4.11. Geomechanics

Geomechanics encompasses the mechanical behavior of soil, rock, and aggregate materials in both their natural and man-made states. Applications for these material models include footings and pilings, tunneling, excavations, seismic events, and compaction or consolidation. Other applications include material failure along stress concentrations or weak regions such as the failure of intact and jointed rock, masonry structures, and the crushing or failure behavior of concrete.

The material behavior is characterized by an initial elastic response followed by plastic deformation and unloading from the plastic state recovers the elastic deformation. The plasticity is a result of the microscopic behavior of the material particles and includes shear loading that causes particles to move past one another, changes in void or fluid content that result in volumetric plasticity, and exceeding the cohesive forces between the particles or aggregates.

Also see:

4.11.1. Understanding the Material Models for Geomechanics

Material behavior is defined by rate-independent plasticity. The general concepts for the models are given in Understanding the Plasticity Models.

Several of the yield surfaces for the geomechanics materials are defined by multiple surfaces. The elastic domain is the interior of the intersection of the surfaces, resulting in a continuous but non-smooth composite yield surface. Yielding and plastic deformation on one of the yield surfaces is the same as that of other rate-independent plasticity models. Plasticity and yielding can also occur at the intersection of two or more yield surfaces, resulting in plastic behavior that is a combination of the multiple yield surfaces [1].

When yielding occurs on multiple yield surfaces, the plastic strain increment is given by:

where is the increment in plastic strain, ranges over the set of active yield surfaces, is the magnitude of the plastic strain increment, and is the flow potential.

The hardening or softening of the yield surfaces is a function of the hardening variable given by:

where is the increment in the hardening variable and is the hardening modulus.

The incremental equations for rate-independent multisurface plasticity result in a system of nonlinear equations for the plastic strain increment, hardening variable increment and stress. One of two well-known integration methods, commonly called return mapping methods, is used to solve these equations: (1) the closest-point projection; and (2) the cutting-plane method. The details for these methods can be found in [2], but other considerations apply.

The closest-point projection is the more accurate method but generally has a smaller radius of convergence and can diverge when the stress state is near the intersection of two or more yield surfaces. This method returns a consistent material tangent, generally resulting in rapid convergence of the Newton-Raphson iterations. For multisurface yielding, however, the consistent tangent can be incorrect when the activity of the yield surfaces changes over an increment; although this behavior has no effect on the accuracy of the converged solution, it can affect the convergence rate. The closest-point projection is used to solve all of the geomechanics models except for the Mohr-Coulomb and jointed rock models.

While the cutting-plane method tends to be more stable, it is less accurate, and does not have a consistent material tangent (instead using the material elastic tangent). The accuracy of the model can be checked heuristically by reducing the strain increments to determine if the solution changes. Use of an elastic material tangent reduces the rate of convergence of the Newton-Raphson iterations for the global system of equations, but does not affect the accuracy of the converged solution. The cutting-plane method is used to solve Mohr-Coulomb and jointed rock models.

4.11.1.1. Guidance for Using the Geomechanics Material Models

Simulations involving the geomechanics materials can sometimes present solution difficulties:

  • Cutting-plane algorithm

    The Mohr-Coulomb and jointed rock materials use the cutting-plane algorithm to solve the system of nonlinear equations that determine the material state at each element integration point. The algorithm tends to be more robust for these materials because of the non-smooth yield surface; however, it also results in an elastic material tangent, reducing the convergence rate of the Newton-Raphson iterations for the global system of equations. It may be necessary therefore to increase the maximum number of equilibrium iterations (NEQIT). Also, the elastic material tangent affects the results of any subsequent analyses that depend on the material stiffness (such as a perturbation analysis).

  • Material models that use the closest-point projection

    When the plastic flow direction is not associated with the stress direction, the material tangent is unsymmetric. This unassociated flow occurs when the plastic potential is different from the yield surface. For such cases, the convergence rate can be increased by using the unsymmetric solver (NROPT).

  • Plastic behavior and failure

    The plastic behavior for some of the geomechanical materials represents the sudden failure typical of geotechnical and aggregate materials. The behavior includes perfect plasticity and non-smooth or continuous softening, where further loading does not experience any resistance to deformation. The behavior can cause convergence difficulty for the global Newton-Raphson iterations.

    The best method for overcoming the problem is to design a simulation that includes only localized yielding, with surrounding structure that prevents uncontrolled deformation of the areas of the mesh that have failed or softened. In some cases where the material has failed, stabilization or using a transient analysis can prevent uncontrolled deformation. In general, displacement boundary conditions that are applied or transferred through the mesh to material that has failed are more easily solved than similar force boundary conditions.

    Material failure can often result in a localized area of high deformation (representing the total failure of the material). The localization is a smeared representation of a crack, shear band,or other failure, but the details of the localization often depend on the simulation setup (including mesh size, substep size or loading rate, and material tangent formulation). While the predicted failure region might be accurate, the details of the localization are often unreliable.

4.11.1.2. Material Solution Option

Softening in the stress-strain response, often used to represent failure in geomechanics materials, can cause convergence difficulty for the global Newton-Raphson iterations. To help overcome convergence issues, you can use the material elastic tangent instead of the consistent material tangent.

A consistent tangent generally results in fewer Newton-Raphson iterations. When using an elastic tangent, it may be necessary to increase the maximum number of Newton-Raphson iterations (NEQIT).

For each of the geomechanics materials (Cam-clay soil, Mohr-Coulomb, jointed rock, Drucker-Prager concrete, and Menetrey-Willam concrete), a data table with the material solution option (TBOPT = MSOL) defines the form of the material tangent. The default value is used if the parameter is not redefined (or is defined as zero) in the material data table (TB).

Table 4.4: Material Solution Parameter

ConstantPropertyDefault ValueRange
C1 Elastoplastic tangent flag Material-dependent

1 = Consistent tangent

2 = Elastic tangent


The default value for the Cam-clay, Drucker-Prager, and Menetrey-Willam models returns the consistent algorithmic tangent. For the Mohr-Coulomb and jointed rock materials, the elastic tangent is the default material tangent.

Example 4.30: Drucker-Prager Concrete Material Using Elastic Tangent

/prep7
TB,CONCR,1,,,MSOL
TBDATA,1,2     

4.11.1.3. Nomenclature

Following are the common symbols used in the geomechanical theory documentation:

SymbolDefinition SymbolDefinition
Plastic strain Initial pressure for porous elasticity
Magnitude of plastic strain increment Cohesion
Plastic potential Residual cohesion
Hardening variable or swell index Friction angle
Hardening modulus Residual friction angle
Yield function Dilatancy angle
Residual yield function Tensile strength
Pressure Residual tensile strength
Modified stress invariant Residual strength coupling flag
Stress Failure plane orientation angles
Von Mises effective stress Compressive strength
2nd principal invariant of the stress tensor Tensile strength
3rd principal invariant of the stress tensor Biaxial compressive strength
Effective shear stress Tension dilatancy parameter
Mean stress Tension dilatancy parameter points
Normal stress Compression dilatancy parameter
Cam-clay yield surface hardening and softening variable Plastic strain at uniaxial compressive strength
Initial porosity Plastic strain at transition point
Porosity Ultimate effective plastic strain
Load angle Softening plastic strain points
Hardening/softening function Relative stress at onset of nonlinear hardening
Tensile hardening/softening function Residual compressive relative stress
Compressive hardening/softening function Residual tensile relative stress
Yield surface shape parameter Residual relative stress at transition
Initial yield surface size parameter Residual tensile relative stress points
Minimum yield surface size parameter Plastic strain limit in tension
Parameter for slope of the critical state line Mode I fracture energy in tension
Yield surface shape parameter Mode I fracture energy in compression
Plastic slope parameter----

4.11.1.4. References

  1. Simo, J. C., Kennedy, J. G., & Govindjee, S. (1988). Non-smooth multisurface plasticity and viscoplasticity. Loading/unloading conditions and numerical algorithms. International Journal for Numerical Methods in Engineering. 26(10), 2161-2185.

  2. Simo, J. C. & Hughes, T. J. R. (1998). Computational Inelasticity. New York: Springer.

  3. Roscoe, K. & Burland, J. B. (1968). On the generalised stress-strain behaviour of wet clay. Engineering Plasticity. 169(1), 535-609.

  4. Wood, D. M. (1990). Soil Behaviour and Critical State Soil Mechanics. Cambridge: Cambridge University Press.

  5. Bazant, Z. P. & Oh, B. H. (1983). Crack band theory for fracture of concrete. Materials and Structures. 16(3), 155-177.

4.11.2. Cam-clay

The modified Cam-clay plasticity model [3] is based on the critical state concept and is commonly used for soil simulation. The critical state concept is a material phenomenon common in the shear deformation of soils where the stress and volume remain constant after a critical state of deformation, and further loading causes increasing plastic strain but no increase in volume or stress. The Cam-clay plasticity model, combined with porous elasticity, models the effect of voids on the elastic behavior of the material.

The following topics for the Cam-clay model are available:

Also see Sequential Construction of an Embankment on a Clay Layer in the Technology Showcase: Example Problems.

4.11.2.1. Yield Surface

The yield surface is a function of pressure and a modified stress invariant :

(4–79)

where , , and are material parameters. The modified Cam-clay model uses an associated flow rule, so the flow potential is the same as the yield function and plastic flow is normal to the yield surface.

Figure 4.35: Modified Cam-clay Yield Surface Plotted in Principal Stresses

Modified Cam-clay Yield Surface Plotted in Principal Stresses

Pressure is a function of elastic volumetric strain and void ratio and is defined by the porous elasticity model. The modified stress invariant is:

where is the von Mises effective stress

and

where and are the second and third principal invariants of the stress tensor, respectively, and is a material parameter that modifies the shape of the yield surface.

In the following figure, the yield surface is plotted in the octahedral plane (the intersection of the yield surface with a plane of constant pressure and plotted in principal stress space):

Figure 4.36: Modified Cam-clay Yield Surface in the Octahedral Plane

Modified Cam-clay Yield Surface in the Octahedral Plane

gives a circular yield surface and decreasing values cause the surface to become more triangular. To ensure that the yield surface is convex, is required.

In the following figure, the yield surface is plotted as effective stress versus pressure (q-p plane):

Figure 4.37: Modified Cam-clay Yield Surface in the q-p Plane

Modified Cam-clay Yield Surface in the q-p Plane

The critical state is then defined by the intersection of the yield surface with a line of slope . The intersection, or critical state, is at the maximum value of effective stress.

Stress states above the critical state line are over-consolidated, and stress states below the line are normally-consolidated.

The plastic strain is normal to the yield surface so that yielding for over-consolidated states results in a decreasing plastic volumetric strain, and yielding for normally-consolidated states results in an increase in plastic volumetric strain. At the intersection of the critical state line with the yield surface, yielding causes no change in the plastic volume.

The part of the yield surface above the critical state line is called the dry yield surface and the yield surface below it is the wet yield surface. The parameter modifies the shape of the yield surface, and separate values are used for the wet and dry surfaces. On the dry yield surface, is proportional to the cohesion. For no cohesion, and the yield surface includes no tensile stress states. As increases, the elastic range for tensile stress states increases. For typical soil behavior, increases by only a few percent, whereas for high cohesion materials such as metal foams, can increase significantly.

Similar to a yield stress, is a hardening or softening variable controlling the size of the yield surface. The initial size is defined by the parameter , and the hardening/softening behavior is given by:

where is the initial porosity and is the swell index defined with the porous elasticity parameters. is a parameter for plastic slope from the assumed linear relationship between void ratio and natural log of the pressure during plastic deformation .

Figure 4.38: Linear Relationship Between Void Ratio and Natural Log of Pressure

Linear Relationship Between Void Ratio and Natural Log of Pressure

Because is the slope of this line during elastic loading, must be larger than and is typically 3 to 6 times the value of .

The following figure shows the behavior of , where hardening occurs when the material is consolidating due to yielding above the critical state, and softening occurs when the materials is dilating due to yielding below the critical state:

Figure 4.39: Hardening and Softening Behavior

Hardening and Softening Behavior

When defining an initial stress state, the stress should be inside or on the yield surface. If the initial stress is outside the yield surface, is adjusted so that the initial stress state is on the yield surface. Softening of can reduce the size of the yield surface and cause numerical difficulties. To prevent the yield surface from becoming too small, the minimum size is set via the parameter .

4.11.2.2. Defining the Cam-clay Model

To use the Cam-clay model:

  1. Define the porous elasticity material model (TB,PELAS,,,,POISSON) and required constants.

  2. Define the Cam-clay plasticity parameters (TB,SOIL,,,,CAMCLAY).

  3. Specify the required Cam-clay material model constants (TBDATA).

  4. Define the initial stress state () (INISTATE).

Table 4.5: Cam-clay Model Constants

ConstantMeaningPropertyUnitRange
C1 Plastic slope parameter--
C2 Slope of critical state line--
C3 Initial yield surface sizeForce/Length2
C4 Minimum yield surface sizeForce/Length2
C5 Dry part of yield surface modifier--
C6 Wet part of yield surface modifier--
C7 Anisotropic yield surface parameter--

Define temperature- or field-dependent data for the data tables via the TBTEMP or TBFIELD commands, respectively.

Example 4.31: Defining the Cam-clay Model

/prep7

! Porous elasticity
Kappa = 0.0024
NU0    = 0.279
pt_el = 5.835
E0    = 0.34 
p0    = 69

TB,PELAS,1,,,POISSON
TBDATA,1, Kappa, pt_el, NU0, E0

! Cam-clay Plasticity
Lambda_F = 0.014
Mc       = 1.24
a0       = 35
ah_min   = 0.35
Beta_Dry = 1
Beta_Wet = 1 
Ks       = 1

TB,SOIL,1,,,CAMCLAY
TBDATA,1, Lambda_F, Mc, a0, ah_min  
TBDATA,5, Beta_Dry, Beta_Wet, Ks

/solu
!define initial stress state
INISTATE,set,dtyp,stre
INISTATE,defi,all,,,,-p0,-p0,-p0,0,0,0

4.11.3. Mohr-Coulomb

Aggregate materials such as soil, rock and concrete begin to plastically deform when the shear stress exceeds the internal friction resistance between the material particles. The friction resistance is a function of the normal force between the particles. Use the Mohr-Coulomb material model to represent such aggregate materials.

The following topics about the Mohr-Coulomb model are available:

Also see Active and Passive Lateral Earth Pressure Analysis and Suction Pile Analysis in the Technology Showcase: Example Problems .

4.11.3.1. Understanding the Mohr-Coulomb Model

4.11.3.1.1. Yield Surface

The model defines yielding when the combination of normal stress and shear stress reaches the cohesion of the material particles. Figure 4.40: Mohr-Coulomb Yield Surface as Shear vs. Normal Stress shows the yield criteria plotted as a function of the shear and normal stress. Yielding occurs when the shear stress reaches this criterion:

where is the shear stress, is the cohesion, is the normal stress, and is the inner friction angle.

Figure 4.40: Mohr-Coulomb Yield Surface as Shear vs. Normal Stress

Mohr-Coulomb Yield Surface as Shear vs. Normal Stress

The friction angle is proportional to the stress required to shear particles past one another. Materials such as loose sand, in which the particles can easily move past one another, have relatively low friction angle.

Generalization to any state of stress gives the Mohr-Coulomb yield surface:

where:

and the stress invariants are:

Figure 4.41: Mohr-Coulomb Yield Surface in Principal Stresses

Mohr-Coulomb Yield Surface in Principal Stresses

The Mohr-Coulomb yield surface can be combined with a tension-failure surface to limit the material strength in tension. The tension yield surface is defined by a Rankine yield surface:

where is the tensile strength.

Figure 4.42: Mohr-Coulomb with Tension-Failure Yield Surface in Principal Stress

Mohr-Coulomb with Tension-Failure Yield Surface in Principal Stress

After the material yields, the inner friction angle , cohesion , and tensile strength harden or soften as a function of the Mohr-Coulomb and tension hardening variables, and . These hardening variables evolve by summing the incremental values for each substep and the incremental values are given by:

The material parameters are then given by:

where , , and are the initial values and , , and are user-defined hardening/softening scaling functions.

At zero plastic deformation, the scaling functions are equal to 1.0. It is good practice to define the value in the input so that the scaling functions are interpolated as expected for small values of plastic deformation. If any scaling function is defined, the undefined scaling functions are assumed constant .

If no scaling functions for the yield surface parameters are defined, then after initial yield, the yield surfaces reduce to the residual yield surfaces. If any parameter-scaling functions are defined, the residual parameters are ignored.

The residual Mohr-Coulomb yield surface is:

where is the residual cohesion and is the residual inner friction angle. To ensure a decrease in the yield strength, the residual yield surface constants should satisfy:

Otherwise, the residual cohesion is redefined as:

The residual Rankine yield surface is:

Tension cutoff is defined at the intersection of the Mohr-Coulomb and Rankine yield surfaces:

When the tension cutoff is defined, the residual Mohr-Coulomb yield surface parameters should satisfy:

Otherwise, the residual cohesion is redefined as:

The residual strength of the Mohr-Coulomb surface and the tension-failure surfaces can be coupled so that when yielding occurs on one surface, the other yield surface is reduced to the residual strength yield function. The value of the flag defines the coupling:

4.11.3.1.2. Plastic Flow Potentials

Three plastic flow potentials are available for the Mohr-Coulomb material model.

The Mohr-Coulomb flow potential is:

where is the dilatancy angle. This is the only flow potential available when no scaling functions are defined for the yield surface parameters.

The Mohr-Coulomb potential is not smooth, requiring the material solution algorithm to use the cutting-plane return algorithm.

A smooth potential is obtained from the Mohr-Coulomb potential by fixing the value of :

  • Setting gives a potential surface that intersects the Mohr-Coulomb surface at the tension meridians (intersection of the Mohr-Coulomb surface with the positive principal stress axes):

  • Setting gives a potential surface that intersects the Mohr-Coulomb surface at the compression meridians (the intersection of the Mohr-Coulomb surface with the negative principal stress axes):

A smooth potential enables the use of the closest-point projection method.

Using a smooth potential requires the material scaling parameters. If none are defined, the parameters are constant.

Figure 4.43: Flow Potentials in the Octahedral Plane

Flow Potentials in the Octahedral Plane

Similar to the yield surface parameters, the dilatancy angle can be defined to evolve based on the scaling:

where is the initial value of the dilatancy angle and is the dilatancy angle scaling function.

When the Mohr-Coulomb material includes the Rankine tension failure surface, two plastic flow potentials for yielding on this surface are available. The potential is selected based on the flow potential used for the Mohr-Coulomb yield surface.

When the Mohr-Coulomb potential is used with the Mohr-Coulomb surface, the tension failure potential is the Rankine tension failure surface.

When one of the smooth potentials is used with the Mohr-Coulomb surface, the plastic flow potential for the tension failure surface is:

where are the ordered principal stresses.

4.11.3.2. Defining the Mohr-Coulomb Model

To use the Mohr-Coulomb model:

  1. Define the isotropic or anisotropic elastic behavior (via MP commands or via the elastic material data table (TB,ELASTIC)).

  2. Define the Mohr-Coulomb yield surface and flow potential (TB,MC,,,,BASE).

  3. Input the model constants (TBDATA).

Table 4.6: Mohr-Coulomb Model Constants

ConstantMeaningPropertyUnitRange
C1 Initial inner friction angleDegrees
C2 Initial cohesionForce/Length2
C3 Dilatancy angleDegrees
C4 Residual inner friction angleDegrees
C5 Residual cohesionForce/Length2

To define the optional tension Rankine yield surface, define the material data table (TB,MC,,,,RCUT) and input the following constants (TBDATA):

Table 4.7: Tension Rankine Yield Surface Constants (Optional)

ConstantMeaningPropertyUnitRange
C1 Initial tensile strengthForce/Length2
C2 Residual tensile strengthForce/Length2

To define the optional residual strength coupling, define the material data table (TB,MC,,,,RSC) and input the following constant (TBDATA):

Table 4.8: Residual Strength Coupling Constants (Optional)

ConstantMeaningPropertyUnitRange
C1 Residual strength coupling flag--

To define the optional flow potential, define the material data table (TB,MC,,,,POTN), then define the flow type (TBDATA):

Table 4.9: Plastic Potential (Optional)

ConstantMeaningPropertyUnitRange
C1Plastic potential Type

0, 1, 2

Type = 0 – Mohr-Coulomb potential surface (default).
Type = 1 – Smoothed compression-meridian potential surface.
Type = 2 – Smoothed tension-meridian potential surface.

To define the optional scaling function for the inner friction angle, define the material data table (TB,MC,,,,FRICTION) and define the points (TBPT):

Table 4.10: Friction Angle Scaling Function (Optional)

ConstantMeaningPropertyUnitRange
X Hardening variable--
Y Scaling function--

To define the optional scaling function for the cohesion, define the material data table (TB,MC,,,,COHESION) and define the points (TBPT):

Table 4.11: Cohesion Scaling Function (Optional)

ConstantMeaningPropertyUnitRange
X Hardening variable--
Y Scaling function--

To define the optional scaling function for the tension strength, define the material data table (TB,MC,,,,TENSION) and define the points (TBPT):

Table 4.12: Tension Strength Scaling Function (Optional)

ConstantMeaningPropertyUnitRange
X Hardening variable--
Y Scaling function--

To define the optional scaling function for the dilatation angle, define the material data table (TB,MC,,,,DILATATION) and define the points (TBPT).

Table 4.13: Dilatation Angle Scaling Function (Optional)

ConstantMeaningPropertyUnitRange
X Hardening variable--
Y Scaling function--

Define temperature- or field-dependent data for the data tables via the TBTEMP or TBFIELD commands, respectively.

Example 4.32: Defining the Mohr-Coulomb Model

/prep7
! ELASTIC CONSTANTS
MP,EX,1,20.0E9 
MP,NUXY,1,0.2

! Mohr Coulomb yield surface

phi=30         ! friction angle 
c=8.7e6        ! cohesion
psi=15         ! dilatancy angle
phir=20        ! residual friction angle
cr=0.8*c       ! residual cohesion

sigt=3e6       ! tensile strength
sigtr=0.8*sigt ! residual tensile strength


TB,MC,1,,,BASE
TBDATA,1,phi,c,psi,phir,cr

TB,MC,1,,,RCUT
TBDATA,1,sigt,sigtr

TB,MC,1,,,RSC
TBDATA,1,0


Example 4.33: Defining the Mohr-Coulomb Model with Softening

/prep7
YoungsModulus= 20.0E9
NU    =0.2

mp,ex,1,YoungsModulus
mp,prxy,1,NU

! nonlinear parameters for intact rock
phig=30       ! friction angle
cg=8.7e6      ! cohesion
psig=15       ! dilatancy angle
sigt=3e6      ! tension failure

TB,MC,1,,,BASE
TBDATA,1,phig,cg,psig

TB,MC,1,,,RCUT
TBDATA,1,sigt

! plastic potential
TB,MC,1,,,POTN
TBDATA,1,1

! softening functions
TB,MC,1,,,friction
TBPT,, 0.0 ,1
TBPT,, 1E-4,35/30

TB,MC,1,,,dilatation
TBPT,, 0.0 ,1
TBPT,, 1E-4,0.9

TB,MC,1,,,cohesion
TBPT,, 0.0 ,1
TBPT,, 1E-4,0.95
TBPT,, 1E-3,0.75

4.11.4. Jointed Rock

Geologic and aggregate materials can have inhomogeneous behavior that causes weakness, represented as joints, along planes within the material. Examples of such planes are stratification planes, and geometric stress concentrations or regions of strain localization.

The behavior of the planes is defined by an anisotropic Mohr-Coulomb yield surface that depends on the normal and shear stress on the face of the planes. The planes are embedded in a base material that can be modeled as either elastic, or elastic-plastic with an isotropic Mohr-Coulomb yield surface.

4.11.4.1. Mohr-Coulomb Base Material

The base material for the jointed rock model can be either elastic, or elastic-plastic defined via the Mohr-Coulomb model. The Mohr-Coulomb base material includes the Mohr-Coulomb yield surface and a tension-cutoff yield surface.

If the Mohr-Coulomb base material is not defined, the program considers the base material to be elastic, and the only plastic behavior occurs in the joints.

4.11.4.2. Anisotropic Mohr-Coulomb Joints

Failure of the joints is defined via an anisotropic Mohr-Coulomb failure model that can be combined with tensile failure. The anisotropic Mohr-Coulomb yield function defines the yield surface on a failure plane in terms of the resolved normal and shear stress on the plane:

where is the magnitude of the resolved shear stress on the failure plane, is the normal stress on the failure plane, is the failure plane friction angle, and is the failure plane cohesion.

After the failure plane yields, the residual yield surface of the failure plane is given by:

The flow potential for the failure plane is:

A tension cutoff can be defined to limit the normal stress on the plane. The tension-cutoff yield surface is:

where is the tensile strength of the failure plane.

The residual tensile strength yield surface is:

The following figure shows the anisotropic Mohr-Coulomb yield surface for a plane with normal :

Figure 4.44: Joint Failure Surface Superimposed on Mohr-Coulomb with Tension-Cutoff Yield Surface

Joint Failure Surface Superimposed on Mohr-Coulomb with Tension-Cutoff Yield Surface

The failure plane yield surface is superimposed on the base material yield surface, a Mohr-Coulomb model with a tension cutoff.

The following figure shows the composite yield surface in the tension region of the stress space in which the failure crops off a portion of the yield surface of the base material:

Figure 4.45: Composite Yield Surface of Jointed Rock with Mohr-Coulomb Base Material

Composite Yield Surface of Jointed Rock with Mohr-Coulomb Base Material

The joints are oriented relative to the element coordinate system. By default, the plane has a normal in the z direction. The plane can be reoriented by defining angles and , where the plane is first rotated about the negative z axis by angle and then rotated about the plane's y axis by angle . The orientation of the failure plane normal is:

The following figure shows an example of how a failure plane is reorineted:

Figure 4.46: Reorienting the Failure Plane

Reorienting the Failure Plane

The plane has a local coordinate system initially aligned with the element coordinate system. It is reoriented by rotation about the z axis, then rotation about the plane v axis.

The failure surfaces of the base material and the joints can be coupled by defining the residual strength coupling flag , as follows:

4.11.4.3. Defining the Jointed Rock Model

4.11.4.3.1. Part 1: Defining the Base Material

To use the jointed rock material model with an elastic base:

  1. Define the isotropic or anisotropic elastic behavior (via MP commands or via the elastic material data table (TB,ELASTIC)).

  2. Proceed to Part 2: Defining the Joints.

    In this case, the base model uses the elastic constants (from MP or TB,ELASTIC), and the only plastic behavior occurs in the joints.

To use the jointed rock material model with an elastic-plastic (Mohr-Coulomb) base:

  1. Define the isotropic or anisotropic elastic behavior (via MP commands or via the elastic material data table (TB,ELASTIC)).

  2. Define the jointed rock material data table (TB,JROCK,,,,BASE).

  3. Input the required base material constants (TBDATA).

Table 4.14: Jointed Rock Constants for the Base Material

ConstantMeaningPropertyUnitRange
C1 Initial inner friction angle for base materialDegrees
C2 Initial cohesion for base materialForce/Length2
C3 Dilatancy angle for base materialDegrees
C4 Residual inner friction angle for base materialDegrees
C5 Residual cohesion for base materialForce/Length2

To define the optional tension Rankine yield surface for the base material, define the material data table (TB,JROCK,,,,RCUT) and input the following constants (TBDATA):

Table 4.15: Tension Yield Surface for the Base Material (Optional)

ConstantMeaningPropertyUnitRange
C1 Initial tensile strengthForce/Length2
C2 Residual tensile strengthForce/Length2

To define the optional residual strength coupling, define the material data table (TB,JROCK,,,,RSC) and input the following constant (TBDATA):

Table 4.16: Residual Strength Coupling for the Base Material (Optional)

ConstantMeaningPropertyUnitRange
C1 Residual strength coupling flag--

4.11.4.3.2. Part 2: Defining the Joints

You can define up to four joints. Each joint requires failure plane yield and flow-potential parameters (TB,JROCK,,,,FPLANE). Defining this table creates a new set of joint parameters.

After defining a new joint, you can then specify the optional failure plane tensile strength (TB,JROCK,,,,FTCUT) and the failure plane orientation (TB,JROCK,,,,FORIE).

To initiate a new joint set and specify the failure plane yield and flow-potential parameters:

  1. Define the material data table (TB,JROCK,,,,FPLANE).

  2. Input the required joint constants (TBDATA).

Table 4.17: Jointed Rock Constants for the Joint

ConstantMeaningPropertyUnitRange
C1 Inner friction angle for jointDegrees
C2 Initial cohesion for jointForce/Length2
C3 Dilatancy angle for jointDegrees
C4 Residual inner friction angle for jointDegrees
C5 Residual cohesion for jointForce/Length2

To define the optional tension yield surface for the joint, define the material data table (TB,JROCK,,,,FTCUT) and input the following constants (TBDATA):

Table 4.18: Tension Yield Surface for the Joint (Optional)

ConstantMeaningPropertyUnitRange
C1 Initial tensile strength for jointForce/Length2
C2 Residual tensile strength for jointForce/Length2

To define the optional orientation of the failure plane for the joint, define the material data table (TB,JROCK,,,,FORIE) and input the following constants (TBDATA):

Table 4.19: Orientation of the Joint Failure Plane (Optional)

ConstantMeaningPropertyUnit
C1 Reorientation of joint about the negative z axisDegrees
C2 Reorientation of joint about the new y axisDegrees

Define temperature- or field-dependent data for the data tables via the TBTEMP or TBFIELD commands, respectively.

Example 4.34: Defining the Jointed Rock Model with an Elastic-Plastic (Mohr-Coulomb) Base

/prep7
! ELASTIC CONSTANTS
MP,EX,1,20.0E9 
MP,NUXY,1,0.2

! Base Mohr Coulomb yield surface

phi=30         ! friction angle 
c=8.7e6        ! cohesion
psi=15         ! dilatancy angle
phir=20        ! residual friction angle
cr=0.8*c       ! residual cohesion

sigt=3e6       ! tensile strength
sigtr=0.8*sigt ! residual tensile strength


TB,JROCK,1,,,BASE
TBDATA,1,phi,c,psi,phir,cr

TB,JROCK,1,,,RCUT
TBDATA,1,sigt,sigtr

TB,JROCK,1,,,RSC
TBDATA,1,0

! Joint 1

phi_j=25        ! friction angle 
c_j=1E6         ! cohesion 
psi_j=25.0      ! dilatancy angle 
phir_j=25       ! residual friction angle 
cr_j=0.8*c_j    ! residual cohesion 

alpha_j=-90     ! negative rotation about element Z-axis
beta_j=45       ! positive rotation about plane Y-axis

TB,JROCK,1,,,FPLANE
TBDATA,1,phi_j,c_j,psi_j,phir_j,cr_j

TB,JROCK,1,,,FORIE
TBDATA,1,alpha_j,beta_j

! Joint 2

phi_j=25        ! friction angle 
c_j=1E6         ! cohesion 
psi_j=25.0      ! dilatancy angle 
phir_j=25       ! residual friction angle 
cr_j=0.8*c_j    ! residual cohesion 

sigt_j=0.5E6    ! tensile strength 
sigtr_j=0.1E6   ! residual tension strength 

alpha_j=-45     ! negative rotation about element Z-axis
beta_j=45       ! positive rotation about plane Y-axis

TB,JROCK,1,,,FPLANE
TBDATA,1,phi_j,c_j,psi_j,phir_j,cr_j

TB,JROCK,1,,,FTCUT
TBDATA,1,sigt_j,sigtr_j

TB,JROCK,1,,,FORIE
TBDATA,1,alpha_j,beta_j

4.11.5. Drucker-Prager Concrete

Single-surface Drucker-Prager models often do not represent the large differences in tensile and compressive behavior of concrete. To represent the weak tension behavior, you can use either of the following:

  • A second Drucker-Prager yield surface

  • A composite surface consisting of a Rankine tension failure surface and a Drucker-Prager surface in compression

The Drucker-Prager yield surfaces are similar to the linear form of the Extended Drucker-Prager yield surface.

Joints represent failure along planes of weakness, such as geometric stress concentrations or inhomogeneous material regions. The Drucker-Prager concrete model can be combined with the anisotropic Mohr-Coulomb model to represent joints in the concrete material. (The anisotropic Mohr-Coulomb model is the same as the jointed rock failure plane model.) The composite Drucker-Prager and Rankine surface cannot be used with joints.

The following topics about the Drucker-Prager concrete material model are available:

Also see Load-Limit Analysis of a Reinforced Concrete Slab in the Technology Showcase: Example Problems.

4.11.5.1. Yield Surfaces and Flow Potentials

4.11.5.1.1. Tension and Tension-Compression Surfaces

For loading in tension and tension-compression, you can use either a Drucker-Prager yield surface or a Rankine tension failure surface to define the yield condition.

Drucker-Prager Yield Surface

The tension and tension-compression Drucker-Prager yield surface is given by:

where and are constants defined by the uniaxial tensile strength and uniaxial compressive strength :

and are hardening/softening functions in compression and tension, which depend on the stress and hardening variables . For a definition of and the specific forms of the hardening/softening functions, see Hardening, Softening and Dilatation (HSD) Behavior.

where is the compression dilatancy parameter.

For the Drucker-Prager yield surface in tension and tension-compression, the flow potential is:

Rankine Tension Failure Surface

The Rankine tension failure surface is useful for representing the brittle tensile behavior of concrete. The surface defines yielding when the largest principal stress exceeds the tensile strength and is given by:

where is the tensile strength and

The stress invariants are:

The flow potential is the Rankine surface when there is only one positive principal stress. When more than one principal stress is positive, a smooth approximation of the Rankine surface is used as the potential surface, resulting in nonassociated flow.

4.11.5.1.2. Compression Surface

For compressive loading, the Drucker-Prager yield surface is:

where the constants and are related to the biaxial compressive strength and uniaxial compressive strength by:

The flow potential for yielding in compression is:

where the compression dilatancy parameter.

Figure 4.47: Composite Surface with Drucker-Prager Compression and Drucker-Prager Tension Yield Surfaces

Composite Surface with Drucker-Prager Compression and Drucker-Prager Tension Yield Surfaces

Figure 4.48: 2D Yield Surfaces Showing Drucker-Prager and Rankine Surfaces

2D Yield Surfaces Showing Drucker-Prager and Rankine Surfaces

4.11.5.2. Defining the Drucker-Prager Concrete Model

To use the Drucker-Prager concrete model:

  1. Define the isotropic or anisotropic elastic behavior (via MP commands or the elastic material data table (TB,ELASTIC)).

  2. Define the Drucker-Prager concrete yield-strength parameters (TB,CONCR,,,,DP).

  3. Optional: Define the Rankine tension parameter (TB,CONCR,,,,RCUT).

The following table shows the Drucker-Prager yield strength parameters (TB,CONCR,,,,DP):

Table 4.20: Drucker-Prager Concrete Model Constants

ConstantMeaningPropertyUnitRange
C1 Uniaxial compressive strengthForce/Length2
C2 Uniaxial tensile strength

Note:  This constant is not used if a Rankine tension table is defined.

Force/Length2
C3 Biaxial compressive strengthForce/Length2

To use the Rankine tension surface, define the material data table (TB,CONCR,,,,RCUT) and input the uniaxial tension strength:

Table 4.21: Rankine Surface Constant

ConstantMeaningPropertyUnitRange
C1 Uniaxial tensile strengthForce/Length2

When using the composite Drucker-Prager with Drucker-Prager tension surface, dilatancy parameters are optional. The composite Drucker-Prager with Rankine surface requires a compression dilatancy parameter.

If you define dilatancy parameters, the following considerations apply:

To define dilatancy parameters, define the material data table (TB,CONCR,,,,DILA) and input the following constants (TBDATA):

Table 4.22: Drucker-Prager Dilatancy Parameters

ConstantMeaningPropertyUnitRange
C1 Tensile dilatancy parameter

Note:  This constant is not used if a Rankine tension table is defined

--
C2 Compression dilatancy parameter--

4.11.5.2.1. Hardening, Softening and Dilatation (HSD) Behavior

When using the Drucker-Prager tension surface, you can use any of the HSD models. When using the Rankine tension surface, the linear, exponential and fracture energy HSD models are available. If no HSD model is defined, the program uses perfect plasticity.

The hardening and softening behavior of the yield surfaces is defined by and , respectively. In the case of the steel reinforcement HSD model, you can also specify the evolution of the tensile dilatation .

The and functions evolve based on the hardening variable . The general form of the hardening variable evolution is:

where:

= number of active yield surfaces
= strength parameter
= hardening or softening function value
= magnitude of the plastic strain increment for each active yield surface.

Except in special cases, the hardening variable is different from the equivalent plastic strain.

To help overcome mesh-dependent softening behavior, the exponential softening model for the tension yield function and the fracture energy model for the compression yield function are normalized with an effective element length . For 3D elements, the effective element length is the cube root of the integration point volume. For 2D elements, it is the square root of the integration point volume. For 1D elements, the effective element length is the integration point volume.

With the HSD models, the total energy dissipated in a localized failure or crack band approaches the area-specific fracture energy [5].

4.11.5.2.1.1. Exponential HSD Model (TB,CONCR,,,,HSD2)

The hardening yield function in compression, , for is given by the power-hardening-law function (Equation 4–82).

The softening yield function in the range is given by:

The softening yield function in the range is given by:

The yield function in tension, , is given by an exponential softening function where the volumetric energy dissipated in softening is proportional to the Mode I area-specific fracture energy in tension :

(4–80)

where:

where is the effective element length and E is the Young's Modulus, giving the following relation for energy dissipated during softening of the tension yield function:

To limit the softening, set the relative residual stress level to give perfect plasticity when the yield function (Equation 4–80) is less than .

Figure 4.49: Exponential Softening in Compression and Tension

Exponential Softening in Compression and Tension

To define exponential softening, define the material data table (TB,CONCR,,,,HSD2) and input the following constants (TBDATA):

Table 4.23: Exponential HSD Constants

ConstantMeaningPropertyUnitRange
C1 Plastic strain at uniaxial compressive strength--
C2 Plastic strain at transition from power law to exponential softening--
C3 Relative stress at start of nonlinear hardening--
C4 Residual relative stress at --
C5 Residual compressive relative stress--
C6 Mode I area-specific fracture energyForce/Length
C7 Residual tensile relative stress--

4.11.5.2.1.2. Steel Reinforcement HSD Model (TB,CONCR,,,,HSD4)

The yield function in compression, , is given by the nonlinear hardening function and a linear softening function, with the ultimate effective plastic strain defined as .

The yield function in tension, , is defined by a multilinear law for the residual relative strength of steel reinforcement. The multilinear softening is defined by the relative stress vs. effective plastic strain points defined in this figure:

Figure 4.50: Steel Reinforcement Softening in Compression and Tension

Steel Reinforcement Softening in Compression and Tension

The dilatancy factor in tension, , is defined by a multilinear function. The points defining the dilatancy factor as a function of effective plastic strain are shown in this figure:

Figure 4.51: Steel Reinforcement Dilatancy in Tension

Steel Reinforcement Dilatancy in Tension

The parameter for tension dilatancy (entered in the dilatancy table (TB,CONCR,,,,DILA)) is not used. The compression dilatancy is constant.

To define the steel reinforcement HSD model, define the material data table (TB,CONCR,,,,HSD4) and input the following constants (TBDATA):

Table 4.24: Steel Reinforcement HSD Constants

ConstantMeaningPropertyUnitRange
C1 Plastic strain at uniaxial compressive strength--
C2 Relative stress at onset of nonlinear hardening--
C3 Residual compressive relative stress--
C4 Softening plastic strain point 1 --
C5 Softening plastic strain point 2--
C6 Softening plastic strain point 3--
C7 Residual tensile relative stress point 1--
C8 Residual tensile relative stress point 2--
C9 Residual tensile relative stress point 3--
C10 Initial tensile dilatancy parameter--
C11 Tensile dilatancy parameter point 1--
C12 Tensile dilatancy parameter point 2--
C13 Tensile dilatancy parameter point 3--

4.11.5.2.1.3. Fracture Energy HSD Model (TB,CONCR,,,,HSD5)

The yield function in compression, , is given by a nonlinear hardening function and a volumetric fracture energy based softening function. The hardening from to is given above in Equation 4–82. Softening begins to occur at and the yield function is:

(4–81)

where:

where is the effective element length and is the Young's Modulus.

Then, , giving a constant volumetric energy dissipated during softening proportional to the Mode I area-specific fracture energy in compression .

The relative residual stress level can be set to limit the softening and give perfect plasticity when the yield function (Equation 4–81) is less than .

The yield function in tension, , is given by the exponential softening function described in Exponential HSD Model (TB,CONCR,,,,HSD2).

Figure 4.52: Fracture Energy Softening in Compression and Tension

Fracture Energy Softening in Compression and Tension

To define fracture energy softening, define the material data table (TB,CONCR,,,,HSD5) and input the following constants (TBDATA):

Table 4.25: Fracture Energy HSD Constants

ConstantMeaningPropertyUnitRange
C1 Plastic strain at uniaxial compressive strength--
C2 Relative stress at onset of nonlinear hardening--
C3 Residual compressive relative stress--
C4 Mode 1 area-specific fracture energy in compressionForce/Length
C5 Mode 1 area-specific fracture energy in tensionForce/Length
C6 Residual tensile relative stress--

4.11.5.2.1.4. Linear HSD Model (TB,CONCR,,,,HSD6)

The yield function in compression, , is given by a nonlinear hardening function and a linear softening function. The relative stress level at the onset of nonlinear hardening is with a hardening yield function:

(4–82)

At , the peak compression strength is reached and softening starts with:

At , the relative stress level is the residual value.

The yield function in tension, , is given by a linear softening function. The relative yield stress is equal to 1 at initial yielding, decreases to the relative residual stress when the effective plastic strain is , and is constant for .

Figure 4.53: Linear Softening in Compression and Tension

Linear Softening in Compression and Tension

To define linear softening, define the material data table (TB,CONCR,,,,HSD6) and input the following constants (TBDATA):

Table 4.26: Linear HSD Constants

ConstantMeaningPropertyUnitRange
C1 Plastic strain at uniaxial compressive strength--
C2 Ultimate effective plastic strain in compression --
C3 Relative stress at onset of nonlinear hardening--
C4 Residual compressive relative stress--
C5 Plastic strain limit in tension--
C6 Residual tensile relative stress--

4.11.5.2.2. Representing Weak Joints in Concrete

You can combine the Drucker-Prager concrete model with failure planes to represent weak joints in the concrete. The failure planes are represented by anisotropic Mohr-Coulomb yield surfaces on oriented planes. The process is described in Defining the Joints in the Jointed Rock material model discussion.


Important:  You can define joints in the Drucker-Prager concrete model only with the elastic, perfectly plastic base model with associated flow (TB,CONCR,,,,DP). Joints are not applicable when specifying the optional Drucker-Prager dilitancy parameters (TB,CONCR,,,,DILA).


You can define up to four joints. Each joint requires failure plane yield- and flow-potential parameters (TB,CONCR,,,,FPLANE). Each joint definition creates a new set of joint parameters.

After defining a new joint, you can then specify the optional failure plane tensile strength (TB,CONCR,,,,FTCUT) and the failure plane orientation (TB,CONC,,,,FORIE).

To initiate a new joint parameter set and specify the failure plane yield and flow-potential parameters:

  1. Define the material data table (TB,CONCR,,,,FPLANE).

  2. Input the required joint constants (TBDATA).

Table 4.27: Jointed Rock Constants for the Joint

ConstantMeaningPropertyUnitRange
C1 Inner friction angle for jointDegrees
C2 Initial cohesion for jointForce/Length2
C3 Dilatancy angle for jointDegrees
C4 Residual inner friction angle for jointDegrees
C5 Residual cohesion for jointForce/Length2

To define the optional tension yield surface for the joint, define the material data table (TB,CONCR,,,,FTCUT) and input the following constants (TBDATA):

Table 4.28: Tension Yield Surface for the Joint (Optional)

ConstantMeaningPropertyUnitRange
C1 Initial tensile strength for jointForce/Length2
C2 Residual tensile strength for jointForce/Length2

To define the optional orientation of the failure plane for the joint, define the material data table (TB,CONCR,,,,FORIE) and input the following constants (TBDATA):

Table 4.29: Orientation of the Joint Failure Plane (Optional)

ConstantMeaningPropertyUnit
C1 Reorientation of joint about the negative z axisDegrees
C2 Reorientation of joint about the new y axisDegrees

4.11.5.2.3. Example Drucker-Prager Concrete Definitions

Example 4.35: Defining the Drucker-Prager Concrete Model with Dilitancy Parameters

/prep7
! ELASTIC CONSTANTS
MP,EX,1,20.0E9 
MP,NUXY,1,0.2

! Base Drucker-Prager concrete 
Rc=30.0E6        ! uniaxial compressive strength
Rt=3.0E6         ! uniaxial tensile strength
Rb=36.0E6        ! biaxial compressive strength

delta_t=0.25 ! dilatancy factor tension
delta_c=1.0  ! dilatancy factor compression

! Linear hardening/softening 
kappa_cm=0.0025-Rc/20.0E9
kappa_cr=0.0025
omega_ci=0.33
omega_cr=0.1

kappa_tr=0.0005
omega_tr=0.2

TB,CONCR,1,,,DP
TBDATA,1,Rc,Rt,Rb

TB,CONCR,1,,,DILA
TBDATA,1,delta_t,delta_c

TB,CONCR,1,,,HSD6
TBDATA,1,kappa_cm,kappa_cr,omega_ci,omega_cr,kappa_tr,omega_tr

Example 4.36: Defining the Drucker-Prager Concrete Model with Failure Joints

/prep7
! ELASTIC CONSTANTS
MP,EX,1,20.0E9 
MP,NUXY,1,0.2

! Base Drucker-Prager concrete 
Rc=30.0E6        ! uniaxial compressive strength
Rt=3.0E6         ! uniaxial tensile strength
Rb=36.0E6        ! biaxial compressive strength

TB,CONCR,1,,,DP
TBDATA,1,Rc,Rt,Rb

! Joint 1

phi_j=25        ! friction angle 
c_j=1E6         ! cohesion 
psi_j=25.0      ! dilatancy angle 
phir_j=25       ! residual friction angle 
cr_j=0.8*c_j    ! residual cohesion 

alpha_j=-45     ! negative rotation about element Z-axis
beta_j=45       ! positive rotation about plane Y-axis

TB,CONCR,1,,,FPLANE
TBDATA,1,phi_j,c_j,psi_j,phir_j,cr_j

TB,CONCR,1,,,FORIE
TBDATA,1,alpha_j,beta_j

! Joint 2

phi_j=25        ! friction angle 
c_j=1E6         ! cohesion 
psi_j=25.0      ! dilatancy angle 
phir_j=25       ! residual friction angle 
cr_j=0.8*c_j    ! residual cohesion 

sigt_j=0.5E6    ! tensile strength 
sigtr_j=0.1E6   ! residual tension strength 

alpha_j=-90     ! negative rotation about element Z-axis
beta_j=45       ! positive rotation about plane Y-axis

TB,CONCR,1,,,FPLANE
TBDATA,1,phi_j,c_j,psi_j,phir_j,cr_j

TB,CONCR,1,,,FTCUT
TBDATA,1,sigt_j,sigtr_j

TB,CONCR,1,,,FORIE
TBDATA,1,alpha_j,beta_j


4.11.6. Menetrey-Willam

The Menetrey-Willam constitutive model [1] is based on the Willam-Warnke yield surface [2], incorporating dependence on three independent invariants of the stress tensor.

The Willam-Warnke surface is similar to the Mohr-Coulomb surface, but without the sharp edges that can cause difficulty in the Mohr-Coulomb surface stress solution. It also shares some characteristics with the Drucker-Prager model and can model similar materials. The Menetrey-Willam model, however, is generally better for simulating the behavior of bonded aggregates such as concrete.

4.11.6.1. Yield Surfaces and Flow Potentials

The material parameters defining the yield function are:

  • Yield strengths in uniaxial tension ()

  • Uniaxial compression ()

  • Biaxial compression ()

Parameter hardening and softening is defined by:

where is a material parameter, and and are the compression and tension-hardening/softening functions that depend on the compression or tension-hardening variables and .

The yield surface in Haigh-Westergaard stress coordinates is given by:

where and are functions of the material parameters and the hardening softening functions

and

The Haigh-Westergaard stress coordinates are:

with the first principal invariant of the stress tensor, and and the second and third principal invariants of the deviatoric stress tensor:

The flow potential is:

where:

and is the dilatancy angle. The flow potential gives unassociated plastic flow and results in an unsymmetric consistent material tangent.

Figure 4.54: Yield Surface for the Menetrey-Willam Model

Yield Surface for the Menetrey-Willam Model

4.11.6.2. Defining the Menetrey-Willam Model

Define the isotropic or anisotropic elastic behavior via MP commands or via the elastic table (TB,ELASTIC).

To define the Menetrey-Willam yield strength parameters, define the material data table (TB,CONCRETE,,,,MW) and input the following constants (TBDATA):

Constant Meaning Property Units Range
C1 Uniaxial compressive strengthForce/Length2 >
C2 Uniaxial tensile strengthForce/Length2 > 0
C3 Biaxial compressive strengthForce/Length2 >

To define the dilatancy angle, define the material data table (TB,CONCRETE,,,,DILA) and input the following constant (TBDATA):

Constant Meaning Property Units Range
C1 Dilatancy angleDegrees

4.11.6.3. Hardening and Softening Functions

The hardening-softening behavior of the yield surfaces is defined by the hardening-softening functions and . These functions depend on the compression- and tension-hardening variables and which evolve according to the work hardening expressions: [4,5]

where and are compression and tension weight functions given by:

Specifying the hardening-softening behavior is optional. If a hardening-softening model is not specified, the material is considered to be elastic-perfectly plastic.

To help overcome mesh-dependent softening behavior, the fracture-energy model is normalized with an effective element length :

  • 3D elements – = cube root of the integration-point volume

  • 2D elements – = square root of the integration-point volume

  • 1D elements – = integration-point volume

With the fracture-energy model, the total energy dissipated in a localized failure or crack band approaches the area-specific fracture energy [3].

4.11.6.3.1. Linear Softening

The yield function in compression, , is given by a nonlinear hardening function and a linear softening function. The relative stress level at the start of nonlinear hardening is , with a hardening yield function of:

(4–83)

At , the peak compression strength is reached and softening starts with:

(4–84)

At , the relative stress level is the residual value .

The yield function in tension, , is given by a linear softening function. The relative yield stress is equal to 1 at initial yielding, decreases to the relative residual stress when the effective plastic strain is , and is constant for .

Figure 4.55: Linear Softening in Compression and Tension

Linear Softening in Compression and Tension

To specify linear softening, define the material data table (TB,CONCRETE,,,,HSD6) and input the following constants (TBDATA):

Constant Meaning Property Units Range
C1 Plastic strain at uniaxial compressive strength-
C2 Ultimate effective plastic strain in compression -
C3 Relative stress at start of nonlinear hardening-
C4 Residual compressive relative stress-
C5 Plastic strain limit in tension-
C6 Residual tensile relative stress-
4.11.6.3.2. Exponential Softening

The hardening yield function in compression, , for is given by the power hardening law function in Equation 4–83. The softening yield function in the range is given by:

and for :

The yield function in tension, , is given by an exponential softening function where the volumetric energy dissipated in softening is proportional to the mode I area specific fracture energy in tension :

(4–85)

where:

where is the effective element length and is the Young’s Modulus, giving the following relation for energy dissipated during softening of the tension yield function:

To limit the softening, set the relative residual stress level to give perfect plasticity when the yield function calculated via Equation 4–85 is less than .

Figure 4.56: Exponential Softening in Compression and Tension

Exponential Softening in Compression and Tension

To define exponential softening, define the material data table (TB,CONCRETE,,,,HSD2) and input the following constants (TBDATA):

Constant Meaning Property Units Range
C1 Plastic strain at uniaxial compressive strength-
C2 Plastic strain at transition from power law to exponential softening-
C3 Relative stress at start of nonlinear hardening-
C4 Residual relative stress at -
C5 Residual compressive relative stress-
C6 Mode I area-specific fracture energyForce-length/length2
C7 Residual tensile relative stress-

4.11.6.4. Example Menetrey-Willam Definitions

Example 4.37: Defining the Menetrey-Willam Concrete Model

/prep7
! ELASTIC CONSTANTS
MP,EX,1,20.0E9 
MP,NUXY,1,0.2

! base Menetrey-Willam concrete 
Rc=30.0E6        ! uniaxial compressive strength
Rt=3.0E6         ! uniaxial tensile strength
Rb=36.0E6        ! biaxial compressive strength

psi=10           ! dilatancy angle

! linear hardening softening 
kappa_cm=0.0025-Rc/20.0E9
kappa_cr=0.0025
omega_ci=0.33
omega_cr=0.1

kappa_tr=0.0005
omega_tr=0.2

TB,CONCRETE,1,,,MW
TBDATA,1,Rc,Rt,Rb

TB,CONCRETE,1,,,DILA
TBDATA,1,psi

TB,CONCRETE,1,,,HSD6
TBDATA,1,kappa_cm,kappa_cr,omega_ci,omega_cr,kappa_tr,omega_tr

4.11.6.5. References

  1. Menetrey, P. (1994). Numerical analysis of punching failure in reinforced concrete structures. Diss. Ecole Polytechnique Federale de Lausanne, Lausanne.Infoscience. Web.

  2. Willam, K. J. & Warnke, E. P. Constitutive models for the triaxial behavior of concrete. Seminar on Concrete Structures Subjected to Triaxial Stresses. International Association for Bridge and Structural Engineering. 19, 1-30.

  3. Bazant, Z. P. & Oh, B. H. (1985). Microplane model for progressive fracture of concrete and rock. Journal for Engineering Mechanics. 111, 559-582.

  4. Jirasek, M. (2000). Numerical modeling of deformation and failure of materials. Lecture Notes.

  5. Chen, W. F. (1994). Constitutive equations for engineering materials. Plasticity and Modeling (Vol. 2). Amsterdam. Elsevier Publishing.