On Numerically Accurate Finite Element Solutions
On Numerically Accurate Finite Element Solutions
BROWN UNIVERSITY
PROVIDENCE, R. I.
I) O SPERMUS '-4 00
ON NUMERICALLY ACCURATE
FINITE ELEMENT SOLUTIONS
IN THE FULLY PLASTIC RANGE ~ U
J. C. NAGTEGAAL, D. M. PARKS,
AND J. R. RICE
Reproduced by l r
NATIONAL TECHNICAL
INFORMATION SERVICE
US Department of Commerce v E tne
Springfield, VA. 22151 ; 0
Technical Report
NASA NGL 40-002-080/14 to the
National Aeronautics and Space Administration
March 1974
NOTICE
by
Abstract
elastic-plastic materials exhibit much too stiff a response in the fully plastic
range. This is most striking for the perfectly plastic material idealization,
in which case a limit load exists within conventional small displacement gradient
assumptions. However, finite element solutions often exceed the limit load by
substantial amounts, and in some cases have no limit load at all. It is shown
cal two and three-dimensional finite elements are highly constrained at or near
A general criterion for testing a mesh with topologically similar repeat units
is given, and the analysis shows that only a few conventional element types and
arrangements are,or can be made suitable for computations in the fully plastic
range. Further, a new variational principle, which can easily and simply be
accurate computations to be made even for element designs that would not normally
be suitable. Numerical results are given for three plane strain problems, namely
pure bending of a beam, a thick-walled tube under pressure, and a deep double
edge cracked tensile specimen. These illustrate the effects of various element
designs and of the new variational procedure. An appendix extends the discussion
1. Introduction
In recent years, the finite element method has been employed for analysis
and three-dimensional problems, inaccurate results are often obtained. The problem
limit load exists, which can sometimes be calculated exactly or bounded from above
exhibit
by the kinematical theorem, while the finite element solution often seems to
no limit load at all, but rather a steadily rising load-displacement curve attaining
values far in excess of the true limit load. Similarly, for strain-hardening
materials having typical values for the plastic tangent modulus of two or more
orders of magnitude less than the elastic modulus, the finite element solution
Later we shall see examples of this for a beam in pure bending and for a
punch problem. In the case of a plane strain extrusion problem [3], when a
-3
bilinear constitutive law with slight plastic work-hardening (da/de-- H = 4.4x10-3 E;
p
E , the Young's modulus) was incorporated, a finite element extrusion pressure of
1.5 times a slip-line solution (using the same yield stress as the linearly har-
dening model) was attained with overall billet displacements only of order six
times the displacement obtained at the slip line limit load. This indicates that
the computed stiffness in the fully plastic range far exceeds what would be ex-
illustrates some of the peculiarities which have been encountered by the authors
-2-
containing cracks. The finite element meshes shown in fig. 2 were used to
solve the single-edge notched plane strain tensile specimen subject to both
uniform end displacement [4] (i.e., no end rotation) and the static equivalent
of a mid-ligament concentrated force. These two loading conditions have the same
analytical limit load, namely, twice the yield stress in shear on the net section.
As can be seen in fig. 1, the overall load-deflection curves of the two loading
conditions are quite different in the fully plastic regime. The solution for
the mid-ligament loading clearly indicates the correct limit load, while the
constant end displacement solution exceeds limit load, with the load continuing
A cause of these problems relates to the fact that the deformation state
for the usual material idealization, deformation increments at limit load will
to be satisfied.
f dS
S elem.. elem. ij13ijdVelem
precisely, where .ij is the stress rate following from the prescribed con-
stitutive law in terms of the current stress a.. and strain rate s.. within
* * 1 (1.2)
=
1j ij 3 13 kk
be expressed as
where
K = E/3(1-2v) (1.4)
is the elastic bulk modulus and skk is the dilatational strain increment.
T.u.dS = +
K(kk)2] dV elem
s..e.. (1.5)
S elem. Velem.
where e.. is the deviatoric strain increment. In the vicinity of the limit
load, the term s..e.. will tend to vanish pointwise, but, as follows from
plastic normality, will never be negative. Hence, from (1.5), this implies that
In order for a limit load to exist for the discretized finite element
In the next section it will be shown that, except for plane stress
formations of which typical finite element grids are capable. When they are
or no limit loads at all, will result. On the other hand, even if no limit
is, indeed, observed) that the slope of the load-deflection curve is reduced
substantially from its initial elastic value at load levels near the theo-
retical limit load. This, however, does not allow an accurate inference of
coefficient, the terminal slope of the load-deflection curve will not necessarily
limit conditions are approached because the effective shear modulus then tends
to vanish, whereas fluids and rubber-like solids behave incompressibly from the
start because the bulk modulus is very large. The essential problem is the
same in both cases, however, in the sense that the incompressibility require-
2. Analysis
In the previous section, it has been observed that elements must be capable
of elements.
ments shown in fig. 3. Within each element, the displacement increments are
of the form
where the vectors {a), {b}, etc., are expressed in terms of the nodal veloci-
ties and coordinates. Now, the incompressibility constraint for plane strain has
the form
S +E _ x - : 0 (2.2)
xx yy axx y
The fact that {d} = {O) means that at limit load each element has strain
increments which are constant throughout the element. It is then evident from
displacement continuity that all elements marked * in fig. 3 must have the
same value of Exx , whereas all elements marked t must have the same value
and column, E and must be the same in every element of the grid.
by- c8 - by + Cy8 = 0 ,
xy x y yx y x
b6 - d -b6 + dy = 0 , (2.5)
xy xy yx yx
dxy - Cxy - dyYx + Cy6 = 0
bx = ABx + B , b C8x - AS
dx = A x + B6 , d = C6x - A6 .
F
-7-
U = D + Ax + By
(2.7)
u = E + Cx - Ay
This has catastrophic effects for an arbitrary grid. Consider, for instance,
the grid in fig. 4, and specify the displacement increments of the three nodes
the extensional strain along AB in element I , and because the element must
deform with a constant strain of zero dilatation, we must then specify the com-
the extensional strain along BC in element II , we must also specify the dis-
F are then also determined. Continuation of the argument then furnishes, that
the strain increment will be constant throughout the grid. It should be noted
that this argument holds only if the element boundaries do not form a straight
line through the body; over such a line the shear strain increment can be dis-
continuous.
Examples of the unreasonable constraints enforced by pointwise incompress-
a result of perfectly-plastic limit analysis: namely, that while the limit load
is unique, the deformation field at limit load need not be unique. An acceptable
limit field for both loading conditions, giving the correct limit load, is that
of concentrated deformation on 450 lines from the crack tip to the surface, as
shown in fig. 5a. An alternative limit field for the mid-ligament loading is
shown in fig. 5b. This field consists of constant strain increments within the
region bounded by the two 450 lines from the crack tip. The rest of the body is
rigid, so the overall effect is that of a rotation of the two rigid regions about
the crack tip. In fact, this was exactly the velocity field obtained at limit
Since the deforming regions do so uniformly, and the 450 lines coincide with
with the constant end-displacement boundary condition of the other loading. The
same constant strain pattern cannot develop, and an ever-worsening overestimate
Although the problem has now been analyzed sufficiently for four-noded iso-
of the solution if the mesh is refined. Refinement of the mesh will have two
-9-
opposing effects:
1) It will increase the number of nodes, and since each node represents
of degrees of freedom.
2) On the other hand, it will increase the number of elements, and since
It is clear that convergence will occur only if the total number of degrees of
freedom increases faster than the total number of constraints. This furnishes a
adequate to furnish accurate limit loads and flow fields. It should be noted
that the total number of constraints is not always equal to the number of ele-
ments times the number of constraints per element. Sometimes elements can be
example of which will be discussed in the next section. For the moment, however,
to the total number of elements in the grid when it is refined. Consider a body,
the inner angle formed at the node by the two adjacent element boundaries. For
each type of element, the sum of these nodal angles is a given number. Consider,
for instance, the six-noded triangle of fig. 6. Each of the nodal angles at the
midpoints of the sides is equal to 7 radians, whereas the sum of the angles
8a 82 and 83 is also equal to r radians. Hence the sum of all nodal angles
types of elements. For generality, let us assume that the sum of the
-10-
consists of p elements, then the sum of all nodal angles of all elements is
equal to
e = pe . (2.8)
The sum of all nodal angles can also be calculated in a different way.
Suppose the mesh has k nodes inside the body and Z nodes on the boundary.
The sum of the nodal angles around each interior node is equal to 2n radians,
whereas the sum of the nodal angles at a boundary node is equal to 7-y radians,
where y is a small angle due to the curvature of the boundary. In this way,
pn = 2k + Z + 2c - , (2.10)
or, alternatively,
P 2 £ 2c-4
k n nk nk (2.11)
Now if the mesh is refined, p, k and t will all increase; hence the last
term will vanish. Moreover, if the mesh is refined uniformly, k (the number
of interior nodes) will increase approximately as the square of £ (the number
lim ( )
2
k+
k4m k n (2.12)
-11-
The sum of the nodal angles in one element is equal to 2w radians, and hence
nodes will become equal to the number of elements. Since each node represents
upon each element, the total number of degrees of freedom is only 2/3 of the
total number of constraints. Hence convergence will not in general occur, which
reaffirms the previously obtained result. Similar derivations can be made for
other planar elements as well as for axisymmetric elements, since (2.12) holds
as well for this class. The results are displayed in Table I. The range of value
of constraints per element from 6 to 8 given for the 8-noded quadrilateral depends
is not unique, since the sum of the solid angles of a polyhedron is not a con-
stant. Therefore one has to determine this ratio for a specific arrangement of
elements. For some common regular grids, and thus of any grid which is equi-
valent, the results are displayed in Table II. It may be assumed that these
The results in Tables I and II indicate that none of the usual finite
elements, except the 6-noded plane strain triangle and, to a much lesser degree,
and thus limit load, behavior. Before going into a more systematic approach to
solve the problem, let us consider what can be obtained by a special arrange-
one constraint. Thus this method will work only for elements for which the ratio
Let the coordinates of the node i be x. , and let the node have a dis-
placement increment . The area A of element I before the displacement
increment has taken place is then equal to
A )
x-( -x)] k , (3.1)
-13-
where x and * denote the usual vector and scalar products respectively, and
k is the unit vector perpendicular to the plane of the elements. Similarly one
When one neglects the second order terms, the incompressibility constraint
(x -x ) x (u -u = (x 1 -x ) x (u -u ) (3.4)
2 -o 1 ~o 1 -o 2 -o
(x -x ) x (u -u ) (x 2 -x ) x (u -u ) , (3.5)
3 ~o -2 -o ~o 3 -o
(x -x ) x (u u ) (X -x x (u -u ) (3.6)
x -x-
- ( )
X-xxx ( - ) x (u -ux (3.7)
- -o 3 O -2 -o
Ik7oI -2 'o
-14-
for (3.6)
Ix -x
-(x -x )x(u-u ) (x -x )x(u -u (3.8)
2 o 3 o 3 o ~4 o
x xx -x
2
(3-
X-x )X( - ) =- (x'-x)x(u4 -uo) ' (3.9)
-l 1'4-ool
or, alternatively,
(x
-2-x
-oo)x( ~ -u~o ) = (x -X)x( -u ) . (3.10)
3 ~o 4 "o
1 2 -xo II x,3
3-x
,o
I
With the same argument as was used for (3.7) and (3.8) this yields
finite element programs. The groups of four elements have all the generality in
carried out without changes to an already existing program. -The principal short-
coming of the method is that it can be applied only to problems of plane strain,
for which another alternative (the 6-noded triangular element) already exists.
Even this alternative can be slightly improved by arranging the 6-noded elements
similar to the 3-noded ones. The ratio of degrees of freedom to constraints then
-25-
It should also be noted that the degrees of freedom available need not
actually be incorporated into the finite element solution. For example the
example, [7])
As yet no method has been devised to arrange the lO0-noded tetrahedra such
It has been discussed before that an absolute requirement for the existence
has also been demonstrated that only a few of the conventional element types are,
goal can be obtained by taking care that the dilatation is governed by fewer pa-
rameters than in the conventional elements. Clearly, this procedure may be applied
only to higher order elements for which the conventional number of constraints per
functions that have this property; in fact, if the arrangement in fig. 7 is taken
as a single element, and the displacement of node 0 is chosen such that the di-
latation in all subelements I through IV is the same, one has obtained such a
function. This special field must, however, be considered as a "lucky guess", and
a general approach to construct similarly simple fields for other elements does
principle in which the dilatational strain increment and the displacement incre-
admits the displacements and the stresses as independent variables. The validity
of the present principle will, however, be proved here independently of the H-R
principle.
-17-
proved in the usual way. By taking the first variations in u. and * one finds
and since
a + O
s ij K 6 ,j (4.3)
The first two terms express the virtual work principle, and
hence imply that
aij satisfy the continuing equilibrium equations in V and force rate boundary
conditions on ST , in the usual manner. The last term provides the additional
result that * = ;,k
-18-
H[u,=hJ [ + 2
vhekk - v(l-2v)h JdV - j Ti dS ' (4.5)
T
where p is the shear modulus, v is Poisson's ratio, and where the "mean
pressure function: is given by
akk
2p(l+v) (4.6)
Apart from its specialization to linear elasticity, the fact that ;kk enters
Herrmann's principle quadratically is an essential difference. Indeed, this does
not allow for it the same simple procedures that can be based on the
present
principle, and that allow the easy adaption of existing, conventional
stiffness
programs to it.
in the last column of Table I and a "three" in the last column of Table II. This
readily leads to a desired number of constraints per element. For instance, for
be one, and for the 8-noded quadrilateral elements, it would be three. Similarly,
for the 8-noded cube in Table II, one constraint would be desired. Similar desired
Now let us, for the moment, restrict the discussion to those elements for
which the desired number of constraints is equal to one. As has been discussed
It is, on the other hand, necessary for convergence that at least a constant
S= = constant . (4.7)
I = + K 1 2
I
a=1 V
[W'(e) dV-
S T.u.dS (4.8)
In practically all cases, the bulk modulus r will be constant within an element;
V- 1= ,kdV . (4 10)
ta a ukka
aV
W'e ) +
t (V- u kkdVa )2 dV - u dS . (4.11)
a=1 Vk a S
a a T
i - e..ij +6
p
SW(j..)dV I T.u.dS , (4.13)
U=1 V ST
a T
where W represents the total rate potential (i.e., 6W(C) = ij..6..) , but
evaluated for the modified strain increment i.. . It should be observed that the
hence no additional variables (i.e., mean pressure) have to be used. It should also
be noted that the functional is identical to the usual rate potential functional,
with the exception of the relation between strain and displacement increments. As
was mentioned before, this makes it extremely simple to adapt an already existing
program for this method. One need only rewrite the matrix -convertingnodal param-
similar to the present one, in both cases implemented by changing the strain-
-21-
displacement relationship.
For element types for which the desired numer of constraints is larger than
one, the situation is only slightly more complicated. Consider for instance the
8-noded quadrilateral plane strain element. For this element type, the desired
0 = V-1 n ,kdV
a a V
where Ix a and I are the moments of inertia around the x and yo axes.
A similar procedure can be followed for the 8-noded quadrilateral axisymmetric
element and the 20-noded cube. In each case c.. is defined analogously to (4.12)
elements the desired number of constraints does not make such a simple choice
possible. If we did not already know that the six-noded plane strain triangle
-22-
was suitable for incompressible analysis, we would conclude, for example, that
the desired number of constraints is two, which does not correspond to a complete,
occur in the fully plastic range, where the deformation gradients may no longer be
infinitesimal, a rigorous finite strain finite element formulation may in some cases
a formulation is given in Appendix II. It may also be noted that the present varia-
tional principle, eq. (4.1), may be put in a mean stress form by letting K4 = q ,
say, be the variable, and this is suitable for strictly incompressible materials
(K w)
-23-
5. Numerical Examples
in specific numerical solutions, plane strain beam bending, the thick-walled plane
strain tube under internal pressure, and the tensile analogue of the Prandtl punch
For the first type of problem, the kinematic requirement of pure bending that
plane sections remain plane was enforced, so it was necessary to examine only one
row of elements across the beam thickness. Ten square 4-noded isoparametric ele-
ments were placed across the thickness. Displacement boundary conditions were
imposed, and the total moment was computed from the calculated nodal tractions.
0.3. The problem was also solved by applying identical boundary conditions to
The results are shown in fig. 8. The limit moment for plane strain, assuming
limit 2 (5.1)
where a is the tensile yield stress and h is the beam thickness. The beam
As can be seen, the ordinary isoparametric element does not find a limit load
at all; all elements across the thickness reach yield, and the moment-curvature
curve continues to rise. The steepest curve shown in fig. 8 is the case for which
-24-
one stress state per element, calculated at the element centroid, is used to
represent the stress state throughout the element. The next steepest curve
The third steepest curve is the result of using the constant dilatation
quadrilateral element with one stress state per element. Finally, the two
essentially coincident curves which actually find the correct limit load are
the results of the crossed triangles and the constant dilatation quadrilaterals,
calculating and storing the stress state within each element at each Gaussian
integration station.
is a necessary but not sufficient condition for a limit load to exist in the
finite element model. In fact, it is also necessary that, at limit load, all
to vary within the element, be pointwise normal to a single stress state on the
yield surface. When the stress state was calculated and stored at the Gaussian
vanish in the fully plastic range for both the regular and constant dilatation iso-
parametrics. In fact, the difference between the terminal slopes of the moment-
curvature relations for one vs. four stress measures per element was identical for
Thus, for this particular problem, the relative magnitudes of the errors in
the terminal load-deflection curve associated with failure to meet the necessary
-25-
of deviatoric
conditions of pointwise incompressibility and pointwise normality
The same problem was also solved for the case of linearly hardening material
In this
behavior, with a hardening modulus of one-hundredth the elastic modulus.
calculable. The results are shown in fig. 9, where it is seen that the ordinary
fully
isoparametric element formulation exhibits much too stiff a response in the
dilatation
slopes of the moment-curvature curves exhibited by both the constant
in error from
quadrilaterals and the crossed triangles were less than one percent
The beam-bending problems were also solved with ten irregularly-shaped con-
stant dilatation quadrilaterals across the thickness, and the results were es-
sentially unaffected.
The thick-walled plain strain tube under internal pressure proved a less
of a tube with outer diameter to inner diameter ratio of two was modeled with
five quadrilateral elements, the nodes having equal radial spacing. The problem
was also solved with the same quadrilaterals formed from four triangles with the
Radial displacement increments were prescribed at the two nodes on the inner dia-
direction. Pressures were inferred in the virtual work sense from the nodal forces
The trend in load-displacement curves obtained was the same as in the beam-
-26-
given deformation were greatly reduced. In fact, the maximum difference in any
computed load at any point in the deformation was less than four percent. The
load obtained when "steady state" conditions were reached was within a few percent
of the known limit load, and the pressure-expansion curves agreed closely with
that obtained by Hodge and White [9] using finite differences and presented
graphically in [10].
tensile specimen, shown in fig. 10, was solved using a very fine mesh of first
analogue of Prandtl's punch problem (see, for example, [101). In this case, the
for the von Mises yield criterion, where ao is again the uniaxial tension
yield strength.
By symmetry, only one-fourth of the DEN specimen was analyzed. The specimen
width to ligament width ratio was W/b = 10 . The loading was accomplished by
imposing increments of constant end displacement at the upper end of the specimen.
The load was computed both from the nodal tractions along the top row of nodes
and from the average stresses.in the top row of elements. Both methods gave
essentially the same results.
fail to find the correct limit load, and, in fact, the load-deflection curve con-
tinues to rise at roughly constant rate, exceeding the limit value by approximately
-27-
25% at the point for which computation was stopped. Indeed, this large error
given by extrapolating the linear elastic loading line to the limit load.
metric elements up to roughly 60% of the limit load, at which point it begins
to decrease in slope more rapidly. As the deformation proceeds into the fully
plastic region the difference between the two load-deflection curves increases
rapidly. At the point at which computation was stopped, the load was approxi-
mately 3% less than the analytical limit load, and the tangent modulus load-
deflection slope was roughly 0.1% of its elastic value, versus an apparently
Acknowledgement
The authors are grateful for financial support of this work by the U. S.
References
El] J. H. Argyris, Continua and Discontinua, Proc. ist Conf. Matrix Methods
Struct. Mech., Wright-Patterson Air Force Base, Ohio, October (1965).
[10) W. Prager, and P. G. Hodge, Jr., Theory of Perfectly Plastic Solids, Dover,
New York, (1968).
[11] R. Hill, Some Basic Principles in the Mechanics of Solids Without a Natural
Time, J. Mech. Phys. Solids, 7, pp. 209-225, (1959).
[14] R. Hill, on Constitutive Inequalities for Simple Materials, I and II, J. Mech.
Phys. Solids, 16, pp. 229-242 and 315-332, (1968).
-29-
Ratio Ratio
Constraints Nodes Dea.Freedom
Element Type Element Elements Constraints
3 1S2-node 2/3
quadrilateral
linear strain
3 2 4/3
triangle
8-node 6 to 8 3 1 to 3/4
quadrilateral
3 1/2 1/3
5 1 2/5
4-
x 6 2 2/3
S> 9 3 . 2/3
5 constant strain
cubes in regular
lattice
8-node
So risoparametric
cubes in 7 1 3/7
oregular lattice
5 linear strain
tetrahedra in 4 7/5 21/20
cube; cubes in
regular lattice
20-node
isoparametric > 16 4
cubes in 16 3/4
r~'-0 regular
lattice
Figure Titles
2) Finite element meshes used to solve the mid-ligament loading and constant
end displacement single edge cracked tensile specimens.
5a) Acceptable limit mechanism for plane strain single edge notched tensile speci-
men subject to either mid-ligament loading or constant end displacement.
finite
12) Three-dimensional block of rectangular eight-node isoparametric brick
elements.
1.4
LIMIT LOAD - -
1.2 -
ID-LIGAMENT
1.0 CONSTANT END
N 1.0 DISPLACEMENT
DLOADING
NET
F/ LOAD POINT END LOAD POINT
F,u
00 0.8
h h
0.4 - - h - h_
2 2
0.2
0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0
Figure 1.
EU
u.
E LOAD POINT/. 0 h
UEND
4h
3h
h h
Figure 2.
-35-
yX
Figure 3.
B
Figure 4.
-36-
I RIGID
450 450
- - = -yy= CONST.
xx yy
II I I I RIGID
F
(a) (b)
Figure 5.
-37-
-'i-
ea
7.
7r4
a
2
Figure 6.
Figure 7.
REGULAR ISOPARAMETRICS.
1.4 - ONE STRESS/ELEMENT
STRESS AT EACH INT. STA.
I CROSSED TRIANGLES
AND
0.8. - CONSTANT DILATATION ISOPARAMETRICS
0.6
1.5
H b
1.0 H/W=3
W/b= 10
0.5
0.1 I 1
I I 1 I I I
0 I 2 3 4 5 6 7 8 9 10
Figure
Figure 10.
-41-
Figure l1.
Figure 12.
-42-
ments that c and c must be the same for every element of the grid, and,
xx yy
if the 4-node isoparametrics are arbitrarily skewed quadrilaterials, then the
shear strain increment rate E must also be the same in every element.
xy
for some other typical mesh configurations in both two and three dimensions to
demonstrate the unrealistic constraints which are enforced upon admissible dis-
Consider the array of triangular constant strain elements shown in fig. 11,
show that x and E are the same in every element of such a band.
xx yy
elements, shown in fig. 12. The displacement rate field within such an element
can be expressed as
u
x
u = {a}) + {b) x + {c) y + {d) z + {e) xy + {f} yz
requires that {h) = 0 and that the other terms be constrained such that the
xx = P - ay + YZ
y = q - ax - yz (AI.3)
ZZ = -(p+q) + ax + ay
straint that zz has the same value at all points along a line extending in the
where A(x) , B(y) , and C(z) are each continuous functions of their respective
all displacements must be zero. Then cxx and cyy vanish on this plane and,
A(x) - B(y) = 0
B(y) - c(0) = 0
which requires that A(x) = B(y) = C(O) , and hence the most general admissible
S = - = C(z) - C(0)
y XX (AI.6)
E; = 0 .
zz
boundary, then
= = = 0 (AI.7)
xx yy zz
throughout.
-45-
admit the existence of a constitutive rate potential, Hill [ll] has shown that
the solution of boundary value problems can be obtained from the following varia-
S U(D .)dV - dV - vd 0 .
(AII.1)
Here V is the current (deformed) volume, and ST is the portion of the current
boundary on which nominal traction rates T.1 , based on the current state as the
reference state, are prescribed. The rate of deformation tensor is D.. , a..
av. ij 13
is the Cauchy (true) stress tensor, and - is the velocity gradient with respect
V au
Ti aD . - ijkg D * (AII.2)
1]
Here Tijk£ are the instantaneous moduli appropriate to the adopted measure of
stress rate, and are possibly dependent on the direction of D.. . The stress rate
1)
V
itself, rij , is the Jaumann, or co-rotational, rate of Kirchhoff stress, based on
a reference state that has been chosen to coincide instantaneously with the current
state, and Kirchhoff stress t.. is defined as a.. times the ratio of reference
state to current material density, so that Tij = oij instantaneously. Note that
the existence of U requires that moduli have the symmetry
-46-
and this ensures a symmetric incremental stiffness matrix for the corresponding
finite element equations. The class of materials which admit rate potentials
include all elastic materials for which a strain energy function exists, and also
all elastic-plastic materials which satisfy the normality rule when phrased in
which admits the rate potential U is obtained by relating the Jaumann rate of
form:
e l+v V v V
ij E 1T E ij kk
, , V (AII.4)
D?. ij k£ kk
13 4hr-2
and
-2 3
for continued plastic loading. Here T = Tr .. and h is the slope of the
2 ijij
stress-logarithmic plastic strain curve for simple tension. Note that in this
Thus equations (AII.4) can be inverted, and the resulting Jaumann rates of
Thus the deviatoric and hydrostatic Jaumann rates of Kirchhoff stress have
functional dependence only on the deviatoric and hydrostatic parts of the rate
obtain
Because this integrand has the same functional form as the conventional
The finite element implementation of the finite strain version of the varia-
element stiffness matrix being the sum of the stiffness corresponding to the modi-
fied rate potential and the "initial stress" stiffness, which corresponds to the