Vous êtes sur la page 1sur 14

International Journal of Plasticity,

Vol. 9, pp. 907-920, 1993

0749-6419/93 $6.00 + .00


Copyright 1993PergamonPressLtd.

Printed in the USA

A DEFORMATION THEORY OF PLASTICITY


BASED ON MINIMUM WORK PATHS
KWANSOOCn-ONG and OWEN RICHMOND
Aluminum Company of America
(Communicated by Kenneth Neale, Universit6 de Sherbrooke)

A b s t r a c t - A deformation theory o f plasticity is proposed wherein the deformation paths for

material elements are assumed and the plastic work becomes dependent on displacements. Among
the infinite possible ways to assume deformation paths, one has been chosen that has several
advantages when materials harden isotropically. Earlier, this path was shown to require the minimum work path to achieve a desired strain. Here, a mathematical description o f a constitutive law of deformation plasticity is developed based upon this path for rigid-plastic and for
elastoplastic materials. The proposed deformation theory provides a convenient theoretical basis
for FEM applications involving analysis, and especially design, o f forming processes.

I. INTRODUCTION

In the early development of the theory of plasticity of metals, two types of theories were
proposed: flow and deformation. Flow theories eventually became dominant because
they reflected the observed dependence of metal behavior on deformation path. Deformation theory, which did not reflect this dependence, was considered accurate only in
limited cases where materials deform in special ways; therefore, use of the theory dwindled. In 1959, Budiansky introduced a new interpretation of the deformation theory for
yield surfaces with edges (vertexes). After this work, major applications of the deformation theory developed in the area of instability analysis: the works of Hutchinson
[1974], St6ren and Rice [1975], and Neale [1989] among others.
Recently, deformation theory has been used in formulating finite element modeling
(FEM) codes for a variety of engineering problems. Applications of deformation theory in FEM are so versatile that they include designing as well as analyzing forming processes. The range of application includes rigid-plastic and elastoplastic materials. In
analysis codes the deformation theory is applied incrementally, and geometry and material properties are updated at discrete steps without the use of numerical integration
(Wang [1982]; Braudel et al. [1986]; Yang & Kim [1986]; Germain et al. [1989]). This
differs from the conventional flow formulations that use numerical integration to update
geometry and material properties. In design codes, the deformation theory is applied
in a single step. These codes provide information about formability of products and optimized process parameters in the preliminary design stages (Chung & Richmond [1992b,
1993]). There have been some efforts to apply the deformation theory in a single step
to analyze forming processes. However, this approach might not be appropriate for general forming because the deformation theory with a single step does not properly reflect
the dependence of metal behavior on deformation path (Levy et al. [1978]; Chung & Lee
[1984]; Sklad [1986]).
To provide a theoretical basis for application of deformation theory to FEM, a special deformation theory of plasticity is developed in this article for both rigid-plastic and
907

908

K. Cmm~ and O. RICHMOND

elastoplastic materials. The theory covers arbitrary anisotropic materials that harden isotropically. In this deformation theory, the deformation paths of material elements are
specified, so that the plastic work becomes dependent on displacements. A mathematical description of a constitutive law for the deformation theory is then derived from
flow theory using the assumed deformation path. Among the infinite number of ways
to assume a deformation path, one has been assumed that has several advantageous features. The specific assumed path is the minimum plastic work path in homogeneous
deformation. The constitutive laws derived from the minimum work path are different
from those of the conventional deformation theory. Earlier, Ponter and Martin [1972]
presented theoretical work that connects the deformation theory and the minimum work
path, but without explicit details of the minimum work path or derivations of the constitutive laws.
Derivations of the deformation theory begin with a brief review of minimum work
paths and their kinematics for rigid-plastic materials in Section II. In Section III, constitutive laws are derived for smooth yield conditions based on both the yield surface
and its associated dissipation surface in rigid-plastic materials. In Section III, a one-step
minimum work path is assumed in order to obtain results that are useful for design purpose FEM codes. In Section IV, the incremental deformation theory is obtained by discretizing a deformation process into multistep minimum work paths. These results are
useful for analysis purpose codes. In Section V, it is shown that the results derived for
smooth yield conditions are also valid for yield conditions having sharp edges. The
deformation theory is compared with hyperelasticity and extended to elastoplasticity in
Sections VI and VII, respectively.
II. KINEMATICS OF M I N I M U M WORK PATHS

Requirements for achieving minimum work paths in homogeneous deformations are


well documented in the works by HILL [1957], NADAI [1963], D ~
[1978], and HILL
[1986]. In these works, it was found that the minimum work path is achieved if materials deform in such a way as to satisfy two conditions: first, the set of three principal
axes of stretching is fixed with respect to the material; second, the logarithms of the
stretches remain in a fixed ratio. HILL [1986] proved that these two conditions are valid
for any convex yield condition with isotropic hardening. Recently, CnuN6 and RICHMOND [1992a] showed that both of the conditions are required only when the yield surface is smooth enough to have a unique normal direction (in such a case, the minimum
work path is uniquely defined). When the yield surface has sharp edges, no unique normal direction is defined at these corresponding stress states, so that the two conditions
are only partially required (in such a case, multiple minimum work paths are possible).
For the Tresca yield condition, for example, only the maximum stretch axis is required
to be fixed with respect to the material, and no other major conditions are necessary.
For isotropic materials, the fixed material lines may be chosen arbitrarily; for anisotropic
materials, they are more restricted. Because of the differences in the minimum work
path, the deformation theory is discussed separately for smooth yield surfaces and for
pointed yield surfaces.
The minimum plastic work paths can be quantified in equations. The deformation gradient tensor F describes a deformation between a final and an initial shape. F is defined,
by polar decomposition, as
F(t) = R(t).U(t),

(1)

Deformationtheory of plasticity

909

where U is the right stretch tensor, which is symmetric, R is the orthogonal rotation tensor, and t is time. The three principal directions of U, because of its Lagrangian nature,
refer to three material lines that are perpendicular to each other in the original configuration. Eqn (1) represents a physical deformation process that is composed of an initial stretch, U, of principal material lines followed by a rigid body rotation, R. It is a
description of the total amount of deformation and the total rigid body rotation required
by two prescribed geometries, but the history of deformation is only implicit.
When decomposed into principal values and directions, U and the logarithmic strain
tensor eL are defined as
U(t) = Q(t).,I (t).Qt(t),

eL(t) = Q ( t ) . / j ( t ) - Q t ( t ) ,

(2)

where the superscript t stands for transpose. In eqn (2), ,I and ~ are diagonal tensors
consisting of principal values (,~i and ~i ~--- In Ai), and Q is an orthogonal rotation tensor consisting of the principal directions of U. Note that eL is also a Lagrangian quantity. When the same principal material lines are kept constant during deformation, the
principal material lines of l~l and I~/~coincide with those of U obtained for a final shape
at a time tf; i.e.
U(t) = Q(tf).A(t).Qt(ty),

U(t) = Q ( t f ) . A ( t ) . Q t ( t y )

0 < t < tf

(3)

~L(t) = Q ( t f ) . ~ ( t ) . Q t ( t f ) ,

EL(t) = Q ( t f ) . / j ( t ) . Q t ( t f )

0 < t < tf.

(4)

and

Let us consider a strain rate g, which is defined as


g = Rt.D.R

= (l)-U-l)s,

(5)

where D is the rate of deformation tensor, and the subscript s stands for the symmetric part of a tensor. The quantity g is considered to be the value D measured with respect
to the coordinate system that rotates by R ( t ) , so that the values are invariant with
respect to R. (For this reason, ~ is called the "rotationless" strain rate by STOREN and
RICE [1975].) When the same principal material lines are kept constant during deformation, the principal material lines of l~l and U coincide as shown in eqn (3). Consequently,
in such cases, the principal directions of g are stationary and are aligned with the principal material lines of U(tf), regardless of arbitrary rotation R (t), whereas the principal directions of D might vary continuously; i.e.
~(t) = gL(t) = Q(t I) . ~ ( t ) . Q t ( t f ) .

(6)

In fact, when the same principal materials lines are kept constant, g also becomes a
Lagrangian quantity whose principal directions refer to the fixed principal material lines
and whose principal values mean the logarithmic (or true) strain rates of the principal
material lines.
The minimum work path for a smooth yield function is achieved if deformation follows the description
~(t) = ~L(t) = &(t)"eL(tf),

e(t) = eL(t) = or(t)"eL(tf),

(7)

910

K. CHt~NCand O. RICHMOND

where a (t) is a function of time t, satisfying fos& d t = 1, and & is always positive. In
eqn (7), the principal material lines are fixed, and the ratios of the principal true strain
rates are fixed during deformation. Note that the condition/; = & ( t ) . g is not sufficient
to describe the minimum work path, because Q can be time-dependent. Also, the condition in eqn (6) alone is not sufficient, because the fixed ratios of the principal true
strain rates are not imposed. The minimum work path described in eqn (7) allows arbitrary rotations, R ( t ) , of the principal material lines in eqn (1). In rigid-plasticity, materials are considered incompressible; however, the fixed ratios of principal values of e ( t )
are permissible even under the incompressibility condition. When a yield surface is
smooth so that normal directions are uniquely defined on the surface, ~;(t) is fully prescribed for a given o~(t). When a yield surface has a pointed vertex at a desired stress
state, g ( t ) is not prescribed completely.
In eqn (7), e was obtained by integration of ~;, following a prescribed deformation
path; g is inherently path-dependent. In contrast to this, the "logarithmic" strain tensor gL is defined for any two geometries, and the value is independent of deformation
history: ~L is path-independent. However, ~ becomes equivalent to ~L if the principal
material lines are fixed during deformation. The difference between g and gL is similar to the difference between the deformation theory derived in this article and the original deformation theory (and hyperelasticity). For the new deformation theory, a path
is prescribed as it is for ~, whereas the original deformation theory (and hyperelasticity) is independent of deformation history, as is IzL. Besides, there is another major difference between these deformation theories as is shown in the following section.
IlL RIGID-PLASTIC DEFORMATION THEORY FOR SMOOTH YIELD SURFACES

The effective yield stress, 6, and the associated effective strain rate, ~, are defined in
the Cauchy stre_.ss space, a, and in the rate of plastic deformation space, D, respectively.
Because 6 and 6 are materially embedded properties, they are invariant with respect to
any rotation by "objectivity." Therefore, they can be written in terms of ~ and its conjugate stress, 7 (defined as r = R t . o . R ) , simply by replacing D and cr with ~ and r,
respectively; i.e.

6(0) = ~(7),

~(O) = ~(~),

(8)

where ~ and ~ are the effective stress and the associated effective strain rate defined in
the r and ~ spaces, respectively. Note that eqn (8) expresses "objectivity," but not "isotropy." The definitions of g and ~ are considered first-order homogeneous functions here
for isotropically hardening materials.
By applying the m i n i m u m work path in eqn (7), the minimum effective strain is
obtained as
~(tf) =

~(~) d t = ~(~L)

& dt = ~(GL(tf))

(9)

if the yield surface is smooth I and materials harden isotropically. Isotropic hardening
throughout this article implies that the definitions of g and ~ do not change during deforINote that eqn (9) is also valid for pointed yield surfaces even though the minimum work path expressed
in eqn (7) is not rigorously required for such cases. This is because the set of multiple minimum work paths
for pointed yield surfaces contains the minimum work path expressed in eqn (7).

Deformation theory of plasticity

911

mations. Eqn (9) shows that the total effective strain g in the minimum work path is
obtained from ~ in eqn (8) simply by replacing g with 8L (tf). Eqn (9) is equivalent to
= &(t).~(ty),

~;(t) = a ( t ) . g ( t f ) .

(10)

The deformation theory is derived by applying the relations in eqns (7) and (10) to the
flow theory.
The constitutive equation for the flow theory of incompressible rigid-plastic materials is obtained from the normality rule, which is

D = ,~ aa(,,)
ao

_ ~, aa'(,,')
0o'

'

(11)

where o ' is the deviatoric part o f o. By "objectivity," eqn (1 l) is equivalent to


_. 0~
_. 0~'
t = e ~ -- e Or'

(12)

where r ' is the deviatoric part of r. Eqns (11) and (12) satisfy the relation o f the rate
of plastic work density per unit volume, fi,:
fi, = o . D = o ' . D = r.[; = r " g = 66 = ~ .

(13)

An explicit representation of the constitutive relation for deformation plasticity is now


obtained from eqn (12) by applying eqns (7) and (10); i.e.

0~

g = ~ ; ~ ~ = ~ : -a'
0r' '

R.E.R, g:

aa _ aa-0o
0o' "

(14)

As a conjugate quantity for #, ~ is derived from eqn (13) for a specified effective stress:
6 forms a yield surface in stress space, and 6 forms an associated surface in strain rate
space, which is called the "dissipation surface" by ZIEGLER [1977]. Consequently, the
normality rule can be expressed for a dissipation surface in the rate of plastic deformation space as

,,' = a' ~a0D'

,., = ~, o~
0~"

(15)

For the minimum work path, eqn (15) becomes

z' = ~,0~:
0~'

o' = R ~,0~: .R t
0g

(16)

after eqns (7) and (10) have been applied. Eqns (16) show alternative forms of the constitutive laws of the deformation theory defined for the dissipation surface. In eqns (14)
and (16), the direction of r ' is fixed if materials harden isotropically, while a ' is in an
arbitrary direction depending on R. (Because ~ is independent of hydrostatic pressure
for incompressible materials, r is not completely determined). The secant slope H is
defined as W g ( = 0 / ~ ) .

912

K. CHUNG and O. RICHMOND

Eqns (14) and (16) represent constitutive relations for deformation plasticity when the
minimum work path is assumed. Not only do these equations satisfy the objectivity rule,
but also R is arbitrary, which allows arbitrary rigid-body rotations during deformation.
The equations describe a deformation theory for large deformations and arbitrary anisotropic materials that harden isotropically. Comparison of the derived equations and the
original flow rules shows that the constitutive relations in deformation plasticity are
obtained from the normality rule simply by replacing a', D, and ~ with r', e and ~,
respectively. When the Mises yield condition is assumed, eqns (14) and (16) become
a' =

(17)

(R.e.R').

Even though the original deformation theory proposed by HENCKY [1924] for Mises
materials and small deformations is path-independent, the equation is known to comply with the flow theory when a ' is proportional under the isotropic hardening condition. Because r ' is proportional in the theory based on the minimum work path,
although a ' is arbitrary, there is a difference between the equations of the two deformation theories. When isotropic materials harden isotropically, for example, the original deformation theory complies with the flow theory for simple shear; however, the
deformation theory discussed here does not comply because simple shear is not the minimum work path. The equations of the two theories become identical when R = I (identity tensor), as is the case for pure shear.
When the new deformation theory is applied as a one-step process for a whole deformation, it is called the total deformation theory. FEM formulation based on the total
deformation theory is a useful tool for process design. In a design code, materials are
assumed to deform in a prescribed manner that is considered ideal, and the code provides information about the formability of products. The minimum work path is especially useful because formability of many metals might be assumed optimum in such a
path. For special forms of the yield function, a code based on such paths can also provide information about optimum process parameters. Detailed discussions on this subject can be found elsewhere ( C n u ~ o 8~ RICHMOND [1992,1993]).

IV. INCREMENTAL DEFORMATION THEORY FOR MULTIPLE STEPS

The deformation theory derived in Section III can be applied to a multi-step process
where a whole process is discretized into several finite steps: the incremental deformation theory. Discretization is an inevitable procedure in the numerical analysis of a process, and a deformation path is assumed during discretized time steps. Conventional
approaches involve numerical integration schemes that usually assume various paths at
element nodes. Employing the incremental deformation theory introduces the minimum
work path during the discretized step, and no integration scheme is necessary; thus the
incremental deformation theory is useful in developing FEM codes to analyze processes.
Derivation o f the incremental deformation theory, which follows the same procedure
as for the total deformation theory, is summarized below.
In the multi-step process, consider an i-th step deformation during a discretized time
period, t 0 - t < to + At, as
F(t)

= Afi'Af

i-1 ' ' '

Af I = A f i ' F ( t o )

(Ari'Ani)

-F(to),

(18)

Deformation theory of plasticity

913

or equivalently as
F(t) = Ari(t) (Qi.AAi(t). (Qi)t)'F(to).

(19)

In eqn (19), Q~ and A A t are an orthogonal tensor and a diagonal tensor consisting of
the principal directions and stretches of Au i, respectively, which are based on the configuration at t = to. The incremental logarithmic strain Ag~ is obtained from A u ~ as
A~it ( t ) = Q i . At~i( t ) . (Qi),

(20)

A/j i = ln(AAi).

(21)

where

The minimum work path for smooth yield conditions is applied to each discretized
step: the minimum work path at each discretized step is achieved when the same principal material lines are fixed and the ratio of principal logarithmic strain rates is kept
constant during each step (see CrIoNc a: RICHMOND [1992a]). Now, we define g~ as
~i =

d( Al~i)/dt = ( A r i ) t . D . (Ari).

(22)

Note the difference between the definitions of ~ in eqns (22) and (5). The definition in
eqn (5) is valid when the principal stretch directions are fixed all the way from the start
to the end.
The minimum work conditions are satisfied when deformation satisfies the following relation:
~i(t ) = e.iL ( t ) = &i(t)"AeiL(to + A t ) ,

Aei(t) = Aeit(t) = ai(t)"Agit(to + At),

(23)
where cti(t) is a function of time, satisfying Jto
~to+4t.oLi dt = 1, and & i is always positive.
Eqn (23) shows that the principal material lines are fixed, and the ratios of the principal true strain rates are fixed during the discretized steps. Eqn (23) is similar to eqn (7).
In fact, eqn (23) is general, and eqn (7) is obtained from eqn (23) when a whole process
is considered as one step.
The minimum effective strain at the i-th step is also similar to eqn (9); i.e.
['to +At

A~(to + At) = |
a,o

~(6) dt = zl~i(Ae~L(to + At))

(24)

or, equivalently,
~ = &t(t).A~(to + At),

A~(t) = u~(t).Agi(to + At)

(25)

if the material hardens isotropically along the minimum work path during the interval.
If the material hardens anisotropically, AU is different for each step because A~;
evolves.
The constitutive laws for the incremental deformation theory are obtained by apply-

914

K. CmJN~ and O. RICHMOND

ing the relations in eqns (23) and (25) to the flow theory, and following the same procedures as in Sections III. Now, eqns (14) and (15) become
Aa i = ZIgi ~9'
3(?')'
~#7 = g g i a ( r , ) i ,

(r'

),

O,j~. ~

--BAli,

(26)

respectively. The secant slope, H, is defined as zaf/A~:(=A#/A~) in the incremental case.


The incremental deformation theory for rigid-plastic materials is illustrated in Fig. 1(a),
for discrete steps. Note that there is a discontinuous shift o f the direction o f r ' between

A
I
i AE i+1
I

/_~ _

/d

\\
~- i+1

/'

\\

I,

?'

~i-1
(a)

,i + 1/~--.....
~

/,~C ip

\\\\

(b)
Fig. 1. Schematic representations of FEM applications of the incremental deformation theory in the deviatoric stress ( r ' ) space for (a) rigid-plastic (b) etastoplastic materials.

Deformation theory of plasticity

915

the steps. The value of Q in eqn (19) is constant within each step, but it is different for
different steps. The change in the magnitude of r ' is controlled by the slope, H. In viscoplastic applications, A ~ / A t approximates ~, and it has different values for different
incremental steps to simulate changes of strain rate during deformation processes.
V. R I G I D - P L A S T I C D E F O R M A T I O N T H E O R Y FOR P O I N T E D YIELD SURFACES

Derivations of the constitutive laws for the total and incremental deformation theories in Sections III and IV are valid for smooth yield surfaces, where minimum work
paths are uniquely defined. When a yield surface has a pointed vertex at a desired stress,
multiple minimum work paths are possible and the condition in eqn (7) is not fully
required. However, the final equations in Sections III and IV are valid even for multiple minimum work paths, and this is discussed here. Only the total deformation theory
is considered because the case of the incremental deformation theory is similar to the
case of the total deformation theory.
The case with a pointed yield surface is illustrated in Fig. 2. In Fig. 2(a), point A has
multiple normals (therefore, multiple/~) on the yield surface defined in the deviatoric
stress space. A corresponding dissipation surface in the strain rate space is shown in
Fig. 2(b). On the surface between B and C, the normal direction is commonly in the
direction of r'. All the multiple ~ at A (or, between B and C) have the same w( = r ' .g),
so that there are multiple minimum work paths (e.g. the dotted paths are typical minimum work paths between A and A'). For each minimum work path, eqn (7) is not
required because the ratios of components of ~ are not fixed.

(a)

(b)
Fig. 2. Schematicdemonstrations of (a) a piecewisesmooth yield surface having multiple normal directions
at A, and (b) a correspondingrate of effective strain surface.

916

K. CHUNG and O. RICHMOND

Validation of eqn (16) is obtained f r o m the relation

m = f r'.a

=f

(27)

ae,

where m is the minimum work per unit volume. In eqn (27), m is a potential function
for each minimum work path because the values of r ' - g are the same for all * in Fig. 2;
i.e.

r.

Om

.
.
0~

dm Og
.
dg 0~

~'

Og
0~

(28)

Therefore, eqn (16) is valid for each of the multiple minimum work paths. Also, the following relation is obtained f r o m eqn (28):
r'-e = ?'~.a
Oa

= ~'g

(29)

after the Euler theorem is used.


Now, in order to verify that eqn (14) is also valid for multiple minimum work paths,
consider the following relation:

~(-_f , dt = f e Or'
~' dt) = K Or'

(30)

where K is a scalar quantity. Eqn (30) is possible because one of the multiple normal
directions 0 g ' / 0 r ' is parallel to each of ~, which is the value at the terminal state available from the multiple minimum work paths. The value of K is obtained from eqn (29)
after eqn (30) is substituted: K = g. Therefore, eqn (14) is valid for each of E available
at A and one of the multiple normal directions 0 ~ ' / 0 r ' which is parallel to it.
VI. COMPARISON WITH HYPERELASTICITY

The constitutive relation for hyperelasticity is known to be

o-

- -

det(F)

0o~(C)

F ' - - ' F

0C

t,

(31)

where ~0(C) is a strain energy function per unit initial volume and C is the Green deformation tensor. To compare this with deformation plasticity, this equation is rewritten
with respect to o~(sL) as follows. (Note that o~ is not a function of g, because E is pathdependent.) The definition of gL in eqn (2) suggests that ~ = U = exp(gL); therefore,
eqn (31) becomes equivalent to
r -

2
det(F)

exp(I;L)-

0o~(C(~L))
0C

exp(~L)t.

(32)

For incompressible hyperelasticity, eqn (32) becomes


r ' = exp(~L)

2.&o(C(~D)
exp(L) t.
0C

(33)

Deformation theory of plasticity

917

Comparison of eqn (33) with eqn (16) shows that deformation plasticity is similar to
incompressible hyperelasticity because a current stress state is determined by a current
strain state (note that 8 in eqn (16) is equivalent to tL for the minimum work path).
Even though the constitutive equations are similar, there is a major difference between
the present deformation theory of plasticity and hyperelasticity: there is no assumption
of a deformation path in hyperelasticity (as was also the case in the original deformation theory), whereas the present deformation theory of plasticity assumes a minimum
work path.
VII. ELASTO-PLASTICDEFORMATIONTHEORY
In elastoplastic deformation theory, the total deformation gradient, F, becomes
F = Fe'F p

= (Re.ue) (RP.U p)

(34)

if the theory by LEE [1969] is followed. The superscripts e and p stand for elastic and
plastic deformations, respectively. In the deformation theory for elastoplastic materials, plastic deformation is assumed to follow the minimum work path under isotropic
hardening, and the principal material lines for plastic deformation are assumed to coincide with the principal material lines for elastic deformation. Under this assumption,
eqn (34) becomes
F = (R e . R P ) (LIe . U P ) ,

(35)

where the principal material lines of ~e (= (R p)t. U e" R p) and U p are the same. Therefore, the following relation is obtained:
= 6~ + 8 p.

(36)

The constitutive equation for the elastoplastic deformation theory is obtained from
eqn (36) when E~ and gP are substituted in the elastic and plastic constitutive equations,
respectively. The plastic part is obtained from eqn (14). For the elastic part, three choices
might be considered: hyperelasticity, Cauchy elasticity, and hypoelasticity. When hyperelasticity is considered, G~ is obtained from eqn (32). When Cauchy elasticity is considered, the elastic part is obtained from the relation
r = r(E~).

(37)

When hypoelasticity is considered, the elastic part is derived from the hypoelastic relation, which is linear; i.e.
b = K - D e,

(38)

where K is a fourth order tensor. In eqn (38), the notation ~ stands for an objective
Cauchy stress rate. # was originally defined as the Jaumann rate of a. However, a proper
definition of an objective Cauchy stress rate became controversial recently, as shown
in the works by Dmr~ES [1979] and LEE e t al. [1983] among others. Here, the objective
rate proposed by Dienes is used purely for explanation purposes. Then, b ( =# - ~. o +
o.o~, where oJ = R . R t) becomes R.,/-.R t, so that eqn (38) becomes
~" = K ' ~ e -

(39)

918

K. CHUNG and O. RICHMOND

Because ~ is path-dependent, so is the integration of eqn (39). When the principal material lines for elastic deformation are fixed, eqn (39) becomes
~" = K.e~.

(40)

When the classical Hooke's law is used for the elastic part, and eqn (40) for Mises materials is used for the plastic part, the constitutive equation for small deformation and
small rotation becomes the original deformation theory proposed by HENCKY [1924].
In elastoplasticity, elastic and plastic deformations are responses to the same stress:
i.e. r = r e = .r p. Therefore, the condition in deformation plasticity that the principal
material lines of elastic and plastic deformations should coincide in the deformation theory is a restriction on material properties; elastic and plastic deformations have the same
principal material lines in response to the same stress. If materials are isotropic both in
elastic and plastic responses, the condition might be satisfied without a problem. If either
the elastic deformation or the plastic deformation is anisotropic, the condition will not
be satisfied, unless the anisotropic responses of both the elastic and the plastic deformations always have the same principal material lines in response to the same stress. If
the principal material lines of elastic and plastic deformations in eqn (34) do not coincide, the additive decomposition of the strain in eqn (36) is not assured, so that the constitutive equation becomes involved.
The incremental deformation theory for elastoplasticity is obtained from the following relation:
D = D e + D p.

(41)

The additive decomposition of the strain rate in eqn (41) is an assumed relation; however, LUBARDAand LEE [1981] showed that this relation is assured from eqn (34) when
materials are isotropic and elastic deformations are small compared to plastic deformations as in metals. When principal material lines and principal values are assumed fixed
for elastic and plastic deformations respectively during a time increment, eqn (41) becomes
At

= A E ie + /tF. ip,

(42)

where Ae ip and A~. ie a r e the increments of the plastic true strain and the elastic strain
at the i-th step, respectively. Incremental application of elastoplasticity for FEM analysis is illustrated in Fig. 2(b). In Fig. 2(b), zig ip is a response to r i, and the relation in
eqn (26) is used. Ae ie is a response to Ar i, and the relation is linear for a small increment; i.e.
A r i = K . Ag~[.

(43)

The linear relation in eqn (43) can be obtained from hyperelasticity, Cauchy elasticity,
or hypoelasticity; for small elastic deformation, the difference between them becomes
trivial.
VIII. S U M M A R Y

Even though the deformation theory of plasticity has been known to be accurate for
the description of mechanical properties of metals only in limited cases, the theory has

Deformation theory of plasticity

919

been found useful in FEM applications for both analysis and design purposes. To provide a theoretical basis for these applications, a special deformation theory has been
developed from flow theory here by assuming a minimum work path and isotropic hardening. The theory was obtained for arbitrary yield conditions including yield conditions
having pointed vertices. In this deformation theory, the objective constitutive equations
were easily obtained from flow theory simply by replacing variables without changing
the forms of equations. The "rotationless" strain rate g and its work-conjugate stress
7 were found to be convenient for this purpose. The equations of the deformation theory were obtained for an incremental approach as well as for a total approach. The
deformation theory for elastoplastic materials was covered as well as that for rigid-plastic materials.
Acknowledgement-The authors thank Dr. R.L. Mallett for his critical review of this article.

REFERENCES

1924
1957
1959
1963
1969
1972
1974
1975
1977
1978
1978
1979
1981
1982
1983
1984
1986

1986
1986
1986
1989
1989

HENCKY, H., "Zur Theorie Plastischer Deformationen und der Hierdurch Hervorgerufenen Nachspannunger," Zeits. Angew. Math. u. Mech., 4, 323.
HILL,R., "Stability of Rigid-Plastic Solids," J. Mech. Phys. Solids, 6, 1.
BUDIANSKY,B., "A Reassessment of Deformation Theories of Plasticity," J. Appl. Mech., 26, 259.
NADAI,A., Theory of Flow and Fracture of Solids, Vol. 2, McGraw-Hill, New York, pp. 96-105.
LEE, E.H., "Elastic-Plastic Deformation at Finite Strains," J. Appl. Mech., 36, 1.
PONTER,A.R.S. and MARTIr~, J.B., "Some Extremal Properties and Energy Theorems for Inelastic
Materials and Their Relationship to the Deformation Theory of Plasticity," J. Mech. Phys. Solids,
20, 281.
HUTCHINSON,J.W., "Plastic Bucking," Adv. Appl. Mech., 14, 67.
STOREN,S., and RICE, J.R., "Localized Necking in Thin Sheets," J. Mech. Phys. Solids, 23, 421.
ZmGLER,H., An Introduction to Thermomechanics, North-Holland Publishing Company, Amsterdam, pp. 287-290.
DAMAMME,G., "M6canique des Milieux Continus.-Minimum de la D6formation G6n6ralis~e d'un
l~16ment de Mati&e, pour les D6formation Passant d'un I~tat Initial ~t un l~tat Final Donn6s," Comtes
Rendus, Academic Society of Paris, 287, 895.
LEVY,S., SHIN,C.F., WILKINSON,J.P.D., STINE,P., and MCWILSON, R.C., "Analysis of Sheet Metal
Forming to Axisymmetric Shapes," in Formability Topics--Metallic Materials, ASTM STP 647, American Society of Testing and Materials, Philadelphia, pp. 238-260.
DIENES,J.K., "On the Analysis of Rotation and Stress Rate in Deforming Bodies," Acta Mechanica, 32, 217.
LUBARDA,V.A., and LEE, E.H., "A Correct Definition of Elastic and Plastic Deformation and Its
Computational Significance," J. Appl. Mech., 48, 35.
WANC,N.M., "A Rigid-Plastic Rate-Sensitive Finite Element Procedure for Sheet Metal Forming Processes," in Proc. International Conference on Numerical Methods in Industrial Forming Processes,
Pineridge Press, Swansea, U.K., pp. 797-806.
LEE, E.H., MALLETT, R.L., and WERTHEIMER,T.B., "Stress Analysis for Anisotropic Hardening in
Finite-Deformation Plasticity," J. Appl. Mech., 50, 554.
CrloNC, K., and LEE, D., "Computer-Aided Analysis of Sheet Material Forming Processes," in Proc.
1st International Conference of Technology in Plasticity, Vol. 1, pp. 660-662.
BRAUDEL,H.J., ABOtlAF, M., and CHENOT, J.L., "An Implicit and Incrementally Objective Formulation for Solving Elastoplastic Problems at Finite Strain by the F.E.M. Application to Cold Forging," in MATTIASSON,K. et al. (ed.) Proc. 2nd International Conference on Numerical Methods in
Industrial Forming Processes (NUMIFORN'86), Balkema, Rotterdam, pp. 255-260.
HILt, R., "Extremal Paths of Plastic Work and Deformation," J. Mech. Phys. Solids, 34, 511.
SKI.AD,M.P., "The Numerical Analysis of the Influence of the Material Hardening Behavior on Strain
Distribution in Drawing Sheet Metal Parts of Complex Shape," in Proc. Congress of International
Deep Drawing Research Group, pp. 464-467.
YANG,D.Y., and Kna, Y.J., "A Rigid-Plastic Finite-Element Formulation for the Analysis of General Deformation of Planar Anisotropic Sheet Metal and its Application," Int. J. Mech. Sci., 28, 825.
GES.MAn'LY., CI-ItlNG, K., and WAGONER,R.H., "A Rigid-Visco-Plastic Finite Element Program for
Sheet Metal Forming Analysis," Int. J. Mech. Sci., 31, 1.
NEALE,K.W., "Numerical Analysis of Sheet Metal Wrinkling," in THOMPSON, E.G. et al. (ed.) Proc.

920

1992a
1992b
In press

K. CHUNG and O. RICHMOND

3RD INTERNATIONALCONFERENCEIN INDUSTRIALFORMINGPROCESSES(NUMIFORN'89), BALKEMA,


ROTTERDAM, PP. 501-505.
CHUNG, K., and RICHMOND, O., "Ideal Forming, Part l: Homogeneous Deformation with Minimum Plastic Work," Int. J. Mech. Sci., 34, 575.
CHUNG, K., and RICHMOND, O., "Ideal Forming, Part II: Sheet Forming with Optimum Deformation," Int. J. Mech. Sci., 34, 617.
CHONG, K., and RICHMOND,O., "Mechanics of Ideal Forming," J. Appl. Mech.

Aluminum Company of America


Alcoa Center
Pennsylvania, USA

(Received 2 December 1991; in final revised form 24 May 1993 )

Vous aimerez peut-être aussi