U.S. Department of Transportation
Federal Highway Administration
1200 New Jersey Avenue, SE
Washington, DC 20590
202-366-4000


Skip to content
Facebook iconYouTube iconTwitter iconFlickr iconLinkedInInstagram

Federal Highway Administration Research and Technology
Coordinating, Developing, and Delivering Highway Transportation Innovations

Report
This report is an archived publication and may contain dated technical, contact, and link information
Publication Number: FHWA-HRT-05-062
Date: May 2007

Users Manual for LS-DYNA Concrete Material Model 159

PDF Version (1.49 KB)

PDF files can be viewed with the Acrobat® Reader®

Chapter 2. Theoretical Manual

Elastic Update

Concrete is typically assumed to be isotropic; therefore, Hooke's Law is used for the elastic stress-strain relationship. Hooke's law depends on two elastic constants: the bulk modulus (K) and the shear modulus (G).

Plastic Update

Following an initial elastic phase, concrete will yield and possibly fail, depending on the state of stress (or type of test being simulated). The yield stresses are defined by a three-dimensional yield surface. The functional form of the yield surface is discussed in the next section. This section briefly discusses the plasticity algorithm responsible for setting the initial yield stresses.

At each time step, the stress is updated from the strain rate increments and the time step via an incremental form of Hooke's Law (an elastic increment). This updated stress is called the trial elastic stress and is denoted σijT. If the trial elastic stress state lies on or inside the yield surface, the behavior is elastic, and the plasticity algorithm is bypassed. If the trial elastic stress state lies outside the yield surface, the behavior is elastic-plastic (with possible damage, hardening, and rate effects), and the plasticity algorithm returns the stress state to the yield surface. This elastic-plastic stress is called the inviscid stress, and is denoted σijP.

Details of the return algorithm are well documented.(20,21) For those knowledgeable about plasticity, the algorithm enforces the plastic consistency condition with associated flow. For efficiency, the algorithm employs subincrementation, rather than iteration, to ensure accurate return of the stress state to the yield surface. Subincrementation is invoked when the current strain increment exceeds a maximum strain limit specified by the user, or defaulted by the model.

The associative return algorithm predicts dilation of the concrete after the yield surface is engaged in the tensile and low confining pressure regimes. Modeling dilation is one reason for employing a sophisticated plasticity algorithm rather than a simple Mises return algorithm. Simple return algorithms do not model dilation.

Yield Surface

The concrete model is a cap model with a smooth or continuous intersection between the failure surface and hardening cap. The general shape of the yield surface in the meridonal plane is shown in Figure 15 and Figure 16. This surface uses a multiplicative formulation to combine the shear (failure) surface with the hardening compaction surface (cap) smoothly and continuously. The smooth intersection eliminates the numerical complexity of treating a compressive 'corner' region between the failure surface and cap. This type of model is often referred to as a smooth cap model or as a continuous surface cap model (CSCM).

Figure 15. Illustration. General shape of the concrete model yield surface in three dimensions. This plot shows the yield surface plotted in principal stress space. It is a closed surface with a discontinuous apex when the pressure is tensile, and a rounded capped surface when the pressure is compressive. The cross section changes from triangular in tension to irregular hexagonal then circular in compression.

Figure 15. Illustration. General shape of the concrete model yield surface in three dimensions.

Figure 16. Illustration. General shape of the concrete model yield surface in two dimensions in the meridonal plane. This plot shows the yield surface plotted as shear strength versus pressure. It is a closed surface with a discontinuous apex when the pressure is tensile, and a rounded capped surface when the pressure is compressive. The shear surface is approximately linear until it intersects the elliptical cap. The intersection is continuous and smooth.

Figure 16. Illustration. General shape of the concrete model yield surface in two dimensions in the meridonal plane.

The original formulation of the smooth cap model was developed as a function of two stress invariants.(22) The developer extended the formulation to three stress invariants and verified the three-invariant formulation by comparing smooth-cap results with analytical results for the axisymmetric compression of a Mohr-Coulomb medium around a circular hole.(20,23)

The yield surface is formulated in terms of three stress invariants because an isotropic material has three independent stress invariants. The model uses J¢-the first invariant of the stress tensor, J′2-the second invariant of the deviatoric stress tensor, and 3-the third invariant of the deviatoric stress tensor. The invariants are defined in terms of the deviatoric stress tensor, Sij and pressure, P, as shown in Figure 17:

Figure 17. Equation. Stress invariant J subscript 1, J prime subscript 2, and J prime subscript 3. The first invariant of the stress factor J subscript 1 is equal the product of three and P. The second invariant of the deviatoric stress tensor J prime subscript 2 is equal to the product of one-half, the deviatoric stress tensor S subscript lowercase IJ, and the deviatoric stress tensor S subscript lowercase IJ. The third invariant of the deviatoric stress tensor J prime subscript 3 is equal to the product of one-third, the deviatoric stress tensor S subscript lowercase IJ, the deviatoric stress tensor S subscript lowercase JK, and the deviatoric stress tensor S subscript lowercase KI.

Figure 17. Equation. Stress invariant J 1, J2, and J3.

The three invariant yield function is based on these three invariants, and the cap hardening parameter, κ, as shown in Figure 18:

Figure 18. Equation. Yield function lowercase F. Yield surface lowercase F as a function of J subscript 1, J prime subscript 2, J prime subscript 3, and kappa is equal to J prime subscript 2 minus the product of the squared Rubin strength reduction factor R superscript 2, the squared shear failure surface function F subscript lowercase F, and the cap failure surface function F subscript lowercase C.

Figure 18. Equation. Yield function f.

Here Ff is the shear failure surface, Fc is the hardening cap, and ℜ is the Rubin three-invariant reduction factor. Multiplying the cap ellipse function by the shear surface function allows the cap and shear surfaces to take on the same slope at their intersection, as discussed in subsequent paragraphs.

Trial elastic stress invariants are temporarily updated via the trial elastic stress tensor, σT. These are denoted J1T, J¢2T, and J¢3T. Elastic stress states are modeled when f (J1T, J¢2T, J¢3T, κT) < 0. Elastic-plastic stress states are modeled when f (J1T, J¢2T, J¢3T, κT) > 0. In this case, the plasticity algorithm returns the stress state to the yield surface so that f (J1P, J¢2P, J¢3P, κP) = 0.

Shear Failure Surface. The strength of concrete is modeled by the shear surface in the tensile and low confining pressure regimes. The shear surface F subscript f is defined along the compression meridian as shown in Figure 19:

Figure 19. Equation. Shear failure surface function F subscript lowercase F. Shear failure surface F subscript lowercase F as a function of the first invariant stress factor J subscript 1 is equal to alpha minus the product of lambda times the exponential function raised to the power of the product of negative beta and the first invariant of the stress tensor J subscript 1, plus the product of theta times the first invariant of the stress tensor J subscript 1.

Figure 19. Equation. Shear failure surface function Ff.

Here the values of α, β, l, q are selected by fitting the model surface to strength measurements from TXC tests conducted on plain concrete cylinders (and then adjusting these parameters to account for compaction and damage). The TXC data are typically plotted as principal stress difference versus pressure. The principal stress difference (axial stress minus confining stress) is equal to the square root of 3J¢2. The shear surface is shown in Figure 20, Figure 21, and Figure 22 for typical concrete values.

Figure 20. Graph. Schematic of shear surface. The Y-axis is the shear failure surface function F subscript lowercase F in units of megapascals. It ranges from 0 to 150 megapascals. The X-axis is pressure in units of megapascals. It ranges from negative 10 to 60 megapascals. One curve is plotted which increases in a slightly nonlinear manner from position negative 6,0 to position 60,146. The curve is continuous and smooth.

psi = 145.05 MPa

Figure 20. Graph. Schematic of shear surface.

Figure 21. Graph. Schematic of two-part cap function. The Y-axis is the square root of the cap surface function F subscript lowercase C. It is unitless and ranges from negative 0.2 to 1.2. The X-axis is pressure in megapascals. It ranges from negative 10 to 60 megapascals. One curve is plotted, and one part of the curve is one-quarter of an ellipse, beginning at position 0, 1 and ending at position 47,0. The second part of the curve is a straight line which runs along the lowercase Y equals 0 axis following completion of the quarter ellipse.

psi = 145.05 MPa

Figure 21. Graph. Schematic of two-part cap function.

Figure 22. Graph. Schematic of multiplicative formulation of the shear and cap surfaces.

psi = 145.05 MPa

Figure 22. Graph. Schematic of multiplicative formulation of the shear and cap surfaces.

Cap Hardening Surface. The strength of concrete is modeled by a combination of the cap and shear surfaces in the low to high confining pressure regimes. More importantly, the cap is used to model plastic volume change related to pore collapse (although the pores are not explicitly modeled). The initial location of the cap determines the onset of plasticity in isotropic compression and uniaxial strain. The elliptical shape of the cap allows the onset for isotropic compression to be greater than the onset for uniaxial strain, in agreement with shear enhanced compaction data. Without ellipticity, a "flat" cap would produce identical onsets. The motion of the cap determines the shape (hardening) of the pressure-volumetric strain curves via fits with data. Without cap motion, the pressure-volumetric strain curves would be perfectly plastic.

The isotropic hardening cap is a two-part function that is either unity or an ellipse, as shown in Figure 21. When the stress state is in the tensile or very low confining pressure region, the cap function is unity: yield strength via the equation in Figure 18 is independent of the cap. When the stress state is in the low to high confining pressure regimes, the cap function is an ellipse: yield strength depends on both the cap and shear surface formulations. The two-part cap function is defined as shown in Figure 23:

Figure 23. Equation. Cap failure surface function F subscript lowercase C. The hardening cap surface F subscript lowercase C as a function of J subscript 1 and kappa, is equal to one minus the quotient of the product of the first invariant of the stress tensor minus the element length L as a function of kappa K and the absolute value of the first invariant of the stress tensor J subscript 1 minus the element length L as a function of kappa K, all plus the first invariant of the stress tensor J subscript 1 minus the element length L as a function of kappa K, all over the product of 2 and the squared difference of the current cap location as a function of kappa K minus the element length L as a function of kappa K.

Figure 23. Equation. Cap failure surface function Fc.

where L(κ)is defined as shown in Figure 24:

Figure 24. Equation. L of kappa. The element length L as a function of kappa K is equal to the piecewise function where the function is equal to kappa K when kappa K is greater than kappa not K subscript 0, and equal to kappa not K subscript 0 otherwise.

Figure 24. Equation. L of kappa.

The equation in Figure 23 is equal to unity for J1 £ L(κ). The equation in Figure 23 describes the ellipse for J1 > L(κ). The intersection of the shear surface and the cap is at J1 = κ. κ0 is the value of J1 at the initial intersection of the cap and shear surfaces before hardening is engaged (before the cap moves). The equation in Figure 24 restrains the cap from retracting past its initial location at κ0.

A simpler, but less complete, way of writing the equations in Figure 23 and Figure 24 is shown in Figure 25:

Figure 25. Equation. Simple cap failure surface function F subscript lowercase C. The hardening cap surface F subscript lowercase C as a function of J subscript 1 and kappa, is equal to the piecewise function where the function is equal to 1 minus the quotient the squared difference of the first invariant of the stress tensor J subscript 1 minus kappa K, all over the squared difference of the current cap location minus kappa K when the first invariant of the stress tensor J subscript 1 is greater than or equal to kappa K, and the function is equal to 1 otherwise.

Figure 25. Equation. Simple cap failure surface function Fc.

The intersection of the cap with the J1 axis is at J1 = X(κ). This intersection depends on the cap ellipticity ratio R, where R is the ratio of its major to minor axes, as shown in Figure 26:

Figure 26. Equation. X as a function of kappa. The current cap location X as a function of kappa K is equal to the sum of L as a function of kappa K and the product of the cap aspect ratio R, the shear failure surface F subscript lowercase F, and the element length L as a function of kappa K.

Figure 26. Equation. X as a function of kappa.

The cap moves to simulate plastic volume change. The cap expands (X(κ) and κ increase) to simulate plastic volume compaction. The cap contracts (X(κ) and κ decrease) to simulate plastic volume expansion, called dilation. The motion (expansion and contraction) of the cap is based on the hardening rule, shown in Figure 27:

Figure 27. Equation. Plastic volume strain epsilon superscript lowercase P subscript lowercase V. The plastic volumetric strain epsilon superscript lowercase P subscript lowercase V is equal to the product of the maximum plastic volume strain W and the quantity of one minus the exponential function to the power of the difference between the product of the negative cap linear parameter D subscript 1 and the quantity of the difference between the current cap location X minus the initial cap location X subscript 0 all minus the product of the quadratic shape parameter and the squared quantity of the difference between the current cap location X minus the initial cap location X subscript 0.

Figure 27. Equation. Plastic volume strain ε pv.

Here ε pv is the plastic volume strain, W is the maximum plastic volume strain, and D1 and D2 are model input parameters. X0 is the initial location of the cap when κ = κ0.

The five input parameters (X0, W, D1, D2, and R) are obtained from fits to the pressure-volumetric strain curves in isotropic compression and uniaxial strain. X0 determines the pressure at which compaction initiates in isotropic compression. R, combined with X0, determines the pressure at which compaction initiates in uniaxial strain. D1 and D2 determine the shape of the pressure-volumetric strain curves. W determines the maximum plastic volume compaction.

Rubin Scaling Function. Concrete fails at lower values of J¢2 (principal stress difference) for TXE and TOR tests than it does for TXC tests conducted at the same pressure. This was previously demonstrated in Figure 1 through Figure 4. This situation indicates that concrete strength depends on the third invariant of the deviatoric stress tensor, J¢3. When viewed in the deviatoric plane, a three invariant yield surface is triangular or hexagonal in shape, as shown in Figure 28.

The Rubin scaling function ℜ determines the strength of concrete for any state of stress relative to the strength for TXE.( 24) Strengths like TXE and TOR are simulated by scaling back the TXE shear strength by the Rubin function: ℜFf. The Rubin function ℜ is a scaling function that changes the shape (radius) of the yield surface in the deviatoric plane as a function of angle Beta hat as shown in Figure 28. This shape may be a circle (Drucker-Prager, Maximum Octahedral Shear Stress, von Mises), a hexagon (Mohr-Coulomb), or an irregular hexagon-like shape (Willam-Warnke) in which each of six sides is quadratic (rather than linear) between the TXC and TXE states.

For comparison, a two-invariant model cannot simultaneously model different strengths in TXC, TXE, and TOR. When viewed in the deviatoric plane, the two-invariant yield surface is a circle, as shown in Figure 28. A two-invariant formulation is modeled with the Rubin function equal to ℜ=1 at all angles around the circle. This means that the TXC, TOR, and TXE strengths are modeled the same.

Figure 28. Illustration. Example two- and three-invariant shapes of the concrete model in the deviatoric plane. This figure shows a circle, which is the two stress invariant surface, enclosing an irregular hexagonal shaped surface, which is the three stress invariant surface. Also shown are the meridian positions for triaxial compression, torsion, and triaxial extension. In triaxial compression, the two-invariant circle and three-invariant hexagonal coincided at one apex of the hexagonal. In triaxial extension, the largest difference exists between the two-invariant circles and three-invariant hexagonal. Also shown is an arrow running along the radius of the circle from the center to the two-invariant surface indicating that this distance is the value of the shear surface function F subscript lowercase F. Also shown is a second, shorter arrow running from the center to the three-invariant surface, indicating that this distance is the value of the Rubin scaling function R times the shear surface function F subscript lowercase F.

Figure 28. Illustration. Example two- and three-invariant shapes of the concrete model in the deviatoric plane.

The angle is confined to the range -p/6 < Beta hat< p and is related to the invariants J¢2 and J¢3, as shown in Figure 29:

Figure 29. Equation. Angle beta hat in deviatoric plane. The sine of three times the angle in deviatoric plane beta hat is equal to J hat subscript 3, which is equal to the quotient of the product of 3, the square root of 3, and the third invariate of the deviatoric stress tensor J prime subscript 3 all over the product of 2 and the second invariant of the deviatoric stress tensor raised to the power of 3 halves.

Figure 29. Equation. Angle beta hat in the deviatoric plane.

J hat subscript 3 is a normalized invariant which remains in the range -1 < J hast subscript 3 < 1. For the standard laboratory tests just discussed, the values of Beta hat and J hat subscript 3 are shown in Figure 30:

Figure 30. Equation. Relationship between beta hat and J hat. The angle in deviatoric plane beta hat is equal to the quotient of pi over 6, when the third invariant of the deviatoric stress tensor is equal to 1, for TXC. The angle in deviatoric plane beta hat is equal to 0, when the third invariant of the deviatoric stress tensor is equal to 0, for TOR. The angle in deviatoric plane beta hat is equal the quotient of negative pi over 6, when the third invariant of the deviatoric stress tensor is equal to negative 1, for TXE.

Figure 30. Equation. Relationship between beta hat and J hat.

The form of the Rubin scaling function is shown in Figure 31:

Figure 31. Equation. Rubin scaling function R. Rubin scaling function R is equal to the quotient of the sum of negative lowercase B subscript 1 and the square root of the difference of lowercase B subscript 1 squared minus the product of 4, lowercase B subscript 2, and lowercase B subscript 0, all over the product of 2 and lowercase B subscript 2.

Figure 31. Equation. Rubin scaling function ℜ.

The value of ℜ depends on the state of stress through the angle Beta hat, and on experimentally determined values (fits to data) for Q1 and Q2 as functions of pressure. Strength in TOR is modeled as Q1Ff. Strength in TXE is modeled as Q2Ff.

The Rubin three-invariant formulation is implemented because it is more flexible in fitting data than the more commonly used Willam-Warnke formulation. Four example fits of the Rubin formulation are listed below.

1.      Most general fit: the shape of the yield surface in the deviatoric plane transitions with pressure from triangular, to irregular hexagonal, to circular. Input values for all eight parameters: αλ, l1, b1, q1 and α2, l2, b2, q2 are shown in Figure 32.

Figure 32. Equation. Most general form for Q subscript 1 and Q subscript 2. The Rubin scaling torsion function Q subscript 1 is equal to the shear surface torsion constant alpha subscript one minus the product of lambda subscript one and the exponential function of the product of negative beta subscript 1 and J subscript 1, all added to the product of theta and J subscript 1. The Rubin scaling triaxial extension function Q subscript 2 is equal alpha subscript 2 minus the product of lambda subscript 2 and the exponential function of the product of negative beta subscript 2 and J subscript 1, all added to the product of theta subscript 2 and J subscript 1.

Figure 32. Equation. Most general form for Q1 and Q2.

2.     Mohr-Coulomb fit: a straight line fit between the TXE and TXC states. The strength ratios are estimated from the Mohr-Coulomb friction angle as shown in Figure 33.

Figure 33. Equation. Mohr-Coulomb form for Q subscript 1, Q subscript 2. The Rubin scaling torsion function Q subscript 1 is equal to the quotient of the product of the square root of three and the Rubin scaling triaxial extension function Q subscript 2 all over the sum of one and the Rubin scaling triaxial extension function Q subscript 2.

Figure 33. Equation. Mohr-Coulomb form for Q1, Q2.

3.     Two-parameter fit: the strength ratios Q1 and Q2 remain constant with pressure. Input α1 and α2, with all other Rubin parameters set equal to zero. This allows the user to model the yield surface as an irregular (bulging) hexagonal shape in the deviator plane.

4.     Willam-Warnke fit: select Q2 as a constant or as a function of pressure. Fit Q1 to the Willam-Warnke TOR surface, as shown in Figure 34.

Figure 34. Equation. Willam-Warnke form for Q subscript 1. The Rubin scaling torsion function Q subscript 1 is equal to the quotient of sum of the product of the square root of three and the quantity 1 minus the square of the Rubin scaling triaxial extension function Q subscript 2, added to the product of the quantity of the product of 2 and the Rubin scaling triaxial extension function Q subscript 2 minus 1 and the square root of the sum of the product of 3 and the quantity of 1 minus the square of the Rubin scaling triaxial extension function Q subscript 2, added to the product of 5 and the square of the Rubin scaling triaxial extension function minus the product of 4 and the Rubin triaxial extension function, all over the sum of the product of 3 and the squared quantity of 1 minus the square of the Rubin scaling triaxial extension function Q subscript 2 added to the squared quantity of 1 minus the product of 2 and the Rubin triaxial extension function.

Figure 34. Equation. Willam-Warnke form for Q1.

Currently, the eight input parameters, which define Q1 and Q2, set the shape of the three-invariant yield surface when the pressure is compressive, but not when the pressure is tensile. When the pressure is tensile, the model automatically sets Q1 = 0.5774 and Q2 = 0.5. These values simulate a triangular yield surface in the deviatoric plane, and cannot be overridden by the user. With a triangular yield surface, the strengths attained in uniaxial, equal biaxial, and equal triaxial tensile stress simulations are approximately equal.

For a smooth transition between the tensile and compressive pressure regions, the user should take care to set Q1 = 0.5774 and Q2 = 0.5 at zero pressure. This is accomplished by setting α1 - λ1 = 0.5774 and α2 - λ2 = 0.5774.

Damage Formulation

Concrete exhibits softening (strength reduction) in the tensile and low to moderate compressive regimes. Softening is modeled via a damage formulation. Without the damage formulation, the cap model predicts perfectly plastic behavior for laboratory test simulations such as direct pull, unconfined compression, TXE, and TXE. This behavior is not realistic. Although perfectly plastic response is typical of concrete at high confining pressures, it is not representative of concrete at lower confinement and in tension.

The damage formulation models both strain softening and modulus reduction. Strain softening is a decrease in strength during progressive straining after a peak strength value is reached. Modulus reduction is a decrease in the unloading/loading slopes typically observed in cyclic unload/load tests. The damage formulation is based on the work of Simo and Ju, shown in Figure 35.(25) Strain softening and modulus reduction are demonstrated in psi = 145.05 MPa

Figure 36 for the concrete model.

Figure 35. Equation. Damaged stress sigma subscript lowercase IJ superscript lowercase D. Sigma subscript lowercase IJ superscript lowercase D is equal to the product of sigma subscript lowercase IJ superscript lowercase VP and the quantity one minus lowercase D.

Figure 35. Equation. Damaged stress σ dij.

Here d is a scalar damage parameter that transforms the stress tensor without damage, denoted σ vp, into the stress tensor with damage, denoted σ d. The damage formation is applied to the stresses after they are updated by the viscoplasticity algorithm. The damage parameter d ranges from zero for no damage to 1 for complete damage. Thus 1 − d is a reduction factor whose value depends on the accumulation of damage. The effect of this reduction factor is to reduce the bulk and shear moduli isotropically (simultaneously and proportionally).

Damage initiates and accumulates when strain-based energy terms exceed the damage threshold. Damage accumulation via the parameter d is based on two distinct formulations, which is called brittle damage and ductile damage.

Figure 36. Graph. This cap model simulation demonstrates strain softening and modulus reduction. The Y-axis is stress in megapascals. It ranges from 0 to 50 megapascals. The X-axis is strain in percent. It ranges from 0 to 1.5 percent. One main curve is plotted. It increases linearly with initial loading modulus E to peak strength lowercase F prime subscript lowercase C at about 47 megapascals and 0.27 percent strain. This is followed by softening, which is a nonlinear reduction in strength to 2 megapascals and 1.5 percent strain. The curve also unloads and reloads during softening. The slope of the unloading and reloading curve is the initial loading modulus E times the quantity 1 minus damage lowercase D. Hence the unloading and reloading slope is less than the initial loading slope. Also shown is a dashed straight horizontal line initiating at peak strength, and continuing at 47 megapascals to 1.5 percent strain. This dashed line represents the behavior the model would have without simulating damage. The difference between the dashed line without simulating damage and the softening curve with damage is the damage parameter lowercase D times lowercase F prime subscript lowercase C. The difference between the softening curve and the horizontal X-axis is lowercase F prime subscript lowercase C times the quantity 1 minus lowercase D.

psi = 145.05 MPa

Figure 36. Graph. This cap model simulation demonstrates strain softening and modulus reduction.

Brittle Damage. Brittle damage accumulates when the pressure is tensile. It does not accumulate when the pressure is compressive. Brittle damage accumulation depends on the maximum principal strain, εmax, as shown in Figure 37:

Figure 37. Equation. Brittle damage threshold tau subscript lowercase B. Tau subscript lowercase B is equal to the square root of the product of E and epsilon superscript 2 subscript max.

Figure 37. Equation. Brittle damage threshold τb.

Here τ b is an energy-type term that depends on the accumulation of total strain via εmax. Brittle damage initiates when τ b exceeds an initial threshold r0b.

Ductile Damage. Ductile damage accumulates when the pressure is compressive. It does not accumulate when the pressure is tensile. Ductile damage accumulation depends on the total strain components, εij, as shown in Figure 38:

Figure 38. Equation. Ductile damage threshold tau subscript lowercase D. Tau subscript lowercase D is equal to the square root of the product of one-half, sigma subscript lowercase IJ, and epsilon subscript lowercase IJ.

Figure 38. Equation. Ductile damage threshold τd.

Here τd is an energy-type term. The stress components σij are the elasto-plastic stresses (with kinematic hardening) calculated before application of damage and rate effects. Therefore, this strain-energy term does not represent the true strain energy in the concrete. Ductile damage initiates when τ d exceeds an initial threshold r0d.

Damage Threshold. Brittle and ductile damage initiate with plasticity. This effectively means that the initial damage surface is coincident with the plastic shear surface. Therefore, a distinct damage surface is not defined by the user. Damage initiates at peak strength on the shear surface where the plastic volume strain is dilative. Damage does not initiate on the cap where plastic volume strain is compactive.

One exception to initiation of damage with initiation of plasticity is when rate effects are modeled via viscoplasticity. With viscoplasticity, the initial damage threshold is shifted (delayed), as shown in Figure 39:

Figure 39. Equation. Viscoplastic damage threshold lowercase R subscript 0. General initial damage threshold lowercase R subscript 0 is equal to the product of lowercase R superscript lowercase s and the quantity of 1 plus the quotient of the product of Young’s general modulus E, the effective strain rate epsilon dot, and eta, all over the product of the initial damage before activation of rate effects lowercase R superscript lowercase s and the square root of Young’s general modulus E.

Figure 39. Equation. Viscoplastic damage threshold r0.

Here rs is the damage threshold before application of viscoplasticity, and r0 is the shifted threshold with viscoplasticity. When η is greater than zero (rate effects are modeled), the initial damage threshold is scaled up by the term in brackets. Hence, damage initiation is delayed while plasticity accumulates. This shift requires no input parameters and is hardwired into the model based on the viscoplastic theory previously discussed.

Damage accumulates when either the brittle or ductile energy term, generically called τn, exceeds the current damage threshold, rn. Here the subscript 'n' indicates the nth time step. Once damage initiates, the value of the damage threshold increases. The new threshold at time step n + 1, denoted rn+1, is set equal to the exceeded value of τ . If the previous value of τ did not exceed the previous threshold (an increment without damage), the threshold does not increase. Mathematically this is expressed as shown in Figure 40:

Figure 40. Equation. Incremental damage threshold, small R subscript lowercase N plus 1. Lowercase R subscript lowercase N plus 1 is equal to the maximum of either lowercase R subscript lowercase N and tau subscript lowercase N.

Figure 40. Equation. Incremental damage threshold, small rn+1.

Hence each strain energy term (brittle or ductile) must increase in value above its previous maximum in order for damage to accumulate. When energy remains constant or decreases, damage temporarily stops accumulating. The description given here is effectively that of an expanding damage surface.

Softening Function. As damage accumulates, the damage parameter d increases from an initial value of zero, towards a maximum value of 1. Damage accumulates with τ according to the following functions, shown in Figures 41 and 42:

Brittle Damage

Figure 41. Equation. Brittle damage small D of tau. Lowercase D as a function of tau is equal to the product of the quotient of 0.999 over D, and the quantity of the quotient of the sum of 1 and D all over the sum of 1 plus the product of D and the exponential of negative C times left parenthesis tau minus lowercase R subscript 0 right parentheses.

Figure 41. Equation. Brittle damage small d of tau.

Ductile Damage

Figure 42. Equation. Ductile damage small D of tau. Lowercase D of tau is equal to the product of the quotient of dmax over B and the quantity of the quotient of the sum of 1 and B all over the sum of 1 and the product of B and the exponential of negative A times left parenthesis tau minus small lowercase R subscript 0 right parentheses.

Figure 42. Equation. Ductile damage small d of tau.

An alternative softening function is suggested in appendix A. The parameters A and B or C and D set the shape of the softening curve plotted as stress-displacement or stress-strain. The parameter dmax is the maximum damage level that can be attained. It is set equal to approximately 1 in the tensile and low confining pressure regimes. Brittle damage is set to 0.999 to avoid computational difficulties associated with zero stiffness at a value of 1. At moderate confining pressures, it is less than 0.999, in agreement with TXE data with residual strength. It is set to less than 0.999 by the formula shown in Figure 43:

Figure 43. Equation. Variation of dmax with stress invariant ratio. If the quotient of the square root of the product of 3 and the second invariant of the deviatoric stress tensor J prime subscript 2 all over the first invariant of the stress tensor J subscript one end is less than 1, then dmax is equal to the quantity raised to the power of 1.5 of the quotient of the square root of the product of 3 and the second invariant of the deviatoric stress tensor J prime subscript 2 over the first invariant of the deviatoric stress tensor J subscript 1.

Figure 43. Equation. Variation of dmax with stress invariant ratio.

The power 1.5 was chosen by the developer based on examination of single element simulations and could be included as a user-supplied input parameter at a later date. The maximum damage level also varies with rate effects, as shown in Figure 44. The nondimensional term in parentheses is a stress invariant ratio that is equal to 1 in unconfined compression and less than 1 for stress states with confinement, as shown in Figure 45.

Figure 44. Equation. Variation of dmax with rate effects. Dmax is equal to the product of dmax and the maximum of either 1 or the quantity to the power of 1.5 of the sum of 1 and the quotient of the product of Young’s general modulus E, epsilon dot, and eta, all over the product of the initial damage before activation of rate effects lowercase R superscript lowercase S and the square root of Young’s general modulus.

Figure 44. Equation. Variation of dmax with rate effects.

Figure 45. Graph. Schematic representation of four stress paths and their stress invariant ratios. The Y-axis is principal stress difference, which is also listed at the square root of the quantity 3 times J prime subscript 2. The X-axis is the pressure invariant J subscript 1. No units are given for either axis. Also shown is one curve representing the shape of the combined shear surface and cap. Four dashed lines are shown which eminate from the origin to the shear surface. One line intersects the shear surface in the tensile pressure regime. It represents the stress path for unconfined tensile stress. Its value is negative 1. The second line is vertical and intersects the shear surface in pure shear stress. It value is 0. The third line intersects the shear surface in compressive pressure. It represents the stress path for unconfined compression stress. Its value is 1. The fourth line is bilinear. It is initially coincident with the X-axis in the compressive pressure regime, then bends and intersects the shear surface. It represents the stress path for triaxial compression simulations. Its value is greater than 1. The value listed is a nondimensional stress invariant ratio used in many of the equations. The numerator is J subscript 1. The denominator is the square root of the quantity 3 times J prime subscript 2.

Figure 45. Schematic representation of four stress paths and their stress invariant ratios.

In addition to reducing the maximum damage level with confinement (pressure), the compressive softening parameter, A, may also be reduced with confinement. The formulation is shown in Figure 46:

Figure 46. Equation. Reduction of A with confinement. The softening parameter A is equal to the product of A and the quantity superscript pmod of the sum of dmax and 0.001.

Figure 46. Equation. Reduction of A with confinement.

Here pmod is a user-specified input parameter. Its default value is 0.0. Input positive value of pmod reduces A when the maximum damage is less than 0.999; otherwise A is unaffected by pmod. Thus, it is only active at moderate confinement levels.

The maximum increment in damage that can accumulate over a single time step is 0.1 (10 percent). This maximum increment is set internally to avoid excessive damage accumulation over a single time step. Excessive damage accumulation can lead to unstable behavior. This 0.1 damage increment was chosen by the developer based on examination of numerous multielement simulations and could be included as a user-supplied input parameter at a later date.

Regulating Mesh Size Sensitivity. If the equations shown in Figure 41 and Figure 42 are used as is to model softening, the softening behavior would be mesh size dependent. This behavior means that different mesh refinements would produce different computational results, typically with the greatest damage accumulation in the smallest elements. This behavior is undesirable and is the result of modeling smaller fracture energy in the smaller elements. The fracture energy is the area under the stress-displacement curve in the softening regime. Direct use of the equations shown in Figure 41 and Figure 42 is demonstrated in the concrete model evaluation report for direct pull and unconfined compression of concrete cylinders.(1) The calculations without softening regulation demonstrate that convergence of the solution is not attained as the mesh is refined to a reasonable element size of about 19 to 38 millimeters (mm) (0.75 to 1.5 inches) for concrete.

It is desirable for a computational solution to converge as the mesh is refined. Regulatory methods promote convergence by reducing or eliminating element-to-element variation in fracture energy. The fracture energy is a property of a material, and special care must be taken to treat it as such. Several possible approaches are available for regulating mesh size dependency. One approach is to manually adjust the damage parameters as a function of element size to keep the fracture energy constant. However, this approach is not practical because the user would need to input different sets of damage parameters for each size element. A more automated approach is to include an element length scale in the model. This is done by passing the element size through to the material model and internally calculating the damage parameters as a function of element size. Finally, viscous methods for modeling rate effects have also been proposed (in the literature) to regulate mesh size dependency. However, if rate-independent calculations are performed, then viscous methods will be ineffective.

To regulate mesh size sensitivity, the concrete model maintains constant fracture energy regardless of element size. This is done by including the element length, L (cube root of the element volume), and a fracture energy type term, Gf , in the softening parameter A of the equation in Figure 41, or C of the equation in Figure 42. The most general way to maintain constant fracture energy is to derive an expression for the fracture energy by integrating the analytical stress-displacement curve as shown in Figure 47:

Figure 47. Equation. Fracture energy integral for G subscript lowercase F. G subscript lowercase F is equal to the integral from small X subscript 0 to infinity of the product of the quantity 1 minus small lowercase D and lowercase F prime integrated over the derivative of lowercase X.

Figure 47. Equation. Fracture energy integral for Gf.

Here x is the displacement and x0 is the displacement at peak strength, f¢ . The fracture energy is defined in terms of the damage softening parameters (with dmax = 1) and element length by substituting either damage formulation (equations in Figure 41 or in Figure 42) into and performing the integration. To accomplish the integration, the damage threshold difference (τ - r0) must be specified. The damage threshold difference is dependent on whether brittle or ductile damage is modeled. Brittle softening (P < 0) and ductile softening (P > 0) are regulated separately, because brittle damage accumulation is modeled differently from ductile damage accumulation.

For brittle damage, the relationship between the fracture energy, Gf, the softening parameters, C and D, the initial damage threshold, r0b, and element size, L, is shown in Figure 48:

Figure 48. Equation. Brittle damage fracture energy G subscript lowercase F. G subscript lowercase F is equal to the product of small lowercase R subscript small lowercase B and 0, the element length L, the quantity of the quotient of the sum of 1 and the softening parameter D all over the product of the softening parameters C and D, and log of the quantity of the sum of 1 and the softening parameter D.

Figure 48. Equation. Brittle damage fracture energy Gf.

This expression was obtained by integrating the equation in Figure 47 using the definition of the damage threshold difference shown in Figure 49:

Figure 49. Equation. Brittle damage threshold difference tau minus small R subscript 0 lowercase B. Tau minus the general initial damage threshold small lowercase R subscript 0 is equal to the product of the square root of Young’s general modulus E and the quantity of the quotient of the instantaneous displacement X minus the instantaneous displacement at peak strength X subscript 0, all over the element length L.

Figure 49. Equation. Brittle damage threshold difference τ minus small r0b.

Rearranging the equation in Figure 48 gives the softening parameter C shown in Figure 50:

Figure 50. Equation. Brittle softening parameter C. C is equal to the product of small lowercase R subscript lowercase B and 0, the element length L, the quantity of the quotient of the sum of 1 and the softening parameter D all over the product of G subscript lowercase F and the softening parameter D, times the log of the quantity of the sum of one and D.

Figure 50. Equation. Brittle softening parameter C.

For ductile damage, the relationship between the fracture energy, Gf, the softening parameters, A and B, the initial damage threshold, r0d, and element size, L, is shown in Figure 51:

Figure 51. Equation. Ductile damage fracture energy G subscript lowercase F. G subscript lowercase F is equal to the addition of two large terms. The first term is 2 times lowercase R subscript 0 and lowercase C times L times 1 plus B times the log of 1 plus lowercase B, all divided by A times B. The second term is 2 times L times 1 plus B, all divided by A squared, times the integral from 0 to infinity, integrated over the derivative of lowercase Y, of lowercase Y exponential negative lowercase Y divided by the quantity of 1 plus B times exponential negative lowercase Y. Where lowercase Y equals negative A times the difference between the square root of lowercase X and the square root of lowercase X subscript 0, all times the square root of yield function lowercase F prime, divided the square root of L.

Figure 51. Equation. Ductile damage fracture energy Gf.

This expression was obtained by integrating the equation in Figure 47 using the definition of the damage threshold difference shown in Figure 52:

Figure 52. Equation. Ductile damage threshold difference tau minus small R subscript 0 lowercase D. Tau minus small lowercase R subscript 0 equals the difference between the square root of small X and the square root of small X subscript 0, all times the square root of yield function lowercase F prime, divided the square root of L.

Figure 52. Equation. Ductile damage threshold difference τr0d.

The integral on the right of the equation in Figure 51 reduces to a dilogarithm, which is not solvable in closed form. Therefore, its value is internally calculated by the LS-DYNA concrete model during the initialization phase. Its value depends on the input parameter B and is termed dilog.

Rearranging the equation in Figure 51 gives the softening parameter A shown in Figure 53:

Figure 53. Equation. Ductile softening parameter A. A equals negative lowercase B plus the square root of lowercase B squared minus 4 times lowercase C, all divided by 2. Lowercase B equals the numerator negative 2 times the quantity 1 plus B times the log of 1 of B times L times tau subscript 0 superscript Ductile, all divided by the denominator B times G subscript lowercase F superscript Ductile. Lowercase C equals negative 2 times L times the quantity 1 plus B times symbol dilog, all divided by G subscript lowercase F superscript Ductile.

Figure 53. Equation. Ductile softening parameter A.

In summary, to regulate mesh size dependency, the concrete model requires input values for B and Gfc rather than for A and B. Similarly, the concrete model requires input values for D and Gft rather than for C and D. When the brittle damage threshold is attained, the concrete material model internally solves the equation in Figure 48 for the value of A based on the initial element size, the initial brittle damage threshold, the brittle fracture energy, and a user-specified input value for B. When the ductile damage threshold is attained, the concrete material model internally solves the equation in Figure 51 for the value of C based on the initial element size, the initial ductile damage threshold, and the ductile fracture energy, and a user-specified input value for D.

Refer to the concrete model evaluation report, which discusses cylinder compression calculations conducted with regulation of the softening response.(1) The calculations conducted in tension demonstrate that brittleness increases with mesh refinement until convergence is attained. Conversely, the calculations conducted in compression demonstrate that brittleness decreases with mesh refinement until convergence is attained.

Specifying the Fracture Energy. The user specifies three distinct fracture energy values. These are the fracture energy in uniaxial tensile stress, Gft, pure shear stress, Gfs, and uniaxial compressive stress, Gfc. The model internally selects the brittle or ductile fracture energy from equations that interpolate between the three fracture energy values as a function of the stress state. The stress state is defined by a nondimensional stress invariant ratio call trans. The interpolation is as shown in Figures 54 and 55:

Figure 54. Equation. Brittle damage threshold G subscript lowercase F superscript Brittle. If the pressure is tensile, meaning J subscript 1 is less than 0, then G subscript lowercase F superscript Brittle equals G subscript lowercase F and lowercase S plus parameter trans times the difference between G subscript lowercase FT and G subscript lowercase FS, where parameter trans equals the negative J subscript 1 divided by the square root of the quantity 3 times J prime subscript 2, all to the power of parameter pwrt. The parameter trans is limited to the minimum of 1 or trans. The parameter trans is also limited to the maximum of 0 or trans.

Figure 54. Equation. Brittle damage threshold Gf Brittle.

Figure 55. Equation. Ductile damage threshold G subscript lowercase F superscript Ductile. If the pressure is compressive, meaning J subscript 1 is greater than or equal to 0, then G subscript lowercase F superscript Ductile equals G subscript lowercase F and lowercase S plus parameter trans times the difference between G subscript lowercase FC and G subscript lowercase FS, where parameter trans equals the positive J subscript 1 divided by the square root of the quantity 3 times J prime subscript 2, all to the power of parameter pwrt. The parameter trans is limited to the minimum of 1 or trans. The parameter trans is also limited to the maximum of 0 or trans.

Figure 55. Equation. Ductile damage threshold Gf Ductile.

Here trans is the interpolation parameter whose value ranges between 0 in pure shear stress to 1 in uniaxial tensile or compressive stress. The interpolation depends on two user-specified input parameters. These are pwrt for the tensile-to-shear transition and pwrc for the shear-to-compression transition.

Fracture Energy with Rate Effects. When rate effects are modeled with viscoplasticity, the user has the option to increase the fracture energy as a function of the rate effect. This is accomplished via the repow parameter as shown in Figure 56:

Figure 56. Equation. The fracture energy with rate effects, G subscript lowercase F superscript lowercase VP. G subscript lowercase F superscript lowercase VP equals G subscript lowercase F times a quantity to the power of repow. The quantity is E times dot epsilon times eta, divided by small lowercase R superscript s times the square root of E, all plus 1.

Figure 56. Equation. The fracture energy with rate effects, Gvpf .

Here Gf is either the brittle or ductile fracture energy calculated from the user-specified input values, and The fracture energy with rate effects, G subscript lowercase F superscript lowercase VP subscript lowercase f is the value that is scaled up with rate effects. A value of repow = 1 is recommended. With a value of 1, the increase in fracture energy with rate effects is approximately proportional to the increase in strength with rate effects. With a value of repow = 0, constant fracture energy is maintained independent of rate effects. Refer to appendix B of the concrete model evaluation report for a review of calculations conducted with repow ranging between 0 and 1.(1) They indicate that repow = 0 tends to model a response that is more brittle than measured in bridge rail impact tests. The recommended range is between 0.5 and 1.

Tracking Damage. Two distinct damage parameters are tracked. One parameter is the ductile damage parameter, denoted d d. The ductile damage parameter increases in value whenever the ductile damage formulation is active (pressure is compressive) and τ exceeds the current damage threshold. The value of the ductile damage parameter never decreases, even temporarily.

The other parameter is the brittle damage parameter d b. The brittle damage parameter increases in value whenever the brittle (pressure tensile) damage formulation is active and τ exceeds the current damage threshold. When inactive, the brittle damage parameter is temporarily set equal to zero in order to model stiffness recovery with crack closing. In other words, brittle damage drops to zero (stiffness is recovered) whenever the pressure switches from tensile to compressive. The maximum value of d b is recovered when the brittle formulation becomes active again (when the pressure becomes tensile again).

A user-specified input parameter, called recov, is available to control stiffness recovery. It is by default zero, which means 100 percent recovery of stiffness and strength when pressure becomes compressive. A value of 1 would provide no recovery of stiffness and strength; hence brittle damage remains at its maximum level. Values between 0 and 1 model partial recovery. Its implementation is shown in Figure 57:

Figure 57. Equation. Default damage recovery of lowercase D of tau subscript lowercase T. If the pressure is compressive, then lowercase D of tau subscript lowercase T equals parameter recov times lowercase D of tau subscript lowercase T.

Figure 57. Equation. Default damage recovery of d of τt .

The damage parameter applied to the six stresses is equal to the current maximum of the brittle or ductile damage parameter: d=max(d b, d d).

An option is also available to control stiffness recovery as a function of volumetric strain as well as pressure, as shown in Figure 58:

Figure 58. Equation. Optional damage recovery of lowercase D of tau subscript lowercase T. If the volumetric strain and pressure are compressive, then lowercase D of tau subscript lowercase T equals parameter recov times small lowercase D of tau subscript lowercase T.

Figure 58. Equation. Optional damage recovery of d of τt.

To select this option, recov is specified by the user with an initial value between 10 and 11, rather than between 0 and 1. When recov is 10 or greater, a flag is internally set to base stiffness recovery on volumetric strain as well as pressure. In addition, the concrete model internally subtracts 10 from recov to obtain a value between 0 and 1 for use in the above formulation.

Element Erosion. An element loses all strength and stiffness as d® 1. To prevent computational difficulties, such as mesh tangling and shooting nodes, element erosion is available as a user option. An element erodes when d > 0.99 and the maximum principal strain is greater than a user-supplied input value.

Rate Effects Formulation

Rate effects formulations are implemented to model an increase in strength with increasing strain rate such as that previously shown in Figure 13. The rate effects formulations are applied to the plasticity surface, the damage surface, and the fracture energy. This section discusses rate effects as they apply to the plasticity surface. Rate effects as they apply to the damage surface and fracture energy were previously discussed in the section on damage formulation above.

Viscoplastic Formulation. The viscoplastic algorithm is applied to the yield surface. The implementation is based on the Simo et al. extension of the commonly used Duvaut-Lions formulation to cap models.(26) The Simo et al. extension requires one rate effects parameter, denoted by η. This parameter is called the fluidity coefficient and is a user-specified input parameter.

The basic viscoplastic update algorithm is simple to implement. At each time step, the algorithm interpolates between the elastic trial stress and the inviscid stress (without rate effects) to set the viscoplastic stress (with rate effects), as shown in Figure 59:

Figure 59. Equation. Viscoplastic stress update for sigma subscript lowercase IJ superscript lowercase VP. Sigma subscript lowercase IJ superscript lowercase VP equals the quantity 1 minus gamma, times sigma subscript lowercase IJ superscript T plus gamma times sigma subscript lowercase IJ superscript P, where gamma equals delta lowercase T divided by eta, all divided by the quantity of 1 plus delta lowercase T divided by eta.

Figure 59. Equation. Viscoplastic stress update for σvpij.

This interpolation depends on the fluidity coefficient, η, and the time step, Δt. When η = 0, the inviscid stress is attained so that the solution is independent of strain rate. As η ® ¥, the elastic trail stress is attained at each and every time step. This corresponds to the absence of plastic flow. Therefore, plastic flow decreases as rate effects increase. At each time step, the viscoplastic stress is bounded between the current rate-independent stress and the elastic trial stress. The viscoplastic algorithm allows the viscoplastic stress state to lie outside the yield surface.

The model's flexibility in fitting high strain rate data is improved by converting the single parameter viscoplastic formulation to a two-parameter formulation. This is a simple modification, shown in Figure 60, in which:(27)

Figure 60. Equation. Two-parameter eta. Eta equals eta subscript 0 divided by dot epsilon to the power of lowercase N.

Figure 60. Equation. Two-parameter η.

The two input parameters are η0 and n. The user can fit rate effects data at two strain rates, instead of at one strain rate, which provides a better fit to the data.

The behavior modeled by the viscoplastic update for direct pull and unconfined compression simulations is as shown in Figure 61:

Figure 61. Equation. Dynamics strengths, lowercase F prime subscript T superscript dynamic and lowercase F prime subscript C superscript dynamic. Lowercase F prime subscript T superscript dynamic equals lowercase F prime subscript T plus E times dot epsilon times eta. Lowercase F prime subscript C superscript dynamic equals lowercase F prime subscript C plus E times dot epsilon times eta.

Figure 61. Equation. Dynamic strengths, f ′T dynamic, and f ′C dynamic.

Here it is shown that the dynamic strength (viscoplastic) is equal to the static strength (inviscid) plus a dynamic overstress equal to E Epsilon η where E is Young's modulus and Epsilon is the effective strain rate. The effective strain rate depends on all six strain components as shown in Figure 62:

Figure 62. Equation. Effective strain rate dot epsilon. Dot epsilon equals the square root of two-thirds of the following quantity: the square of dot epsilon subscript lowercase X minus dot epsilon subscript lowercase V, plus the square of dot epsilon subscript lowercase Y minus dot epsilon subscript lowercase V, plus the square of dot epsilon subscript lowercase Z minus dot epsilon subscript lowercase V, plus the square of dot epsilon subscript lowercase XY, plus the square of dot epsilon subscript lowercase XZ, plus the square of dot epsilon subscript lowercase YZ.

Figure 62. Equation. Effective strain rate Epsilon.

The overstress modeled in tension is the same as that modeled in compression. This means that at high strain rates, the dynamic-to-static strength ratio in tension is about a factor of 10 greater than the dynamic-to-static strength ratio in compression, because the tensile strength is about one-tenth the compressive strength. Hence, the viscoplastic model simulates larger rate effects in tension than in compression, consistent with the data trend previously shown in Figure 14.

Viscoplastic Input. Although the basic two parameter formulation models more significant rate effects in tension than compression, the user may still wish to fit tensile and compressive strain rate data with different viscoplastic fluidity parameters. This is done with four user-specified input parameters. These are η0t and nt for fitting uniaxial tensile stress data, and a separate set, η0c and c, for fitting the uniaxial compressive stress data.

For stress states between uniaxial tensile stress and uniaxial compressive stress, the fluidity parameter is interpolated as a function of stress invariant ratio. The interpolation is shown in Figures 63 and 64:

Figure 63. Equation. Variation of fluidity parameter eta in tension. If the pressure is tensile, meaning J subscript 1 is less than 0, then eta equals eta subscript lowercase s plus parameter trans times the difference between eta subscript lowercase T and eta subscript lowercase S, where parameter trans equals the negative J subscript 1 divided by the square root of the quantity 3 times J prime subscript 2, all to the power of parameter pwrt. The parameter trans is limited to the minimum of 1 or trans. The parameter trans is also limited to the maximum of 0 or trans.

Figure 63. Equation. Variation of fluidity parameter η in tension.

Figure 64. Equation. Variation of fluidity parameter eta in compression. If the pressure is compressive, meaning J subscript 1 is less than 0, then eta equals eta subscript lowercase S plus parameter trans times the difference between eta subscript lowercase C and eta subscript lowercase S, where parameter trans equals the positive J subscript 1 divided by the square root of the quantity 3 times J prime subscript 2, all to the power of parameter pwrt. The parameter trans is limited to the minimum of 1 or trans. The parameter trans is also limited to the maximum of 0 or trans.

Figure 64. Equation. Variation of fluidity parameter η in compression.

Here, the effective ηt , ηs , and ηc are the fluidity parameters in uniaxial tensile stress, shear stress, and uniaxial compressive stress. They are determined from five input parameters as shown in Figure 65:

Figure 65. Equation. Effective fluidity parameters eta subscript lowercase T, eta subscript lowercase C, and eta subscript lowercase S. Eta subscript lowercase T equals eta subscript 0 lowercase T, divided by dot epsilon to the power of lowercase N subscript lowercase T. Eta subscript lowercase C equals eta subscript 0 lowercase C, divided by dot epsilon to the power of lowercase N subscript lowercase C. Eta subscript lowercase S equals parameter Srate times eta subscript lowercase T.

Figure 65. Equation. Effective fluidity parameters, ηt, ηc, and η s.

The effective fluidity parameter in shear is determined from a single input scaling parameter, Srate.

This viscoplastic model may predict substantial rate effects at high strain rates (Epsilon hat > 100). To limit rate effects at high strain rates, the user may input overstress limits in tension (overt) and compression (overc). These input parameters limit calculation of the fluidity parameter as shown in Figure 66:

Figure 66. Equation. Overstress limit of eta. If the quantity E times dot epsilon times eta is greater than the input parameter, then eta equals parameter divided by the product E times dot epsilon.

Figure 66. Equation. Overstress limit of η.

where over = overt when the pressure is tensile, and over = overc when the pressure is compressive.

Kinematic Hardening

In unconfined compression, the stress-strain behavior of concrete typically exhibits nonlinearity and dilation prior to the peak. This behavior was previously demonstrated in Figure 5 and Figure 10. This type of behavior is modeled with an initial shear yield surface, NHFf , which hardens until it coincides with the ultimate shear yield surface, Ff. Two input parameters are required. One parameter, NH, initiates hardening by setting the location of the initial yield surface as a fraction of the final yield surface. Reasonable values are approximately 0.7 < NH £ 1.0. A second parameter, CH, determines the rate of hardening (amount of nonlinearity).

The state variable that defines the translation of the yield surface is known as the back stress and is denoted by αij. The incremental backstress is Δαij. The value of each back stress component is zero upon initial yielding and reaches a maximum value at ultimate yield. The total stress is updated from the sum of the initial yield stress (sigma subscript lowercase IJ superscript KH) plus the back stress (see Figures 67 and 68):

Figure 67. Equation. Back stress alpha subscript lowercase IJ superscript lowercase N plus 1. Alpha subscript lowercase IJ superscript lowercase N plus 1 equals alpha subscript lowercase IJ superscript lowercase N plus delta alpha subscript lowercase IJ.

Figure 67. Equation. Back stress α ij n + 1.

Figure 68. Equation. Updated stress with hardening, sigma subscript lowercase IJ superscript P lowercase N plus 1. Sigma subscript lowercase IJ superscript P lowercase N plus 1 equals sigma subscript lowercase IJ superscript KH lowercase N plus 1 plus alpha subscript lowercase IJ superscript lowercase N plus 1.

Figure 68. Equation. Updated stress with hardening, σP ij n+1.

The hardening rule defines the growth of the back stress. Hardening rules are typically based on stress or plastic strain. The model bases hardening upon stress to ensure that the translating shear surface coincides with the ultimate surface (lack of proper translation is a problem with plastic strain based rules). This process is accomplished by defining the incremental back stress, as shown in Figure 69:

Figure 69 Equation. Incremental back stress, delta alpha subscript lowercase IJ. Delta alpha subscript lowercase IJ equals C subscript H times G subscript alpha times the difference between sigma subscript lowercase IJ superscript P minus alpha subscript lowercase IJ times delta dot epsilon times delta lowercase T.

Figure 69. Equation. Incremental back stress, Δαij.

Here CH is the user input parameter that determines the rate of translation, Gα is a function that properly limits the increments, and sigma subscript lowercase IJ superscript P - αij are the elastoplastic stress components that determines the direction of translation for each component. Also included in the formulation is the effective strain rate increment, delta dot epsilon, and the time step, Δt. These terms are internally calculated by the material model and LS-DYNA, and are included to keep the hardening response independent of time step, time step scale factor, and strain increment.

The input parameter CH is the rate of translation used in unconfined compression. In the brittle regime (and for pure shear stress) the rate of translation is internally increased to 10CH. For stress states with low confinement, the rate of translation transitions between the brittle value and the input value are shown in Figures 70 and 71:

Figure 70. Equation. Brittle rate of translation C subscript H superscript Brittle. If the pressure is tensile, meaning J subscript 1 is less than 0, then C subscript H superscript Brittle equals 10 times C subscript H.

Figure 70. Equation. Brittle rate of translation CH Brittle.

Figure 71. Equation. Ductile rate of translation C subscript H superscript Ductile. If the pressure is compressive, meaning J subscript 1 is greater than or equal to 0, then C subscript H superscript Ductile equals C subscript H superscript Brittle plus the parameter trans times the difference between C subscript H and C subscript H superscript Brittle, where the parameter trans equals J subscript 1 divided by the square root of the quantity 3 times J prime subscript 2, all to the power of pwrc. The parameter trans is limited to the minimum of 1 and trans. The parameter trans is limited to the maximum of 0 and trans.

Figure 71. Equation. Ductile rate of translation CH Ductile.

The function Gα restricts the motion of the yield surface so that it cannot translate outside the ultimate surface.(23) The functional form of Gα is determined from the functional form of the yield surface. It is defined as shown in Figure 72:

Figure 72. Equation. The limiting function G subscript alpha. G subscript alpha equals 1 minus the square of a numerator divided by a denominator. The numerator is bar alpha subscript lowercase IJ times the quantity left parenthesis S subscript lowercase IJ plus bar alpha subscript lowercase IJ divided by 2 right parentheses. The denominator is the square of the Rubin function R, times the square of F subscript lowercase F, times F subscript lowercase C, all evaluated at J subscript 1 superscript P, minus the quantity the square of N subscript H, times the square of the Rubin Function R, times the square of F subscript lowercase F, times F subscript lowercase C, all evaluated at the difference between J subscript 1 superscript P minus alpha subscript lowercase P. With bar alpha subscript lowercase IJ equals alpha subscript lowercase IJ minus the summation of alpha subscript lowercase IJ from1 to 3, with the summation divided by 3.

Figure 72. Equation. The limiting function Gα.

The value of the limiting function is Gα = 1 at initial yield, because αij = 0 at initial yield. The value of the limiting function is Gα = 0 at ultimate yield, because the numerator equals the denominator (for the main term in brackets) at ultimate yield. Thus, Gα limits the growth of the back stress as the ultimate yield surface is approached. The developer has chosen to square the term in brackets based on review of the behavior of single element simulations. The square power could be replaced with a user-defined input parameter at a later date.

Each term in the denominator warrants a verbal description. The term on the left is the J subscript 2 value of the ultimate yield surface, evaluated at the pressure invariant with backstress (before application of damage and viscoplasticity). The term on the right is the d value of the initial yield surface, evaluated at the pressure invariant without backstress (also before application of damage and viscoplasticity).

Use of the kinematic hardening formulation modifies the shear surface definition, as shown in Figure 73:

Figure 73. Equation. Modified shear failure surface F subscript lowercase F. F subscript lowercase F as a function of J subscript 1 equals N subscript H times the quantity, left parenthesis alpha minus lambda times the exponential of negative beta times J subscript 1, plus theta times J subscript 1 right parenthesis.

Figure 73. Equation. Modified shear failure surface, Ff.

If kinematic hardening is not requested (CH and NH are set to zero), then the value of NH is internally reset to 1.0. In this way, the original (ultimate) shear surface is recovered.

Model Input

Ease of use is an important consideration for the roadside safety community. Many users have significant experience performing LS-DYNA analyses, but they are less experienced in material modeling topics. Often, users lack the necessary time, data, and material modeling experience to accurately fit a set of material model parameters to data. The ability of a material model to simulate real world behavior not only depends on the theory of the material model, but on the fit of the material model to laboratory test data. The current version of concrete material model 159 has up to 37 input parameters, with a minimum of 19 parameters that must be fit to data. The model has been made easy to use by implementing a set of standardized material properties for use as default material properties.

Default material model parameters are provided for the concrete model based on three input specifications: the unconfined compression strength (grade), the aggregate size, and the units. The parameters are fit to data for unconfined compression strengths between about 20 and 58 MPa (2,901 to 8,412 psi), with emphasis on the midrange between 28 and 48 MPa (4,061 and 6,962 psi). The unconfined compression strength affects all aspects of the fit, including stiffness, three-dimensional yield strength, hardening, and damage. The fracture energy affects only the softening behavior of the damage formulation. Softening is fit to data for aggregate sizes between 8 and 32 mm (0.3 and 1.3 inches).

The suite of material properties was primarily obtained through use of the CEB-FIP Model Code.(11) This code is a synthesis of research findings and contains a thorough section on concrete classification and constitutive relations. Various material properties, such as compressive and tensile strengths, stiffness, and fracture energy, are reported as a function of grade and aggregate size.

Previous | Table of Contents | Next

Federal Highway Administration | 1200 New Jersey Avenue, SE | Washington, DC 20590 | 202-366-4000
Turner-Fairbank Highway Research Center | 6300 Georgetown Pike | McLean, VA | 22101