Elastic Plastic Notes
Elastic Plastic Notes
Elastic Plastic Notes
Los Alamos National Laboratory, an affirmative action/equal opportunity employer, is operated by the University of California for the U.S. Department of Energy
under contract W-7405-ENG-36. By acceptance of this article, the published recognizes that the U.S. Government retains a nonexclusive, royalty-free license
to publish or reproduce the published form of this contribution, or to allow others to do so, for U.S. Government purposes. Los Alamos National Laboratory
requests that the publisher identify this article as work performed under the auspices of the U.S. Department of Energy. Los Alamos National Laboratory
strongly supports academic freedom and a researchers right to publish; as an institution, however, the Laboratory does not endorse the viewpoint of a
publication or guarantee its technical correctness.
LA-UR-03-0047 March 20, 2003
1 Kinematic Variables 1
1.1 Multiplicative Decomposition . . . . . . . . . . . . . . . . . . 2
1.2 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Transpose, Inverse and Transpose-Inverse . . . . . . . . . . . . 5
1.4 Strain Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.5 Volumetric and Deviatoric Components . . . . . . . . . . . . . 9
2 Hyper-Elasticity 11
2.1 Isotropic Material . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2 Illustrative Elastic Models . . . . . . . . . . . . . . . . . . . . 14
2.2.1 Linear Elasticity . . . . . . . . . . . . . . . . . . . . . 14
2.2.2 Non-linear Isotropic Model . . . . . . . . . . . . . . . . 15
2.3 Conservation Equations . . . . . . . . . . . . . . . . . . . . . 17
2.3.1 Eulerian Form . . . . . . . . . . . . . . . . . . . . . . . 17
2.3.2 Lagrangian Form . . . . . . . . . . . . . . . . . . . . . 18
2.4 Kinematic Equations . . . . . . . . . . . . . . . . . . . . . . . 20
2.5 Dissipation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.6 Wave Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.7 Shock Jump Conditions . . . . . . . . . . . . . . . . . . . . . 31
5 Uniaxial Strain 51
5.1 Stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
5.2 Yield Condition . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.3 Plastic Strain Rate . . . . . . . . . . . . . . . . . . . . . . . . 54
6 Wave Structure 58
6.1 Weak Purely Elastic Wave . . . . . . . . . . . . . . . . . . . . 59
6.2 Split Elastic-Plastic Wave . . . . . . . . . . . . . . . . . . . . 60
6.3 Strong Plastic Wave . . . . . . . . . . . . . . . . . . . . . . . 61
6.4 Visco-Elastic Wave . . . . . . . . . . . . . . . . . . . . . . . . 62
6.5 Partly and Fully Dispersed Waves . . . . . . . . . . . . . . . . 63
6.6 Entropy Anomalies . . . . . . . . . . . . . . . . . . . . . . . . 64
Acknowledgement 79
References 79
Index 83
1 Kinematic Variables
Elastic flow is formulated in terms of a time-dependent mapping from
the body reference frame to the spatial laboratory frame
(, t) : B S , ~ t) ,
xi = i (X,
U ~ t) = T ~ S .
~ (X, (1.1)
(X,t)
t
Alternatively, the particle velocity can be expressed as a map from the spatial
frame to the tangent space of the spatial frame
~ 1 T~x S .
~u(~x, t) = U
F = Fe Fp , Fi = (Fe )i A (Fp )A .
from the deformation gradient are not satisfied. Nevertheless, we assume the
multiplicative decomposition of the deformation gradients and consider the
mappings only as a convenient heuristic.
There is still a question with non-uniqueness of the decomposition; i.e.,
F = (Fe Q1 )(QFp ) for any invertible Q : T P T P. The non-uniqueness
is similar to the question of frame indifference discussed in a later section.
The difficulty is resolved in part by formulating plasticity in terms of total
strain and plastic strain rather than the deformation gradient alone. In
addition, the choice of the plastic strain rate is needed to define the plasticity
model.
1.2 Notation
~ =
E .
X
Then a vector in TX~ B may be written as
~ = U E
U ~ ,
where the summation over repeated indices, upper and lower, is assumed.
~ that are
Thus, it is the components of a vector relative to the basis E
denoted with superscripts.
Similarly, we denote the dual basis elements in T X~ B by
~ = dX ,
E
V~ = V E
~ .
U 7 (U 0 ) = (F) U
V 7 (V 0 ) = (FT ) V ,
The metrics are independent of time but may vary from point to point in B or
S. The matrices G and gij are symmetric, consequently, hW~ , V~ i = hV~ , W
~i
and hw,
~ ~v i = h~v , wi.
~
where ~x = (X,~ t) or X
~ = 1 (~x, t). Moreover, these operators satisfy the
following identities
F1 F = IT ~
X
B and F F1 = IT xS
~
FT FT = IT X~ B and FT FT = IT ~x S
The inverse of the derivative map is simply the derivative of the inverse
map, F1 = D(1 ). The transpose can be defined in a co-coordinate free
manner by the relation
~ )~x = (FT ~v , W
(~v , FW ~ ) ~ = v Fi W ,
X i
where ~v T ~x S and W
~ T ~ B. Instead of the transpose, ref. [11] uses the
X
adjoint F , defined by the relation
~ , ~v i~x = hW
hFW ~ , F~v i ~ ,
X
C = (FT g) F ~ t) = (gij ) Fi Fj
C (X, : TX~ B T X~ B
E = 12 (C G) ~ t) = 1 (C G ) : T ~ B T ~ B .
E (X, (1.3)
2 X X
Transforming the strain tensor from the Lagrangian frame to the spatial
frame, E 7 FT EF1 , results in the Eulerian or Almansi strain tensor
Ep = 21 (Cp G) , ~ t) : T ~ B T ~ B ,
(Ep ) (X, (1.5)
X X
where heuristically
Cp = (Fp T G)
e F ,
p p
~ t) : T ~ B T ~ B
(Cp ) (X, X X
where heuristically
1
be = (Fe G
e ) 1 F T ,
e e (be )ij (~x, t) : T ~x S T~x S
= FCp 1 FT
e = ee + e p .
Similarly, the elastic strain tensor can be transformed to the body frame
Ee = FT ee F , Ee (X, ~ t) : T ~ B T ~ B (1.8)
X X
= 21 (C Cp )
= 12 (C G) 12 (Cp G)
E = Ee + Ep .
Many theories of plasticity start out with the super-position assumption in-
dependent of the multiplicative decomposition. All theories are constructed
to have the same small strain limit. However, theories can differ greatly in
the large strain or non-linear regime.
Remarks:
1.2 One could also define elastic and plastic strains in the intermediate
plastic frame as
eee = 12 [(Fe T g) e Fe G]
e , ~ t) : T ~ P T ~ P
eee (X, X X
The ratio J = VV0 of the specific volume in the spatial frame to the specific
volume in the reference frame is determined from the deformation gradient
by the relation
~ t) det F F = det G1 C
J2 (X,
1
= det [G + 2E]G
= det IT X~ B + 2EG1
For small strain, kEkG1 1, (see Remark 1.5 for definition of norm
and Remark 1.4 for definition of trace) we can write
J2 = 1 + 2 Tr E G1 + O 2 . (1.9)
The determinant factors enter because (det G)1/2 d3 X is the volume element
in B, and (det g)1/2 d3 x is the volume element in S. When the metrics are
normalized, det(g) = det(G) = 1, the specific volume ratio can be expressed
simply as J = det(F). However, the metric still is needed in Eq. (1.9) to
relate J directly to the strain tensor.
The strain tensor can be split into the sum of a volumetric component
and a deviatoric component, (see Remark 1.6 for definition of deviator)
E = 13 Tr E G1 G + devG (E) .
b1 = FT GF1 : T S T S
Transforming the Eulerian strain to the Lagrangian frame gives the decom-
position
E = 13 Tr E C1 C + devC (E) .
The contraction of upper and lower indices is invariant under any transfor-
mation. Hence the trace of an operator on T B or T B, such as the tensor
G1 E : T B T B or the tensor EG1 : T B T B, is invariant under a
change of basis.
The metric is needed in order for the norm to be independent of the choice
of basis. The norm also can be express in terms of the trace,
kEk2G1 = Tr E GT 1
EG 1
.
2 Hyper-Elasticity
A material model is called elastic if the stress is a function of the strain.
Some engineering models are defined in terms of the stress rate as a function
of stress, strain and strain rate. These models have the property that the
stress depends on the strain path. Material models for which the stress can
not be expressed in terms of only the current strain state are called hypo-
elastic. In order to have a thermodynamic description an energy function
is needed such that the stress is the derivative of the energy with respect to
the strain. Materials models with such an energy function are called hyper-
elastic.
Frame indifference requires that the energy function is invariant for all
isometries on T S, i.e., under the transformation F 7 Q F for all Q : T S
T S such that
QT g Q = g .
As a consequence the energy can depend on the material deformation only
through the strain tensor E and not on the deformation gradient by itself.
E = E(E, Ep , ) ,
where is the specific entropy. Furthermore, we assume the thermody-
namic relation
dE = V0 S : dE V0 Sp : dEp + T d , (2.1)
where T is the temperature, and S and Sp are the elastic and plastic stress
tensors in the Lagrangian frame. The : denotes contraction over a pair of
tensor indices, e.g., S : dE = Tr(ST dE) = S dE .
The second Piola-Kirchhoff stress tensor and the corresponding plas-
tic stress tensor are given by
E E
S = 0 , ~ t) =
S (X, 0 : T X~ B TX~ B
E E
E ~ t) = 0 E : T ~ B T ~ B
Sp = 0 , (Sp ) (X, X X
Ep (Ep )
or alternatively as [2]
E
= 2 .
g
Since the strain E is a symmetric matrix, it follows that both the second
Piola-Kirchhoff stress S and the Cauchy stress ij are symmetric matrices.
Remark 2.1 The symmetry of the stress tensor is required to ensure local
angular momentum conservation. This condition assumes that individual
atoms or molecules within a material have no angular momentum.
= P g 1 + devg1 () ,
This formula enables us to transform the stress deviator between the Eulerian
and Lagrangian frames.
G1 C : T B T B , (F F) = G C .
I3 det(G1 C) = J2
and using Lemma A.1, the stress can be expressed as [23, p. 261, Eq. (7.1.108)]
! ! !
E E E E
1
VS
2 0
= + I1 G1 G1 C G1 + I3 C1
I1 I2 I2 I3
! ! !
1 E 1 E E E
2
V = I3 g + + I1 b bgb
I3 I1 I2 I2
Two example models for an isotropic elastic material are described. The
first model is known as linear elasticity. The second model is for a non-linear
isotropic material. For each model, the specific energy and the corresponding
stress tensors are presented.
= J1 F S FT
V0 V0
= K0 Tr (b e) b + 2 G0 devb (b e b) ,
V V
where K0 = 0 + 32 G0 , is the bulk modulus, and G0 is the shear modulus.
In the small strain limit b = g 1 + O kekg1 , and the Cauchy stress
reduces to
= K0 Tr g 1 e g 1 + 2 G0 devg1 g 1 e g 1 + O kek2g1 .
With the identity matrix for the spatial metric, gij = ij , this equation
corresponds to the standard form for the stress of a linearly elastic isotropic
material.
Remark 2.3 Linear elasticity is a reasonable description of many metals
at a stress low compared to the yield strength. It is frequently used in
engineering applications.
Despite its appearance, we note that the deviator term is a symmetric matrix,
due to the identity
1 1 1 1 1 1 1 1
G EC = 2
G (C G)C = 2
G C .
2.5 Even if a material is initially isotropic, the material does not remain
isotropic for non-hydrostatic deformations. The stress is still determined by
the isotropic specific energy for the reference state. However, the elastic
tensor for a particular strain state need not be isotropic. This is important
outside the domain of linear elasticity, i.e., large elastic strains, though plastic
deformation may mitigate the anisotropy of the deformed state by limiting
the non-hydrostatic component of the elastic strain tensor.
+ uj =0
t ;j
ui + ui uj ij =0 (2.5)
t ;j
h1 i
i
h i
2 ui u + E + 12 ui ui + E uj ui ij =0
t ;j
~ (ui~ei ) = i i k
u + ik u = ui ;i
xi
In addition, the covariant derivative of a tensor with contravariant indices is
ij j
ij ;k = k
+ kpi
pj + kp ip ,
x
and a tensor with covariant indices
p p
eij;k = eij ik epj jk eip .
xk
In particular, the divergence of the Cauchy stress is
ij j
ij ;j = j
i
+ kj kj + kj ik .
x
a a
The Christoffel symbol is symmetric in the lower indices, i.e., bc = cb .
The curvature tensor is defined in terms of the Christoffel symbol by
a a
db cb
ra bcd = a e
+ ce a e
db de cb .
xc xd
The non-commutivity of the covariant derivative,
ua ;bc ua ;cb = ra bcd ud ,
is characterized by the curvature tensor.
Remark 2.6 The Christoffel symbol is not a tensor, i.e., for the transformed
metric C = FT gF, the transformed Christoffel symbol 6= (F1 ) i jk
i
Fj Fk .
The flow equations can be transformed from the Eulerian frame to the La-
grangian frame by applying the Piola identity. Define the Piola transform
of ~y T S 7 Y~ T B by
Y = J (F1 ) i y i . (2.7)
Y ; = J y i ;i , (2.8)
~ t) | X
where (, t) = {~x = (X, ~ }. As a consequence of Piola identity
and Gauss law
I Z Z I
dA N Y~ = 3
d XY
; = 3
d xy i
;i = da n ~y .
() ()
This enables the surface terms in the integral form of the conservation laws
to be transformed readily between frames. Then re-expressing the conserva-
tion laws as differential equations leads to the Lagrangian form of the flow
equations
d
J J (F1 ) i U i =0
dt ;
d
0 U i Pi ; =0 (2.9)
dt
d 1
i i
0 U i U + E U i P =0
dt 2 ;
where d
=
= ~ is the convective derivative, and
+ ~u
dt t X t ~
~
x
E ~ t) = Fi S : T ~ B T ~ S
P = g 1 = FS , Pi (X, X X
F
is the first Piola-Kirchhoff stress tensor. In addition, mass conservation
is trivial, dtd 0 = 0, and has been replaced by volume conservation. Equa-
tion (2.9a) is derived by applying the Piola Identity to Lemma A.4. We
note that the first Piola-Kirchhoff stress is a two-point tensor and is not
represented by a symmetric matrix.
2 i i ~
0 = P D .
t2 X
This is well defined for smooth flows but not for discontinuous flows, because
the right hand side is a nonlinear function of X i . Discontinuities, known
as shock waves, commonly arise in the solution of nonlinear wave equations
even when the initial data is smooth. The strain tensor is introduced as an
independent variable in order to obtain a system of quasi-linear PDEs (i.e.,
first order and linear in all derivatives) and use the theory of hyperbolic
PDEs to regularize the continuum model.
In Lagrangian form, the evolution of the deformation gradient
d ~ , d i
F = DU F= Ui , (2.10)
dt dt X
i i 2
F =
F =
i . (2.11)
X X X X
From Eq. (2.10) it follows that if the compatibility conditions are satisfied
at some initial time then the compatibility conditions will be satisfied for
all time. Consequently, Eq. (2.10) is equivalent to 3 independent equations
and subsume the volume conservation equation (2.9a). The effective system
of PDEs has 7 independent equations. Hence, physically there are 7 wave
families. With Eq. (2.11) the full system has 6 additional physically irrelevant
waves. Later we show that the full system is indeed hyperbolic and that the
physical waves can be identified with 3 acoustic wave families (one quasi-
longitudinal and two quasi-transverse) in both the forward and backward
directions and a linearly degenerate mode for the advection of entropy.
Remark 2.8 When the continuum equations are discretized for numerical
simulations, the compatibility conditions may not be satisfied exactly. The
problem is analogous for incompressible fluid flow to enforcing the condition
that the divergence of the velocity vanishes. If the error in the compatibility
conditions increases with time of run then the simulations can be seriously
inaccurate or numerical instabilities may develop. Part of the problem is that
vorticity in the flow can lead to stirring on a fine scale and when resolution
of is lost, the discretization errors can become large.
The Eulerian form for the evolution of the deformation gradient is ob-
tained by applying the chain rule to Eq. (2.10)
d ~ u) f , d i
f = (~ f = ui ;j f j , (2.12)
dt dt
where f(~x, t) = F 1 : T~1 (~x,t) B T~x S. Plohr and Sharp [16] have shown
that this evolution equation may be recast in conservation form.
1 i
J f + J1 f i uj = J f u 1 j i
. (2.13)
t ;j ;j
Proof. From the mass conservation Eq. (2.5a) and Eq. (2.12), the left hand
side of Eq. (2.13) can be expressed as
1 i d i
J f + J1 f i uj = J1 f
t ;j dt
= J1 f j ui ;j
Apply the Piola identity Eq. (2.8) to the second term on the right hand side,
we find
J f = J1 ; = 0 .
1 j
;j
Alternatively, the analog of Eq. (2.10) for the inverse map leads to the
evolution equation for f 1 [28]
(f 1 ) i + (f 1 ) j uj =0, (2.130 )
t ;i
2.5 Dissipation
d
E = V :d . (2.17)
dt
Substituting the relation for the external work Eq. (2.16) and the rate
of change of internal energy Eq. (2.17) into the thermodynamic identity
Eq. (2.1) gives a constraint on the plastic strain rate
2.11 The time derivative of the Eulerian strain can be expressed as a Lie
derivative, de
dt
= L~u (e). In effect, the time derivative is defined as the
The wave speeds for a hyperbolic system of PDEs are determined from
the eigenvalues of the flux matrix. Because of the complications that arise
from the compatibility relations, we shall use the Lagrangian equations and
then transform to the spatial frame.
A few preliminary definitions are needed. The fourth rank stiffness ten-
sor is defined by
S 2E
C = 0 .
E E E
It follows from the definition that the stiffness tensor has the following sym-
metry properties
0 0 0 0 0 0 0 0
C =C =C =C .
The stiffness tensor allows small changes in the stress to be related to small
changes in the strain
0 0
S = C : E , (S) = C (E)0 0 .
Hence,
d d
S = C: E
dt dt
= C : (FT dF) .
where
1 dS
= F FT
J dt
d ~ u) (~
~ u)T + Tr g 1 d ,
= (~
dt
dE
e = FT F1
dt
d ~ u)T e + e(~
~ u) .
= e + (~
dt
Consequently, the time derivative of the Cauchy stress can be expressed as
d
+ = c : d , (2.19)
dt
where
cijpq = cijpq + 1 ip jq
g + g +g + gip jq jp iq jp iq
ij g pq ,
2
and h i
= 1 ~ u) (~
(~ ~ u)T .
2
Remark 2.12 The form of the time derivative of the Cauchy stress given
by Eq. (2.19) is frame indifferent. The form of is known as Truesdells
objective derivative. There are other forms for objective time derivatives of
tensors, see for example [5] and [11, 1.6]. They are related to the choice
of frame in which the stiffness tensor is specified. The form of objective
derivative is important for numerical simulations based on hypo-elastic con-
stitutive models, especially when the shear component of the stiffness tensor
is assumed constant. A constant stiffness tensor is a useful approximation
for small strain but in general is not consistent with hyper-elasticity [1, 24].
The inverse of the stiffness tensor is called the compliance tensor and
allows small changes in the strain to be related to small changes in the stress
0 0
E = S : S , (E) = S0 0 (S) ,
where
0 0
C S0 0 = 1
2
+ .
The symmetry requirement, i.e., the symmetric identity operator on the right
hand side, complicates the computation of S from C. Details are given in
Appendix B.
An additional tensor, closely related to the stiffness tensor, is the acous-
tic tensor
P ij 0 0
A = g 1 , A = Fi 0 F j 0 C + g ij S
F
The acoustic tensor is a two-point rank-four tensor. However, in the spatial
frame, the acoustic tensor,
ipjq ij ipjq
a = J1 Fp Fq A =c + g ij pq ,
is a one-point rank-four tensor.
In the previous subsection we found that one characteristic is associated
with the advection of entropy. To find the other wave families, we take the
entropy to be constant and drop the energy equation. We analyze the system
of 12 PDEs consisting of the three momentum equation (2.9b) and the nine
kinematic equations (2.10)
d i i
U A j
0
Fj = 0
dt X
d i
F Ui =0
dt X
Here, the metric is assumed constant and the covariant derivative reduces to
the ordinary derivative in flat space.
We look for traveling wave solutions of the form W ~ = W~ () where
~ Ct, C is the wave speed, the direction of propagation N
= hN , Xi
is normalized to satisfy hN , N i = 1, and the vector of unknowns is
~ = (Fj =1 , Fj , Fj , U i )T ,
W with i, j = 1, 2, 3 .
=2 =3
The wave speeds are the solution to the equation obtained by setting the
determinant of L to zero. By adding multiples of the first three columns to
the fourth column the determinant equation reduces to
0 = det(L)
CI 0 0 0
0 CI 0 0
= det
0 0 CI 0
i =1 i =2 i =3
N A j N A j N A j 0 CI C 1 A(N )
= C 6 det A(N ) 0 C 2 I
i i
where A(N ) j = N A j N . Thus, there are six trivial wave speeds, C = 0,
and three pairs of non-trivial waves, C, determined by the equation
i 2 i
det A(N ) j 0 C j =0. (2.20)
We show that the trivial waves are ruled out by the compatibility condi-
tion Eq. (2.11), and that each pair of non-trivial waves corresponds to an
acoustic wave family.
First we examine the non-trivial waves. Let U ~ be an eigenvalue of A(N )
with eigenvalue 0 C 2 . Then a solution to the linearized equations is given by
~ =U
U ~ W ()
~0 + U
Fi = (F0 )i N (Ui /C)W ()
for any smooth function W (). We note that the deformation gradient
caused by the wave Fi = N (Ui /C)W () satisfies the compatibility
condition Eq. (2.11) and is physically admissible.
~ = 0 and
The eigenvector corresponding to the eigenvalue C = 0 have U
the change in the deformation gradient satisfying
i
j
N A j F =0, for i = 1, 2, 3. (2.21)
Since there are three linear equations for nine unknowns, Fj , there are
six families of solutions corresponding to the six fold degeneracy of the zero
Remarks:
2.15 We note that for zero stress, the leading order expansion of the specific
energy is
E = 21 E : C : E + O kEk3G1 .
isotropic material drops from [(K+ 43 G)/]1/2 in the elastic regime to (K/)1/2
in the plastic regime. In addition, since plastic flow is dissipative, traveling
waves in the plastic regime can be expected to decay, i.e., rather than an
elastic wave w = w(n ~x ct), a traveling plastic wave would have the form
w = w(n ~x ct) exp(t/ ).
M (U i ) = (Pi ) N
M 1
U Ui
2 i
+ E = (Ui Pi ) N (2.23)
M (Fi ) = 0 (U i ) N
where M = 0 Us is the mass flux through the shock front, and Us is the
shock velocity in the body frame. The jump of a variable U across the shock
front is denoted (U ) = U1 U0 . In addition, the propagation direction is
normalized to satisfy N N = 1.
Using the relation
where a bar over a variable denotes the average, i.e., A = 12 (A0 + A1 ), the
jump in the strain is given by
j i
h i
(E ) = 1
2
(Fi ) gij F + F gij (Fj )
i i
M (E ) = 21 0 (Ui ) N F + F N (2.24)
~ ) and is the analog of a gas dynamic shock wave. The quasi-transverse
(U
waves satisfy (U ~ ) N (U
~ ) and are associated with shear stress.
The Hugoniot equation can be used to show that for weak shocks the entropy
change is third order. However, in contrast to the scalar gas dynamic case,
the Hugoniot equation is not sufficient by itself to determine the shock locus.
Remark 2.19 Using the relation
P = FS + 14 (F)(S)
M (ui ) = ( ij ) nj
M 1
u ui
2 i
+ E = (ui ij ) nj (2.26)
M (f i ) = ( f j ui ) nj
where M = (us n ~u) is the mass flux through the shock front, us is the
shock velocity in the spatial frame and the propagation direction is normal-
ized to satisfy ni ni = 1. It can be shown using the Piola identity that the
Lagrangian and Eulerian form of the jump conditions are equivalent [16].
Ee = Ee (E Ep , ) . (3.2)
e 1 C
By Lemma A.5 the invariants of Cp 1 C are the same as those of G e .
e
Consequently, we take
Ee = Ee (Cp 1 C, ) . (3.3)
d b b
Y dC Y dCp
b
dt
Y = C : dt + Cp
: dt .
Plastic flow occurs when the contribution from the total strain rate, dtd C,
raises the value of the yield function to the yield strength. For many ap-
plications rate independent plasticity models are used. In these models,
the plastic strain rate dtd Cp is chosen such that the yield function is never
exceeded, i.e.,
d b
dt
Y 0 , if Yb = Y .
This determines the magnitude of Cp , though the direction must be specified
by a model.
Remark 3.2 The term rate independent refers to the lack of an explicit
time constant. For simple problems, such as uniaxial strain in which the
total strain increases monotonically, the plastic dissipation depends only on
the strain and is independent of strain rate.
As an extension of elastic flow, rate independent plasticity has two prob-
lems. From Eqs. (2.14) and (2.15) the total strain rate is dtd C = 2 FT dF.
Hence the plastic strain rate is no longer equal to a source term, but depends
on dij = 12 (ui;j + uj;i ). Consequently, just because the elastic flow equations
are hyperbolic, it does not follow that the extended system of PDEs remains
hyperbolic. In fact for some choices of the direction of the plastic strain rate
and yield function, the rate independent equation have elliptic regions. The
initial value problem for these models can be regularized by adding shear
layers when the solution runs into an elliptic region [20]. Another problem
arises even for rate independent plasticity models in which the extended sys-
tem of PDEs remains hyperbolic. The equation for plastic strain is not in
conservation form. Hence there are not sufficient jump conditions to de-
fine strong shock waves. However, fully dispersed weak shocks (for which
the plastic dissipation provides the required shock entropy) are well defined.
Consequently, a rate independent plasticity model only provides solutions for
the limited class of initial value problems which do not lead to strong shocks.
Here, we consider rate dependent plasticity. In these models the value
of the yield function is allowed to exceed the yield strength. The plastic
strain rate serves to relax the yield function back to the yield surface. A rate
dependent plasticity model is determined by specifying a plastic strain rate
such that three conditions are satisfied:
The system of PDEs for rate dependent plasticity is hyperbolic whenever the
elastic system of PDEs is hyperbolic.
This is because the plastic rate equation is in characteristic form. Since
source terms do not affect the flux matrix, the wave speeds of the augmented
elastic-plastic system are the same as the original elastic system with added
degeneracy for the convective wave speed (which contains the entropy wave
family). Hence, the augmented system has real wavespeeds. To remain hy-
perbolic, a complete set of eigenvectors of the flux matrix are required. De-
spite the degeneracy in the eigenvalues, the eigenvectors do remain complete.
To see this we note that the characteristic form of the PDEs is determined
by the left eigenvalues of the flux matrix. Adding rate equations do not
affect the characteristic form of the original system and hence the eigenvec-
tors of the acoustic wave families. Furthermore, the eigenvectors of the rate
equations depends on the directions associated with the new variables and
are linearly independent of the eigenvectors associated with the wave fami-
lies of the original system. Consequently, the elastic-plastic flow equations
are hyperbolic. The same argument applies if additional internal degrees of
freedom, along with rate equations for their evolution, are added in order to
describe the response of a material more realistically. Isotropic or kinematic
work hardening parameters are examples of other internal degrees of freedom
that are commonly added.
As a consequence of the plastic strain rate equation, the physically inter-
esting plastic waves are partly dispersed, i.e., a composite wave consisting
of a discontinuous shock followed by a narrow layer in which the yield func-
tion relaxes to the yield surface. Unlike fluid dynamics, the end state of a
plastic wave depends on the source term for the plastic strain rate and is
not fully determined by the Hugoniot jump conditions. For special cases,
such as uniaxial strain, the jump conditions plus the yield condition suffice
to determine the end state. But in general the jump in the plastic strain
across a plastic wave can only be determined by integrating the ODEs for
the profile of a traveling wave. This is similar to what occurs for the system
of PDEs describing reactive flow with more than one species and multiple
reactions. The change in equilibrium composition for flow through a nozzle,
such as a jet engine, depends on the reaction rates relative to the flow rate.
Remark 3.3 Though strong shock waves are well define for rate dependent
plasticity, the relaxation zone may be very narrow. Problems can occur in
numerical simulations when the cell size is insufficient to resolve the wave
profile.
An associated flow rule relates the plastic strain rate to the yield func-
tion. Though the plastic strain can be based on non-associative flow rules
(see for example [21]), here we consider only the flow rule obtained from
what is referred to as the maximum dissipation postulate. In the rate in-
dependent case, C is varied over the yield surface, Y = Y(C,
b Cp ), to find an
extremum in the plastic dissipation, Eq. (2.18), for fixed Cp and dtd Cp . The
extremum occurs when (see [22] and references therein)
Sp d Yb
: Cp = , (3.4)
C dt C
where is a Lagrange multiplier which is determined from the constraint
that C lies on the yield surface. For the moment, let us assume that S C
p
d Y
Cp = . (3.5)
dt Sp
This equation has the interpretation that the plastic strain rate is normal to
the yield surface when the yield function is expressed in terms of the plastic
stress.
In addition, we assume that for fixed Cp the interior of the yield surface,
{Sp | Y(Sp , Cp ) Y }, is convex and contains the point Sp = 0. From
the geometry of the yield surface, it can be shown that the extremum of
d Y
Sp : Cp = Sp : ,
dt Sp
Y
and Sp
is an outward normal, it follows that > 0. In the rate independent
d
case, setting dt
Y = 0 determines the Lagrange multiplier to be
Y Sp d
S p
: C : dt C
= Y Sp Y Y Y
. (3.6)
Sp
: C :
p Sp
+ Cp
: S p
From the existence of an extremum and > 0, it follows that the denomi-
nator is non-zero and negative.
Remarks:
3.4 The associated flow rule applies when the yield surface is smooth. For
some simple models, such as the maximum shear stress criterion of Tresca,
the yield surface has kinks or points at which the normal is discontinuous.
Models for crystal plasticity assume the plastic strain is due to slip along
preferred crystal planes. This also results in a yield surface with corners. At
corners the gradient of the yield surface is ill defined. The plastic strain rate
can be defined as a convex linear combination of the normals to the adjacent
facets.
3.6 The compatibility condition for the existence of a mapping can be ex-
pressed in terms of the strain tensor rather than the deformation gradient
Eq. (2.11). In general, Ep does not satisfy the compatibility condition. Con-
sequently, p and hence Fp can not exist. For crystal plasticity, at the molec-
ular scale, one expects p to be discontinuous due to dislocations and grain
boundaries. On the macroscopic scale the discontinuities are smeared out us-
ing some coarse grain averaging procedure to give a continuous plastic strain
tensor. The effect of the coarse graining can be thought of as converting
the discontinuities to source terms that prevent the compatibility relations
from being satisfied. Consequently, Ep should be thought of as an internal
degree of freedom that characterizes some aspects of a material not included
in elastic models. The plastic mapping may be used as a heuristic to moti-
vate the structure of the model equations but should not be taken literally
and proofs for properties of the model must not rely on the existence of p
or even Fp .
To generalize to the rate dependent case, we assume that any level set of
the yield function satisfies the geometric conditions of a yield surface (i.e.,
the interior of Y = Y1 for any Y1 > Y is convex and contains the point
Sp = 0), and that the scale factor is a state function, = (C, Cp , ) > 0,
to be specified by the plasticity model. We note that has dimensions of
inverse time, and in order that there is no plastic flow in the elastic regime
we require = 0 for Y(Sp , Cp ) Y . The construction of a rate dependent
model based on the maximum dissipation postulate guarantees that plasticity
is dissipative and that plastic flow relaxes the value of the yield function
toward the yield surface.
Remarks:
3.7 The condition that every level set of the yield function satisfy the geo-
metric conditions for a yield surface is a weak assumption. The same condi-
tion is needed to account for work hardening.
where J2p = det(G1 Cp ). With this functional form the energy is independent
of the plastic volume and it is straight forward to show that the plastic stress
is deviatoric, i.e., devCp 1 (Sp ) = 0.
The plastic strain rate of Eq. (3.5) is volume preserving when the yield
function has the form
Y = Y(devCp 1 (Sp ) , Cp ) . (3.8)
However, to verify the plastic strain rate satisfies the required properties
of a plasticity model, we need to reexamine the steps in the derivation of
Eq. (3.5). In particular, the assumption that SC
p
is invertible is not true
when Sp is deviatoric.
With the volume preserving form of the yield function, Eq. (3.8), the
extremum condition for the dissipation, Eq. (3.4), can be written as
Sp d Y
: Cp =0.
C dt Sp Cp
We note that the term in parenthesis is deviatoric. Provided that the null
space of S
C
p
is limited to the deviatoric component, we can conclude that
Eq. (3.5) remains valid for the volume preserving plasticity.
Remark 3.9 For porous materials or materials with cracks, the volume
preserving assumption of the plastic deformation is not even approximately
valid. Thus, the volume preserving constraint is not a physical law, but
merely an approximation that leads to a useful description of many materials,
in particular, metals.
ii. A yield function Y(Sp , Cp ), and a rule for the direction of the plastic
strain rate, such as the associated flow rule, Eq. (3.5).
iii. A scalar function (C, Cp , ) which specifies the rate at which the yield
function relaxes toward the yield surface.
1
energy depends on the invariants of J2/3 p Cp C, the plastic stress is related
to the elastic stress deviator as follows.
Lemma 4.1. The component of the plastic stress from the elastic energy
and the Piola-Kirchhoff stress are related by
devCp 1 S0p Cp = devC1 (S) C . (4.2)
1
Proof. Let I1 ,I2 , I3 be the invariants of J2/3
p Cp C Applying the chain rule,
the stress tensors are given by
Ee I1 Ee I2 Ee I3
20 S0p =
I1 Cp I2 Cp I3 Cp
Ee I1 Ee I2 Ee I3
20 S = + +
I1 C I2 C I3 C
From the definition of the invariants, it is straight forward to show that the
invariants satisfies
! !
Ij Ij
devCp 1 Cp = devC1 C.
Cp C
The stress stress corresponding for the model, Eq. (4.1), is given by
S = P (V, ) J C1 + G0 (Jp /J)2/3 devC1 Cp 1 ,
G0 (4.3)
= P (V, ) g 1 + (Jp /J)2/3 devg1 (be ) ,
J
where P = V
Ehydro is the hydrostatic pressure, and be = FCp 1 FT . In
addition, the plastic stress is given by
Sp = G0 (Jp /J)2/3 devCp 1 Cp 1 CCp 1
(4.4)
= G0 (Jp /J)2/3 devI Cp 1 C Cp 1 .
Remarks:
4.1 It is interesting to note that Eq. (4.3) for the stress can be expressed as
S = P (V, ) J C1 + 2 G0 (Jp /J)2/3 devC1 Cp 1 Ee C1 .
This is equivalent to Eq. (2.4) with Ee substituted for the strain E and J2/3
p Cp
substituted for the metric G.
where the shear energy eshear is a function of only the strain. This is a
simple, thermodynamically consistent and computationally efficient way to
add strength to a hydrostatic equation of state. Moreover, the material
temperature is determined from the hydrostatic equation of state in the usual
manner. In practice there is no data with which to define the temperature
any better. On the other hand, the shear modulus is expected to decrease
as T % Tmelt .
4.3 A more general functional form for the stress can be obtained in terms
of the Helmhotz free energy, e (Cp 1 C, T ) = Ee (Cp 1 C, ) T . Letting the
shear modulus be a function of T and V ,
1/3
e = hydro (V, T ) + 12 V0 G(V, T ) (I1 I3 3) ,
and there are contributions to the entropy from both the shear strain and
the specific volume. In this case the Cauchy stress is given by
h i
1
= P (V, T ) + 1
2
2/3
3 (Jp /J) Tr(Cp C) V0 G(V, T ) g 1
V
G(V, T )
+ (Jp /J)2/3 devg1 (be ) .
J
We note that a consequence of the dependence of G on V is an additional
hcontribution to the pressure.
i The additional pressure term is proportional to
3 (Jp /J)2/3 Tr(Cp 1 C) , i.e., the norm of the deviator of the elastic strain.
As is shown in a later section, the strain deviator is limited by the yield
condition, and the additional pressure term is O ([Y /G]2 ). Typically, Y is
1 percent of G, and the additional pressure term is neglected.
The simplest yield criterion is known as the von Mises yield condition.
It defined by the yield function
q
3
Y(Sp , Cp ) = dev 1 (Sp )
. (4.5)
2
Cp Cp
Substituting Eqs. (4.2) and (2.2), the yield function in the spatial frame can
be written as q
Y = 32 J kdevg1 ()kg . (4.5)
q
Remark 4.4 The factor of 32 in the yield function is standard and is
chosen to simplify formula involving uniaxial strain. For uniaxial strain with
the identity spatial metric, the deviator of the strain tensor has the form
dev(e) = 3el diag[ 2, 1, 1 ], and its norm is kdev(e)k2 = 23 2el . For
a simple isotropic material dev() = 2 G0 dev(e), and the yield condition
reduces to el 2 YG0 .
Remark 4.5 The plastic strain rate equation (4.9) can be expressed as
d
Cp FT g devg1 () g)F (Cp 1 C)1
dt
Aside from the factor Cp 1 C, which to O (Y /G0 ) is proportional to the iden-
tity matrix, this is the Prandle-Reuss form for the plastic strain rate.
d Y Y
3
Cp = 2
Cp devCp 1 (Sp ) Cp . (4.10)
dt Y
d Y Y
Cp = 3
2
G0 (Jp /J)2/3 Cp devI Cp 1 C . (4.11)
dt Y
It can easily be verified that Tr Cp 1 dtd Cp = 0, and hence the plastic strain
rate is volume preserving. Moreover, from Eqs. (4.11) and (4.4) the plastic
dissipation,
d Y Y
2
Sp : Cp = 3
G20 (Jp /J)4/3
devI Cp 1 C
2
dt Y I
is explicitly positive.
For small , if the yield function were substantially to exceed the yield
surface then the plastic strain rate would be very large and would rapidly
increase Cp in order to decrease the stress and drive the yield function back to
the yield surface. However, as the stress approaches the yield surface 0,
i.e., the time constant for relaxation increases and in fact becomes infinite
on the yield surface.
d 1 d Y d 2
= Y = ln Y
dt dt 2 dt
Y d
2 d
= ln Tr devI Cp 1 C ln(J)4/3
2 dt dt
1 d 1
Tr dev C C C C
Y
I p dt p
= 2
1
Tr devI Cp C
(4.12)
1
Tr devI Cp C Cp 1 dtd C
d
2
+ 2 3
ln(J)
dt
devI Cp 1 C
Tr
By Eq. (4.9)
d 1 d
Cp C = Cp 1 Cp Cp 1 C
dt dt
G0
= 32 (Jp /J)2/3 devI Cp 1 C Cp 1 C
Y
Then from Eq. (4.7), to leading order in Y /G0 , the first term in curly brackets
on the right hand side of Eq. (4.12) is
1
Tr devI Cp C d
C 1
dt p
C G0
2 = 32 1 + O (Y /G0 )
Tr
devI Cp 1 C Y
From Eqs. (2.14), (4.3) and (4.6) the second term in curly brackets on the
right hand side of Eq. (4.12) is
Tr devI Cp 1 C Cp 1 d
dt
C G0
1 T
2 = 3 (Jp /J)2/3 Tr FCp C devC1 (S) F d
Tr devI Cp C
1 Y 2
G0
= 3J devg1 () : d 1 + O (Y /G0 )
Y 2
From Lemma A.4 the third term in curly brackets on the right hand side of
Eq. (4.12) is
d ~ ~u .
23 ln J = 23
dt
Combining the three terms, Eq. (4.12) becomes
" #
d G0 G0 2 J
+ 32 = 32 devg (d) : devg1 () + O (Y /G0 ) .
dt Y
This form of equation is analyzed in Appendix D. In the limit 0,
Y = Y + O () and to leading order in Y /G0
2J
= devg (d) : devg1 () + O () . (4.13)
Y
Y 2 Y 2
Since 2Y
= , from Eq. (4.13) and (4.5) we obtain
3
2
J2 kdevg1 ()k2g = Y 2 + 4 J devg (d) : devg1 () + O 2
Consequently, in the limit of small , the stress can be split into an elastic
component and a viscous component
el + vis .
The elastic component el is given by Eq. (4.3) with the elastic strain such
that the elastic stress is restricted to the yield surface, i.e., rate independent
plasticity. The viscous component of the stress is given by
vis = 43 J1 g 1 devg (d) g 1 .
The viscous stress accounts for the leading order effect of plastic flow, namely,
the relaxation of the stress deviator to the yield surface.
The portion of the stress-strain work, Eq. (2.17), from the stress exceeding
the yield surface, i.e., dtd E d : el = d : vis , is dissipative. The viscous
dissipation corresponds to the plastic work from the stress exceeding the
yield surface. This is in additon to the plastic dissipation corresponding to
rate independent plasticity for which the stress is on the yield surface.
Remark 4.6 From the above derivation, the viscous approximation for
plastic flow is only valid when the viscous stress is small, i.e., k vis kg1 Y .
The visco-plastic constitutive model can be represented by a mechanical
system as shown in figure 1. The mechanical system is a composite made
up of three types of elements: (i) Spring with a linear stress-strain relation
= K, or more generally, dtd = K dtd . (ii) Dashpot with stress-strain
relation = dtd . (iii) Slider with stress-strain relation dtd = 0 if < Y .
More elaborate models allow for a history dependent relaxation to the yield
surface. The generalized models are represented by a sequence of springs
(Gi ) and dashpots (i ) elements in parallel. In effect, each spring represents
an additional internal degree of freedom. The sequence of springs and dash-
pots are an approximation to the Laplace transform of the history response
function. They allow empirical models to account for the observed frequency
dependent viscous response of materials such as polymers. The additional
internal degrees of freedom represent continuum variables needed to account
for aspects of the the short wavelength structure that are lost in coarse grain
or homogenized models of heterogeneous materials.
Remarks:
4.7 Plasticity is the underlying mechanism for viscosity in a solid. This
is very different than the viscous mechanisms in liquids and gases. Conse-
quently, the coefficient of viscosity for a solid can be very much larger than
for a liquid or gas; for metals viscous coefficients are typically in the range
of 100 to 1000 Poise, where as the viscous coefficient for liquids and gases is
typically measured in centi-poise.
4.8 Typically, the yield strength decreases with temperature. This is re-
ferred to as thermal softening, and can result in an instability leading to the
formation of shear layers. Morover, the limit Y 0 corresponds to visco-
elastic flow and can be used as a model for liquids. In this case, the material
does support shear strain, but only as a transient response. On a time scale
set by the ratio of the shear viscosity to the shear modulus, the shear strain
relaxes towards zero.
4.9 Viscosity and rate dependent plasticity lead to different shock wave
structures. With rate dependent plasticity, split waves occur consisting of an
elastic precursor followed by fully dispersed plastic wave. With viscosity the
shock can be partly dispersed but would not split.
4.10 For numerical simulations, rate dependent plasticity circumvents the
problems with shock waves due to the equations for rate independent plas-
ticity not being fully in conservation form. However, for a strong shock the
wave profile can be very narrow and cause problems with resolution. In ad-
dition, it should be noted that the small elastic strain approximation is not
valid in the profile of a strong shock.
4.11 The rate for crystal plasticity is determined by the motion of disloca-
tions, see for example [8], [9] and references therein. At low deformation rates
stress (d small) the plastic rate is dominated by thermal activation for the
movement of a pinned dislocation. This gives rise to the phenomena creep.
At high stress rates, in contrast, the plastic strain rate is dominated by the
drag on a dislocation as it moves through the crystal lattice. However, there
remains questions about the rate at which dislocations are nucleated and
grow. In addition, the initial number of dislocations is sensitive to impurities
and sample preparation.
4.12 The measured rise time for shock profiles in many metals can be fit
with a plastic strain rate in which (kdevg1 () Y kg )2 , see [25]. In
this case, the form in which rate independent approximation accounts for
the effect of plasticity would be different than for the Hohenemser-Prager
rate model.
4.13 The observed phenomenolgy of plastic flow is quite rich. The basic
difficulty with modeling the phenomena stems from the fact that materials
are not really homogeneous. As a result, internal variables (in addition to
the plastic strain tensor) are needed to characterize the state of the material,
along with corresponding rate equation for their dynamics. Many choices are
possible for internal variables. Common internal variables range from a sim-
ple isotropic work hardening parameter, to a back-stress tensor for kinematic
work hardening, to elaborate calibrations such as the mechanical threshold
stress (MTS) model of Follansbee & Kocks [3]. The difficulty in developing
plasticity models lies in the fact that the distribution of inhomogeneities (dis-
locations, cracks, grain boundaries), unlike the Maxwell distribution in gas
dynamics, is not necessarily highly peaked. Consequently, simple mean field
theory for coarse grained average quantities is not sufficient. As yet, there
is single no theory which can explain all the phenomena, let alone is agreed
upon by a majority of researchers.
5 Uniaxial Strain
We chose coordinate systems with identity metrics for the body, plastic
and spatial frames; i.e., G, = , , G
e
AB = AB , and gij = ij . Uniaxial
strain reduces the flow to a one-dimensional problem by restricting the total
deformation gradient to have the form
F = diag[ exp(), 1, 1 ] ,
C = diag[ exp(2 ), 1, 1 ] ,
Cp = exp 32 p diag[ exp(2p ), 1, 1 ] .
Since diagonal matrices commute, in this case the elastic left Cauchy-Green
tensor be has the same matrix elements as Cp 1 C (though they act on dif-
ferent spaces). To leading order, the elastic strain can be expressed as
Cp 1 C = J2/3 IT B + 32 e J2/3 diag[ 2, 1, 1 ] + O 2e .
Later we will have use of the trace and the deviator of the elastic strain.
These are given by
Tr Cp 1 C = exp( 23 e ) exp(2e ) + 2 J2/3
= 3 J2/3 + O 2e ,
exp(2e ) 1
devIT B (Cp 1 C) = exp( 23 e ) J2/3 diag[ 2, 1, 1 ]
3
= 23 e J2/3 diag[ 2, 1, 1 ] + O 2e .
5.1 Stress
Eshear = Ee Ehydro
1/3
= 12 V0 G0 I1 I3 3
h i
1
= V0 G0
2
exp( 23 e ) exp(2e ) + 2 3 .
Eshear = 32 V0 G0 2e .
The sameh expression is obtained when the shear energy for linear elasticity,
i2
V0 G0 Tr devI (Ee ) , is specialize to the case of uniaxial strain.
2
Remark 5.2 For uniaxial strain, the strain tensor, E = diag[ J 21 , 0, 0 ],
has only one non-zero diagonal element. The stress tensor is also diagonal
but all three diagonal elements are non-zero. This is in contrast to uniaxial
stress in which only one diagonal component of the stress is non-zero, but
the strain tensor would have all non-zero diagonal elements.
Y 2 = 32 J2 kdevI ()k2I
2
= 32 G20 J4/3
FCp 1 FT 13 Tr Cp 1 C I
I
= G20 [exp(2e ) 1] 2
exp( 43 e )
= 4 2e G20 + O 3e .
Hence,
Y = 2 |e | G0 + O 2e ,
Y
and on the yield surface |e | = 2 G0
.
The plastic strain rate direction for the flow rule associated with the von
Mises yield stress, Eq. (4.8), is
Y G0 2/3
= 3
2
J Cp devI Cp 1 C
Sp Y
1 G0 h i
= 2
exp(2e ) 1 exp( 23 e ) Cp diag[ 2, 1, 1 ] .
Y
This is compatible with the time derivative for the assumed form of Cp
d
Cp = 23 dtd p Cp diag[ 2, 1, 1 ] .
dt
Therefore, the plastic strain rate reduces to a scalar equation. In the plastic
regime, Y > Y ,
d Y Y G0 h
3
i
dt p
= 4
exp(2e ) 1 exp( 23 e )
Y
3 G0 sgn(e )
h i h i2
Y Y Y
=2 |e | 2 G0 exp( 3 G0 ) + O |e | 2 G0
,
where sgn(e ) = 1 or 1 when e > 0 or e < 0, respectively.
The fully non-linear elastic-plastic flow equations for uniaxial strain with
isotropic constitutive relations are summarized below
+ u =0
t x
2 xx
u + u =0
t x
(5.1)
xx
E + E u u =0
t x
p + u p = Rp
t x
xx = P + 0
P = Phydro V, e Eshear
(5.2)
0 = 2 G0
3J
[exp(2e ) 1] exp( 32 e )
n h i o
Eshear = 12 V0 G0 exp( 23 e ) exp(2e ) + 2 3
4 e G0
0 = 3
+ O 2e ,
J
Eshear = 23 V0 G0 2e + O 3e .
The yield function and rate for the plastic strain variable are given by
Y = G0 |exp(2e ) 1| exp( 23 e )
(Y Y )+ G0 h i (5.3)
Rp = 3
4
exp(2e ) 1 exp( 23 e )
Y
where the subscript + denotes the positive part, i.e., f+ = f for f > 0 and
0 otherwise. To leading order the yield function and plastic strain rates are
given by
Y = 2 |e | G0 + O 2e ,
G0 sgn(e ) h i h i2
Rp = 3
2
|e | Y
2 G0 +
exp( 3 YG0 ) +O |e | Y
2 G0
.
Remarks:
5.3 For uniaxial elastic flow, the Hugoniot jump conditions, Eq. (2.26),
reduce to the standard shock equations for a fluid with P replaced by xx .
Moreover, the differential thermodynamic relation, Eq. (2.1), reduces to
de = Sxx dExx + T d = xx dV + T d .
6 Wave Structure
Weak shock waves are in the purely elastic regime, i.e., the plastic strain
rate is zero. The shock profile is discontinuous and corresponds to an ordinary
fluid dynamic shock wave.
As the wave strength increases above the Hugoniot elastic limit shown
in figure 2, the shock speed decreases and the shock wave is unstable. It
breaks up into a split wave. The lead wave is purely elastic and takes the
material upto the yield surface. Plastic flow occurs in the following wave.
The limiting of the shear stress above the yield surface results in a discon-
tinuous decrease in the longitudinal sound speed. Consequently, the plastic
wave speed is less than the speed of the lead wave, which is at the Hugo-
niot elastic limit. We show in a subsequent subsection that the plastic wave
is fully dispersed. That is to say, the plastic flow provides all of the dissi-
pation required by the Hugoniot jump conditions and the plastic profile is
continuous.
As the wave strength is further increased, the speed of the plastic wave
increases until it overtakes the elastic wave. The split waves are then re-
placed by strong plastic waves. The profile of a strong plastic wave is partly
dispersed and consists of a lead shock followed by a continuous relaxation
region to the state on the Hugoniot locus. As seen in figure 5, across the
lead shock the plastic strain is continuous but the elastic strain is discontin-
uous. The plastic flow provides only part of the dissipation required by the
Hugoniot jump conditions.
An intesting special case occurs when the yield strength is zero. This
corresponds to a visco-elastic material. In this case there are no split waves.
Weak waves are fully dispersed as in the plastic wave of a split wave, and
strong waves are partly dispersed as in strong plastic waves. The fully dis-
persed waves have a profile simpilar to what occurs in fluid dynamics when
a viscous stress, Q = x u, is added to the pressure. However, there is an
important difference for the numerics. The visco-elastic model is hyperbolic
and the stable time step is the minimum of the CFLcondition and the time
x
constant of the plastic strain rate, i.e., t < min c
, . In contrast a
viscous shock profile arises from a parabolic model and the stable time step
2
for an explicit algorithm is t < (x)
. As the resolution is increase the vis-
cous time step is proportional to (x)2 , whereas the visco-elastic time step
is proportional to x. Consequently, it is computational less expensive to
resolve shock profiles with a visco-elastic model.
Figure 7: Hugoniot loci in (V, P )-plane. Blue curve is frozen locus, red curve is
equilibrium locus. Dotted and dashed black lines are the Rayleigh line for strong
plastic wave and second shock of split wave, respectively.
The two types of plastic wave profiles, fully- and partly-dispersed, can be
understood by examining the frozen and equilibrium Hugoniot loci shown in
figure 7. For the frozen locus the plastic strain is held fixed, while for the
equilibrium locus the plastic strain is determined by the condition that the
stress lies on the yield surface. Also shown in the figure are the Rayleigh
lines for weak and strong plastic waves. The Rayleigh line is the cord in
the (V, xx )-plane from the initial to final shock states. It follows from the
jump conditions that its slope is proportional to the square of the wave speed.
Moreover, for a steady traveling wave, the projection of the wave profile in
the (V, xx )-plane lies on the Rayleigh line.
For a weak plastic wave, the Rayleigh line (black dotted line in figure 7)
lies between the equilibrium and frozen Hugoniot loci. Consequently, a shock
is not possible and the wave profile must be continuous. In contrast, for a
strong plastic wave, the Rayleigh line (black dashed line in figure 7) intersects
the frozen Hugoniot locus. An analysis of the ODEs for the steady wave
profiles shows that the plastic flow doesnt provide enough dissipation at the
initial state for the profile to follow the Rayleigh line. Instead, the profile
consists of a shock to the point at which the Rayleigh line intersects the
frozen Hugoniot locus followed by a continuous variation along the segment
of the Rayleigh line to the final state on the equilibrium Hugoniot locus.
For the visco-elastic model, the elastic segment on which the frozen and
equilibrium loci overlap shrinks to zero. A fully dispersed wave profile occurs
when the wave speed lies between the equilibrium and frozen sound speeds.
A partly dispersed profile occurs for stronger waves with a wave speed greater
than the frozen sound speed.
model and is fully resolve. The anomaly is in fact part of the solution to the
PDEs. It would not shrink in spatial extent with a finer grid.
Figure 8: Wave profiles of stress, velocity and entropy for impulsively started
piston driven visco-elastic wave.
Figure 10: Wave profiles of stress, velocity and entropy for visco-elastic wave after
it reflects from wall at x = 5.
5 4 3
i.e., the mapping from a Voigt index to a tensor index is given by V (1) =
1, 1; V (2) = 2, 2; V (3) = 3, 3; V (4) = 2, 3; V (5) = 3, 1; V (6) = 1, 2. The Voigt
vector corresponding to the strain tensor is defined as
~E = (E11 , E22 , E33 , 2 E23 , 2 E31 , 2 E12 )T ,
This non-symmetric treatment of stress and strain has side affects on the
compliance matrix. It is not suited to the spectral decomposition of the
elastic tensor. Also, it affects the coefficients in the formulae for the Reuss
average shear modulus.
V (a),V (b)
The stiffness tensor can be expressed as a 6 6 matrix C b
a,b = C
where a, b = 1, 2, 3, 4, 5, 6. Similarly, for the compliance tensor Sa,b = SV (a),V (b) .
b
Both C
b and S b are symmetric matrices. From the symmetry, we obtain
! !
~S = I3 0 b I3 0
C ~E ,
0 2 I3 0 2 I3
It follows that ! !
b = I3
S
0 b 1 I3
C
0
,
1 1
0 I
2 3
0 I
2 3
E = 12 E : C : E = 12 Tr(SE) .
S8 S7 S3
Then symmetric rank-two tensors are mapped into Voigt vectors by the 6 9
transformation matrix
!
I3 0 0
U= 1 I3 1 I3
.
0 2 2
S = UT CU
b E ,
b
6
X
C
b = i~ei ~ei .
b
i=1
Then the spectral decomposition of the rank 4 stiffness tensor, see for example
[27], is given by
6
X
C= i ei ei ,
i=1
1
where sym(m n) = 2
m n + n m and we have used the fact that
h i
|| sym(m n)||2 = 1
2
1 + (m n)2 1 .
1
0 E = (Tr E)2 + G||E||2
2
1
= K(Tr E)2 + G|| dev E||2 ,
2
Then the stress-strain relations have the simple form (which is a special
case of the spectral decomposition with eigenvalue 3 K and fivefold degener-
ate eigenvalue 2 G)
,
C = + G( + ) ,
1 1
1 (B.1)
S, = + ( + ) .
9K 6G 4G
Here for simplicity we have assumed that the metric is the identity matrix,
i.e., G = . The corresponding Voigt matrices have the form
! !
b = Cu 0
C and b = Su 0
S ,
0 C` 0 S`
where the u block represents the indices 1, 2, 3 and the ` block represents
the indices 4, 5, 6. The components of the stiffness matrix are given by
K + 43 G K 32 G K 32 G
2 4 2
K 3 G K + 3 G K 3 G
Cu = and C` = G I3 ,
K 23 G K 23 G K + 34 G
1 1 1 1 1 1
9K
+ 3G 9K 6G 9K 6G
1 1 1 1 1 1
1
Su =
9K 6G 9K
+ 3G 9K
6G
and S` = I3 .
1 1 1 1 1 1
4G
9K
6G 9K
6G 9K
+ 3G
C = N C N = G + ( + G)N N .
In the ambient equilibrium state the acoustic wave speeds, c2 , and the
direction of the displacement vectors are the eigenvalues and eigenfunctions
of C. It is easy to verify that N is an eigenvector with eigenvalue + 2 G =
K + 34 G. This corresponds to a longitudinal sound wave. Furthermore, any
T orthogonal to N is an eigenvector with eigenvalue G. These correspond to
transverse shear waves.
As expected for an isotropic material, the wave speeds are independent
of the direction of propagation. Moreover, in the plane orthogonal to the
propagation direction, there is no preferred direction for the transverse wave.
Consequently, the eigenvalue for a shear wave is doubly degenerate. We also
note that the longitudinal sound speed is greater than the transverse shear
wave speed.
,
Z
d 0 0 ,0 0
hC iV = R 0 R 0 C (RT )0 (RT )0 ,
4
where R is the rotation matrix corresponding to the solid angle . Since the
measure is invariant under the rotation group (Haar measure on a compact
where
K
(C ), =
G 2
(C ), = + + .
3
It follows that the Voigt average of the bulk and shear moduli are given by
1
KV = 9
C
b +C
11
b +C
22
b +2 C
33
b +C
12
b +C
23
b
31
1 b C
GV = 15
C
b +C
11
b +C
22 33
b +C
12
b +C
23
b
31 + 3 C44 + C55 + C66
b b b .
Remarks:
1
B.3 Typically, the Reuss average is written in terms of C
b rather than S.
b
With this notation the last 3 terms in the formula for GR are multiplied by
a factor of 43 instead of 3.
B.4 The sound speeds computed from the Voigt and Reuss averages, c2long =
K + 43 G and c2trans = G, are upper and lower bounds for average material
but not necessarily on the sound speed for a given propagation direction.
The spectral decomposition provides bounds on the latter.
E 2
where > 0 and again C is used for the metric. When CC is replaced by
the acoustic tensor A this is the condition of strong ellipticity. Then can
be identified with the minimum of c2 over all directions and polarizations.
Consequently, for Eq. (C.4), is the minimum over all directions N of 41 (c2
V0 Sn ) where Sn = N S N is the normal stress. Strong ellipticity is a
condition needed for existence of solutions in elasto-statics, for example, see
[11]. It is also sufficient for hyperbolicity.
~ ~ (X,
U ~ t + t) U
~ (X,
~ t)
U = lim ,
t t0 t
T(X,t+t)
~ S ............................................................. T(X,t)
~ S
~ ~ t) (X,
(X, ~ t + t)U
~ (X,
~ t + t) U
~ (X,
~ t)
U = lim ,
t t0 t
is well defined. On a flat space, metric gij = ij , this gives the standard
definition for the convective derivative, dtd = t
+ ~u .
This idea can be generalized to be independent of the reference frame B.
First, for a vector field ~u(~x) T~x S we define a flow, ~u , by the differential
equation
(~x, t) = ~u ~u (~x, t)
t ~u
with the initial condition ~u (~x, 0) = ~x. Then the push-forward maps
the tangent space at ~x onto the tangent space at (~x, t), and the pull-back
maps the tangent space at (~x, t) onto the tangent space at ~x. The Lie
derivative of a tensor field, see e.g., [11, 1.6], such as the Eulerian strain,
is given by
d
L~u e = ( ~u ) (~x, t) ( ~u ) e(~x, t) .
dt
In effect, one transforms into a common vector space in order to take the
derivative and then transforms back to the desired space.
The Lie derivative allows evolution equations to be defined in a mannar
such that frame indifference is satisfied, see [11, Box 6.1, p. 99]. In particular,
we have defined the plastic strain rate in the Lagrangian frame. It can then
be transformed into the Eulerian frame. The theory is physically the same
when viewed in any inertial frame.
Acknowledgement
The author wishes to thank his colleauge Bradley Plohr. Discussion with
him on elastic-plastic flow have greatly contributed to the authors under-
standing and is the basis for these notes.
References
[1] J. K. Dienes, On the analysis of rotation and stress rate in deforming
bodies, Acta Mechanica 32 (1979), 217232. 16, 25
[2] T. C. Doyle and J. L. Ericksen, Nonlinear elasticity, Advances in Me-
chanics IV, Academic Press, 1956. 12
[3] P. S. Follansbee and U. F. Kocks, A constitutive description of the de-
formation of copper based on the use of the mechanical threshold stress
as an internal state variable, Acta Metall. 36 (1988), 8193. 50
[4] L. F. Henderson and R. Menikoff, Triple shock entropy theorem and its
consequences, J Fluid Mech 366 (1998), 179210. 66
[5] W. Herrmann, Elastic-plastic constitutive relastions at large strain,
Shock Compression of Condensed Matter 1991 (S. C. Schmidt, R. D.
[6] R. Hill, The elastic behaviour of a crystalline aggregate, Proc. Roy. Soc.
London A65 (1952), 349354. 74
[12] R. Menikoff, Errors when shock waves interact due to numerical shock
width, SIAM J Sci Comp 15 (1994), 12271242. 64
[14] W. F. Noh, Errors for calculations of strong shocks using artificial vis-
cosity and an artificial heat flux, J Comp Phys 72 (1987), 78120. 66
Index
:, contraction, 12 J, ratio of specific volumes VV0 , 9
A, acoustic tensor, 26 K, bulk modulus, 15
b, left Cauchy-Green tensor, 6 , Lame coefficient, 14
be , elastic tensor, 7 , coef. of dynamic viscosity, 45
B, body frame, 2 kk, norm, 11
T B, tangent space, 2 P , pressure, 13
T B, cotangent space, 2 P, first Piola-Kirchhoff stress, 19
C, right Cauchy-Green tensor, 6 P, plastic frame, 2
Cp , plastic tensor, 7 T P, tangent space, 2
T P, cotangent space, 2
C , stiffness tensor, body frame, 24 , mapping, B S, 1
c, stiffness tensor, spatial frame, 24 , mass density, 17
d, rate of deformation, 22
S, compliance tensor, 25
, Kronecker symbol, 4
S, spatial frame, 2
dev, deviator, 11
T S, tangent space, 2
E, specific energy, 12, 33
T S, cotangent space, 2
Ee , elastic energy, 33
S, second Piola-Kirchhoff stress, 12
Ep , plastic energy, 33
Sp , plastic stress, 12
e, Eulerian strain tensor, 7
, Cauchy stress tensor, 12
ee , elastic tensor, 7
T , temperature, 12
ep , plastic tensor, 8
Tr (), trace, 10
E, Lagrangian strain tensor, 7
~u, particle velocity, spatial frame,
Ep , plastic tensor, 7
1
, small strain parameter, |V VV 0|
,9 ~
0 U , particle velocity, body frame, 1
, entropy, 12
V , specific volume, 9
F, deformation gradient , 2 V0 , value in body frame, 9
Fe , elastic deformation, 2
Fp , plastic deformation, 2
acoustic tensor, A , 26
f = F 1 , 21
adjoint, 6
G, shear modulus, 15
Almansi strain tensor, e, 7
g, spatial frame metric, 4
associated flow rule, 37
G, body frame metric, 4
G,
e plastic frame metric, 7
bulk modulus, K, 15
a
dc , Christoffel symbol, 17
I, identity operator, 5 Cauchy-Green tensor
left, b, 6 invariants
right, C, 6 tensor, 13
characteristic polynomial, 13 inverse, 5
Christoffel symbol, 17 isotropic material, 13
Clausius-Duhem inequality, 23
compatibility condition, 20 Lame coefficient, 14
Lie derivative, 23, 79
compliance tensor, S , 25 linear elasticity, 14, 16
contravariant index, 3 loss of hyperbolicity, 30
convective derivative, d/dt, 19
cotangent space, 3 mapping, , 1
covariant differentiation, 17 metric, 4
covariant index, 3 body frame, G, 4
crystal plasticity, 3840, 50 plastic frame, G,
e 7
curvature tensor, 18 spatial frame, g, 4
multiplicative decomposition, 2
deformation gradient, F, 2
deviator, 11 norm, kk, 11
dilatancy, 30
particle velocity
Elasticity, 11 body frame, U ~, 1
hyper, 11 spatial frame, ~u, 1
hypo, 11 Piola identity, 19
linear, 14 Piola transform, 19
entropy inequality, 23 Piola-Kirchhoff tensor
first, P, 19
Finger tensor, b1 , 7 second, S, 12
flow equations Prandle-Reuss plastic strain rate,
Eulerian, 17 45
Lagrangian, 19 pressure, P , 13
frame indifference, 11
rank-1 convexity, 76
Hohenemser-Prager rate, 45 rate independent plasticity, 35, 38
Hugoniot elastic limit, 58, 60 rate of deformation tensor d, 22
Hugoniot equation, 32 Rayleigh line, 58, 63
hyper-elastic, 11 Reuss average, 74
hyperbolicity condition, 29
hypo-elastic, 11, 38 shear modulus, G, 15
small strain, 16
inner product, 4 sound speed
isotropic, 29
specific energy, E, 12
spectral decomposition, 70
stiffness tensor, C , 24
strain tensor
Eulerian, e, 7
Lagrangian, E, 7
stress tensor
Cauchy, , 12
linear elastic, 16
Piola-Kirchhoff, S, 12
viscous, vis , 48
strong ellipticity, 29, 77
subscript, covariant index, 3
summation convention, 3
super-position of strains, 8, 34
superscript, contravariant index, 3
tangent space, 3
texture, 16
thermodynamic identity, 12
trace, 10
transpose, 5
Tresca yield condition, 38
uniaxial strain, 51
uniaxial stress, 53
viscous stress, 48
Voigt average, 74
Voigt notation, 68
von Mises yield condition, 43
yield condition
von Mises, 43
yield function, 34
yield strength, 34