Sei sulla pagina 1di 16

Computer Methods in Applied Mechanics and Engineering 90 (1991) 869-884

North-Holland

Secant structural solution strategies under element


constraint for incremental damage
Z. Chen
New Mexico Engineering Research Institute, Box 25, University of New Mexico, Albuquerque,
NM 87131, USA

H.L. Schreyer
Department of Mechanical Engineering, University of New Mexico, Albuquerque, NM 87131, USA

Received 1 February 1990

Two major difficulties in the nonlinear analysis of post-peak structural behavior are the occurrence
of an ill-conditioned tangent stiffness matrix around critical points, namely limit and bifurcation points,
and the selection of a suitable constraint on the solution path. As an attempt to make failure prediction
available in a routine manner, new solution strategies are proposed in which a secant structural stiffness
matrix is formulated for incremental damage models, and the solution path is controlled through a
suitable measure of failure at the most severely damaged point in the body. Numerical solutions are
given for plane strain and plane stress problems to show that snap-back and snap-through associated
with shear band formation can be inexpensively predicted.

1. Introduction

Nonlinear stress-strain response features of engineering materials arise from two distinct
modes of microstructural changes: one is inelastic flow and the other is the damaging or
degradation of material properties. Inelastic flow, which is reflected through permanent
deformation, is the consequence of a dislocation process along preferred slip planes as in
metals, or particle motion and rearrangement as in geological materials. Because the number
of bonds between material particles is hardly altered during the inelastic process, the material
stiffness remains insensitive to this mode of microstructural motion, and change of strength is
reflected through the inelastic strain hardening phenomenon. On the other hand, the
nucleation, crushing and coalescence of microcracks and microvoids result in debonding,
which is reflected through changes in material stiffness and strength. In general, both modes
are present and interacting, and the structure is said to fail when macrocracks occur and
propagate through it. Within the framework of continuum theories, elastoplasticity and
damage mechanics are the tools to simulate the inelastic flow and the damaging process,
respectively.
Generally, there are two approaches to formulate nonlinear constitutive laws, namely
functional or total and differential or rate forms [1,2]. In numerical calculations, one prefers
the differential to the total forms of constitutive laws in order to simulate those paths which

0045-7825/91/$03.50 © 1991 Elsevier Science Publishers B.V. All rights reserved


870 Z. Chert. H. L, Schreyer, Secant structural solution strategies

are load dependent. The differential form is approximated by incremental ferms in numerical
algorithms. An iterative procedure is needed to correct for the incremental linearization error.
From the differential form of the stress-strain relation, it is natural to use the tangent instead
of the secant stiffness tensor. The convergence behavior of the incremental-iterative scheme
then depends heavily on the condition number of the tangent structural stiffness matrix, K,
that includes material nonlinearity, geometrical nonlinearity or both. If K becomes ill-
conditioned, as defined by a large value of the ratio of the maximum to the minimum absolute
values of the eigenvalues of K, the numerical solution will converge very slowly or even fail at
some point on the solution path, especially if bifurcation or limit points exist. Here, a
bifurcation point is defined to be a point on the solution path where the governing equation
admits more than one solution. A bifurcation or limit point is characterized by a critical state
at which the tangent stiffness matrix satisfies the instability criterion

K a,, = 0 , (1.1)

where the superscript c denotes the critical state, and Aa c, the vector (eigenmode) of critical
displacement increments. Equation (1.1) identifies the limit point if g c is symmetric, or the
bifurcation point if g~ is not symmetric for which the limit point is obtained by using the
symmetric part of g c. The bifurcation point can occur before or after the limit point is
reached. Structural failure problems such as buckling and failure zones can be attributed to
the existence of such critical points.
Although the concept of elastic or yield limit design has been and is still commonly used in
conventional engineering design procedures, experimental data for specimens loaded beyond
the critical point have been documented for metals [3-5], soils [6], rocks [7] and concrete
[8, 9]. The apparent features typical of the post-limit response include reductions of strength
and stiffness accompanied by a small zone of large deformations, which is often called
softening with localization. In many engineering applications, the post-limit behavior can be
observed. For dynamic cases such as those associated with explosive or seismic disturbances,
the loads are of very short duration so that even if a limit point is exceeded for a particular
structural member, no collapse occurs after the load is removed. In addition, more structural
members are statically indeterminate which also helps to preclude collapse. Since a significant
amount of energy absorption can be associated with the post-limit regime, a realistic
assessment of the nonlinear behavior of engineering structures must include the softening of
material strength and stiffness with localization. Recently, many research efforts have been
conducted to resolve theoretical arguments associated with softening and localization to make
clear the relation between the evolution of localization zone and the post-peak response, and
to simulate the degradation of material properties and localized failure modes [10-18], just to
mention some. Among the models proposed are nonlocal plasticity, local or nonlocal
continuum damage and other micromechanical models based on fracture mechanics. Prelimi-
nary results obtained for one- and two-dimensional sample problems look quite promising
because, if a noniocal model is used, the failure behavior of materials can be simulated and
the results are mesh independent. The accurate predictions of crack propagation, anisotropy
and post-peak dynamic response, however, are still at an infant stage due to the extreme
complexity of the microstructural features of materials. Applications to the three-dimensional
Z. Chert, H.L. Schreyer, Secant structural solution strategies 871

cases will not be feasible until a robust and efficient solution scheme is designed, even with the
use of supercomputers.
As can be found from the literature, a number of softening models have been put forward
to meet the increasing need involving advanced engineering materials. Now the situation has
come to a stage at which the development of sophisticated constitutive models itself might not
be the limiting factor in nonlinear structural analyses. Rather, a lack of suitable experimental
data to identify material parameters and an inability to obtain a converged solution in a
nonlinear numerical analysis pose the major limitations. Because a robust and efficient
solution scheme is a necessity for a model to be used, an effort will be made here to address
the numerical aspects of failure simulation.
Similar to the techniques used to obtain the post-limit response of experimentally loaded
specimens, the load increment cannot be directly prescribed without restrictions in an
incremental-iterative solution scheme if a solution path that is more realistic is to be traced in
the post-limit regime. Instead, a suitable constraint condition must be imposed to guide the
solution path in the right direction. In other words, the post-limit solution relies heavily on the
constraint condition which is a feature also reflected in experiments. Several schemes have
been proposed to circumvent the difficulties associated with critical points. One method is to
delete the singularity of g ~ by monitoring the presence of negative pivots from the decomposi-
tion of the stiffness matrix and replacing/~ with an expression of the type/i "~ + 771 with 77> 0
so that this matrix is 'safely' positive definite in the vicinity of a critical point [19]. A second
method suppresses the iterations around the critical point [20]. Because it is almost impossible
to obtain converged solutions around critical points by simply prescribing external load
increments, it has been suggested that the load level becomes a function of another variable
through a suitable constraint. Direct or indirect displacement control [21-23], arc-length
control [24, 25] and self-adaptive 'hyperelastic' constraints [26] are designed to provide such a
constraint condition through which the load level changes from iteration to iteration. In order
to simulate more efficiently the softening accompanied by localization, efforts have been made
toward modifying arc-length control so that alternative constraint equations can be applied.
One modification is to use the eigenvector of the lowest eigenvalue of the tangent stiffness
matrix as a guide to obtain the appropriate solution [27, 28]. Another one is to minimize a
function of an out-of-balance force vector to seek a dominant direction and to use a line
search procedure along that direction [19, 29].
Although the arc-length control is still commonly used in geometrically nonlinear problems,
and with some modifications in materially nonlinear problems, lack of convergence is
sometimes reported even after the limit or bifurcation point has been traversed. The main
reason might be the fact that in all the solution strategies reviewed above the load increment
in the first iteration has to be prescribed according to some criterion. The result is that for
some step there is a sign change of the load increment at a critical point. The sign change is
accompanied by changes in domains that are behaving highly nonlinearly, and consequently
the numerical solutions may not converge. A new solution scheme has been proposed for
softening problems involving total strain control so as to overcome such difficulties associated
with both localization and the sign change [12, 30]. The basic idea is that for each load step the
point in the body with the maximum value of a suitable measure of strain is identified, and a
linear combination of total strain increments is prescribed at that point. Implicit in the
872 Z. Chen, H.L. Schreyer, Secant structural solution strategies

formulation is the assumption that a nonlocal constitutive equation governs material behavior
with the consequence that the softening zone expands monotonically. The use of a local
constraint provides such an equilibrium path, a path which may not be accessible via
conventional approaches. Since the increment is assumed to be monotonically increasing and
the constraint equation is formulated locally with respect to that point, the load increment is
determined indirectly with the result that the procedure is particularly robust for tracing the
complete softening range whether or not there exists snap-back or snap-through. Since total
strain control is a constraint condition formulated in the deformation field, it is applicable to
both materially and geometrically nonlinear problems.
However, even with this approach, the rate of convergence of numerical solutions is found
to be quite slow for general cases with a smooth hardening and softening model as reflected,
for example, by an effective stress-effective strain relation. The problem is that the tangent
stiffness matrix might already be ill-conditioned even before the critical point is reached. With
constitutive equations based on continuum damage mechanics as a viable alternative for
modeling inelastic behaviors, the question arises as to whether or not the use of the resulting
secant tensor provides a computational advantage over the more conventional tangential
stiffness tensor. To investigate this issue, a simple strain-based elastodamage model together
with a failure-controlled constraint equation is proposed as a computational procedure.
Preliminary results [13] indicate that this continuum damage approach for structural softening
with localization is a fairly robust and efficient method with a reduction in computer time of an
order of magnitude in comparison with the time required to solve similar plasticity problems
based on the conventional tangent modulus. Although the damage formulation yields a total
relationship between stress and strain, the internal variables governing the damage process are
calculated through differential equations. This combination produces a stable and useful
numerical algorithm for nonlinear structural analyses and requires a small amount of computer
memory.
Here, it should be emphasized that the use of a secant stiffness matrix does not exclude the
existence of critical points that must be identified by analyzing the tangent stiffness matrix. In
other words, the theoretical insight associated with the critical points is explicitly obtained
only from the tangent matrix, but the use of the secant matrix provides an efficient solution
algorithm.
In this paper, secant structural algorithms under element constraint are given in a general
format, and different constraint conditions yield different solution strategies. The selection of
a suitable constraint depends on the constitutive model. Based on the general formulation of
continuum damage models, an elementary model is derived and solved by a simple but robust
constitutive equation solver to show that essential features of damage processes are simulated.
Numerical solutions are then given for plane strain and plane stress problems to show that
snap-back and snap-through associated with shear band formation can be inexpensively
predicted.

2. Failure-controlled solution strategies

For the problems considered here, it is assumed that the deformation and deformation
gradients are small, and no inertial effects occur. A corresponding spatially discretized system
Z. Chen, H.L. Schreyer, Secant structural solution strategies 873

can be obtained through the use of the displacement-based FEM as

Ka = ptq*, (2.1)

in which g now denotes the global secant stiffness matrix that is in general nonlinear and a
function of stress or strain history and internal variables, a the vector of nodal displacements,
q* the reference load vector with the assumption that only proportional loading will be
considered and pt represents the magnitude of the external load. With an appropriate choice of
the reference load vector, the total load parameter pt will be positive but the increment Apt
will change sign during the transition from the pre-limit to post-limit regimes.
In order to impose a suitable constraint on the solution path to indirectly determine the
load increment, let

a* = K-lq * , rl = arc = p t a * ' c , (2.2)

where c is a constraint vector and the superscript t indicates the transpose of a vector. The
control parameter r/is assumed to be monoton;~cally increasing, which can be conceived as a
suitable measure of failure phenomena. Thus, /z can be indirectly determined through the
following constraint equation:

pt = rl / a * t c . (2.3)
The above constraint equation is quite general and includes existing solution strategies, such
as direct or indirect displacement control, arc-length control and total strain control, as special
cases. The key difference here, however, is that the secant stiffne:,J matrix is employed
together with the assumption that the control parameter rl is monotonically increasing. Hence,
at a critical point, no singularity will occur and no sign change of the load increment needs to
be prescribed. Based on (2.2) and (2.3), a robust and efficient incrementai-iterative solution
procedure can be constructed.
Let the letters / and J represent the iteration and load-increment indices, respectively. With
the initial value of J - 1 and the secant stiffness at the end of the previous load incremem
denoted by K~j, the outline of the iteration loop for each increment can be written as follows:

1. r / s = r / J - l + A r / J , I=1.
2. a~ = K / ~ l q * .
3. l.tt = r l J / a * t C J and a t = l.tta:;.
4. If the value of the damage function is larger than zero, then there exists damage.
Calculate the internal variables governing the damage process from the consistency
condition.
5. If the measure of the changes in internal variables from iteration to iteration is larger
than a given tolerance, then set 1 = 1 4- 1 and go to Step 2.
6. Update variables. Exit or go to Step 1 for the next increment with J = J + 1.

An alternative procedure (explicit) without iterations can be obtained by deleting Step 5,


but the small increments in the control parameter are required for a nonlinear model.
874 Z. Chert. H.L. Schreyer, Secant structural sohaion strategies

The choice of a suitable constraint vector is crucial for tracing the correct solution path with
the fastest possible rate of convergence. In materially nonlinear problems, this choice depends
on the constitutive model which simulates the specific mode of failure. In order to make the
constraint more sensitive to the change inside the localization zone, the constraint vector is
formulated locally and associated with the point where a measure of failure such as damage is
a maximum so that the evolution of the localization zone can be easily captured. As the zone
of failure propagates, it is entirely feasible, and in fact necessary, to change the constraint
vector.
To illustrate the numerical approach, specific models must be adopted. The following gives
a general formulation for damage models, a constitutive equation solver and example
problems.

3. Damage simulation

Phenomenological and micromechanical based arguments are the two approaches for
establishing constitutive models based on continuum damage mechanics [16, 31-35]. There is
no question that a meaningful constitutive model must be consistent with the micromechanics
of the material considered. However, the discussion of micromechanical features is well
beyond the scope of the paper. Instead, an approach that includes the restrictions of
thermodynamics will be given with applications involving an elementary model to demonstrate
the procedure.

3. I. General formulation
For a purely mechanical case, there exists an internal energy per unit volume, U(e, Ed), or,
after a contact transformation with respect to stress and strain, a Gibbs free energy per unit
volume, G(tr, E'~), in which e and o" are the total strain and stress tensors, respectively, and
E ~ is the secant damage tensor that is a function of internal variables reflecting the damage
process. The Clausius-Duhem inequality can be expressed in terms of either the internal
energy

-(](e, E 'l) + or: e ~>0 (3.1)

or the Gibbs free energy

G(~r, E'~) - e:&>~O (3.2)

for any admissible process. Equation (3.1) leads to a strain-based formulation, while (3.2)
leads to a stress-based formulation. If the displacement-based FEM is adopted for the
semidiscretization in space, the strain-based formulation has a computational advantage over
the stress-based one, although both formulations are related. Hence, (3.1) will be used below
in the formulation of a strain-based damage model which can be solved via a simple but
accurate constitutive equation solver.
Z. Chert, H.L. Schreyer, Secant structural solution strategies 875

From the definition of the internal energy, an alternative form of (3.1) is given by

OU .pd>~O. (3.3)

Standard arguments [36, 37] and the assumption that the damage unloading process is elastic
result in the constitutive equation

o" = O U / O e (3.4)

and the dissipative inequality

_ ( ? t U / OEa) : : Ed >I O . (3.5)

With the assumption of linear elasticity, the internal energy can be expressed as

U = ½e:E~J.e, (3.6)

in which
E ~a = E ~ + E d (3.7)

is the elasticity (elastodamage) secant tensor. E ~ denotes the initial elasticity, while E d reflects
the damage part. The components of E d can be considered as the internal variables for the
damage process. The substitution of (3.6) in (3.4) yields the constitutive equation

o' = £~d: e . (3.8)

From the dissipative inequality (3.5), it follows that

_e:/~a:e>~0. (3.9)

Suppose damage is described by the evolution equation

/~'J = -~bR, (3.10)

where ~o is a monotonically increasing variable which parameterizes the equation and R is a


dimensionless tensor identifying the direction of damage. Then the dissipative inequality
becomes
e:R:e>~O, (3.11)

which is satisfied if R is positive definite. Define a damage function

fd = e ' R : e - S" , (3.12)


876 Z. Chen. H.L. Schreyer, Secant structural solution strategies

in which S is called the hardening-softening function for damage. Both R and S are considered
to be functions of a set of internal variables {I} where each internal variable, I,,, satisfies an
evolution equation of the type

i,, = (bs,,(t, { I ) ) . (3.13)

The associated damage surface is defined by the equation, fd = 0 , which enforces the
condition (3.11). The loading and unloading criteria for damage are then contained in the
following equations:

f d ~<O, tb i> 0 , (3.14)


In other words, loading ((b > 0) can occur only if fd = 0. The damage consistency condition,
f"l = 0, guarantees that the strain remains on the damage surface during the damage process.
From the consistency condition, o) can be determined as

(b = 2 a t : R : e, (3.15)

where

/,~ 0fd
O/ = --1 ~ Sn . (3.16)

For most computational codes, the strain rate is known. Thus, the stress tensor and internal
variables can be obtained for given increments in the strain tensor with the use of strain-based
constitutive equations consisting of (3.8), (3. I0), (3.12) and (3.13) together with the loading/
unloading criteria (3.14). The difference between different elastodarnage models depends on
the specific evolution equations and damage surface. If a stress-based damage function is used,
then the equation corresponding to (3.15) for (b involves the stress rate instead of the strain
rate.
For damage problems, it is a natural consequence to use the secant tensor E ~d to construct a
structural stiffness matrix for computations because of the explicit formulation shown in (3.8).
It is not difficult to derive a corresponding tangent tensor T ~d since it follows from (3.8) that

do" = E "d : de + dE ~d : e . (3.17)

The use of (3.10) and (3.15) yields

T ~d = E ~ d - 2 a R : e ® e : R , (3.18)

so that

do- = T ~d : d e . (3.19)

Hence, a solution scheme can make use of either the tangent or secant tensor although the
Z. Chen, H.L. Schrt:ver, Secant structural sohaion strategies 877

internal variables are calculated through the rate form in order to simulate those paths which
are load dependent with the least possible computer resources. The use of the secant tensor,
however, avoids the singularity associated with critical points.

3.2. Constitutive equation solver


For nonlinear damage models, an incremental-iterative scheme is required to solve the
constitutive equations. Since the rate or differential form is approximated by incremental
forms in numerical calculations, an acceptable incremental-iterative algorithm must satisfy
three basic requirements: (a) consistency with the constitutive equations to be integrated, (b)
numerical stability, and (c) incremental damage consistency condition. Requirements (a) and
(b) are necessary for attaining convergence of the numerical solution as the incremental step
becomes vanishingly small. Requirement (c) is the algorithmic counterpart of fd = 0. Without
further discussion here about computational aspects of damage models, the outline of a
constitutive equation solver developed by Chen and Schreyer [38] is given as follows for
strain-based damage models. Because the increment Ato can be found for given total strain
increments by using either the Newton-Raphson or secant method, and no elastic-predictor
and damage-corrector are needed as in the stress-based formulation, the integration schemes
for the strain-based formulation can be made much simpler and introduce less numerical error
than for the stress-based one.

Algorithm for strain-based damage models


The user inputs model parameters and control variables. For each increment, the states of
strain, stress and internal variables at the end of the previous increment are known and
denoted with the subscript J - 1, and the current strain ~'.~is prescribed. The itcr~tion loop for
solving Ato is identified with the index 1.

1. Check the damage loading/unloading criterion:


lffO[ej, { l } j _ t ] > 0 , then go to step 3. Otherwise, the solution is elastic.
2. Update {1}, E ~t and o'. Exit or go to Step I for the next increment.
3. Damage has occurred. Solve I'd[e j, {1}1] = 0 for Ao~ by using either the Newton-
Raphson or secant method. Go to Step 2 after a converged solution is obtained.

3.3. An elementary model


Experimental results [39] .indicate that the internal strain localization starts as early as 40%
of the limit load for some concrete specimens, which implies that the degradation of material
properties might occur well before the peak strength of material is reached. Thus a damage
model should be able to include both hardening and softening regimes, although inelastic flow
might dominate the hardening phenomenon. Here, a simple damage softening model is
developed and corresponding numerical results are given to show the robustness of the
proposed solution procedure. The extension to the hardening part is straightforward [38].
Either stress- or strain-based models can be formulated. To illustrate the approach, suppose
a stress-based model is proposed and then transferred to a corresponding strain-based one. If
the elastodamage process is assumed to be isotropic, the elastic and damage parts of E ~J can
878 Z. Chert, H.L. Schreyer, Secant structural solution strategies

be expressed in terms of bulk, B, and shear, G, moduli as follows:

E~=3B~P~+2G~P a " Ea=3Ba({I})P~+2Ga({I})P a, (3.20a, b)

where B ~ and G ~"(B d and G a) are the original (damaged) parts of the bulk and shear moduli,
respectively. P~ and pd are the volumetric and deviatoric orthogonal projection tensors so that
the volumetric and deviatoric parts of a second order tensor such as the strain tensor, e" and
e a, respectively, can be found through

E ,, = p, "E, Ea = pa "E. (3.21)

Consider the simple case of a Von Mises surface:

fa -~ 32 6"z o.,I.o.d (3.22)

in which o "a denotes the stress deviator and 6" the effective stress (stress invariant). Because

o.d. o,d = o r : p d . p a . c r = e:E~a:pd:E~d:e


= 4(G" + G'J)2e "Pa:e = 4(G ~ + G°)2e d" e a , (3.23)

the stress-based formulation (3.22) can be changed to a strain-based one as follows:

f a = e :R: e - S" = E" - S : , R = pd , (3.24a, b)

S" = 3 ' / 4 ( G " + Ga) ' ~ = e a : e '' (3.24c, d)

in which ~ denotes the effective strain (strain invariant). The new damage function fd is
obtained by dividing ]'t by 4(G ~ + Ga). The choice of R given in (3.24b) is necessary for the
equality of (3,24a) to hold. It can be argued from both a thermodynamical and physical basis
that the effective strain must be a ,,,onotonically increasing function. Therefore, with the
strain-based formulation, S 2 must al,,;o mcrease monotonically with the consequence that the
numerator in (3.24c) must ~. ........ ,. at a slower rate than the denominator.
The use of (3.10), ~,,~. ,~,) and (3.24b) yields B a = 0 and

2 d; Opd = _ ¢bp,t, (3.25)

namely G d can be considered the internal variable. With the initial condition G d ( O ) = 0 , it
follows that

G ~a=G ~+G d=G~(1-oJ/2G~), 0 <.co<~2G ~. (3.26)

Thus, (3.24a) becomes

f a = ~2_ 3:/2G~(1 _ co/2G~)]2" (3.27)


Z. Chert, H.L. Schrever, Secant structural solution strategies 879

Let 6"t, denote the limit effective stress and define ft, by gL = 6"t./2G¢ and S by S = 6t.(1 -
to/2G~) ~/2. Then it can be shown that
-2
_'~ E L
fdm.~ E-
- (1 - ~o/2G ~) " (3.28)

As can be seen from (3.28), the effective strain will increase with the damage process.
The simple strain-based damage model consisting of (3.8), (3.25) and (3.28) will be used in
the next subsection to predict damage softening with localization, with the material parame-
ters B ~= 24GPa, G ~= 18GPa and ~L = 5 0 M P a which are representative of a class of
geological materials.

3.4. Numerical results


Since the strain-based damage model given above simulates a failure mode of the shear
band type, the element constraint vector should be formulated in the deviatoric strain space as
follows. With the introduction of the volumetric and deviatoric orthogonal projection mat-
rices, [P~] and [PJ], respectively, the volumetric and deviatoric parts of the total strain vector
in a critical element, {e}~, can be written as

(3.29)

Suppose the invariant of the deviatoric strain in the critical element is monotonically
increasing during the damage process. Then the local control parameter r/can be chosen as
follows:

e ~t i t e ~d = ( I.~P d B ~ a , ) t ( I.tP d B~,a :1~) = ,172


(3.30)

and (2.3) is obtained with the constraint vector defined by

c = Bc~P
' aBcea"* / ~ / a ' B,t~ P IB ~ aa" , , (3.31)

in which B~ is the matrix rclating the strain to the nodal displacement components over the
critical element through a differential operator and interpolation functions. It should be noted
that only a few calculations are needed to calculate the constraint vector given by (3.31) since
the constraint is in terms of a local instead of a global formulation.
With the use of the proposed solution procedure and element constraint vector, numerical
solutions are given for the plane strain and plane stress problems. The geometry and notation
for the plane problems are indicated in Fig. 1. The dimensions uscd for the analysis are
L., = 1 m and L,. = 1.5 m. In order to accommodate isochoric deformation, the argument [40]
for the use of a four-node element composed of four triangles is considered to be persuasive,
and therefore the same type of element is employed here. The boundary condition at y = 0
consists of zero vertical displacement at all points and a lateral constraint of zero displacement
at one point that is x = 0 unless specified otherwise. The external load is applied so that the
displacement in the y-direction is the same for all points on the upper surface, y = 1.5 m.
880 Z. Chen, H.L. Schreyer, Secant structural solution strategies

Force
Y

Be = 24 GPa
Ge = 18GPa
oL = 50MPa
ZL = 1.39E-3
A11 = 1.E-4
Lx

Fig. 1. Geometry of the plane problem.

Element mesh configurations I, II and III are defined to be 3 x 3, 5 x 7 and 9 x 13 rectangular


grids, respectively. Normally, solutions to plane strain problems are given because the
two-dimensional displacement formulation implies that the plane strain condition is satisfied
automatically. With the modification that the total strain component in the normal or
z-direction is chosen to yield zero normal stress based on the constitutive relation, however,
solutions to plane stress problems can also be obtained with little additional computational
effort. Both classes of plane problems can have the same geometric configuration in the
two-dimensional space, although there exists a difference of the dimension in the z-direction.
Figure 2 illustrates the loading and unloading of the simple damage model. As can be seen,
no permanent deformation occurs after unloading as expected for a damage model. With

60 I I I I

50

i 40
3O

2O

10

0
0 2 4 6 8 12x10 -3

Strain invariant
Fig. 2, Loading and unloading of the damage model.
Z. Chen, H.L. Schreyer, Secant structural solution strategies 881

AT = 10 -4, and with an initial imperfection defined by setting 5 L = 49 MPa rather than the
prescribed value of 50 MPa in one or more of the triangular elements, Figs. 3-5 give
numerical solutions for plane strain, and Fig. 6 for plane stress. Figure 3(a) shows the final
deformation field under compression for element configuration III with the left and right
triangles in the middle rectangle being weaker than all others, Fig. 3(b), with the left and
lower triangles in the middle rectangle being weak elements, and Fig. 4(a, b) illustrate the
intermediate and final deformation fields under compression with the same weak elements as
in Fig. 3(a) but with lateral constraints at the corners (0, 0), (1 m, 0), (1 m, 1.5 m) and
(0, 1.5 m). As can be observed, failure modes are very sensitive to imperfections and
boundary conditions, which is consistent with experimental results [39]. Figure 5 shows the

2.0 (a) ' ' 2.0 (b) I I m |


b

1.5 1,5

1.0 1.0
|diiiiiii
| N i m m i i l |
i l l l l i d l l l
0,5 0.5 --

o,o 0.0 :l I
-0.5 0.0 0.5 1.0 1.5 -0.5 0.0 0,5 1,0 1.5

x-=is (m) x-axis(m)


Fig. 3. Final deformation fields with different weak elements.

2.0 (a) I I I I I I
2,0 ib )
1.5 1.5

mmmmuummE mimmmnun
HHHnnNp_~ i mmmmm i
1.0 ilillllllli_ll~ 1.0 b
HHINHuDINi im ._ u_mi__ in
>,

0,5
imimammimmmm
iiil
lniUl 0s[ mHi . - iim_m
mmlihidiun

0.0 0.0
-0,5 0.0 0.5 1,0 .5 -0.5 0.0 0.5 1.0 1.5

x-axis (m) x-axis (m)


Fig. 4. Deformed mesh in the post-peak regime for plane strain. (a) Intermediate post-peak. (b) Final post-peak.
882 Z. Chen, H.L. Schreyer, Secant structural solution strategies

1.5 I I I
1.5 I I I

, Mesh III M e s h III


...... Mesh II ...... M e s h II
.... Mesh I .... Mesh I
1.0 m m
"0 1.0 m

10 t,
:. "~,
0 0.5
z 0.5
°eeeeeeew~l I

o.o& 0.5 1.0


I
1.5 2.0
0.0
0.0 0.5 1.0 1.5 2.0

Normalized displacement Normalized displacement

Fig. 5. Plane strain solutions for different meshes, Fig. 6. Plane stress solutions for different meshes.

load-displacement curves for different meshes, where snap-back phenomena are predicted
without difficulty. It should be emphasized that with displacement-controlled devices the
equilibrium path exhibiting snap-back would not actually be followed in a physical situation.
However, if deformations are slow it is plausible that such a path would be followed at later
stages of loading. Numerical solutions for plane stress are similar to those for plane strain, and
failure patterns are quite sensitive to imperfections and boundary conditions. Figure 6
illustrates plane stress solutions for different meshes.

4. Conclusions

The combination of a local constraint and a damage-based constitutive model provides a


particularly simple solution algorithm for problems exhibiting softening and localization. The
procedure has been demonstrated on model problems for which considerably more computer
time is required if a constitutive relation with softening plasticity is assumed. Since damage
mechanics provides microstructural arguments for softening, the proposed approach may be
more physically appropriate than one based on plasticity, with the additional benefit of
numerical efficiency. For each problem, the predicted post-peak mode is only one of an
infinite number and, ultimately, the selection of the appropriate mode must be based on
additional factors such as loading conditions and more sophisticated constitutive equations.
For numerical investigations of these issues, the robust approach proposed here should be
invaluable.

Acknowledgment

The support of this work by the US Air Force Office of Scientific Research is gratefully
acknowledged. The authors are also grateful to M.K. Neilsen for valuable joint discussions.
Z. Chen, H.L. Schreyer, Secant structural solution strategies 883

References

[1] S. Nemat-Nasser, On nonequilibrium thermodynamics of viscoelasticity and viscoplasticity, in: J.J. Delgado
Domingos, M.N.R. Nina and J.H. Whitelaw, eds., Foundation of Continuum Thermodynamics (Macmillan,
London, 1974) 259-282.
[2] C. Truesdell and W. Noll, The non-linear field theories of mechanics, Encyclopedia Phys. 3 (3) (Springer,
Berlin, 19651.
[31 L. Anand and W.A. Spitzig, Initiation of localized shear bands in plane strain, J. Mech. Phys. Solids 28 ( 19801
113-128.
[4] Y.W. Chang and R.J. Asaro, Lattice rotations and localized shearing in single crystals, Arch. Mech. 32 (1980)
369-388.
[5] D. Peirce, R.J. Asaro and A. Needleman, An analysis of nonuniform and localized deformation in ductile
single crystals, Acta Metall. 30 (19821 1087-1119.
[6] I. Vardoulakis, Bifurcation analysis of the triaxial test on sand samples, Acta Mech. 32 (1979) 35-54.
[7] W.R. Wawersik and W.F. Brace, Post-failure behavior of a granite and a diabase, Rock Mech. 3 (19711 61-85.
[8] S.P. Shah and V.S. Gopalaratnam, Softening response of plain concrete in direct tension, J. Amer. Concrete
Inst. 82 (19851 310-323.
[9] J.G.M. van Mier, Strain-softening of concrete under multiaxial loading conditions, Ph.D. Dissertation,
University of Eindhoven, The Netherlands, 1984.
[10] Z.P. Bazant and G. Pijaudier-Cabot, Nonlocal continuum damage, Localization instability and convergence,
J. Appl. Mech. 55 (19881 287-293.
[11] Z. Chen and H.L. Schreyer, Simulation of soil-concrete interfaces with nonlocal constitutive models, J. Engrg.
Mech. 113 (11) (1987) 1665-1677.
[12] Z. Chen, Nonlocal theoretical and numerical investigations of soil-concrete interfaces, Ph.D. Dissertation,
University of New Mexico, USA, 1989.
[13] Z. Chen and H.L. Schreyer, Failure-controlled solution strategies for damage softening with localization, in:
S.P. Shah, S.E. Swartz and M.L. Wang, eds., Micromechanics of Failure of Quasi-Brittle Materials,
Proceedings of the International Conference on Micromechanics of F~.ilure of Quasi-Brittle Materials
(Elsevier, New York, 19901 135-145.
[14] M. Ortiz, An analytical study of the localized failure modes of concrete, Mechanics of Materials 6 (19871
159-174,
[15] H.L. Schreyer and Z. Chen, One dimensional softening with localization, J. Appl. Mech. 53 (4) (19861
791-797.
[16] L.M. Taylor, E.P. Chen and J.S. Kuszmaul, Microcrack-induced damage accumulation in brittle rock under
dynamic loading, Comput. Methods Appl. Mech. Engrg. 55 (19861 301-320.
[17] W.R. Wawersik, J.W. Rudnicki, W.A. Olsson, D.J. Holcomb and K.T. Chau, Localization of deformation in
brittle rock: Theoretical and laboratory investigations, in: S.P. Shah, S.E. Swartz and M.L. Wang, eds.,
Micromeehanics of Failure of Quasi-Brittle Materials, Proceedings of the International Conference on
Micromechanics of Failure of Quasi-Brittle Materials (Elsevier, New York, 19911) 115-124.
[18] S. Yazda:~i and H.L. Schreyer, An anisotropic damage model with dilatation for concrete, Mechanics of
Materials 7 (1988) 231-244.
[19] T. Belytschko and D. Lasry, Localization limiters and numerical strategies for strain-softening materials, in: J.
Mazars and Z.P. Bazant, eds., Cracking and Damage: Strain Localization and Size Effect (Elsevier, New
York, 19891 349-362.
[20] E. Ramm, Strategies for tracing non-linear responses near limit points, in: W. Wunderlich, E. Stein and K.J.
Bathe, eds., Non-Linear Finite Element Analysis in Structural Mechanics (Springer, New York, 19811 68-89.
[21] J.H. Argyris, Continua and discontinua, Proc. 1st Conference in Matrix Methods in Structural Mechanics,
Wright-Patterson Air Force Base, Ohio, USA, 1965, 11-189.
[22] R. de Borst, Non-linear analysis of frictional materials, Ph.D. Dissertation, Delft University of Technology,
The Netherlands, 1986.
[23] G. Poweil and J. Simons, Improved iteration strategy for nonlinear structures, lnternat. J. Numer. Methods
Engrg. 17 (19811 1455-1467.
[24] M.A. Crisfield, A fast incremental/iterative solution procedure that handles 'snap through', Comput. &
Structures 13 (1981) 55-62.
884 Z. Chen, H.L. Schreyer, Secant structural solution strategies

[25] E. Ramm, The Riks/Wempner approach-An extension of the displacement control method in nonlinear
analysis, in: E. Hinton, D.R.J. Owen and C. Taylor, eds., Recent Advances in Non-Linear Computational
Mechanics (Swansea, UK, 1982) 63-86.
1261 J. Padovan and S. Tovichakchaikul, Self-adaptive predictor-corrector algorithms for static nonlinear structural
analysis, Comput. & Structures 15 (1982) 365-377.
127] M.A. Crisfield and J. Wills, Solution strategies and softening materials, Comput. Methods Appl. Mech.
Engrg. 66 (3) (1988) 267-289.
[28] R. de Borst, Computation of post-bifurcation and post-failure behavior of strain-softening solids, Comput. &
Structures 25 (1987) 211-224.
[29] M.A. Crisfield, An arc-length method including line searches and accelerations, Internat. J. Numer. Methods
Engrg. 19 (1983) 1269-1298.
[3(1[ Z. Chen and H.L. Schreyer, A numerical solution scheme for softening problems involving total strain
control, Comput. & Structures 37 (6) (1990) 1043-1050.
[311 H. Horri and S. Nemat-Nasser, Overall moduli of solids with microcracks: Load induced anisotropy, J. Mech.
Phys. Solids 31 (2) (1983) 155-171.
[32] J.W. Ju, On energy-based coupled elastoplastic damage theories: Constitutive modeling and computational
aspects, Internat. J. Solids and Structures 25 (7) (1989) 803-833.
[33] L.M. Kachanov, Introduction to Continuum Damage Mechanics (Martinus Nijhoff, Dordrecht, 1986).
[34] D. Krajcinovic, Damage mechanics, Mechanics of Materials 8 (1989) 117-197.
[35] S. Yazdani and H.L. Schreyer, Combined plasticity and damage mechanics model for plain concrete, J. Engrg.
Mech. 116 (7) (1990) 1435-1450.
[36] B.D. Coleman and W. Noll, The thermodynamics of elastic materials with heat conduction and viscosity,
Arch. Rat. Mech. Anal. 13 (1963) 167-178.
[37] B.D. Coleman and M. Gurtin, Thermodynamics with internal variables, J. Chem. Phys. 47 (1967) 597-613.
[38] Z. Chen and H.L. Schreyer, Formulation and computational aspects of plasticity and damage models for
geological materials with emphasis on concrete, SAND90-7102, Sandia National Laboratories, Albuquerque,
NM, USA, 1990.
[39] M.L. Wang, H.L. Schreyer and C.A. Rutland, Internal deformation measurements with the use of real time
X-r:lys, in: S.P. Shah, S.E. Swartz and M.L. Wang, eds., Micromechanics of Failure of Quasi-Brittle
Materials, Proceedings of the International Conference on Micromechanics of Failure of Quasi-Brittle
Materials (Elsevier, New York, 19911) 81-94.
[401 J.C. Nagtegaal, D.M. Parks and J,C. Rice, On numerically accurate finite element solutions in the fully plastic
range. Comput. Methods Appl. Mech. Engrg. 4 (1974) 153-177.

Potrebbero piacerti anche