based on a recently proposed three field
mixed variational formulation presented in Simo and Armero (1991) and recently improved upon
in Simo, Armero and Taylor (1992). As is demonstrated below, this finite element method is specially
well-suited for localization problems. This approach generalizes to the nonlinear regime the
assumed strain mixed method proposed in Simo and Rifai (1990), which includes the classical
method of incompatible modes as a particular case. The specific structure of the enhanced strain
field advocated in these references is adopted here regardless of any localization criteria, since it
yields a locking-free element in the nearly incompressible regime with improved accuracy for
coarse meshes. Consequently, this approach completely by-passes the need for alternative inter-
polation schemes designed on the basis of costly local localization analyses that monitor (at
quadrature points) the eigenvalues of the acoustic tensor, as advocated in Ortiz et al. (1987), Leroy
and Ortiz (1990) and others.
The formation of an unsymmetric shear band is triggered by the introduction of a small geo-
metric imperfection at a level of 60% of the total height, where the width of the specimen is reduced
to 10.15cm (99% of b) and linearly extended to both ends of the specimen (Fig. 6.1). It is well
known that the numerical results for this type of problems involving softening materials are
considerably dependent on the size of the chosen mesh. Several devices, like the introduction of a
mesh-dependent characteristic length in the constitutive relations, may be used to avoid non-
objective numerical results. A particularly convenient implementation of this idea that yields
excellent numerical results is presented in Oliver (1989). Since the main focus of this subsection
lies on the treatment of large strain plasticity models and their performance in the context of
localization problems, no use of the above mentioned techniques has been made in the displacement-
controlled analyses presented below. (A monotonically increasing vertical displacement u is
imposed on the top of the specimen, as indicated in Fig. 6.1.)
272 C o m p u t a t i o n a l Mechanics 11 (1993)

b = 10.25 crn

JO
~[

10.7 c m

h= 26.7 cm
II
II
II
[I
II
II
(
I
I
b
f
IIII]
IIII 1
IIIIiiii
IIIIiiii
IIIIiiii
IIIIiiii
FIIIili
]

IIIIiii
b = 10.15 c m
I
I I
fllllll[
IIIIiiii
IIIli~il I t
VOL. PL. ST. II IIIIIIL~
I I I I I I I~-1 '"'r i l l;
< 2.3/.,z, E - 03 t IIIft,~T1
I lll.k..->r3~

I
IIII!'L~ N i

li
J 16,0 cm ......:.. .. ::

III Villi

i i!!llll II III
I I ~111111[

Ill! It tlI
II1!1
tlli)ii~
11111111
11112;i[
o 666000 o o o o o o
"= 1.l,06 E - 0 2
P "1 a b
6.1 b= 10.25 cm 6.2

Figs. 6.1 anti 6.2. 6.1 G e o m e t r y and finite element discretization of the soil specimen. 6.2a, b Deformed mesh and distribution
of volumetric plastic strains at u = 0.85 cm (Cam clay model); a deformed mesh, h distribution of volumetric plastic strains

800 I [ I I L i tl II I I I I_k.
II II I t I g-.~
700 II I I I~i-l
II II I I~N---1-1
tl it f.,-,X.~r~
600 II II kA---t-Y1
Z II II I I~'~-I
500 II II
If I 1 ~
400
300 h imperfection
o
200 u n i f o r m solution . . . . . .
I00

H
I
0
0 0.2 0.4 0.6 0.8 1.0 1.2

6.8 vertical top displacement (cm) ilt I III il

400
l tll II
r

6.4a
IllLI
I
Ill
i
II
II
II b 1111 LII' I
I
'll
I
tit
It~
350
300 EQ PL ST < 1 000E-01
Z 250
wkh imperfection
200
lution . . . .
150
I00

N
50
I I , I , I [ [
0
0 0.2 0.4 0.6 0.8 1.0 1.2

6.5 vertical top displacement (cm) > 5 O00E-01

Figs. 6.3-6.5o 6.3 Load-displacement curves from the analysis of a triaxial test on overconsolidated clay ( C a m - c a y model).
6.4 deformed mesh and distribution of equivalent plastic strains at u = 0.85 cm (General plasticity model); a deformed mesh.
b distribution of equivalent plastic strains. 6.5 Load-displacement curves from analyses of a triaxial test on a dense sand
specimen (general plasticity model)
J, C. S i m o a n d G. Meschke: A n e w class of a l g o r i t h m s for classical plasticity 273

The first example is concerned with the analysis of a heavity overconsolidated clay specimen
by means of the fully-implicit version of the return algorithm for the C a m - C l a y model described
in Sect. 4. The soil is assumed to be Weald Clay, for which the C a m - C l a y parameters are taken
from W o o d (1990) as: 2 = 0.088, ~ = 0.031, N = 2.097, M = 0.882, G = 3000kPa. The isotropic
precompression is qo = 138 kPa; the confining pressure is considered by setting t = 6.5 kPa. In case
of severe softening, a lower b o u n d for q, qm~,, is provided by setting qm~. = 20 kPa. Figure 6.2 shows
the deformed configuration (a) and the distribution of the volumetric plastic strains (b) at a top
displacement u = 0.85 cm ( - 3.2~o of h). A p r o n o u n c e d shear band with an inclination of 0 ~ 50 ~
relative to the horizontal plane fully develops. Observe the sharp localization of plastification
within a narrow band containing not more than 2 elements.
Figure 6.3 shows the load-displacement relationship as obtained from analyses of a specimen
with and without geometric imperfection. As soon as the formation of a continuous shear band
starts, the load baring capacity drops rapidly from the level of the uniform solution. It is followed
by a stabilized behavior, characterized by a soft descent of the load-displacement curve. This
behavior is in excellent agreement with experimental observations from sand specimen (Vardoulakis
et al. 1978) and from theoretical considerations (Leroy and Ortiz 1989).
The second example is concerned with the analysis of a dense sand specimen by means of the
general (associative) plasticity model presented in Sect. 5, modified to incorporate a separate
softening mechanism. The yield function now takes the following form:
f(It, J2, Ja, q~(c~), q~(c~)):= [/~(J2, J~)'q] l'J2 + cl Iq~im(/~ + qh)n] ~/2

+ c~[~ - x/c~ Iq,,I"+" - f2(qh) <=O, (6.1)


where the softening law is assumed as

q,(~,) = 2 - tanh(a~ ~ - a~ - 1 =< q, < 0. (6.2)


2 - tanh( - a0,,)
In the numerical example described subsequently we set t/3 = 1 and ~o., = 1. The model parameters
used are: E = 9600kPa, v = 0.21, c~ = 10 -4, c 2 = 0.26, fl = 0.747, t/~ = 3, 1/2 = 1, ~o = 0.016. The
confining pressure is assumed as t = 6.5 kPa.
Since the inclination 0 of the shear band is expected to be greater than the one observed from
the C a m - C l a y model, the above mentioned geometrical imperfection is introduced on the top of
the specimen. Figure 6.4 shows the deformed mesh (Fig. 6.4a) and the distribution of the equivalent
plastic strains gr:= ~ ~ (Fig. 6.4b), respectively, corresponding to a top displacement
Path

u = 0.85 cm. Observe that the magnitude of the equivalent plastic strains exceeds 50~. The inclina-
tion 0 of the shear band is 0 -~ 61.5 ~

EQ PL ST

< 6 O00E-02

[- I..l.4~I ~I ~I ~I ~I ~I ~I ~I ~ ~ ~I ~I ~I ~I ~ ~I ~ ~I ~I ~I ~i ~ ~

~ ~.~.~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~.~-,_~, ~I-I ~ ~ ~

I I [ ] .,]. ,J ~J ~1 ~J ~ ~ ~,l',,,J,,,,l.,,,i_l_l_l '1=1-1

f.t.,t ,~ ,~ ,~ ,! ,4 ,~ ~ , ~ . J a , I J :1 :i :i :i-i d 4 "1

N
i: ~'~.~:~~:5-~:::'. ~-:-:'.K::~E~-::?-L5--'-:....
' . . . . . . .
I ] ,~.~1,1 ,~ ~ ,~ ,J,.!,.t.J.
I " 1 | L9 i i J i i i i , . . . .
a j :1 :~,a n -i -i -,l-i ~i-
i ..... i
. . . . . . . . . . . . . ~. : 1 . ~ . . ~ . ' ~ ~ . ~ , ~ _ ~ . . . . . . . . . . ..j - . ,
9 : ..... ! ..: ! __ ".-'A~,~,,,. . . . . . i,...,.~..:. : . . . I
, ~ , ! i 9 . . .:' . ' ~ ._..]....,
. . . . . . . . . . . ; 9 9 -,4.-. . 9 "T-" - - "
t...a_a.J . . . [ .. , . L. . . . . . . . . . . . . . . . .

> 4 O00E-01

Fig. 6.6. D e f o r m e d m e s h , n o d a l velocities a n d d i s t r i b u t i o n of e q u i v a l e n t plastic s t r a t u s at u = 2.5 inch


274 Computauonal Mechanics 11 (1993)

200
180
o-~ 160
140
120
100
80
60 / ~:;a~tth I Ss:lUiii~n5 . ~ .2 .-~.-(_
40
20
0 , I i I ~ I , I , I ,

0.5 1.0 1.5 2.0 2.5 3,0


6.7 displacement at base of rigid footing (in)

600 I i

500
1
400
Z 4
300 1
o 200 J
1
100
Figs. 6.7-6.9. 6.7 Load-displacement curve from the
0 I I rigid footing problem (associative Drucker Prager model).
5 I0 15 6.8 Finite element discretization of the soil stratum and
the pile. 6.9 Load-displacement-curve from the simu-
@.9 pile penetration (m) lation of pile penetration (Cam-Clay model)

In contrast to the Cam-Clay model, which performs as an elastic-plastic softening material


in case of high overconsolidation, the general plasticity model exhibits pronounced hardening
behavior prior to softening (Fig. 6.5). Consequently, a less sharp localization zone is observed
(Fig. 6.4a).
Figure (6.5) contains the load-displacement curves obtained from the analysis of the specimen
with and without imperfection. Compared to the result from the Cam-Clay model (Fig. 6.3), the
more pronounced softening behavior exhibited by the general plasticity model yields a relatively
steeper descent of the load-displacement curve after formation of the shear band to an asymptotic
lower level of ~ 30~ of the peak level.

6.2 Rigid footin9

This second example, the rigid footing on a stratum of soil, is a classical problem of soil mechanics
extensively discussed and analysed in the literature; see e.g. Zienkiewicz et al. (1975) and the more
recent work of Chen and Mizuno (1990), among many others. The geometry and the material data
are taken from this latter reference. An (associative) Drucker-Prager constitutive model is used
in the simulation, obtained as a special case of the generalized plasticity model presented in Sect. 5
for the following values of the material parameters: E = 30000psi, v = 0.3, cl---e = 0.112 and
t = k/3o~ = 9.22/(3"0.112)= 27.44 psi. The width of the footing is 61.71 inch and the depth of the
stratum is 144 inch. The horizontal extent from the footing center is 349.7 inch. The same assumed
strain mixed finite element methodology used in the preceding example is also employed here. Due
to obvious symmetry considerations, only one half of the stratum is discretized using 476 bilinear
4-node plane strain elements. The weight of the soil is neglected.
9

!i~:::"

m"
P.
i
Oo
0

iiii:i: ==~/ii:
9.! ..~ ~!~ i~
ii~ ) ~ Ni:::i::
9:::.::iiiii~
I !~: :

TENS. VOL. PL. ST. COMP, VOL. PL. S T RADIAL STRESSES AXIAL STRESSES

< 1.000E-03 < -3.000E-02 < -2.200E+02 Figs. 6.10-6.12. 6.10 Distribution of
< -1.200E§
plastic volumetric tensile strains (left
hand side) and compressive strains
(right hand side) in the deformed con-
figuration at u = 2.0m, 8.0m and
13.6 m. 6.11 Distribution of compres-
sive radial stresses in the deformed
m configuration at u = 2.0 m, 8,0 m and
13.6 m. 6.12 Distribution of compres-
sive axial stresses in the deformed
9 3.000E-01 >-3.000E-03 9 -1.000E+00 9247

6.10 6.11
II 6.12
configuration at u = 2.0 m, 8.0 m and
13.6m L/I
276 Computational Mechanics 11 (1993)

Figure 6.6 shows the computed distribution of the equivalent plastic strains gP and the velocity
field in the deformed configuration of the stratum at a settlement of 2.5 inches. Observe the massive
concentration of plastic strains in the vicinity of the corner of the footing (gP > 40~o!) and, in
particular, the narrow slip-line shaped zone of plastification, emanating from this corner. Figure 6.7
shows the load-displacement curve, with the load being related to the area under the footing. The
numerically predicted collapse load, p, = 156.3 psi is higher than the classical Prandtl solution
(143 psi) and lower than the Terzaghi solution (175 psi). (The values for p, obtained by Zienkiewicz
et al. and by Chen and Mizuno are 151 psi and 171 psi, respectively).

6.3 Pile penetration in soil

As a final example we consider the quasi-static, axisymmetric simulation of a pile penetration in


soil, under the assumption of drained conditions. This type of problems clearly involves large
deformations and finite strains. The soil is assumed to be normally consolidated Weald Clay,
represented by the modified Cam-Clay model.
The same model parameters employed in the preceding examples are also used here; i.e.:
2 = 0.088, x = 0.031, N = 2.097, M = 0.882, G = 3000kPa. The preconsolidation pressure is
t = q0 = 87 kPa, along with a threshold value qmin 20 kPa such that q > qmin. The finite element
=

mesh is as depicted in Fig. 6.8. The cylindrical stratum of soil (radius r = 22.4 m, height h -- 20.7 m)
is discretized by 960 axisymmetric crossed-triangles. The pile (radius r = 0.88 m, total length
l = 13.8 m, length of the vertical shaft l = 12.05 m) is assumed to be rigid. The horizontal displace-
ment boundary conditions at the remote boundary of the stratum are assumed fixed. The "centerline',
which has to be given a hypothetical small thickness, as well as the pile are treated as rigid obstacles
within the deformable soil. The contact constraints are enforced using an augmented lagrangian
technique with frictionless contact assumed at the interface between the soil and the pile. We
remark, however, that the consideration of friction does not pose any additional difficulties, and
we refer to Simo and Laursen (1991) for a detailed description concerning the augmented lagrangian
treatment of contact problems involving friction. The normal and tangential penalties have been
chosen as ~N= 109 and ~r = 109, respectively. 200 time steps were used per 1 meter pile advancement.
Figure 6.9 shows the load-displacement curve obtained from the penetration analysis. The
zig-zag form of the curve is induced by the relatively coarse mesh used in the simulation and
caused by the periodical release of nodal forces whenever a node loses contact with the inclined
surface (cone) of the pile. The numerically predicted maximal load at the end of the pile installation
is 516 kN, the corresponding cone pressure is 212.1 kPa. Figures 6.10a-c show the distribution of
volumetric plastic strains at 3 different stages of the penetration process: u = 2 m, u = 8 m and
u = 13.6 m. The left hand side contains tensile, the right hand side compressive plastic volumetric
strains. In the immediate vicinity of the pile, very large tensile plastic strains, connected with
considerable softening of the material adjacent to the pile, are induced during the process of pile
stallation. Qualitatively, this behavior is in good agreement with field observations. Quantitatively,
however, the physical relevance of the numerical predictions in this narrow zone may be question-
able, since the predictive capability of the Cam-Clay model on the supercritical side is rather
limited (see, for instance Wood 1990). Figure 6.11 contains contour-plots of the compressive radial,
Fig. 6.12 of the compressive axial stresses in the three above-mentioned stages of the penetration.
A concentration of compressive axial stresses is observed beneath the cone, high compressive radial
stresses are induced within ~ 3 times of the pile radius lateral to the inclined surface of the cone.
The white shaded zone adjacent to the shaft of the pile again illustrates the severe softening due
to remolding of the soil.

7 Concluding remarks

The methodology described in the preceding sections allows the generalization of any (isotropic)
model of infinitesimal plasticity to the regime of finite strains. As already pointed out, this extension
J. C. Simo and G. Meschke: A new class of algorithms for classical plasticity 277

preserves the form of the yield condition, now formulated in therms of true stresses, and retains
the hyperelastic character of the stress-strain relations. The exponential approximation described
in Sect. 3 possesses the remarkable property of preserving both the structure and algorithmic
moduli of the now standard return mapping algorithms, now reformulated in principal Eulerian
axes. An element of novelty in the application of the preceding methodology to the modified
Cam-Clay model concerns the formulation of semi-explicit integration schemes. This alternative
approach retains the unconditional stability property associated with an implicit treatment of the
flow rule, while achieving symmetry of the algorithmic moduli via an explicit integration of the
hardening law.
The general class of single surface phenomenological plasticity models for granular materials
presented in Sect. 5 is applicable to a number of different geologic materials, including sands,
cohesive soils and rocks. This unified treatment encompasses Most of the standard phenomeno-
logical plasticity models as special cases. In particular, J2-flow theory, Drucker-Prager and
modified Cam-Clay. The restriction to isotropy is the direct result of formulating the yield criterion
in terms of true stresses. We remark that nonlinear isotropic and kinematic hardening rules fit
within this framework.
The nonlinear finite element simulations presented in Sect. 6 demonstrate the general appli-
cability of the proposed algorithmic treatment of classical plasticity by means of three rather
different and representative geomechanical problems. An analysis of shear band formation in soil
specimen, a rigid footing problem and the simulation of pile penetration in soil.

Acknowledgements

Support for this research was partially provided by DNR via a contract with the Port Huaneme Naval Civil Engineering
Laboratory. G. Meschke was supported by the MAX KADE Foundation under Grant No. 2-DJA-616 with Stanford
University. This support and the interest of Dr. J. Malvar and Dr. T. Shugar are gratefully acknowledged.

References

Borja, R. I. (1991): Cam-Clay Plasticity. Part II: Implicit integration of constitutive equation based on a nonlinear elastic
stress predictor. Comput. Meth. Appl. Mech. Eng. 20, 225 240
Borja, R. I.; Lee, S. R. (1990): Cam Clay Plasticity. Part I: Implicit integration ofelasto plastic constitutive relations. Comput.
Meth. Appl. Mech. Eng. 78, 49 72
Chen, W. F..; Mizuno, E. (1990): Nonlinear analysis in soil mechanics-theory and implementation. Devel. Geotech. Eng.
53
Desai, C. S.; Somasundaram, S.; Frantziskonis, G. (1986): A hierarchical approach for constitutive modeling of geologic
materials. Int. J. Num. Meth. Eng. 10, 225-257
Frantziskonis, G.; Desai, C. S.; Somasundaram, S. (1986): Constitutive model for nonassociative behavior. J. Eng. Mech. 112,
9, 932-946
Hill, R. (1950): The mathematical theory of plasticity. Dover editions, U.K.
Lade, P. V.; Kim, M. K. (1988): Single hardening consistitutive model for frictional materials II. Yield criterion and plastic
work contours. Comput. Geotech. 6, 13-29
Lerov, Y.; Ortiz, M. (1989): Finite element analysis of strain localization in frictional materials. Int. J. Num. Anal. Meth.
Geomech. 13, 53-74
Luenberger, D. G. (1984): Linear and nonlin6ar programming. Addison Wesley, Menlo Park
Maenchen, G.; Sacks, S. (1964): The tensor code, in methods of computational physics, 3, 181-210. B. Alder, S. Fernback
and M. Rotenberg Editors. New York: Academic Press
Moreau, J. J. (1977): Evolution problem associated with a moving convex set in a hilbert space. J. Diff. Eq. 26, 347
Oliver, J. (1989): A consistent characteristic length for smeared cracking models. Int. J. Num. Anal. Meth. Eng. 28, 461-474
Ortiz, M.; Leroy, Y.; Needleman, A. (1987): A finite element method for localized failure analysis. Comput. Meth. Appl. Mech.
Eng. 61,189-214
Ortiz, M.; Simo, J. C. (1986): An analysis of a new class of integration algorithms for elastoplastic constitutive relations.
Int. J. Num. Meth. Eng. 23, 353-366
Roscoe, K. H.; Burland, J. B. (1968): On the generalized stress strain behavior of'wet' clay. In: Heyman, J.; Leckie, F. A., (eds.)
Engineering plasticity. Cambridge University Press, 535-609
Simo, J. C.; Taylor, R. L. (1985): Consistent tangent operators for rateindependent elastoplasticity. Comput. Meth. Appl. Mech.
Eng. 48, 101-118
Simo, J. C.; Ju, J.; Pister, K. S.; Taylor, R. U (1988): Assessment of cap model: Consistent return algorithms and rate-dependent
extension. J. Eng. Mech. 114, 2, 191-218
278 Computational Mechanics 11 (1993)

Simo, J. C. (1988): A framework for finite strain elastoplasticity based on the multiplicative decomposition and hyperelastic
relations. Part I: Formulation and Part II: Computational aspects. Comput. Meth. Appl. Mech. Eng. 66, No. 2, 199-219
and 66, No. 3, 1-31
Simo, J. C.; Kennedy, J. G.; Govindjee, S. (1988): Non-smoth multisurface plasticity and viscoplasticity. Loading/unloading
conditions and numerical algorithms. Int. J. Num. Meth. Eng. 26, 2161-2185
Simo, J. C.; Rifai, M. S. (1990): A class of mixed assumed strain methods and the method of incompatible modes. Int. J. Num.
Meth. Eng. 29, 1595-1638
Simo, J. C.; Taylor, R. L. (1991): Finite elasticity in principal stretches; formulation and finite element implementation. Comput.
Meth. Appl. Mech. Eng. 85, 273-310
Simo, J. C.; Miehe, C. (1991): Coupled associative thermoplasticity at finite strains. Formulation, numerical analysis and
implementation. Comput. Meth. Appl. Mech. Eng. 98, 41-104
Simo, J. C. (1991a): Nonlinear stability of the time discrete variational problem in nonlinear heat conduction and elastoplasticity.
Comput. Meth. Appl. Mech. Eng. 88, 111-131
Simo, J. C. (1991b): Algorithms for multiplicative plasticity which preserve the structure of the classical return mappings of
the infinitesimal theory, SUDAM report No. 91-4. In press in comput. Meth. Appl. Mech. Eng.
Simo, J. C.; Hughes, T. J. R. (1991): Plasticity viscoplasticity and viscoelasticity: formulation and numerical analysis aspects,
Springer-Verlag (in press)
Simo, J. C.; Armero, F. (1991): Geometrically nonlinear enhanced strain method and the method of incompatible modes.
Int. J. Num. Meth. Eng. 33, 1413-1449
Simo, J. C.; Laursen, T. A. (1991): An augmented lagrangian treatment of contact problems involving friction. Computers
and Structures, 42, No. 1, 97-116
Vardoulakis, I.; Goldscheider, M.; Gudehus, G. (1978): Formation of shear bands in sand bodies as a bifurcation problem.
Int. J. Num. Anal. Meth. Geomech. 2, 99 128
Wood, D. M. (1990): Soil behavior and critical state soil mechanics. Cambridge University Press
Wilkins, M. L. (1964): Calculation of elastic-plastic flow in Meth. Comput. Phys. 3, 211 272. B. Alder, S. Fernback and M.
Rotenberg editors. New York: Academic Press
Zienkiewicz, O. C.; Humphenson, C.; Lewis, R. W. (1975): Associated and nonassociated viscoplasticity and plasticity in soil
mechanics. G6otechnique 25, 671-689

Communicated by S. N. Atluri, February 18, 1992

View publication stats

You might also like