The smeared crack concrete model in
Abaqus/Standard:
provides a general capability for modeling concrete in all types of
structures, including beams, trusses, shells, and solids;
can be used for plain concrete, even though it is intended primarily
for the analysis of reinforced concrete structures;
can be used with rebar to model concrete reinforcement;
is designed for applications in which the concrete is subjected to
essentially monotonic straining at low confining pressures;
consists of an isotropically hardening yield surface that is active
when the stress is dominantly compressive and an independent “crack detection
surface” that determines if a point fails by cracking;
uses oriented damaged elasticity concepts (smeared cracking) to
describe the reversible part of the material's response after cracking failure;
requires that the linear elastic material model (see
Linear Elastic Behavior)
be used to define elastic properties; and
cannot be used with local orientations (see
Orientations).
See
Inelastic Behavior
for a discussion of the concrete models available in
Abaqus.
Reinforcement in concrete structures is typically provided by means of
rebars, which are one-dimensional strain theory elements (rods) that can be
defined singly or embedded in oriented surfaces. Rebars are typically used with
metal plasticity models to describe the behavior of the rebar material and are
superposed on a mesh of standard element types used to model the concrete.
With this modeling approach, the concrete behavior is considered
independently of the rebar. Effects associated with the rebar/concrete
interface, such as bond slip and dowel action, are modeled approximately by
introducing some “tension stiffening” into the concrete modeling to simulate
load transfer across cracks through the rebar. Details regarding tension
stiffening are provided below.
Defining the rebar can be tedious in complex problems, but it is important
that this be done accurately since it may cause an analysis to fail due to lack
of reinforcement in key regions of a model. See
Defining Reinforcement
for more information regarding rebars.
Cracking
The model is intended as a model of concrete behavior for relatively
monotonic loadings under fairly low confining pressures (less than four to five
times the magnitude of the largest stress that can be carried by the concrete
in uniaxial compression).
Crack Detection
Cracking is assumed to be the most important aspect of the behavior, and
representation of cracking and of postcracking behavior dominates the modeling.
Cracking is assumed to occur when the stress reaches a failure surface that is
called the “crack detection surface.” This failure surface is a linear
relationship between the equivalent pressure stress, p,
and the Mises equivalent deviatoric stress, q, and is
illustrated in
Figure 5.
When a crack has been detected, its orientation is stored for subsequent
calculations. Subsequent cracking at the same point is restricted to being
orthogonal to this direction since stress components associated with an open
crack are not included in the definition of the failure surface used for
detecting the additional cracks.
Cracks are irrecoverable: they remain for the rest of the calculation (but
may open and close). No more than three cracks can occur at any point (two in a
plane stress case, one in a uniaxial stress case). Following crack detection,
the crack affects the calculations because a damaged elasticity model is used.
Oriented, damaged elasticity is discussed in more detail in
An inelastic constitutive model for concrete.
Smeared Cracking
The concrete model is a smeared crack model in the sense that it does not
track individual “macro” cracks. Constitutive calculations are performed
independently at each integration point of the finite element model. The
presence of cracks enters into these calculations by the way in which the
cracks affect the stress and material stiffness associated with the integration
point.
Tension Stiffening
The postfailure behavior for direct straining across cracks is modeled with
tension stiffening, which allows you to define the strain-softening behavior
for cracked concrete. This behavior also allows for the effects of the
reinforcement interaction with concrete to be simulated in a simple manner.
Tension stiffening is required in the concrete smeared cracking model. You can
specify tension stiffening by means of a postfailure stress-strain relation or
by applying a fracture energy cracking criterion.
Postfailure Stress-Strain Relation
Specification of strain softening behavior in reinforced concrete generally
means specifying the postfailure stress as a function of strain across the
crack. In cases with little or no reinforcement this specification often
introduces mesh sensitivity in the analysis results in the sense that the
finite element predictions do not converge to a unique solution as the mesh is
refined because mesh refinement leads to narrower crack bands. This problem
typically occurs if only a few discrete cracks form in the structure, and mesh
refinement does not result in formation of additional cracks. If cracks are
evenly distributed (either due to the effect of rebar or due to the presence of
stabilizing elastic material, as in the case of plate bending), mesh
sensitivity is less of a concern.
In practical calculations for reinforced concrete, the mesh is usually such
that each element contains rebars. The interaction between the rebars and the
concrete tends to reduce the mesh sensitivity, provided that a reasonable
amount of tension stiffening is introduced in the concrete model to simulate
this interaction (Figure 1).
The tension stiffening effect must be estimated; it depends on such factors
as the density of reinforcement, the quality of the bond between the rebar and
the concrete, the relative size of the concrete aggregate compared to the rebar
diameter, and the mesh. A reasonable starting point for relatively heavily
reinforced concrete modeled with a fairly detailed mesh is to assume that the
strain softening after failure reduces the stress linearly to zero at a total
strain of about 10 times the strain at failure. The strain at failure in
standard concretes is typically 10−4, which suggests that tension
stiffening that reduces the stress to zero at a total strain of about
10−3 is reasonable. This parameter should be calibrated to a
particular case.
The choice of tension stiffening parameters is important in
Abaqus/Standard
since, generally, more tension stiffening makes it easier to obtain numerical
solutions. Too little tension stiffening will cause the local cracking failure
in the concrete to introduce temporarily unstable behavior in the overall
response of the model. Few practical designs exhibit such behavior, so that the
presence of this type of response in the analysis model usually indicates that
the tension stiffening is unreasonably low.
Fracture Energy Cracking Criterion
As discussed earlier, when there is no reinforcement in significant regions
of a concrete model, the strain softening approach for defining tension
stiffening may introduce unreasonable mesh sensitivity into the results.
Crisfield (1986) discusses this issue and concludes that Hillerborg's (1976)
proposal is adequate to allay the concern for many practical purposes.
Hillerborg defines the energy required to open a unit area of crack as a
material parameter, using brittle fracture concepts. With this approach the
concrete's brittle behavior is characterized by a
stress-displacement response rather than a
stress-strain response. Under tension a concrete
specimen will crack across some section. After it has been pulled apart
sufficiently for most of the stress to be removed (so that the elastic strain
is small), its length will be determined primarily by the opening at the crack.
The opening does not depend on the specimen's length (Figure 2).
Implementation
The implementation of this stress-displacement concept in a finite element
model requires the definition of a characteristic length associated with an
integration point. The characteristic crack length is based on the element
geometry and formulation: it is a typical length of a line across an element
for a first-order element; it is half of the same typical length for a
second-order element. For beams and trusses it is a characteristic length along
the element axis. For membranes and shells it is a characteristic length in the
reference surface. For axisymmetric elements it is a characteristic length in
the r–z plane only. For cohesive
elements it is equal to the constitutive thickness. This definition of the
characteristic crack length is used because the direction in which cracks will
occur is not known in advance. Therefore, elements with large aspect ratios
will have rather different behavior depending on the direction in which they
crack: some mesh sensitivity remains because of this effect, and elements that
are as close to square as possible are recommended.
This approach to modeling the concrete's brittle response requires the
specification of the displacement
at which a linear approximation to the postfailure strain softening gives zero
stress (see
Figure 2).
The failure stress, ,
occurs at a failure strain (defined by the failure
stress divided by the Young's modulus); however, the stress goes to zero at an
ultimate displacement, ,
that is independent of the specimen length. The implication is that a
displacement-loaded specimen can remain in static equilibrium after failure
only if the specimen is short enough so that the strain at failure,
,
is less than the strain at this value of the displacement:
where L is the length of the specimen.
Obtaining the Ultimate Displacement
The ultimate displacement, ,
can be estimated from the fracture energy per unit area,
,
as ,
where
is the maximum tensile stress that the concrete can carry. Typical values for
are 0.05 mm (2 × 10−3 in) for a normal concrete to 0.08 mm (3 ×
10−3 in) for a high strength concrete. A typical value for
is about 10−4, so that the requirement is that
mm (20 in).
Critical Length
If the specimen is longer than the critical length,
L, more strain energy is stored in the specimen than can
be dissipated by the cracking process when it cracks under fixed displacement.
Some of the strain energy must, therefore, be converted into kinetic energy,
and the failure event must be dynamic even under prescribed displacement
loading. This implies that, when this approach is used in finite elements,
characteristic element dimensions must be less than this critical length, or
additional (dynamic) considerations must be included. The
analysis input file processor
checks the characteristic length of each element using this concrete model and
will not allow any element to have a characteristic length that exceeds
.
You must remesh with smaller elements where necessary or use the stress-strain
definition of tension stiffening. Since the fracture energy approach is
generally used only for plain concrete, this rarely places any limit on the
meshing.
Cracked Shear Retention
As the concrete cracks, its shear stiffness is diminished. This effect is
defined by specifying the reduction in the shear modulus as a function of the
opening strain across the crack. You can also specify a reduced shear modulus
for closed cracks. This reduced shear modulus will also have an effect when the
normal stress across a crack becomes compressive. The new shear stiffness will
have been degraded by the presence of the crack.
The modulus for shearing of cracks is defined as ,
where G is the elastic shear modulus of the uncracked
concrete and
is a multiplying factor. The shear retention model assumes that the shear
stiffness of open cracks reduces linearly to zero as the crack opening
increases:
where
is the direct strain across the crack and
is a user-specified value. The model also assumes that cracks that subsequently
close have a reduced shear modulus:
where you specify .
and
can be defined with an optional dependency on temperature and/or predefined
field variables. If shear retention is not included in the material definition
for the concrete smeared cracking model,
Abaqus/Standard
will automatically invoke the default behavior for shear retention such that
the shear response is unaffected by cracking (full shear retention). This
assumption is often reasonable: in many cases, the overall response is not
strongly dependent on the amount of shear retention.
Compressive Behavior
When the principal stress components are dominantly compressive, the
response of the concrete is modeled by an elastic-plastic theory using a simple
form of yield surface written in terms of the equivalent pressure stress,
p, and the Mises equivalent deviatoric stress,
q; this surface is illustrated in
Figure 5.
Associated flow and isotropic hardening are used. This model significantly
simplifies the actual behavior. The associated flow assumption generally
over-predicts the inelastic volume strain. The yield surface cannot be matched
accurately to data in triaxial tension and triaxial compression tests because
of the omission of third stress invariant dependence. When the concrete is
strained beyond the ultimate stress point, the assumption that the elastic
response is not affected by the inelastic deformation is not realistic. In
addition, when concrete is subjected to very high pressure stress, it exhibits
inelastic response: no attempt has been made to build this behavior into the
model.
The simplifications associated with compressive behavior are introduced for
the sake of computational efficiency. In particular, while the assumption of
associated flow is not justified by experimental data, it can provide results
that are acceptably close to measurements, provided that the range of pressure
stress in the problem is not large. From a computational viewpoint, the
associated flow assumption leads to enough symmetry in the Jacobian matrix of
the integrated constitutive model (the “material stiffness matrix”) such that
the overall equilibrium equation solution usually does not require unsymmetric
equation solution. All of these limitations could be removed at some sacrifice
in computational cost.
You can define the stress-strain behavior of plain concrete in uniaxial
compression outside the elastic range. Compressive stress data are provided as
a tabular function of plastic strain and, if desired, temperature and field
variables. Positive (absolute) values should be given for the compressive
stress and strain. The stress-strain curve can be defined beyond the ultimate
stress, into the strain-softening regime.
Uniaxial and Multiaxial Behavior
The cracking and compressive responses of concrete that are incorporated in
the concrete model are illustrated by the uniaxial response of a specimen shown
in
Figure 3.
When concrete is loaded in compression, it initially exhibits elastic
response. As the stress is increased, some nonrecoverable (inelastic) straining
occurs and the response of the material softens. An ultimate stress is reached,
after which the material loses strength until it can no longer carry any
stress. If the load is removed at some point after inelastic straining has
occurred, the unloading response is softer than the initial elastic response:
the elasticity has been damaged. This effect is ignored in the model, since we
assume that the applications involve primarily monotonic straining, with only
occasional, minor unloadings. When a uniaxial concrete specimen is loaded in
tension, it responds elastically until, at a stress that is typically 7%–10% of
the ultimate compressive stress, cracks form. Cracks form so quickly that, even
in the stiffest testing machines available, it is very difficult to observe the
actual behavior. The model assumes that cracking causes damage, in the sense
that open cracks can be represented by a loss of elastic stiffness. It is also
assumed that there is no permanent strain associated with cracking. This will
allow cracks to close completely if the stress across them becomes compressive.
In multiaxial stress states these observations are generalized through the
concept of surfaces of failure and flow in stress space. These surfaces are
fitted to experimental data. The surfaces used are shown in
Figure 4
and
Figure 5.
Failure Surface
You can specify failure ratios to define the shape of the failure surface
(possibly as a function of temperature and predefined field variables). Four
failure ratios can be specified:
The ratio of the ultimate biaxial compressive stress to the ultimate
uniaxial compressive stress.
The absolute value of the ratio of the uniaxial tensile stress at
failure to the ultimate uniaxial compressive stress.
The ratio of the magnitude of a principal component of plastic strain at
ultimate stress in biaxial compression to the plastic strain at ultimate stress
in uniaxial compression.
The ratio of the tensile principal stress at cracking, in plane stress,
when the other principal stress is at the ultimate compressive value, to the
tensile cracking stress under uniaxial tension.
Default values of the above ratios are used if you do not specify them.
Response to Strain Reversals
Because the model is intended for application to problems involving
relatively monotonic straining, no attempt is made to include prediction of
cyclic response or of the reduction in the elastic stiffness caused by
inelastic straining under predominantly compressive stress. Nevertheless, it is
likely that, even in those applications for which the model is designed, the
strain trajectories will not be entirely radial, so that the model should
predict the response to occasional strain reversals and strain trajectory
direction changes in a reasonable way. Isotropic hardening of the “compressive”
yield surface forms the basis of this aspect of the model's inelastic response
prediction when the principal stresses are dominantly compressive.
Calibration
A minimum of two experiments, uniaxial compression and uniaxial tension, is
required to calibrate the simplest version of the concrete model (using all
possible defaults and assuming temperature and field variable independence).
Other experiments may be required to gain accuracy in postfailure behavior.
Uniaxial Compression and Tension Tests
The uniaxial compression test involves compressing the sample between two
rigid platens. The load and displacement in the direction of loading are
recorded. From this, you can extract the stress-strain curve required for the
concrete model directly. The uniaxial tension test is much more difficult to
perform in the sense that it is necessary to have a stiff testing machine to be
able to record the postfailure response. Quite often this test is not
available, and you make an assumption about the tensile failure strength of the
concrete (usually about 7%–10% of the compressive strength). The choice of
tensile cracking stress is important; numerical problems may arise if very low
cracking stresses are used (less than 1/100 or 1/1000 of the compressive
strength).
Postcracking Tensile Behavior
The calibration of the postfailure response depends on the reinforcement
present in the concrete. For plain concrete simulations the stress-displacement
tension stiffening model should be used. Typical values for
are 0.05 mm (2 × 10−3 in) for a normal concrete to 0.08 mm (3 ×
10−3 in) for a high-strength concrete. For reinforced concrete
simulations the stress-strain tension stiffening model should be used. A
reasonable starting point for relatively heavily reinforced concrete modeled
with a fairly detailed mesh is to assume that the strain softening after
failure reduces the stress linearly to zero at a total strain of about 10 times
the strain at failure. Since the strain at failure in standard concretes is
typically 10−4, this suggests that tension stiffening that reduces
the stress to zero at a total strain of about 10−3 is reasonable.
This parameter should be calibrated to a particular case.
Postcracking Shear Behavior
Combined tension and shear experiments are used to calibrate the
postcracking shear behavior in
Abaqus/Standard.
These experiments are quite difficult to perform. If the test data are not
available, a reasonable starting point is to assume that the shear retention
factor, ,
goes linearly to zero at the same crack opening strain used for the tension
stiffening model.
Biaxial Yield and Flow Parameters
Biaxial experiments are required to calibrate the biaxial yield and flow
parameters used to specify the failure ratios. If these are not available, the
defaults can be used.
Temperature Dependence
The calibration of temperature dependence requires the repetition of all the
above experiments over the range of interest.
Comparison with Experimental Results
With proper calibration, the concrete model should produce reasonable
results for mostly monotonic loadings. Comparison of the predictions of the
model with the experimental results of Kupfer and Gerstle (1973) are shown in
Figure 6
and
Figure 7.
Elements
Abaqus/Standard
offers a variety of elements for use with the smeared crack concrete model:
beam, shell, plane stress, plane strain, generalized plane strain,
axisymmetric, and three-dimensional elements.
For general shell analysis more than the default number of five integration
points through the thickness of the shell should be used; nine thickness
integration points are commonly used to model progressive failure of the
concrete through the thickness with acceptable accuracy.
Output
In addition to the standard output identifiers available in
Abaqus/Standard
(Abaqus/Standard Output Variable Identifiers),
the following variables relate specifically to material points in the smeared
crack concrete model:
CRACK
Unit normal to cracks in concrete.
CONF
Number of cracks at a concrete material point.
References
Crisfield, M.A., “Snap-Through
and Snap-Back Response in Concrete Structures and the Dangers of
Under-Integration,” International Journal for
Numerical Methods in
Engineering, vol. 22, pp. 751–767, 1986.
Hillerborg, A., M. Modeer, and P. E. Petersson, “Analysis
of Crack Formation and Crack Growth in Concrete by Means of Fracture Mechanics
and Finite Elements,” Cement and Concrete
Research, vol. 6, pp. 773–782, 1976.
Kupfer, H.B., and K. H. Gerstle, “Behavior
of Concrete under Biaxial Stresses,” Journal
of Engineering Mechanics Division,
ASCE, vol. 99853, 1973.