Output of Section

Two types of section output are available.

First is the time history output /TH/SECTIO consisting of the sum of the force and moments acting on the section. The output can be in the global system or the local system with the most commons output being the variable groups, GLOBAL, LOCAL, and CENTER.

The second type of output is displacements and optional force and moment for every node in the section written to the section output data SC01 file. This file can then be read as an imposed displacement applied to a second cut section model. RD-E: 5400 Cut Methodology is an example of the cut section methodology that can be used. In this case, a full model is run with ISAVE =1 or 2 to save displacement and optionally the resultant section forces/moments in the file file_nameSC01.

Next, the second cut model (submodel) is created with a section defined using the option ISAVE =100 or 101 to read the displacement of the nodes the section file. The section’s local system defined by the three nodes node_ID1, node_ID2, and node_ID3 or frame_ID must be the same one used when saving the data (ISAVE =1 or 2) and reading the data (ISAVE =100 or 101).

If ISAVE =2 is used in the full model and ISAVE =101 is used in the cut model, then PRADIOS outputs the difference between resultant section forces and moments in full and cut model in the section (/TH/ SECTIO).

To improve solution times and decrease memory needed, it is recommended to set ISAVE =0 if no cut modeling methodology is needed.

Filter

When reading and applying the displacements to a cut section model, the displacement can be filtered using an exponential moving average filter. For example, the filtered displacement in the x-direction would be:

\[x_{f}=\alpha x(t)+(1-\alpha)x(t-dt) \tag{49}\]

Where,

\(x_{f}\) Filtered displacement

\(\alpha\) Exponential moving average filtering constant

\(x(t)\) Unfiltered displacement at the current time

\(x(t-dt)\) Displacement at the previous time step

Recommendations are:

\(\alpha =\frac{2\pi dt}{T}\) for filtering -3dB

\(\alpha =\frac{2\pi dt}{\sqrt{3T}}\) for filtering -6dB

Where,

\(T\) Filtering period

\(dt\) Model time step

The filtering period \(T = 10\Delta t\) is often used making \(\alpha =\frac{2\pi}{10}\) = 0.62832 for a -3dB filter.

Materials

Different material tests could result in different material mechanic character.

The typical material test for metal is tensile test. With this test strain-stress curve, yield point, necking point and failure point of material could be observed.

../_images/imyuoage1.jpg

Figure 131: Force (F) and Length (l) are Measured

Engineer strain-stress curve could be generated by:

\[\sigma_{e}=\frac{F}{S_{0}} \tag{50}\]
\[c_{e}=\frac{\Delta I}{I_{0}} \tag{51}\]

Where,

\(S_{0}\) Section area in the initial state

\(I_{0}\) Initial length

In this Force-elongation curve or engineer stress-strain curve, three points are important.

  1. Yield point: where material begin to yield. Before yield you can assume material is in elastic state (the Young’s modulus E could be measured) and after yield, material plastic strain which is nonreversible.

    • Some material in this test will first reach the upper yield point (ReH) and then drop to the lower yield point (ReL). In engineer stress-strain curve, lower yield stress (conservative value) could be taken.

    • Some material can not easily find yield point. Take the stress of 0.1 or 0.2% plastic strain as yield stress.

  2. Necking point: where the material reaches the maximum stress in engineer stress-strain curve. After this point, the material begins to soften.

  3. Failure point: where material failed.

../_images/image249.jpg

Figure 132:

\(\mathbf{R_{m}}\) Maximum resistance

\(\mathbf{F_{max}}\) Maximum force

\(\mathbf{R_{eH}}\) Upper yield level

\(\mathbf{R_{eL}}\) Lower yield level

\(\mathbf{A_{g}}\) Uniform elongation

\(\mathbf{A_{gt}}\) Total uniform elongation

\(\mathbf{A_{t}}\) Total failure strain

True stress-strain curve which is requested in most materials in PRADIOS, except in LAW2, where both engineer stress-strain and true stress-strain are possible to input material data.

In Figure 133, find engineer stress-strain curve (blue) by using:

\[\sigma_{tr}=\sigma_{e}\text{exp}\Big{(}\varepsilon_{tr}\Big{)} \tag{52}\]
\[\varepsilon_{tr}=\text{ln}\big{(}1+\varepsilon_{e}\big{)} \tag{53}\]

The result is true stress-strain curve (red). Plastic true stress-strain curve is shown in green, which plastic strain begin from 0. This green plastic true stress-strain curve is what you need, as in LAW36, LAW60, LAW63, and so on.

../_images/image250.jpg

Figure 133:

The true stress-strain curve is valid until the necking point of the material. After the necking point, the material curve has to be defined manually for hardening. Using a different material law, PRADIOS will extrapolation the true stress-strain curve to 100%.

  • Linear extrapolation: If stress-strain curve is as function input (LAW36), then stress-strain curve is linearly extrapolated with a slope defined by the last two points of the curve. It is recommended that the list of abscissa value be increased to a value greater than the previous abscissa value.

  • Johnson-Cook: After necking point, Johnson-Cook hardening is one of the most commonly used to extrapolate the true stress-strain curve.

\[\sigma_{y}=a+b\varepsilon_{pn}\]

However, it may overestimate strain hardening for automotive steel, In this case, combination of swift-voce hardening is more accurate.

  • Swift and Voce: After necking point, use one of the following equations to extrapolate the true stress-strain curve.

    Swift model

    \(\sigma_{y}=A\left(\varepsilon_{p}+\varepsilon_{0}\right)^{n}\)

    \(A\text{ and } n \text{ are positive.}\)

    Voce model

    \(\sigma_{y}=k_{0}+Q\left[1-\exp\left(-B\varepsilon_{p}\right)\right]\)

    \(k_{0,}Q and B\) are positive

Combination of Switch and
Voce model (LAW84 and LAW87)

\(\sigma_{y}=\alpha\big{[}A\big{(}\overline{\varepsilon}_{p}+\varepsilon_{0}\big{)}^{n}\big{]}+(1-\alpha)\big{\{}{k_{0}+Q\big{[}1-exp\big{(}-B\overline{\varepsilon}\{p}\big{)}\big{]}\big{\}}\)

../_images/image351.jpg

Figure 134:

Here, α is weight of Swift hardening and Voce hardening. Here one Compose script as example to fit the Swift hardening parameters \(A,\varepsilon_{0}, n\) and Voce hardening parameters \(k_{0,} Q, B\) with input stress-strain curve.

See Also
/MAT/LAW84 (Starter)
/MAT/LAW87 (BARLAT2000) (Starter)

Hyperelastic Materials

Hyperelastic materials are used to model materials that respond elastically under very large strains. These materials normally show a nonlinear elastic, incompressible stress strain response which returns to its initial state when unloaded.

../_images/i567ymage5.jpg

Figure 135:

Hyperelastic materials are a specific case of Cauchy material where the stress only depends on the current deformation. Meyer’s kinetic theory describes how these materials consist of flexible chain like structure which rotates and straightens when deformed. This led to the theory that a strain energy density function could be written as a function of the deformation.3 The strain energy density can then be differentiated to obtain the stress strain behavior of the material.

The most common hyperelastic materials are elastomers or rubbers. Some properties of elastomers include.2

  • The material is nearly ideally elastic and deformation is reversible with stress being a function onlyof current strain and independent of history or rate of loading, if deformed at constant temperature or adiabatically.

  • The material incompressible and strongly resists volume changes. The bulk modulus which is aratio of volume change to hydrostatic component of stress is comparable to that of metals.

  • The material is very weak in shear with a shear modulus 10-5 times small than most metals.

  • The material is isotropic its stress-strain response is independent of material orientation.

Element Property Recommendations

When using hyperelastic material laws, there are some recommended element property settings. When using solid elements, it is always better to mesh with 8 node /BRICK elements, if possible. If not, then /TETRA4 or /TETRA10 elements can be used. Recommended /PROP/SOLID for 8 nodes brick are, \(I_{smstr}\) =10, \(I_{cpre}\) =1, with \(I_{solid}\) =24. If hourglassing occurs, then \(I_{solid}\) =17 with \(I_{frame}\) =2 can be used.

  1. Treloar, L. R. G. “The elasticity and related properties of rubbers.” Reports on progress in physics 36, no. 7 (1973): 755

  2. Bower, Allan F. Applied mechanics of solids. CRC press, 2009

Ogden Materials

Hyperelastic materials can be used to model the isotropic, nonlinear elastic behavior of rubber, polymers, and similar materials. These materials are nearly incompressible in their behavior and can be stretched to very large strains.

In PRADIOS, material laws LAW42, LAW62, LAW69, LAW82, and LAW88 utilize different strain energy density functions of the Ogden material model5 to model hyperelastic materials.6

Material Definition

Stretch (also called stretch ratio) \(\lambda\) is the ratio of final length and initial length. It is used for materials with large deformations. For a cube in tension:

\(\varepsilon_{1}=\frac{\Delta I}{I_{01}}\) Engineering strain (also called nominal strain) in direction 1

\(\lambda_{1}=\frac{I_{1}}{I_{01}}\) Strength in direction 1

../_images/image666.jpg

Figure 136:

Thus, strain and stretch are related as:

\[\lambda= 1 + \varepsilon \tag{113}\]

Principal stretch \(\lambda_{i}\) can be used to describe the volumetric deformation by calculating the relative volume, J, computed as:

\[J=\frac{V}{V_{0}}=\frac{I_{1}\cdot I_{2}\cdot I_{3}}{I_{01}\cdot I_{02}\cdot I _{03}}=\lambda_{1}\cdot\lambda_{2}\cdot\lambda_{3} \tag{114}\]

For an incompressible material the volume should not change and thus =1 and thus, the stretch can be calculated for the following material tests.

  • Uniaxial test:

\(\lambda_{1}=\lambda\text{ and }\lambda_{2}=\lambda_{3}2=\frac{1}{\lambda}\)

  • Biaxial test:

\(\lambda_{1}=\lambda_{2}=\lambda and (\lambda_{3}=\lambda^{-2})\)

  • Planer (shear) test:

\(\lambda_{1}=\lambda; \lambda_{3}=1 and (\lambda_{2}=\lambda^{-1})\)

/MAT/LAW42 (Ogden)

This material model defines a hyperelastic, viscous, and incompressible material specified using the Ogden, Neo-Hookean, or Mooney-Rivlin material models. This law is generally used to model incompressible rubbers, polymers, foams, and elastomers. This material can be used with shell and solid elements.

LAW42 uses the following strain energy density representation of the Ogden material model.

\[W\big{(}\lambda_{1},\lambda_{2},\lambda_{3}\big{)}=\sum_{p=1}^{5}\frac{u_{p}}{\alpha_{p}}\big{(}\vec{\lambda}_{1}\alpha_{p}+\vec{\lambda}_{2}\alpha_{p}+\vec{\lambda}_{3}\alpha_{p}-3\big{)}+\frac{K}{2}\big{(}J-1\big{)}^{2}\]

Where,

\(W\) Strain energy density

\(\lambda_{1}\) \(i_{th}\) principal engineering stretch

\(J\) Relative volume defined as: \(J=\lambda_{1}\cdot\lambda_{2}\cdot\lambda_{3}=\frac{\rho_{0}}{\rho}\)

\(\vec{\lambda}_{i}=J^{\frac{1}{3}}\lambda_{i}\) Deviatoric stretch

\(\alpha_{p} and mu_{p}\) Material constants coefficient pairs.

Up to 5 material constant pairs can be defined.

The initial shear modulus and bulk modulus (K) are given by:

\[\mu=\frac{\sum_{p=1}^{5}\mu_{p}\cdot\alpha_{p}}{2} \tag{58}\]

and

\[K=\mu\cdot\frac{2\big{(}1+v\big{)}}{3\big{(}1-2v\big{)}} \tag{59}\]

Where, \(v\) is the Poisson’s ratio and is only used for computing the bulk modulus.

Material Parameters

Parameters \(\alpha\) and \(\mu\) must be chosen so that initial shear modulus is:

\[\mu=\frac{\sum_{p=1}^{5}\mu_{p}\cdot\alpha_{p}}{2}>0 \tag{60}\]

For material stability, it is required that each material constant pair

\[\mu_{p}\cdot\alpha_{p}>0 \tag{61}\]

In general, the Ogden model can be used for strains up to 700%. The number of terms material pairs, \(\alpha\) and \(\mu\), needed depends on the range of experimental data that is fit and curve fitting accuracy

desired. In practice, 3 material pairs fit most data. If the material pairs are not known for a particular material, then a curve fit of uniaxial test data can be done in PRADIOS using LAW69 or via separate fitting software.

Neo-Hookean Model

A simple case of the Ogden material model is the Neo-Hooken model represented using the following equation for the strain energy density function:

\[W=C_{10}(I_{1}-3) \tag{62}\]

Where,

\(I_{1}\) The first invariants of the right Cauchy-Green Tensor

\(C_{10}\) Material constant

This representation can be derived from the LAW42 Ogden strain energy density function when:

\(\mu_{1}=2\cdot C_{10}:\alpha_{1}=2, and\mu_{2}=\alpha_{2}=0\)

The Neo-Hookean model is a simple model that is typically only accurate for strains less than 20%.

Mooney-Rivlin Model

A slightly more complex case of the LAW42 Ogden material model is the Mooney-Rivlin model, which can be represented using the following equation for the strain energy density function:

\[W=C_{10}(I_{1}-3)+C_{0}(I_{2}-3) \tag{63}\]

Where,

\(I_{1}\) and \(I_{2}\) The first and second invariants of the right Cauchy-Green Tensor

\(C_{10}\) and \(C_{01}\) Material constants

This representation can be derived from the LAW42 Ogden strain energy density function when:

\(\mu_{1}=2\cdot C_{10},\mu_{2}=-2\cdot C_{01}, \alpha_{1}=2, and \alpha_{2}=\cdot 2\)

Mooney-Rivlin constants are available from a material supplier or testing company. If they are not available, then a curve fit of uniaxial test data can be done in PRADIOS using LAW69 or via separate fitting software. The Mooney-Rivlin material law is accurate for strains up to 100%.

Poisson’s Ratio and Material Incompressibility

If a material is truly incompressible, then \(v\) = 0.5. However, in practice is not possible to use because that would result in an infinite bulk modulus, an infinite speed of sound, and thus an infinitely small solid element Time Step.

\[K=\mu\cdot\frac{2\left(1+v\right)}{3\left(1-2v\right)}=\mu\cdot\frac{2\left(1+v\right)}{3\left(1-2*0.5\right)}=\infty \tag{64}\]

The effect of different Poisson’s ratio input can be seen in Figure 137. The largest difference in the results is at higher amounts of strain. The results will match the test data better when (v=0.4997) but this results in a time step that is 4 times lower than (v=0.495). Thus, to balance the computation time and accuracy it is recommended to use (v=0.495) for incompressible rubber material.

The effect of Poisson’s ratio and Bulk modulus are similar in other Ogden material law.

../_images/image7.jpg

Figure 137:

Higher values of the Poisson’s ratio may lead to a very small time step or divergence for explicit simulations.

In LAW42, material incompressibility is provided by using a penalty approach, which calculates the pressure proportional to a change in density.

\[P=K\cdot Fscale_{blk}\cdot\mathrm{f}_{blk}(J)\cdot(J-1) \tag{65}\]

Where,

\(K\) Bulk modulus

\(J=\frac{V}{V_{0}}=\frac{m\rho_{0}}{m\rho}=\frac{\rho_{0}}{\rho}\) Relative volume which simplifies to relative density if mass is constant

\(\mathrm{f}_{blk}(J)\) Bulk coefficient scale factor versus relative volume function

\(Fscale_{blk}\) Abscissa scale factor for function \(F_{blk}(J)\)

The bulk modulus (K) of hyperelastic materials is generally a very high value which provides the needed pressure-resistance to maintain the incompressibility condition (J=1). But if a material starts to compress (j < 1) then the bulk modulus can be increased by including the fct_IDblk input function which allows the scaling of the bulk coefficient value as a function of . By default, there is no scaling and; thus, if the function identifier is zero and the value of the bulk scaling function is equal to 1. It is advisable to output (/ANIM/BRICK/DENS) and review the material density of LAW42 components to make sure that the density variation is small, that is the value of J is close to 1 and the material is incompressible.

../_images/imag1e8.jpg

Figure 138: Bulk Modulus Scale Factor Function fct_IDblk

Viscous (Rate) Effects

Viscous (rate) effects are modeled in LAW42 using a Maxwell model, which can be described in a simplified manner as a system of \(\eta\) springs with stiffness’ \(G_{i}\) and dampers \(\eta_{i}\):

../_images/Untitled.png

Figure 139: Maxwell Model

The Maxwell model is represented using Prony series inputs (\(G_{i},\tau_{i})\). The hyperelastic initial shear modulus \(\mu\) is the same as the long-term shear modulus \(G_{\infty}\) in the Maxwell model, and \(\tau_{i}\) is the relaxation time:

\[\tau_{i}=\frac{\eta_{i}}{G_{i}} \tag{66}\]

The \(G_{i} and \tau_{i}\) values must be positive.

/MAT/LAW62 (VISC_HYP)

A hyper visco-elastic material law in PRADIOS that can be used to model polymers and elastomers.

The hyperelastic behavior in this material law is defined using the following strain energy density function:

\[W(\lambda_{1},\lambda_{2},\lambda_{3})=\sum_{i=1}^{N}\frac{2\mu_{i}}{\alpha_{i}!}\Big{(}\lambda_{1}\alpha_{i}+\lambda_{2}\alpha_{i}+\lambda_{3}\alpha_{i}-3+ \frac{1}{\beta}(J^{-\alpha_{i}\beta}-1)\Big{)} \tag{67}\]

Where,

\(W\) Strain energy density

\(\lambda_{i}\) \(i_{th}\) principal stretch

\(J\) Relative volume defined in Equation 67

\(\beta=\frac{v}{(1-2v)}\)

\(\alpha_{i}\) and \(\mu_{i}\) Material constants coefficient pairs.

Up to 5 material constant pairs can be defined.

Poisson’s ratio must be (0<v<0.5). This law can be used to model compressible or sometimes called hyperfoam materials by defining a low Poisson’s ratio value.

image1 Note: The \(\mu_{i}\) material coefficients are different, but can be converted using:

\[\mu_{i}^{LAW62}=\frac{\mu_{i}^{LAW42}\cdot a_{i}^{LAW42}}{2} \tag{68}\]

Viscous (Rate) Effects

Viscous (rate) effects are modeled in LAW62 using a Maxwell model which can be described in a simplified manner as a system of n springs with stiffness’ \(G_{i}\) and dampers \(\eta_{i}\).

../_images/Untitled.png

Figure 140: Maxwell Model

The Maxwell model is represented using a Prony series with inputs. The initial shear modulus is:

\[G_{0}=\sum_{i=1}^{N}\mu_{i} \tag{69}\]

The sum of \(\mu_{i}\) should be greater than 0.

\[G_{0}=G_{\infty}+\sum_{i}G_{i} \tag{70}\]

The stiffness ratio is:

\[\gamma_{\infty}=\frac{G_{\infty}}{G_{0}}=1-\sum_{i}\gamma_{i} \tag{71}\]
\[\gamma_{i}=\frac{G_{i}}{G_{0}} \tag{72}\]

With,

\[\gamma_{i}\in\left[0,\,1\right]\sum_{i}\gamma_{i}<1 \tag{73}\]

and the ground shear modulus

\[G_{0}=G_{\infty}+\sum_{i}G_{i} \tag{74}\]

The relative time, \(\tau_{i}\) must be positive:

\[\tau_{i}=\frac{\eta_{i}}{G_{i}} \tag{75}\]

image1 Note: When viscosity is included, the shear modulus in LAW62 is the initial shear modulus \(G_{0}=\sum_{i=1}^{N}\mu_{i}\) which includes viscosity, but in LAW42 the shear modulus is the long-term shear modulus, which does not include viscosity \(G_{\omega}=\frac{\sum_{p=1}^{5}\mu_{p}\cdot a_{p}}{2}\).

/MAT/LAW69

This law, like /MAT/LAW42 (OGDEN) defines a hyperelastic and incompressible material specified using the Ogden or Mooney-Rivlin material models. Unlike LAW42, where the material parameters are input, this law computes the material parameters using test data from a uniaxial engineering stress-strain curve.

This material can be used with shell and solid elements.

The strain energy density formulation used depends on the law_ID:

  • law_ID = 1 (Ogden law):

\[W\big{(}\lambda_{1},\lambda_{2},\lambda_{3}\big{)}=\sum_{p=1}^{5}\frac{\mu_{p}}{a_{p}}\big{(}\lambda_{1}a_{p}+\lambda_{2}a_{p}+\lambda_{3}a_{p}-3\big{)}+\frac{K}{2}\big{(}J-1\big{)}^{2} \tag{76}\]
  • law_ID = 2 (Mooney-Rivlin law):

\[W=C_{10}\big{(}I_{1}-3\big{)}+C_{0}\big{(}I_{2}-3\big{)} \tag{77}\]

Material Parameters

After reading the stress-strain curve (fct_ID1), PRADIOS calculates the corresponding material parameter pairs using a nonlinear least-square fitting algorithm. For the classic Ogden law, (law_ID=1), the calculated material parameter pairs are and where the value of p is defined via the N_pair input.

The maximum value is N_pair=5 with a default value of 2. Usually no more than N_pair=3 is needed for a good fit.

For the Mooney-Rivlin law (law_ID =2), the material parameter \(C_{10}\) and \(C_{01}\) are calculated for the LAW42 Ogden law can be calculated using this conversion:

\(\mu_{1}=2\cdot C_{10},\mu_{2}=-2\cdot C_{01},\alpha_{1}=2, and (\alpha_{2}=)-2\)

The minimum test data input should be a uniaxial tension engineering stress strain curve. If uniaxial compression data is available, the engineering strain should increase monotonically from a negative value in compression to a positive value in tension. In compression, the engineering strain should not be less than -1.0 since -100% strain is physically not possible.

To improve the quality of the nonlinear least square fit, it is recommended that:

  • The experimental data curve represents a smooth monotonically increasing function with uniform distribution of abscissa points. The number of data points in the experimental data curve should be greater than the number of parameter pairs (N_pair).

  • The engineering strain is negative in compression and positive in tension. For compression test data, the engineering strain should be greater than -1.0 (100% compression maximum) but tension only stress strain data can also be used.

  • If N_pair ≥ 3, then the test data should cover at least 100% of the tensile strain and/or 50% of the compressive strain.

  • N_pair should not be set to a very large value to avoid instabilities in the fitting procedure.

This material law is stable when \(\mu_{p}\alpha_{p}>0\) (with (p)=1,…5) is satisfied for parameter pairs for all loading conditions. By default, Radios tries to fit the curve by accounting for these conditions (\(I_{check}=2)\). If a proper fit cannot be found, then Radios uses a weaker condition (\(I_{check}=1\):), which ensures that the initial shear hyperelastic modulus (\(\mu\)) is positive.

To determine how well the calculated material parameters represent the input test data, the PRADIOS Starter outputs an “averaged error of fitting” value which is recommended to not exceed 10%. For visual comparison, the stress-strain curve calculated from the strain energy density and calculated material parameters is also output by the PRADIOS Starter.

Due to the friction involved in a uniaxial compression test, it is usually more accurate to take equal biaxial tension test data and convert it to uniaxial compressive data using these formulas 7 which are valid for incompressible materials.

\[\varepsilon_{c}=\frac{1}{\left(\varepsilon_{b}+1\right)^{2}}-1 \tag{78}\]
\[\sigma_{c}=\sigma_{b}\left(1+\varepsilon_{b}\right)^{3} \tag{79}\]

Where,

\(\varepsilon_{c}\) Uniaxial engineering compressive stain

\(\varepsilon_{b}\) Equal biaxial engineering tension strain

\(\sigma_{c}\) Uniaxial engineering compressive stress

\(\sigma_{b}\) Equal biaxial engineering tension stress

Material Incompressibility

Material LAW69 uses the same method to maintain incompressibility as LAW42. For additional information, refer to Poisson’s Ratio and Material Incompressibility in LAW42.

Viscous (Rate) Effects

/VISC/PRONY must be used with LAW69 to include viscous effects. Alternatively, LAW69 could be used to extract the Ogden or Mooney-Rivlin parameters and then those parameters can be used in LAW42 with viscosity added.

/MAT/LAW82

This material model defines a hyperelastic, and incompressible material specified using the Ogden, Neo-Hookean, or Mooney-Rivlin material models. This law is generally used to model incompressible rubbers, polymers, foams, and elastomers.

This material can be used with shell and solid elements. As compared to LAW42 or LAW62, this law uses a different Ogden strain energy density formulation given in Equation 80. The LAW82 strain energy density formulation matches what is used in some other finite elements solver’s hyperelastic model and; thus, the material parameters for this form of the Ogden strain energy density are sometimes available from material suppliers or other sources.

\[W=\sum_{i=1}^{N}\frac{2\mu_{i}}{\alpha_{i}^{2}}\big{(}\overline{\lambda}_{i}\alpha_{i}+\overline{\lambda}_{2}\alpha_{i}+\overline{\lambda}_{3}\alpha_{i}-3\big{)}+\sum_{i=1}^{N}\frac{1}{D_{i}}\big{(}J-1\big{)}_{2i} \tag{80}\]

Where,

\(W\) Strain energy density

\(N\) Number of material constants \(\alpha_{i},\mu_{i}\) and \(D_{i}\)

\(\overline{\lambda}_{i}=J^{\frac{1}{3}}\lambda_{i}\) Deviatoric stretch

\(J\) Relative volume as defined in Equation 56

The initial shear modulus:

\[\mu=\sum_{i=1}^{N}\mu_{i} \tag{81}\]

The Bulk Modulus is calculated as \(K=\frac{2}{D_{1}}\) based on these rules:

  • If \(v=0\), \(D_{1}\) should be entered.

  • If \(v\neq 0\), \(D_{1}\) input is ignored and will be recalculated and output in the Starter output using:

\[D_{1}=\frac{3(1-2v)}{\mu(1+v)} \tag{82}\]
  • If \(v=0\) and \(D_{1}\) =0, a default value of \(v=0.475\) is used and \(D_{1}\) is calculated using Equation 82

Neo-Hookean Model

Like LAW42, LAW82 can also be simplified to a Neo-Hooken model by using:

\(\mu_{1}=2\cdot C_{10},\alpha_{1}=2) and \mu_{2}=\alpha_{2}=0\)

Mooney-Rivlin Model

Like LAW42, LAW82 can also be simplified to a Mooney-Rivlin model by using:

\(\mu_{1}=2\cdot C_{10},\mu_{2}=2\cdot C_{01},\alpha_{1}=2 and \alpha_{2}=-2\)

Viscous (Rate) Effects

/VISC/PRONY must be used with LAW82 to include viscous effects.

Drücker Condition Stability Check

In LAW42 and LAW69, the Drücker stability is automatically calculated by the PRADIOS Starter.

The Drücker stability condition checks if the change in the Kirchhoff stress corresponding to the infinitesimal change in the logarithmic strain (true strain) satisfies the following inequality.

\[\sum_{i=1}^{3}d\tau_{i}de\_{i}>0 \tag{83}\]

Where, i =1,2,3 principal direction.

With the change in logarithmic strain

\[de_{i}=\frac{d\lambda_{i}}{\lambda_{i}} \tag{84}\]

\(d\tau_{i}=J\cdot d\sigma_{i}\) The change of Kirchhoff stress

\(d\tau=\mathbf{D}:d\varepsilon\) Relationship between Kirchhoff stress and logarithmic strain

The Drücker stability condition will be:

\[\sum d\varepsilon :\mathbf{D}:d\varepsilon > 0 \tag{85}\]

Here \(\mathbf{D}\) is tangential material stiffness matrix and it is also the slope of stress-strain curve:

\[\tag{86}\]
\[\begin{split}\mathbf{D} =\left [ \begin{matrix} D_{11}&D_{12}&D_{13} \\ D_{21}&D_{22}&D_{23} \\ D_{31}&D_{32}&D_{33} \end{matrix}\right]\end{split}\]

For a stable material, it requests tangential material stiffness \(\mathbf{D}\) be positive (slope of stress-strain curve is positive). The tangential material matrix \(\mathbf{D}\) is positive if following conditions satisfied:

\[I_{1}=tr(\mathbf{D})=D_{11}+D_{22}+D_{33}>0 \tag{87}\]
\[I_{2}=D_{11}D_{22}+D_{22}D_{33}+D_{33}D_{11}-D_{23}-D_{13}{}^{2}-D_{12}{}^{2}>0 \tag{88}\]
\[I_{3}=\det(\mathbf{D}`)>0 \tag{89}\]

The Kirchhoff stress for Ogden model is:

\[\tau_{i}=\sum_{p}\mu_{R}\Big{[}\overline{\lambda}_{\alpha p}-\frac{1}{3}\big{(}\overline{\lambda}_{I,p}+\overline{\lambda}_{2,p}+\overline{\lambda}_{3,p}\big{)}\Big{]}+K\big{(}J^{2}-J\big{)}\]

Since \(D_{ij}=\frac{\partial\tau_{i}}{\partial\lambda_{j}}\), then for a given Ogden parameter \(\alpha_{p}\), \(\mu_{p}\) with conditions \(I_{1}>0, I_{2}>0 and I_{3}>0\), the strain range of material in Drucker stability could then be calculated.

../_images/image-10.jpg

Figure 141:

The Drücker stability criterion calculates the strain range where the material model will remain stable given a set of material parameters. This stability check cannot be made for every deformation but instead is commonly used to check material stability under uniaxial, biaxial and planar strain loading.

For example, using the following Ogden parameters:

\(\mu_{1}\) = 13.99077258830 \(\alpha_{1}\) = 3.78819292935039

\(\mu_{2}\) = 9.13454532223 \(\alpha_{2}\) = 7.17617341059

\(\mu_{3}\) = 8.904655103235 \(\alpha_{3}\) = 7.27028137148

Then the Drücker stability will be automatically checked in PRADIOS Starter, and results printed in Starter output file *0.out. This shows the strain at which instability can occur for the given Ogden parameters:

CHECK THE DRUCKER PRAGER STABILITY CONDITIONS
       -----------------------------------------------
     MATERIAL LAW = OGDEN (LAW42)
     MATERIAL NUMBER = 1
      TEST TYPE = UNIXIAL
       COMPRESSION:   UNSTABLE AT A NOMINAL STRAIN LESS THAN -0.3880000000000
       TENSION:       UNSTABLE AT A NOMINAL STRAIN LARGER THAN 0.9709999999999
      TEST TYPE = BIAXIAL
       COMPRESSION:   UNSTABLE AT A NOMINAL STRAIN LESS THAN -0.2880000000000
       TENSION:       UNSTABLE AT A NOMINAL STRAIN LARGER THAN 0.2780000000000
      TEST TYPE = PLANAR (SHEAR)
       COMPRESSION:   UNSTABLE AT A NOMINAL STRAIN LESS THAN -0.3680000000000
       TENSION:       UNSTABLE AT A NOMINAL STRAIN LARGER THAN 0.5829999999999

image1 Note: For a Neo-Hookean material with \(C_{10}!>!0 (or \mu_{1}!>!0)\), the material is always stable and thus no critical value is found by the Drucker stability check.

For a Mooney-Rivlin material, the Drucker stability should be checked since \(C_{01} or \mu_{2}\) could be negative, which leads material instability.

/MAT/LAW88

This law utilizes a tabulated uniaxial tension and compression engineering stress and strain test data at different strain rates to model incompressible materials. It is only compatible with solid elements.

The material is based on the following Ogden’s strain energy density function but does not require curve fitting to extract material constants like most other hyperelastic material models. 8

\[W=\sum_{i=1}^{3}\sum_{j=1}^{m}\frac{\mu_{j}}{\alpha_{j}}\Big{(}\lambda_{i}\alpha_{j}-1\Big{)}+{K(J-1-ln J)} \tag{91}\]

Instead, this law determines the Ogden function directly from the uniaxial engineering stress strain curve tabulated data.

Unlike other Ogden material laws, the Bulk Modulus must be input from either test data or extracted from Starter output of the LAW69 Ogden curve fit. When comparing results between LAW42 or LAW69 to LAW88, the same bulk modulus must be used.

Unloading Behavior

Unloading can be represented using an unloading function or by providing hysteresis and shape factor inputs to a damage model based on energy.

If using the damage model, the loading curves are used for both loading and unloading and the unloading stress tensor is reduced by:

\[\sigma=\big{(}1-D\big{)}\sigma \tag{92}\]

with

\[D=\big{(}1-Hys)\bigg{(}1-\bigg{(}\frac{W_{cur}}{W_{\max}}\bigg{)}^{Shape}\bigg{)} \tag{93}\]

Where,

\(W_{cur}\) Current energy

\(W\_{\max}\) Maximum energy corresponding to the quasi-static behavior

Hys and Shape Input by user

If an unloading curve is provided, these options are available:

Tension Loading and Unloading

= 0

../_images/image1-1.jpg

Figure 142:

Loading use loading function fct_IDLi

Unloading use unloading function fct_IDunL

= 1

../_images/image1-2.jpg

Figure 143:

Loading and unloading all use loading function fct_IDLi

= -1

../_images/image1-3.jpg

Figure 144:

Loading use loading function fct_IDLi

  • Tension

Unloading use unloading function fct_IDunL

  • Compression

Unloading use loading fct_IDLi

Viscous (Rate Effects

Strain rate effects can be modeled by including engineering stress strain test data at different strain rates. This can be easier than calculating viscous parameters for traditional hyperelastic material models.

Conclusion

Make sure to use the material law that best fits the test data available.

For example, if minimal test data is available and the strains are not too large than the LAW42 NeoHookean Model could be used. If the loading state is known, then it is important to have test data that represents that stress state and make sure the material model fits that test data.

  1. Ogden, R. W., and Non-linear Elastic Deformations. “Ellis Horwood.” New York (1984)

  2. Miller, Kurt. “Testing Elastomers for Hyperelastic Material Models in Finite Element Analysis” Axel Products, Inc., Ann Arbor, MI (2017). Last modified April 5, 2017

Arruda-Boyce (/MAT/LAW92)

LAW92 describes the Arruda-Boyce material model, which can be used to model hyperelastic behavior. This model is based on the statistical mechanics of a material with a cubic representative volume element containing eight chains along the diagonal directions.

It assumes that the chain molecules are located on the average along the diagonals of the cubic in principal stretch space.

Material Parameters

The strain energy density function is:

\[W=\mu\sum_{l=1}^{5}\frac{c_{i}}{(\lambda_{m})^{2i-2}}({l_{1}^{i}}-3^{i})+\frac{1}{D}\big{(}\frac{J^{2}-1}{2}+\ln(J)\big{)} \tag{94}\]

The material constant, \(c_{i}\) are:

\(c_{1}=\frac{1}{2}, c_{2}=\frac{1}{20}, c_{3}=\frac{11}{1050}, c_{4}=\frac{19}{7000}, c_{5}=\frac{519}{673750}\)

\(\overline{I_{1}}=\overline{\lambda_{1}^2}+\overline{\lambda_{2}^2}+\overline{\lambda_{3}^2}\) First strain invariant

\(\lambda_{i}\) \(i_{th}\) principal engineering stretch

A material with LAW92 can be defined in two different ways:

  • Parameter Input

    Shear modulus, bulk modulus and strain stretch (\(\mu,D,\lambda_{m})\)

    Where, only the above 3 parameters with clear physical meaning are necessary to define the material.

    \(\mu\) is shear modulus at zero strain.

\[D=\frac{2}{K} \tag{95}\]

Where,

\(K\) Bulk coefficient at zero strain

\(\lambda_{m}\) Defines the limit of stretch

  1. Axel Products, Inc. “Compression or Biaxial Extension”, Ann Arbor, MI (2017). Last modified November 12, 2008

  2. Kolling, S., P. A. Du Bois, D. J. Benson, and W. W. Feng. “A tabulated formulation of hyperelasticity with rate effects and damage.” Computational Mechanics 40, no. 5 (2007): 885-899

    Also called locking stretch. It specifies the beginning of the hardening phase in tension (locking strain in tension). Default = 7.0.

../_images/image1-5.jpg

Figure 145:locking Stretch \(\lambda_{m}\)

In parametric input, Poisson’s ratio is computed as:

\[v=\frac{3K-2\mu}{6K+2\mu} \tag{96}\]
  • When using function input, Poisson ratio and Itype must be defined. Itype defines which type of engineering stress strain test data that is being used as input.

../_images/image1-6.jpg

Figure 146: Itype = 1: Uniaxial data test

../_images/image1-7.jpg

Figure 147: Itype = 2: Equibiaxial data test

../_images/image1-8.jpg

Figure 148: Itype = 3: Planar data test

Poisson’s Ratio and Material Incompressibility

If function input is defined, then parameters \(\mu,D,\lambda_{m}\) are ignored and PRADIOS will calculate the material constant by fitting the input function. A nonlinear least squares algorithm is used to fit the ArrudaBoyce parameters by PRADIOS. The curve fitting is performed using the assumption that Poisson’s value is close to 0.5, which means the material is incompressible. Similar to the other hyperelastic material models, Poisson ratio values closer to 0.5 result in high bulk modulus and a lower timestep. For a good balance between incompressibility and a reasonable timestep, a Poisson’s ratio value of 0.495 is recommended.

The material fitting information can be found in the Starter output file (*0000.out).

../_images/image1-9.jpg

Figure 149: LAW92 function example

The fitting error and fitted material parameters are printed in the Starter output file.

../_images/image2-0.jpg

Figure 150:

Viscous (Rate Effects

/VISC/PRONY must be used with LAW92 to include viscous effects.

Yeoh (/MAT/LAW94)

LAW94 is a hyperelastic material model that can be used to describe incompressible materials.

The strain energy density function of LAW94 only depends on the first strain invariant and is computed as:

\[W=\sum_{i=1}^3\left[C_{i0}(I_{i}-3)^i+\frac{1}{D_{i}}(J-1)^2i\right] \tag{97}\]

Where,

\(\overline{I_{1}}=\overline{\lambda_{1}^2}+\overline{\lambda_{2}^2}+\overline{\lambda_{3}^2}\) First strain invariant

\(\overline{\lambda_{i}}=J^{-\frac{1}{3}}\lambda_{i}\) Deviatoric stretch

The Cauchy stress is:

\[\sigma_{i}=\frac{\lambda_{i}}{J}\frac{\partial W}{\partial\lambda_{i}}\]

Material Parameters

For incompressible materials with =1 only and are input and the Yeoh model is reduced to a NeoHookean model.

  1. Arruda, E. M. and Boyce, M. C., 1993, “A three-dimensional model for the large stretch behavior of rubber elastic materials”, J. Mech. Phys. Solids, 41(2), pp. 389–412

\(C_{10}C_{20}C_{30}\) Material constants specify the deviatoric part (shape change) of the material

\(D_{1,}D_{2,}D_{3}\) Parameters specify the volumetric change of the material

These six material constants need to be calculated by curve fitting material test data. RD-E: 5600

Hyperelastic Material with Curve Input includes a Yeoh fitting Compose script for uniaxial test data. The Yeoh material model has been shown to model all deformation models, even if the curve fit was obtained using only uniaxial test data.

The initial shear modulus and the bulk modulus are:

\[\mu=2\cdot C_{10} \tag{99}\]

and

\[K=\frac{2}{D_{1}} \tag{100}\]

Poisson’s Ratio and Material Incompressibility

LAW94 is available only as an incompressible material model.

If = D1 0, an incompressible material is considered where, v = 0.495 D1 is calculated as:

\[D_{1}=\frac{3(1-2\nu)}{\mu(1+\nu)} \tag{101}\]

Bergstrom-Boyce (/MAT/LAW95)

This law is a constitutive model for predicting the nonlinear time dependency of elastomer like materials. It uses a polynomial material model for the hyperelastic material response and the Bergstrom-Boyce material model to represent the nonlinear viscoelastic time dependent material response.

This law is only compatible with solid elements.

The response of the material can be represented using two parallel networks A and B. Network A is the equilibrium network with a nonlinear hyperelastic component. In Network B, a nonlinear hyperelastic component is in series with a nonlinear viscoelastic flow element, and hence, is time-dependent network.

  1. Yeoh, O. H. “Some forms of the strain energy function for rubber.” Rubber Chemistry and technology 66, no. 5 (1993): 754-771

../_images/image2-1.png

Figure 151:

Material Parameters

The same polynomial strain energy density formulation is used for the hyperelastic components in both networks. In Network B, it is scaled by a factor Sb. The strain energy density is then written for the hyperelastic component of the network.

\[W_{A}=\sum_{i+j=1}^{3}C_{ij}\big{(}I_{1}-3\big{)}^i\cdot\big{(}I_{2}-3\big{)}^J+\sum_{i=1}^{3}\frac{1}{D_{i}}(J-1)^2i \tag{102}\]

and

\[W_{B}=S_{b}\cdot W_{A} \tag{103}\]

Where,

\(\overline{I_{1}}=\overline{\lambda_{1}^2}+\overline{\lambda_{2}^2}+\overline{\lambda_{3}^2}\)

\(\overline{I_{2}}=\overline{\lambda_{1}^{-2}}+\overline{\lambda_{2}^{-2}}+\overline{\lambda_{3}^{-2}}\)

\(\overline{\lambda_{i}}=J^{-\frac{1}{3}}\lambda_{i}\)

\(C_{ij}\) and \(D_{i}\) Material Parameters

The hyperelastic component the Cauchy stress is computed as:

\[\sigma_{i}=\frac{\dot{\lambda}_{i}}{J}\frac{\partial W}{\partial\dot{\lambda}_{i}} \tag{104}\]

The total stress is the summer of stress in network A and network B.

../_images/ima-ge22.jpg

Figure 152:

\(\overline{\sigma}=\overline{\sigma_{A}}+\overline{\sigma_{B}}\)

Since \(W_{B}=S_{b}\cdot W_{A}\), then \(\overline{\sigma}_{B}=S_{b}\cdot\overline{\sigma}_{A}\) and total stress is \(\sigma=(1+S_{b})\cdot\overline{\sigma}_{A}\)

For example, in one tensile test. If use Sb, then the stress is 3 times of the one without considering viscous (which means only considered hyperelastic).

../_images/imag-e23.jpg

Figure 153:

For special values of , the polynomial model can be reduced to the following material models.

  • Yeoh:

*J*=0

Where, \(C_{10,}C_{20,}C_{30,}\) are not zero.

../_images/imag-e24.jpg

Figure 154:

  • Mooney-Rivlin:

\(i+j=1\)

Where, \(C_{10}\) and \(C_{01}\) are not zero and \(D_{2}=D_{3}=0.\)

  • Neo-Hookean:

Only \(C_{10}\) and \(D_{1}\) are not zero.

Where,

\(C_{ij}\) and \(D_{i}\) Material parameters which can be calculated by completing a curve fit for quasi-static material test data.

RD-E: 5600 Hyperelastic Material with Curve Input, contains a curve fit example for Mooney-Rivlin and Yeoh material models. can be calculated from the bulk modulus or left blank.

The initial shear modulus and the bulk modulus are computed as:

\[\mu=2\big{(}S_{b}+1\big{)}\big{(}C_{10}+C_{01}\big{)} \tag{105}\]

and

\[K=\frac{2}{D_{1}}\big{(}1+S_{b}\big{)} \tag{106}\]

If the bulk modulus of the material is known, D1 can be calculated, or if D1 =0, an incompressible material is assumed.

Viscous (Rate) Effects

The effective creep strain rate in Network B is given by:

\[\dot{\varepsilon}_{B}^{\nu}=A(\overline{\dot{\lambda}}-1+\dot{\varepsilon}g)^{C}\frac{\overline{\sigma}_{B}}{\tau_{ref}}^{M} \tag{107}\]

Where,

\(\overline{\dot{\lambda}}=\sqrt{\frac{T_{1}}{3}}\)

\(\overline{\sigma}_{B}\) Effective stress in Network B.

\(A, \bar{\varepsilon}, M, C\) , and \(\tau_{ref}\) Input material parameters.

The material constants A, M and C are limited to a specific range of real values as defined in the Reference Guide. If limited data is available, a trial and error method11 could be used to determine these constants. Start with the default values of \(\bar{\xi}, M, C, S_{b}\) =1.6; and A=5. Next, compare model predictions with experimental data for at least one strain rate and adjust A to get a fit for the strain rate data.

  1. Bergström, J. S., and M. C. Boyce. “Constitutive modeling of the large strain time-dependent behavior of elastomers.” Journal of the Mechanics and Physics of Solids 46, no. 5 (1998): 931-954

Elasto-plastic Materials

Johnson-Cook (/MAT/LAW2)

In LAW2 there are three parts to the stress calculation.

../_images/ima-ge25.jpg

Figure 155:

  • Influence of plastic strain

  • Influence of strain rate

  • Influence of temperature change

Material Parameters

There are two ways to input material parameter for LAW2.

  • Iflag=0: Classic input for Johnson-Cook parameter a, b, n is active

  • Iflag=1: New, simplified input with yield stress, UTS (engineering stress), or strain at UTS Iflag= 0

\[\sigma=a+b\cdot\varepsilon_{pn} \tag{108}\]

Where,

a

The yield stress which could be read from material test and converted to true stress.

b and n

The material parameters. Fitting the material stress-strain curve (for example, Laduga Compose script) can result in these two parameter.

Iflag = 1

With this new input, you will need yield stress Ultimate tensile engineer stress (UTS) and engineer strain at necking point. With this new input, PRADIOS automatically calculates the equivalent value for a, b and n.

../_images/image-26.jpg

Figure 156: Tension Test

Strain Rate

Strain rate has a major effect of material character on crash performance in tensile or in fracture. In Johnson-Cook theory, the yield stress is affected directly by the strain rate and is described as:

\[\sigma=\left(a+b\cdot\varepsilon_{p\eta}\right)\left(1+c\ln\frac{\xi}{\xi_{0}}\right) \tag{109}\]

Generally, yield stress increases with increasing the test strain rate. With the strain rate coefficient, c, you can scale the factor of yield stress increase. No effect of strain rate could also be defined, if c =0; or with \(\xi_{0}=10^{30} or \xi\leq\xi_{0}).\)

../_images/image271.jpg

Figure 157:

Temperature Change

Yield stress decreases with increasing temperature. In LAW2 influence is considered with \(\left(1-T^{\*m}\right).\)

\[\sigma=\left(a+b\cdot\varepsilon_{p}n\right)\left(1+c\ln\frac{\dot{\varepsilon}}{\dot{\varepsilon}_{0}}\right)\left(1-T^{\*m}\right) \tag{110}\]

with

\[T^{\*}=\frac{T-T_{r}}{T_{melt}-T_{r}} \tag{111}\]

Where,

\(T_{melt}\) Melt temperature in unit Kelvin.

\(T_{r}\) Room temperature in unit Kelvin.

With T computed as:

\[T=T_{i}+\frac{E_{int}}{\rho C_{p}\left(Volume\right)} \tag{112}\]

Where,

\(E_{int}\) Internal energy.

Change of internal energy will affect the yield stress in Johnson-Cook law.

Hardening Coefficient

Metal deformed up to yield and then generally hardened (yield stress increased). Different materials show different ways of hardening (isotropic hardening, kinematic hardening, etc.). This is also a very important material character (for spring-back).

In LAW2, use option \(C_{hard}\) (hardening coefficient) to describe which hardening model is used for the material. This feature is also available in material LAW36, 43, 44, 57, 60, 66, 73 and 74.

The value of \(C_{hard}\) is from 1 to 0. \(C_{hard}\) =0 for isotropic model, \(C_{hard}\) =1 for kinematic Prager-Ziegler model, or between 1 and 0 for hardening between the above two models.

\(C_{hard}\) = 0: Isotropic Model

In a one dimension case, material strengthens after yield stress. The maximum stress of the last tension is the yield in the subsequent loading, and this new yield stress is the same in subsequent tension and compression.

../_images/image2-8.jpg

Figure 158:

\(C_{hard}\) = 1: Kinematic Prager-Ziegler Model

To model the Bauschinger effect (after hardening in tension, there is softening in a subsequent compression which mean yield in compression is decreased), use kinematic hardening.

../_images/image2-9.jpg

Figure 159:

Elastic Plastic Piecewise Linear Material (/MAT/LAW36)

In LAW36, the numbers of plastic stress-strain curves can be directly defined for different strain rates.

Plastic stress-strain of high strain rate should always be above the lower plastic stress-stain curve.

../_images/image301.jpg

Figure 160:

Young’s Modulus

Young’s modulus can be updated (decreased) in unloading with options fct_IDE, Einf and CE. Using this feature improves the accuracy of spring-back (in unloading phase) for high strength steel. This feature is also available in material LAW43, LAW57, LAW60, LAW74 and LAW78.

  • Use fct_IDE to update the Young’s modulus (fct_IDE ≠ 0):

../_images/im-age31.jpg

Figure 161:

  • Use Einf and CE to update the Young’s modulus (fct_IDE = 0):

../_images/ima-ge32.jpg

Figure 162:

Material Behavior

fct_IDp is used to distinguish the behavior in tension and compression for certain materials (pressure dependent yield). The effective yield stress is then obtained by multiplying the nominal yield stress by the yield factor corresponding to the actual pressure.

../_images/image331.jpg

Figure 163:

See Also
/MAT/LAW2 (PLAS_JOHNS) (Starter)
RD-E: 1101 Elasto-plastic Material Law Characterization

HILL Materials

In PRADIOS material laws LAW32, LAW43, LAW72, LAW73, LAW74, LAW78 and LAW93 use HILL criteria.

HILL Criteria

The typical HILL criteria is:

  • 3D equivalent HILL stress:

\[f=\sqrt{F\big{(}\sigma_{yy}-\sigma_{zz}\big{)}^{2}+G\big{(}\sigma_{zz}-\sigma_{xx}\big{)}^{2}+H\big{(}\sigma_{xx}-\sigma_{yy}\big{)}^{2}+2L\sigma_{yz}^{2}+2M\sigma_{xx}^{2}+2N\sigma_{xy}^{2}} \tag{113}\]
\[=\sqrt{(G+H)\sigma_{xx}^{2}+(F+H)\sigma_{yy}^{2}+(F+G)\sigma_{zz}^{2}-2H\sigma_{xx}\sigma_{yy}-2F\sigma_{yy}\sigma_{zz}-2G\sigma_{zz}\sigma_{xx}+2L\sigma_{yz}^{2}+2M\sigma_{xx}^{2}+2N\sigma_{xy}^{2}}\]
  • Shell element:

\[f=\sqrt{F\sigma_{yy}^{2}+G\sigma_{xx}^{2}+H\big{(}\sigma_{xx}-\sigma_{yy}\big{)}^{2}+2N\sigma_{xy}^{2}}=\sqrt{(G+H)\sigma_{xx}^{2}+(F+H)\sigma_{yy}^{2}-2H\sigma_{xx}\sigma_{yy}+2N\sigma_{xy}^{2}} \tag{114}\]

Where, F, G, H, L, M and N are six HILL anisotropic parameters. For shell elements, only F, G, H and N are the four HILL parameters needed.

In LAW78, the HILL criteria is:

\[\varphi(A)=\frac{1}{G+H}\cdot A_{xx}^{2}-\frac{2r_{0}}{\frac{1+r_{0}}{2H}}A_{xx}A_{yy}+\frac{r_{0}(1+r_{90})}{\frac{r_{90}(1+r_{0})}{F+H}}A_{yy}^{2}+\frac{r_{0}+r_{90}}{\frac{r_{90}(1+r_{0})}{2N}}A_{xy}^{2} \tag{115}\]

There are two ways to determine HILL parameters by using Lankford parameters.

◦ Strain ratio \(r_{00}, r_{45}, r_{90}\) (LAW32, LAW43, LAW72, LAW73)

◦ Yield stress ratio \(R_{1}, R_{22}, R_{33}, R_{12}, R_{13}, R_{23}\) (LAW74, LAW93)

Strain Ratio

The Lankford parameters \(r_{\alpha}\) is the ratio of plastic strain in plane and plastic strain in thickness direction \(\varepsilon_{33}\).

\[r_{\alpha}=\frac{d\varepsilon_{\alpha\neq 2}}{d\varepsilon_{33}} \tag{116}\]

Where, \(\alpha\) is the angle to the orthotropic direction 1.

\(r_{\alpha}\) could be measured with different samples which cut in different angle with orthotropic direction 1. Like \(r_{00}\) measured from tensile test in which the loading direction is along the orthotropic direction 1. \(r_{90}\) measured from tensile test in which the loading is perpendicular to orthotropic direction 1.

The strain ratio is the strain in width direction of sample to strain in thickness direction of sample.

../_images/image3-5.jpg

Figure 164:

In this case, the HILL parameters are:

\[F=\frac{r_{00}}{r_{90}(r_{00}+1)} \tag{117}\]
\[G=\frac{1}{(r_{00}+1)} \tag{118}\]
\[H=\frac{r_{00}}{(r_{00}+1)} \tag{119}\]
\[N=\frac{(1+2r_{45})(r_{00}+r_{90})}{2r_{90}(r_{00}+1)} \tag{120}\]

Here, *G+H*=1.

In LAW32, LAW43, and LAW73, the HILL criteria is:

\[\sigma_{eq}=\sqrt{A_{1}\sigma_{1}^{2}+A_{2}\sigma_{2}^{2}-A_{3}\sigma_{1}\sigma_{2}+A_{1}\sigma_{12}^{2}} \tag{121}\]

\(R=\frac{r_{00}+2r_{4.5}+r_{90}}{4}\) \(H=\frac{R}{1+R}\)

\(A _{1}=H\Big{(}1+\frac{1}{r_{00}}\Big{)}\) \(A_{2}=H\Big{(}1+\frac{1}{r_{90}}\Big{)}\)

\(A_{3}=2H\) \(A_{12}=2H(r_{4.5}+0.5)\Big{(}\frac{1}{r_{00}}+\frac{1}{r_{90}}\Big{)}\)

They all request Lankford parameter (strain ratio) \(r_{00}, r_{4.5}, r_{90}\) and the HILL parameter \(A_{i}\) is automatically computed by PRADIOS.

Yield Stress Ratio

In LAW93, the yield stress ratio used is:

\[R_{ij}=\frac{\sigma_{ij}}{\sigma_{0}} \tag{122}\]

To get yield stress ratio \(R_{ij}\), yield stress in two loading cases need to be measured.

  • Yield stress \(\sigma_{11,}\sigma_{22,}\sigma_{33}\) from tensile test

  • Yield shear stress \(\sigma_{12,}\sigma_{13,}\sigma_{23}\) from shear test

In LAW93, if parameter input is used, then take initial stress parameter \(\sigma_{y}\) as reference yield stress \(\sigma_{0}\). If curve input is used, then take the yield stress from curve as reference yield stress \(\sigma_{0}\).

Four HILL parameters for shell are automatically computed by PRADIOS.

\[F=\frac{1}{2}(\frac{1}{R_{22}^{2}}+\frac{1}{R_{33}^{2}}\frac{1}{R_{11}^{2}}) \tag{123}\]
\[H=\frac{1}{2}(\frac{1}{R_{22}^{2}}+\frac{1}{R_{11}^{2}}-\frac{1}{R_{33}^{2}}) \tag{124}\]
\[\tag{125}\]
\[N=\frac{3}{2R_{12}^{2}} \tag{126}\]

In LAW74, yield stress ratio \(R_{ij}\) is used with yield stress \(\sigma_{11,}\sigma_{22,}\sigma_{33,}\) and \(\sigma_{12,}\sigma_{13}\sigma_{23}\) input directly, and then six HILL parameters for solid are automatically computed by PRADIOS.

\(F=\frac{1}{2}\Bigg{(}\frac{1}{\sigma_{22}^{2}}+\frac{1}{\sigma_{33}^{2}}-\frac{1}{\sigma_{11}^{2}}\Bigg{)}\) \(G=\frac{1}{2}\Bigg{(}\frac{1}{\sigma_{22}^{2}}+\frac{1}{\sigma_{33}^{2}}-\frac{1}{\sigma_{11}^{2}}\Bigg{)}\)

\(H=\frac{1}{2}\Bigg{(}\frac{1}{\sigma_{22}^{2}}+\frac{1}{\sigma_{33}^{2}}-\frac{1}{\sigma_{11}^{2}}\Bigg{)}\) \(L=\frac{1}{2\sigma_{23}^{2}}\)

\(M=\frac{1}{2\sigma_{31}^{2}}\) \(N=\frac{1}{2\sigma_{12}^{2}}\)

For shell element, take M=N and L=N.

Concrete and Rock Materials

In PRADIOS these materials can be used to represent rock or concrete materials.

These materials use a Drücker–Prager yield criterion12, which is a pressure-dependent model for determining whether a material has failed or undergone plastic yielding.

Concrete Material (/MAT/LAW10 and /MAT/LAW21)

Drücker-Prager Yield Criteria

The material has failed or undergone plastic yielding is determined by pressure using:

\[F=\frac{J_{2}}{J_{2}part}-\frac{(A_{0}+A_{1}P+A_{2}P^{2})}{I_{1}part} \tag{127}\]

Where,

\(J_{2}\) Second stress invariant (von Mises stress) of the deviatoric part of the stress and \(P=-\frac{I_{1}}{3}\).

\(I_{1}\) First stress invariant (hydrostatic pressure).

\(I_{1}=\sigma_{1}+\sigma_{2}+\sigma_{3}=-3P\)

\(J_{2}=\frac{1}{6}\bigg{[}\big{(}\sigma_{1}-\sigma_{2}\big{)}^{2}+\big{(}\sigma_{2}-\sigma_{3}\big{)}^{2}+\big{(}\sigma_{3}-\sigma_{1}\big{)}^{2}\bigg{]}=\frac{1}{3}{\sigma_{VM}2}\) in a uniaxial test.

../_images/image3-6.jpg

Figure 165: Drücker-Prager Yield Criteria

A polynomial equation is used to describe the pressure \(A_{0}+A_{1}P+A_{2}P^{2}\) at the Drucker-Prager yield surface of the material:

\[\sigma_{VM}=\sqrt{3\big{(}A_{0}+A_{1}P+A_{2}P^{2}\big{)}} \tag{128}\]

The constants of the polynomial \(A_{0}, A_{1}, A_{2}\) are determined by:

  • If \(F<0, J_{2}<A_{0}+A_{1}P+A_{2}P^{2}\) the material is under yield surface and is in the elastic region.

  • If \(F=0, J_{2}=A_{0}+A_{1}P+A_{2}P^{2}\) and the material is at the yield surface.

  • If \(F>0, J_{2}>A_{0}+A_{1}P+A_{2}P^{2}\) and the material is past the yield surface and has failed.

  • If \(A_{1}=A_{2}=0,\sigma_{VM}=\sqrt{3J_{2}}=\sqrt{3A_{0}}\), which is the von Mises criterion.

../_images/image3-7.jpg

Figure 166:

Pressure Computation

In LAW10, a polynomial equation with input parameters \(C_{0}C_{1}C_{2}C_{3}\) is used to describe the pressure. The pressure can be plotted as a function of volumetric strain.

\[\mu=\frac{\rho}{\rho_{0}}-1 \tag{129}\]
  • If \(P_{ext}=0\), the pressure is \(P=\Delta P\) and the pressure limit is \(P_{\min}=\Delta P_{\min}\).

../_images/imag-e38.jpg

Figure 167: Pressure curve without external pressure

  • If \(P_{ext}\neq 0\), the pressure is shifted by \(P_{ext}, then P=P_{ext}+\Delta P\) and the pressure limit is \(P_{\min}=P_{ext}+\Delta P_{\min}\).

../_images/image3-9.jpg

Figure 168: Pressure curve with external pressure

Here,

\[\begin{split}\Delta P=\left \{ \begin{matrix}\max\{\Delta P_{\min}\ C_{0}+C_{1}\mu+C_{2}\mu^{2}+C_{3}\mu^{3}\}if&\mu\geq 0compression \\ \max\{\Delta P_{\min}C_{0}+C_{1}\mu\}&if&\mu<0traction\end{matrix}\right|\end{split}\]
  • In traction or tension the pressure is linear and limited by \(\Delta P_{\min}\).

  • In compression the pressure is nonlinear also limited by \(\Delta P_{\min}\).

The only difference between the material laws is that in LAW10 the material constants \(C_{0} C_{0} C_{2} C_{3}\) are used to describe the pressure versus volumetric strain (\(P-\mu\) curve). In LAW21 you can describe this curve via function input fct_IDr.

Load and Unload

In LAW10 and LAW21 different loading and unloading paths of the \(P-\mu\) curve can be considered by using the parameters \(\mu_{\max}\) and B.

  • In Tension (\(\mu<0)\)

    • For LAW10, linear loading and unloading with \(P=C_{1}\mu\) (Figure 167).

    • For LAW21, loading is defined using the input function fct_IDr and linear unloading with \(P=K_{1}\mu\).

  • In Compression \(\mu > 0\), for both LAW10 and LAW21:

    • If neither B and \(\mu_{\max}\) are defined, the loading and unloading path are identical.

../_images/im-age40.jpg

Figure 169: Identical loading and unloading for LAW10 and LAW21

  • If either B or \(\mu_{\max}\) is defined:

    1. If only B is defined, \(\mu_{\max}\) is the volumetric strain where the tangent of \(P-\mu\) curve is equal to B with \(B=\frac{dP}{d\mu}\Big{|}_{\mu_{\max}}\).

    2. If only \(\mu_{\max}\) is defined, then B is the tangent of \(P-\mu\) curve at \(\mu_{\max}\). The loading and unloading in compression is:

      • If \(\mu>\mu_{\max}\), loading and unloading path are identical.

      • If \(\mu<\mu_{\max}\), loading and unloading path are different, it is linear unloading with slope B.

../_images/image-41.jpg

Figure 170: Different loading and unloading treatment for LAW10 and LAW21

Concrete Material (/MAT/LAW24)

LAW24 uses a Drücker-Prager criteria with or without a cap in yield to model a reinforced concrete material. This material law assumes that the two failure mechanisms of the concrete material are tensile cracking and compressive crushing.

Concrete Tensile Behavior

In LAW24, the options \(H_{t}, D_{sup}\), and \(\varepsilon_{\max}\) can be used to describe tensile cracking and failure in tension.

../_images/im-age42.jpg

Figure 171: LAW24 tensile loading

In the initial very small elastic phase, the material has an elastic modulus \(E_{c}\).

Once tensile strength, \(f_{t}\) is reached, the concrete starts to soften with the slope \(H_{t}\). The maximum damage factor, \(D_{sup}\), is significant because it enables the modeling of residual stiffness during and after a crack.

../_images/image4-3.jpg

Figure 172: Maximum Damage Factor Effects

The residual stiffness is computed as:

\[E=\left(1-D_{\text{sup}}\right)\cdot E_{c} \tag{131}\]

When there is crack closure, the concrete becomes elastic again, and the damage factor (for each direction) is conserved.

The bearing capacity of concrete in tensile is much lower than in compression. It is normally considered elastic when in tension.

It is recommended to choose a \(D_{sup}\) value close to 1 (default is 0.99999) in order to minimize the current stiffness at the end of the damage and consequently avoid residual stress in tension, which can become very high if the element is highly deformed due to tension. This will happen if the force causing the damage remains.

It is possible to adjust the \(D_{sup} (and H_{t})\) in order to simulate and fit the behavior of concrete reinforced by fibers. The concrete material fails once it reaches the total failure strain \(\varepsilon_{\max}\).

Concrete Yield Surface in Compression

For concrete, the yield surface is the beginning of the plastic hardening zone which is between the failure surface, \(r_{f}\), and the yield surface.

The yield surface is assumed to be the same as the failure surface in the tension zone. In compression, the yield surface is a scaled down failure surface using the factor k(\(\left(\sigma_{m},k_{0}\right))\). The yield in LAW24 for concrete is:

\[f=\frac{r}{J_{2}}-\frac{\text{k}\left(\sigma_{m},k_{0}\right)\cdot r_{f}}{I_{part}}=0 \tag{132}\]
  • For \(I_{cap}\) =0 or 1 (without a cap in yield) the yield curve is:

../_images/ima-ge44.jpg

Figure 173: Drücker-Prager Criteria without a Cap in Yield

  • For \(I_{cap}\) =2 (with cap in yield) the yield is:

../_images/im-age45.jpg

Figure 174: Drücker-Prager Criteria with Cap in Yield

\(r<k\big{(}\sigma_{m},k_{0}\big{)}\cdot r_{f}\) (green area in Figure 174)

The material is under yield in the elastic phase.

\(r\geq r_{f}\) (red area in Figure 174)

The material has failed.

\(k\big{(}\sigma_{m},k_{0}\big{)}\cdot r_{f}<r<r_{f}\) (yellow area in Figure 174)

The material is above yield and below the failure surface which is the plastic hardening phase.

The input parameter \(\rho_{t}\) is the hydrostatic failure pressure in a uniaxial tension test and \(\rho_{c}\) is the hydrostatic pressure by failure in a uniaxial compression test.

The scale factor \(k\big{(}\sigma_{m},k_{0}\big{)}\) is a function of mean stress \(\sigma_{m}\) and can be described as:

  • When \(\sigma_{m}\geq\rho_{t}\) (in tension) the scale factor \(k\big{(}\sigma_{m},k_{0}\big{)}=1)\). In this case, the yield surface equals the failure surface, \(r=r_{f}\).

../_images/imag-e46.jpg

Figure 175: k function in the tension zone

  • In the tension-compression region, \(\rho_{t}>\sigma_{m}\geq\rho_{c,}\) then

\(\mathrm{k}\big{(}\sigma_{m} k_{0}\big{)}=1+\frac{\big{(}1-k_{0}\big{)}\cdot\Big{[}\rho_{t}\big{(}2\rho_{c}-\rho_{t}\big{)}-2\rho_{c}\sigma_{m}+\sigma_{m}\Big{]}}{\big{(}\rho_{c}-\rho_{t}\big{)}^{2}} \mathrm{with} k_{y}\leq k_{0}\leq 1\)

../_images/image4-7.jpg

Figure 176: k function in the Compression-Tension zone

  • The rest of the curve depends on the \(I_{cap}\) option and the different scale factors \(k\big{(}\sigma_{m},k_{0}\big{)}\) used.

    ◦ For \(I_{cap}\) =0 or 1 and \(\sigma_{m}<\rho_{c}\) (in compression), then \(k\big{(}\sigma_{m},k_{0}\big{)}=k_{y}\)

../_images/ima-ge48.jpg

Figure 177: K function in the compression zone

◦ For \(I_{cap}\) =2 (with cap in yield) and \(\rho_{c}<\sigma_{m}<f_{k}\) (in compression), then \(k\big{(}\sigma_{m},k_{0}\big{)}=k_{y}\)

../_images/im-age49.jpg

Figure 178: K funtion in Drücker-Prager criteria without a cap

◦ In \(f_{k}<\sigma_{m}<f_{0}\) (in cap zone)

\(\mathrm{k}\big{(}\sigma_{m},k_{0}\big{)}=k_{0}\bigg{[}1-\bigg{(}\frac{\sigma_{m}-f_{k}}{f_{0}-f_{k}}\bigg{)}^{2}\bigg{]},\mathrm{with} 0\leq k_{0}\leq k_{y}\)

../_images/image5-0.jpg

Figure 179: K funtion in Drücker-Prager criteria without a cap

The material constant \(k_{y}\) should be \(0\leq k_{y}\leq 1\). A higher value of \(k_{y}\) results in a higher yield surface. For example, if \(I_{cap}=2\) (yield with cap), the difference of yield surface between \(k_{y}=0.8\) and \(k_{y}=0.6\) (Figure 180). The default value of \(k_{y}\) in LAW24 is 0.5.

../_images/imag-e51.jpg

Figure 180: After of different K function values

../_images/i-mage52.jpg

Figure 181: Drücker-Prager criteria with different K funtion values

Concrete Plastic Flow Rule in Compression

A non-associated plastic flow rule is used in LAW24. The plastic flow rule is:

\[g=\alpha I_{1}+\sqrt{J_{2}} \tag{133}\]

Where,

\(\alpha\) Plastic dilatancy.

\(\alpha=\frac{\partial g}{\partial I_{1}}\) Governs the volumetric plastic flow.

\(I_{1}\) First stress invariant (hydrostatic pressure).

Experimentally, \(\alpha\) is a linear function of \(k_{0}\):

\[\alpha=\frac{\big{(}1-k_{0}\big{)}\alpha_{y}+\big{(}k_{0}-K_{y}\big{)}\alpha_{f}}{1-K_{y}} \tag{134}\]

If\(k_{0}=K_{y}\) then, \(\alpha=\alpha_{y}\) which means the material is in yield.

If\(k_{0}<K_{y}\) then, \(\alpha\) becomes negative is the cap region.

If\(k_{0}=1\) then, \(\alpha=\alpha_{f}\) which means the material has failed.

The values of \(\alpha_{y}, \alpha_{f}\) are used to describe the material beyond yield, but before failure. It is recommended to use -0.2 and -0.1 for \(\alpha_{y}, \alpha_{f}\) in LAW24. If very small values of \(\alpha_{y}, \alpha_{f}\) are used, there is no volumetric plasticity (no cap region).

Concrete Crushing Failure in Compression

Failure surface is given by:

\[f=r-r\_{f}(\sigma_{mp} \theta)=0 \tag{135}\]

Where, \(r=\sqrt{2J_{2}}\left/\int_{c}\sigma_{m}=I_{1}\right/3f_{c}\). and \(\theta\) is Lode angle, such as:

\[\mathrm{cos}3\theta=\frac{J_{3}}{2}\bigg{(}\frac{3}{J_{2}}\bigg{)}^{3/2} \tag{136}\]

An Ottosen surface is built to design this surface using:

\[r_{f}(\sigma_{m} \theta)=\frac{1}{a}\Big{(}-b+\sqrt{b^{2}-a(\sigma_{m}-c)} \Big{)} \tag{137}\]

Where, \(a, b_{c}, b_{i} and c\) are 4 values which shape the surface and

\[b(b_{c}, b_{p}, \theta)=\frac{1}{2}\Big{[}b_{c}(1-\mathrm{cos}3\theta)+b_{1}(1+\mathrm{cos}3\theta)\Big{]} \tag{138}\]

For concrete, the compression failure curve \(r_{f}\) can be defined with a strength of:

\(f_{t}\) Uniaxial tension (triaxiality is 1/3)

\(f_{c}\) Uniaxial compression (triaxiality is -1/3)

\(f_{b}\) Biaxial compression (triaxiality is -2/3)

\(f_{2}\) Confined compression strength (tri-axial test)

\(s_{0}\) Under confined pressure

The best way to fully determine the 3D failure envelope is to get experimental data for all of these values, \(f_{c} f_{f} f_{b} f_{2} s_{0}\) which are schematically illustrated in Figure 182.

../_images/image5-3.jpg

Figure 182: Failure Parameters that Fully Determine 3D Failure Envelope

Figure 183 and Figure 184 show the points that determine the failure surface.

../_images/im-age54.jpg

Figure 183: Trace of failure surface with planar stress plane

../_images/i-mage55.jpg

Figure 184: Failure trace with several cut plan which are normal to the hydrostatic axis

From these plots that the failure envelope is not a convex surface. Figure 185 shows this behavior.

../_images/image5-6.jpg

Figure 185: Influence of the biaxial compressive strength value with all other characteristic failure points fixed

../_images/image5-7.jpg

Figure 186: Influence of the compressive strength value with all other characteristic failure points fixed

../_images/image5-8.jpg

Figure 187: Influence of the tensile strength value with all other characteristic failure points fixed

In this particular case, the compressive strength is changing but all other ratios are fixed

. This leads to an envelope scaling, as shown in Figure 188.

../_images/image5-9.jpg

Figure 188: Influence of compressive strength value

All other ratios are fixed.

Here with same strength in LAW24, but different confined compression strength \(f_{2}\)

../_images/imag-e60.jpg

Figure 189: Failure envelope on the plane stress surface influenced by the triaxial failure point

\(\left(\sigma_{1,}\sigma_{2,}\sigma_{3}\right)=\left(f_{2,}s_{0,}s_{0}\right)\)

\(f_{c}\) and the ratios \(f_{t}/f_{c}\) and \(f_{b}/f_{c}\) in the :math:`r-sigma_{m} space (used to define the concrete failure) are:

../_images/image6-1.jpg

Figure 190: Different tests (uniaxial tension, uniaxial compression, and biaxial compression) to determine failure curve

Where the failure curve is defined using \(r=\sqrt{2J_{2}}=\sqrt{\frac{2}{3}} \sigma_{YM}\) and \(\sigma_{m}=\frac{I_{1}}{3}\) is the mean stress (pressure), then \(I_{1}\) and \(J_{2}\) are the first and second stress invariants.

The material fails once it reaches the failure curve \(r_{f}\).

Concrete Reinforcement

In PRADIOS there are two different ways to simulate the reinforcement in concrete.

  • One way is to use beam or truss elements and connect them to the concrete with kinematic conditions.

  • Another way is to use the parameters in LAW24 along with the orthotropic solid property /PROP/ TYPE6 to define the reinforced direction. Parameters \(a_{1}, a_{2}, a_{3}\) in LAW24 are used to define the reinforcement cross-section area ratio to the whole concrete section area in direction 1, 2, 3.

\[a_{i}=\frac{Area_{steel}}{Area_{concrete}} \tag{138}\]

Where, \(\sigma_{y}\) is the yield stress of the reinforcement. If steel is used as a reinforcement, then \(\sigma_{y}\) is the yield stress of steel and \(E_{t}\) is the modulus of steel in the plastic phase.

../_images/image6-2.jpg

Figure 191: Stress-Strain Curve of Reinforcement (steel)

Concrete Material (/MAT/LAW81)

LAW81 can be used to model rock or concrete materials.

Drücker-Prager Yield Criteria

LAW81 uses a Drücker–Prager yield criterion where the yield surface and the failure surface are the same. The yield criteria is:

\[F=\frac{q}{J_{2}\,\text{part}}-\frac{\text{r}_{c}(p)\cdot\big{(}\text{ptan}\phi+c\big{)}}{I_{1}\,\text{part}}=0 \tag{140}\]

Where,

\(q\) von Mises stress with \(q=\sigma_{VM}=\sqrt{3J_{2}}\)

\(p\) Pressure is defined as \(p=\frac{1}{3}I_{1}\)

  1. Han, D. J., and Wai-Fah Chen. “A nonuniform hardening plasticity model for concrete materials.” Mechanics of materials 4, no. 3-4 (1985): 283-302

../_images/image-63.jpg

Figure 192: Yield Surface (LAW81)

The yield surface can be described in two parts:

  • The linear part image7 (\(p\leq p_{a})\), where the scale function is \(\mathrm{r}_{c}(p)=1\) which leads to the von Mises stress being linearly proportional to pressure:

\[q=p\mathrm{tan}\phi +c \tag{141}\]

Where,

\(c\) Cohesive and is the intercept of yield envelope with the shear strength.

If \(c\) =0, the material has no strength under tension.

\(\phi\) Angle of internal friction, which defines the slope of the yield envelope.

\(c\) and \(\phi\) are also used to define the Mohr-Coulomb yield surface. The Drücker-Prage yield surface is a smooth version of the Mohr-Coulomb yield surface.

  • The second part image8 (\(p_{a}<p<p_{b})\) of the yield surface simulates a cap limit. An increase of pressure in a rock or concrete material will increase the yield of the material; but, if pressure increases enough, then the rock or concrete material will be crushed. The Drücker-Prager model with the cap limit can be used to model this behavior. The cap limit defined in part image9 and uses the scale function:

\[\mathrm{r}_{c}\big{(}p\big{)}=\sqrt{1-\left(\frac{p-p_{a}}{p_{b}-p_{a}}\right)^{2}} \tag{142}\]

The von Mises stress is:

\[q=\sqrt{1-\left(\frac{p-p_{a}}{P_{b}-\overline{p}_{a}}\right)^{2}}\cdot\left(p\text{tan}\phi+c\right) \tag{143}\]

Where,

\(P_{b}\) Curve is defined using the \(fct_ID_{Pb}\) input

\(p_{a}\) Computed by PRADIOS using the input \(a\) ratio value.

\(p_{a}=\alpha\cdot P_{b}\) with \(0<\alpha <1\).

Where, \(P_{0}\) is the maximum point of yield curve, where \(\frac{\partial F}{\partial p}\left(p_{0}\right)=0\)

If \(p=p_{b}\), then \(r_{\text{c}}\left(p_{b}\right)=0\) and the yield function is then,

\(q=0:\cdot\left(p\text{tan}\phi+c\right)=0\) which means the material is crushed.

The input parameters \(\phi, c, p_{b^{\prime}}\alpha\) need to determine for the Drucker-Prager yield surface. At least four tests are needed to fit these parameters. In the simplest case, uniaxial tension and uniaxial compression can be used to determine the linear part, \(\phi, \text{and} c\). To determine \(p_{b^{\prime}\) and \(\alpha\) biaxial compression tests and compression/compression tests are needed (refer to CC00 and CC01 in RD-E: 4701 Concrete Validation with Kupfer Tests).

../_images/image-66.jpg

Figure 193: Yield Surface of LAW81 Showing Different Load Conditions

For most materials such as metal, the plastic strain increment could be considered normal to yield surface. However, if the plastic strain increment normal to yield surface is used for rock or concrete materials, the plastic volume expansion is overestimated. Therefore, a non-associated plastic flow rule is used in these materials. In LAW81 the plastic flow function G defined as:

  • \(G=q-p\cdot\tan\psi=0 if p\leq p_{a}\)

  • \(G=q-\tan\psi\left(p-\frac{\left(p-p_{a}\right)^{2}}{2\left(p_{0}-p_{a}\right)}\right)=0 if p_{a}<p\leq p_{0}\)

  • \(G=F if p>p_{0}\)

Since the pressure is \(p_{0}\), the yield function F and plastic flow function G are the same and the following condition is fulfilled:

\[G(p_{0})=F(p_{0}) \tag{144}\]
\[\frac{\partial G}{\partial p}|{p_{0}}=\frac{\partial F}{\partial p}|{p_{0}}=0 \tag{145}\]

The pressure \(p\_{0}\) can be calculated using the yield surface where \(\frac{\partial F}{\partial p}| p_{0}=0\). With G defined as:

\[G=q-\tan\psi\left(p-\frac{\left(p-p_{a}\right)^{2}}{2\!\left(p_{0}-p_{a}\right)}\right)=0 \tag{146}\]

The parameter \(\psi\) can be determined using the von Mises stress at pressure, \(p_{0}\) in the function.

../_images/ima-ge67.jpg

Figure 194: Yield Surface of LAW81 with plastic flow

Failure

Ductile

The /FAIL/BIQUAD, /FAIL/JOHNSON, and /FAIL/TAB1 failure models define material failure by relating the plastic strain at failure to the stress state in the material.

These failure models are often used to describe the ductile failure of materials. The state of stress in the material can be defined by using stress triaxiality.

Stress Triaxiality (Normalized Mean Stress)

For ductile materials, the state of stress (compression, shear, tension, etc.) of the material affects the plastic strain value at which the material will fail. An important and useful characteristic to describe the state of stress, stress triaxiality is defined as:

\[\sigma^{*}=\frac{\sigma_{m}}{\sigma_{VM}} \tag{147}\]

Where,

\(\sigma_{m}=\frac{1}{3}\big{(}\sigma_{1}+\sigma_{2}+\sigma_{3}\big{)}\) Mean (hydrostatic) stress

\(\sigma_{VM}=\sqrt{\frac{1}{2}\big{[}\big{(}\sigma_{1}-\sigma_{2}\big{)}^{2}+\big{(}\sigma_{2}-\sigma_{3}\big{)}^{2}+\big{(}\sigma_{3}-\sigma_{1}\big{)}^{2}\big{]}}\) Mises stress

Triaxiality values for some commons stress states can be derived as:

  • In pure tension:

    \(\sigma_{2}=\sigma_{3}=0\) then \(\sigma^{*}=\frac{\sigma_{m}}{\sigma_{VM}}=\frac{1}{3}\)

  • In biaxial compression:

    \(\sigma_{1}=\sigma_{2}\), and \(\sigma_{3}=0\), then \(\sigma^{*}=\frac{\sigma_{m}}{\sigma_{VM}}=-\frac{2}{3}\)

Stress triaxiality for various stress states:

Stress Triaxiality \(\sigma^{*}\) Stress State

\(-\frac{2}{3}\) Biaxial compression

\(-\frac{1}{3}\) Uniaxial compression

\(\textbf{0}\) Pure shear

\(\frac{1}{3}\) Uniaxial tension

\(\frac{1}{\sqrt 3}\) Plain strain

\(\frac{2}{3}\) Biaxial tension

/FAIL/JOHNSON

The Johnson-Cook failure model is often used to describe the ductile failure of metals. It uses a Johnson-Cook equation to define failure strain as a function of stress triaxiality.

In the Johnson-Cook failure model, there are three parts to the failure model;

\[\varepsilon_{f}=\big{[}D_{1}+D_{2} exp(D_{3}\sigma^{*})\big{]}\big{[}1+D_{4}in(\varepsilon^{*})\big{]}\big{[}1+D_{5}T^{*}\big{]} \tag{148}\]

Where,

\(\varepsilon_{f}\) Plastic failure strain

\(\varepsilon^{*}=\frac{\varepsilon}{\varepsilon_{0}}\) Current strain rate divided by the input reference strain rate

\(T^{*}\) Computed in the material law or /HEAT/MAT

Ignoring the influence of strain rate and temperature a plot of the Johnson-Cook failure is:

../_images/image6-9.jpg

Figure 195: Example Plot of a Johnson-Cook Failure Model

Plastic strains above the curve represent material fracture and below the curve no material fracture.

In a simple case where only the triaxiality influence is considered, the failure strain is:

\[\varepsilon_{f}=\mathrm{D}_{1}+\mathrm{D}_{2}\cdot\exp\bigl{(}\mathrm{D}_{3}\cdot\sigma^{\bullet}\bigr{)} \tag{149}\]

Using 3 failure data points from test:

\(\varepsilon_{f}=0.1585\) by uniaxial tension (\(\sigma^{\bullet}=1/3)\)

\(\varepsilon_{f}=0.19\) by pure shear (\(\sigma^{\bullet}=0)\)

\(\varepsilon_{f}=0.2419\) by uniaxial compression (\(\sigma^{\bullet}=-1/3)1\)

The parameters \(\mathrm{D}_{1}, :math:\)mathrm{D}_{2}` and \(\mathrm{D}_{3}\) could be calculated analytically by solving the following equations:

\[\tag{150}\]
\[\begin{split}\left \{ \begin{matrix} 0.1585=\mathrm{D}_{1}+\mathrm{D}_{2}\cdot\exp\bigl{(}\mathrm{D}_{3}\cdot \tfrac{1}{3}\bigr{)} \\ 0.19=\mathrm{D}_{1}+\mathrm{D}_{2}\cdot\exp\bigl{(}\mathrm{D}_{3}\cdot 0\bigr{)} \\ 0.2419=\mathrm{D}_{1}+\mathrm{D}_{2}\cdot\exp\bigl{(}\mathrm{D}_{3}\cdot-\tfrac{1}{3}\bigr{)}\end{matrix}\right|\end{split}\]

Element Failure treatment

A cumulative damage method is used to sum the amount of plastic strain that has occurred in the element using:

\[D=\sum\frac{\Delta\varepsilon_{p}}{\varepsilon_{f}}\geq 1\tag{151}\]

What happens when \(D\geq 1\) depends on the values of element failure flags (\(I_{fail_sh} and I_{fail_so})\) and XFEM formulation flag (\(I_{\textit{xfem}})\). When the XFEM formulation is not used (\(I_{\textit{xfem}})=0\), the following table summarizes the different element failure flag options:

Table 18: Element failure option

Element

Element failure flag

If \(D\geq 1\)

Failure behavior

Shell

\(I_{fail_sh}\) =1 (Default)

In 1 IP or layer

Element deleted

Shell

\(I_{fail_sh}\) =2

In 1 IP or layer

stress tensor set to zero im IP or layer

Shell

\(I_{fail_sh}\) =2

ALL IP or layer

Element deleted

Solid

\(I_{fail_sh}\) =1 (Default)

In 1 IP

Element deleted

Solid

\(I_{fail_sh}\) =2

In 1 IP

stress tensor set to zero im IP

Solid

\(I_{fail_sh}\) =2

ALL IP

stress tensor set to zero in element

Details on the XFEM formulation (\(I_{\textit{xfem}})=1)\), can be found in /FAIL/JOHNSON.

The damage, D, can be plotted in animation files using /ANIM/SHELL/DAMA or /ANIM/BRICK/DAMA. This will show the risk of material damage.

/FAIL/BIQUAD

In PRADIOS, /FAIL/BIQUAD is the most user-friendly failure model for ductile materials. It uses a simplified, nonlinear strain-based failure criteria with linear damage accumulation.

The failure strain is described by two parabolic functions calculated using curve fitting from up to 5 user input failure strains.

By default, /FAIL/BIQUAD (``S-Flag``=1) uses two parabolic curves to describe the plastic failure strain \(\varepsilon_{f}\), as a function of stress triaxiality \(\sigma\). The two parabolic curves use:

\[f_{2}(x)=dx^{2}+ex+f \tag{152}\]

Where,

\(a, b, c, d, e,\) and \(f\) Parabolic coefficients

\(x\) Stress triaxiality

\(f_{1}(x)\) and \(f_{2}(x)\) Plastic failure strain

../_images/ima-e71.jpg

Figure 196: /FAIL/BIQUAD Failure Strain Curve Made of 2 Parabolic

The parabolic coefficients \(a, b, c, d, e,\) and \(f\) are computed by PRADIOS using a curve fit based on the plastic failure strain c1-c5 input values. If the calculated parabolic failure strain curves have negative failure strain values, these negative values will be replaced by a failure strain of 1E-6 which results in a very high damage accumulation and brittle behavior. The results of the curve fit are in the Starter *0000.out file.

   Bi-Quadratic FAILURE
     - - - - - - - - - - - - - -
c1. . . . . . . . . . . . . . . . . . .= 0.2419E+00
c2. . . . . . . . . . . . . . . . . . .= 0.1900E+00
c3. . . . . . . . . . . . . . . . . . .= 0.1585E+00
c4. . . . . . . . . . . . . . . . . . .= 0.1437E+00
c5. . . . . . . . . . . . . . . . . . .= 0.1394E+00
   COEFFICIENTS OF FIRST PARABOLA
     - - - - - - - - - - - - - -
a . . . . . . . . . . . . . . . . . . .= 0.9180E-01
b . . . . . . . . . . . . . . . . . . .= -0.1251E+00
c . . . . . . . . . . . . . . . . . . .= 0.1900E+00
   COEFFICIENTS OF SECOND PARABOLA
     - - - - - - - - - - - - - -
d . . . . . . . . . . . . . . . . . . .= 0.3753E-01
e . . . . . . . . . . . . . . . . . . .= -0.9483E-01
f . . . . . . . . . . . . . . . . . . .= 0.1859E+00

The c1c5 plastic failure strains definitions are:

c1

Plastic failure strain in uniaxial compression

c2

Plastic failure strain in shear

c3

Plastic failure strain in uniaxial tension

c4

Plastic failure strain in plane strain tension

c5

Plastic failure strain in biaxial tension

M-Flag Input Options

Depending on the M-Flag input option, there are three different ways to define the c1-c5 values.

    • M-Flag=0, User-defined Test Data For this case, you must enter c1-c5 which represents the plastic failure strain for the 5 different stress states. Ideally this data would be obtained from test or the material supplier.

    • M-Flag=1-7, Predefined Material Data If failure strain data is not available, you can pick from 7 predefined materials. Figure 197 shows the plastic strain at failure curves for the 7 materials.

image12 Note: The predefined values are supplied for early design exploration and it is your responsibility to verify that their material has the same properties.

../_images/image72.png

Figure 197: Predefined Material Failure Curves

  • M-Flag=99, Plastic Failure Strain Ratio Input, r1-r5

The last input method is to enter the plastic failure strain in uniaxial tension, c3, and plastic failure strain ratios for the other four stress states. These ratios are defined as:

r1

Failure plastic strain ratio, Uniaxial Compression (c1) to Uniaxial Tension (c3), so \(c1=r1\cdot c3\)

r2

Failure plastic strain ratio, Pure Shear (c2) to Uniaxial Tension (c3), so \(c2=r2\cdot c3\)

r4

Failure plastic strain ratio, Plane Strain Tension (c4) to Uniaxial Tension (c3), so \(c4=r4\cdot c3\)

r5

Failure plastic strain ratio, Biaxial Tension (c5) to Uniaxial Tension (c3), so \(c5=r5\cdot c3\)

Using this method, it is easy to change the failure curve by adjusting the single plastic failure strain in uniaxial tension value, c3.

../_images/im-age73.jpg

Figure 198: Changes in Plastic Failure Strain Curve by increasingthe uniaxial tension failure, c3, with the same failure plastic strain ratios

Default Behavior

By default, values different than 0 need to be entered for c1 to c5. However, specific default behaviors exists, in case failure information are missing.

  • In case the material failure behavior is unknown, c1 to c5 are set to 0.0 and the mild steel behavior (M-Flag=1) is used.

  • If only the tensile failure value is known, c3 is defined (c1=c2=c4=c5=0.0). The mild steel behavior is used and scaled by the user- defined c3 value.

  • In case the material behavior is known, M-Flag is defined and c3 can be used to adjust the failure model according the expected tensile failure. The selected material behavior is scaled by the userdefined c3 value.

  • For all other cases, all c1 to c5 are intended to be defined and default value of 0.0 is used.

Element Failure Treatment

A cumulative damage method is used to sum the amount of plastic strain that has occurred at each integration point in the element using:

\[D=\sum\frac{\Delta\varepsilon_{p}}{\varepsilon_{f}}\geq 1 \tag{154}\]

Where,

D Damage

\(\Delta\varepsilon_{p}\) The change in plastic strain of the integration point

\(\varepsilon_{f}\) Plastic failure strain for the current stress triaxiality

In shell elements after an integration point reaches D*=1, the integration point’s stress tensor is set to zero. The element fails and is deleted when the ratio of through thickness failed integration points equals *P_thickfail. In solid elements, the element is deleted when any integration point reaches .

Plane Strain as Global Minimum

The S-Flag=2 option can be used to force the global minimum of the plastic failure strain curve to occur at the plane strain stress triaxiality location, c4. This is accomplished by splitting the second equation into 2 separate quadratic sub-functions.

Modeling Material Instability (Localized Necking)

In materials such as sheet metal, thickness thinning and diffuse necking may appear during tensile loading of the material. This is called localized necking and normally occurs in the stress triaxiality range of \(\frac{1}{3}\leq\sigma^{\bullet}\leq\frac{2}{3}:\)

../_images/image7-4.jpg

Figure 199:

In /FAIL/BIQUAD it is possible to simulate this localized necking using the option, S-Flag=3 and Inst_start. This option uses the same plastic failure strain curve as S-Flag=2 and adds two additional quadratic functions that define a curve that represents the start of localized necking between stress triaxiality \(\frac{1}{3}\) and \(\frac{2}{3}\). The minimum value of this curve is a user-defined value in the Inst_start field and occurs at plane strain tension \(\sigma^{*}=\frac{1}{\sqrt{3}}\). Using this localized necking curve, a second localized necking damage value is calculated and failure due to necking only occurs when all integration points reach \(D=\mathrm{I}\). The localized necking criteria is based on the Marciniak-Kuczynski analysis.13

../_images/im-age75.jpg

Figure 200: Default Failure Strain Curve with additional localized necking curve (blue)

When using S-Flag=1 or 2, the damage accumulation begins once the plastic strain reaches in failure curve (red in Figure 201).

If S-Flag=3is used to describe the localized necking, damage accumulation begins once the plastic strain reaches in the localized necking curve (blue curve in Figure 201). For localized necking, the element is deleted when all integration points reach damage, D*=1, whereas element deletion not due to localized necking is defined by *P_thickfail.

../_images/i-age76.jpg

Figure 201:

Perturbation of the Failure Limit

Due to a materials imperfection or production process, a material’s failure strain may not be exactly the same everywhere and thus very small perturbations of failure limit may exist. When using M-Flag>0 with /PERTURB/FAIL/BIQUAD, a statistical distribution of the failure limit is applied to each element assigned the failure model. This is accomplished by calculating the normal or random distribution of a failure scale factor which is applied to /FAIL/BIQUAD, c3. The two different distributions methods are shown in Figure 202 and Figure 203.

../_images/imag-e77.jpg

Figure 202: Random Distribution Idistri=1, of the failure limit in the Starter 0.out file

../_images/i-mage78.jpg

Figure 203: Normal (Gaussian) Distribution Idistri=2, of the failure limit in the Starter *0.out file

/FAIL/BQUAD uses small perturbations of c3 generated by /PERTURB/FAIL/BIQUAD to scale the entire failure curve using the strain ratio values for the predefined materials or user-defined ratios, r1, r2, r4 and r5, depending on the value of M-Flag.

../_images/imuge104.jpg

Figure 204:

Reference

Tabulated Failure Model /FAIL/TAB1

In PRADIOS, /FAIL/TAB1 is the most sophisticated failure model for ductile material. The plastic failure strain can be defined as a function of: stress triaxiality, strain rate, Lode angle, element size, temperature, and instability strain.

Damage is accumulated based on user-defined functions. The functionality of this failure model will be described starting with the most basic input to the most complex options.

Plastic Failure Strain

Similar to /FAIL/JOHNSON and /FAIL/BIQUAD, it is possible to define a curve that represents the plastic failure strain, \(\varepsilon_{f}\), as a function of stress triaxiality, \(\sigma_{*}\). Unlike /FAIL/JOHNSON and /FAIL/BIQUAD where the failure strain curve is defined using parameters in predefined equations, in /FAIL/TAB1 any number of discrete points can be entered to create an arbitrary function that represents the failure strain curve. This curve is defined using the /TABLE entity and is referenced in table_ID1. This method can be used for shell and solid elements.

  1. Pack, Keunhwan, and Dirk Mohr. “Combined necking & fracture model to predict ductile failure with shell finite elements.” Engineering Fracture Mechanics 182 (2017): 32-51

../_images/i-mage80.jpg

Figure 205: Material Failure Curve Defined using Discrete Points with a Local Maximum

Example /TABLE, dimension=1

Input for failure plastic-strain versus triaxiality using table_ID1.

/TABLE/1/4711
failure plastic-strain vs triaxiality
#dimension
      1
# Triaxiality  Failure_Strain
   -0.7000     0.3386
   -0.6000     0.3068
   -0.5000     0.2794
   -0.4000     0.2558
   -0.3333     0.2419
   -0.3000     0.2355
   -0.2000     0.2180
   -0.1000     0.2029
   0.0000      0.1900
   0.1000      0.1789
   0.2000      0.1693
   0.3000      0.1610
   0.3333      0.1585
   0.4000      0.1539
   0.5000      0.1478
   0.6000      0.1425
   0.7000      0.1380

Strain Rate Dependency

/FAIL/TAB1 can also include the influence of strain rate on material failure. For this case the /TABLE must be defined such that, the first dimension is the function ID for the failure curve and the second dimension is the strain rate where that failure curve is applied.

Example /TABLE, dimension=2

/TABLE/1/4711
failure plastic-strain vs triaxiality and strain rate
#dimension
        2
#   FCT_ID       strain_rate
     3000        1E-4
     3001        0.1
     3002        1.0
/FUNCT/3000
failure plastic-strain vs triaxiality
#   Triaxiality  Failure_Strain
     -0.7000     0.3386
     -0.6000     0.3068
     -0.5000     0.2794
     -0.4000     0.2558
     -0.3333     0.2419
     -0.3000     0.2355
     -0.2000     0.2180
     -0.1000     0.2029
     0.0000      0.1900
     0.1000      0.1789
     0.2000      0.1693
     0.3000      0.1610
     0.3333      0.1585
     0.4000      0.1539
     0.5000      0.1478
     0.6000      0.1425
     0.7000      0.1380
/FUNCT/3001
failure plastic-strain vs triaxiality
#   Triaxiality  Failure_Strain
     -0.7        0.27088
     -0.6        0.24544
     -0.5        0.22352
     -0.4        0.20464
     -0.3333     0.19352
     -0.3        0.1884
     -0.2        0.1744
     -0.1        0.16232
     0           0.152
     0.1         0.14312
     0.2         0.13544
     0.3         0.1288
     0.3333      0.1268
     0.4         0.12312
     0.5         0.11824
     0.6         0.114
     0.7         0.1104

Lode Angle with Solid Elements

For solid elements, the failure strain can also depend on the 3D stress state defined using the Lode angle.

This can be included by adding failure strain as a function of Lode angle parameter in the /TABLE entity referenced by table_ID1. For shells elements, it is only necessary to define the failure strain as a function of stress triaxiality. But for solid elements, it is more accurate to include the failure strain as a function of stress triaxiality and Lode angle.

In PRADIOS, the Lode angle is entered using a normalized and dimensionless Lode Angle parameter \(\zeta\) and is defined here.

The stress state at a point, P, could be expressed with principle stresses (\(\sigma_{1},\sigma_2{},\sigma_{3})\) or could also be expressed using stress invariants \((I_{1},J_{2},J_{3})\). The advantage of using stress invariants is that they are constant and do not depend on the orientation of the coordinate system. In Figure 206, to correctly describe stress state of point P \(\sigma_{1},\sigma_{2},\sigma_{3}\) using stress invariants, the magnitude of OO’ as:

\[\sqrt{3}\sigma_{m}=\frac{\sqrt{3}}{3},I_{1} \tag{155}\]

Where,

\(\sigma\) Mean stress

\(I_{1}\) First stress invariant \(I_{1}=\sigma_{1}\sigma_{2}\sigma_{3}\)

OO’ is in the hydrostatic axis, which means the principle stress in this axis is the same (\(\sigma_{1}=\sigma_{2}=\sigma_{3})\).

OO is the hydrostatic pressure.

../_images/1image81.jpg

Figure 206: Stress State of P

The magnitude of O’P is:

\[\sqrt{2J_{2}}=\sqrt{\frac{2}{3}} \sigma_{VM} \tag{156}\]

Here, \(J_{2}\) is the second invariant of deviatoric stress \(\mathbf{s(s=\sigma-p)}\) with

\(J_{2}=\frac{1}{2}(S_{1}2+S_{2}2+S_{3}2)=\frac{1}{2}\Big{[}\big{(}\sigma_{1}-\sigma_{2}\big{)}^{2}+\big{(}\sigma_{2}-\sigma_{3}\big{)}^{2}+\big{(}\sigma_{3}-\sigma_{1}\big{)}^{2}\Big{]}\).

To identify the point P, the angle in the circular plane must be calculated. This angle is called the Lode Angle \(\theta\):

\[\cos\big{(}3\theta\big{)}=\frac{27}{2}\frac{J_{3}}{\sigma_{VM}^{3}}=\frac{3\sqrt{3}}{2}\frac{J_{3}}{J_{2}^{3/2}} \tag{157}\]

with \(0\leq\theta\leq\frac{\pi}{3}\) and \(J_{3}\) is the third invariant of deviatoric stress calculated as:

\[J_{3}=S_{1}S_{2}S_{3} \tag{158}\]

In /FAIL/TAB1 a normalized and dimensionless Lode Angle parameter \(\zeta\) with the range \((-1\leq\zeta\leq 1)\) is used and defined as:

\[\zeta=\cos\left(3\theta\right) \tag{159}\]
../_images/image83.jpg

Figure 207: Stress state with different lode angle

The special character for lode angle \(\theta\) and load angle parameter \(\zeta\) are:

load angle parameter \(\zeta\)

lode angle \(\theta\)

Strass State

1

0

Uniaxial tension + hydrostatic pressure (triaxial tension or axisymmetric tension)

0

30

Pure shear + hydrostatic pressure (plane strain)

-1

60

Uniaxial compression + hydrostatic pressure (axisymmetric compression)

A failure strain surface can be created from the stress triaxiality and Lode angle failure data.

../_images/image894.jpg

Figure 208: 3D Failure Surface

The material failure surface could be created using the following material tests.

../_images/image8315.jpg

Figure 209: Stress State and Lode Angle for Various Tests

Example /TABLE, dimension=3

Input for failure plastic-strain versus triaxiality, strain rate, and Lode angle using table_ID1

/TABLE/1/4711
failure plastic-strain vs triaxiality and strain rate
#dimension
         3

#---1----|----2----|----3----|----4----|----5----|----6----|----7----|----8----|----9----|---10----|
# FCT_ID                      strain_rate           Lode_angle
    3000                            1E-4                   -1
    3001                             0.1                    0
    3002                             1.0                    1
. . . . . . . . .
../_images/image80-6.jpg

Figure 210: Failure Surface when table_ID1 references a /TABLE with dimension=3

Scaling Failure Strain

Material failure based on temperature and element size can be considering in /FAIL/TAB1 by including functions that scale the failure strain depending on the element size and/or temperature using:

\[\varepsilon_{f}=Xscale1\cdot f(\sigma^{*},\xi,\xi)\cdot factor_{el}\cdot factor_{T} \tag{160}\]

Where,

Xscale1 General scaling factor

\(f actor_{el}\) Scale factor based on element size

\(f actor_{T}\) Scale factor based on temperature

Element Length Dependency

In numerical simulations, the element size will affect the material failure. Using the same failure parameters, a coarse mesh will fail earlier than a fine mesh.

../_images/image871.jpg

Figure 211: Influence of Element Mesh Size on Material Failure in Numerical Simulation

To account for the variation in results based on mesh size, the element size scale factor can be defined to scale the failure strain based on mesh element size. The scale factor defined in Equation 160 is:

\[f actor_{el}=Fscale_{el}\cdot\mathsf{f}_{el}\Big{(}\frac{Size_{el}}{EI_{-}ref} \Big{)} \tag{161}\]

Where,

\(F scale_{el}\) Element size function scale factor

\(\mathsf{f}_{el}\Big{(}\frac{Size_{el}}{EI_{-}ref}\Big{)}\) failure strain scale factor as a function of the normalized element size (defined via \(fct_ID_{el}\)), and EI_ref is the reference element size used to normalize the element size

For example, a mesh size of 2mm was used in a material validation simulation. After the initial validation, the same simulation is reran using EI_ref =2 with different element sizes to determine the correct scaling factor for each element size. Using the results of the second validation, a failure strain scale factor function is constructed as (Figure 212):

../_images/i1mage88.jpg

Figure 212: Example of Element Size Scale Factor Function fct_IDel in /FAIL/TAB1

Temperature Dependency

To account for how the temperature affects material failure, failure strain values defined in Equation 160 can be scaled using a temperature scale factor function:

\[factor_{T}=Fscale_{T}\cdot\mathrm{f}_{T}(T_{start}) \tag{162}\]

Where,

\(F scale_{T}\) Temperature function scale factor

\(\mathrm{f}_{T}(T_{start})\) Failure strain scale factor as a function temperature defined via \(fct_ID_{T}\)

Scaling the failure strain based on temperature works with any material that has /HEAT/MAT defined or materials that include thermos plasticity such as /MAT/LAW2 (PLAS_JOHNS).

In \(fct_ID_{T}\) the temperature is defined relative to the melting and initial temperature.

\[T^{*}=\frac{T-T_{ini}}{T_{melt}-T_{ini}} \tag{163}\]
../_images/im2age89.jpg

Figure 213: Influence of Temperature on Material Failure with fct_IDT in /FAIL/TAB1

Element Failure Treatment

In /FAIL/TAB1 an accumulative damage model is used. The damage can be output for contour plotting using the Engine options, /ANIM/SHELL/DAMA or /ANIM/BRICK/DAMA.

The accumulated damage is calculated by first calculating a damage increment:

\[\Delta D=\frac{\Delta E_{p}}{\varepsilon_{f}}\cdot n\cdot D_{p}(\frac{1}{n})\tag{164}\]

Where,

\(\Delta\varepsilon_{p}\) The change in plastic strain of the integration point

\(\varepsilon_{f}\) Plastic failure strain for the current stress triaxiality

\(D_{p}\) and n Damage parameters

The accumulated damage is:

\[D=\frac{\sum_{\Delta D}}{D_{crit}}\tag{165}\]

Where, \(D_{crit}\) is defined in /FAIL/TAB1 with a recommended between 0 and 1. The elements fail when \(D\geq 1\) which means \(\sum\Delta D\geq D_{crit}\).

../_images/image9-0.jpg

Figure 214: Influence of \(D_{crit} in\) * /FALL/TAB1

It is also interesting to understand the influence of the damage accumulation parameter, in Equation 164.

When *n*=1 damage is linear, otherwise the damage curve is nonlinear.

../_images/image9-1.jpg

Figure 215: Influence of damage parameter *n

Material Instability (Diffuse Necking)

In a tensile test, a material will reach a maximum engineering stress and then the stress will decrease or sometimes called soften. This maximum engineering stress point is called the necking point. After necking, the true stress actually increases, due to the decrease in the cross sectional area. This is called diffuse necking. In sheet metal, thickness thinning or localized necking in diffuse area may appear if the material continues to be loaded in the tensile direction. The diffuse necking normally appears in stress triaxiality ranges of \(0<\sigma^{*}\leq\frac{2}{3}\) and localized necking in the range of \(\frac{1}{3}\leq\sigma^{*}\leq\frac{2}{3}\).

../_images/imag3e92.jpg

Figure 216: Sketch of Diffuse Necking and Localized

In /FAIL/TAB1 it is possible to account for the material instability (diffuse necking) with options table_ID2, or Inst_start and *Fad_exp

The material will start to weaken due to diffuse necking when the instability strain is reached. The reduced stress in the material is:

\[\sigma_{reduced}=\sigma\cdot\left(1-\left(\frac{D_{instability}-inst_start}{1- inst_start}\right)^{Fad_exp}\right)\tag{166}\]

Where,

\[D_{instability}=\sum\frac{\Delta\varepsilon_{p}}{\varepsilon_{f}}\tag{167}\]

with \(\varepsilon_{f}\) being the diffuse necking strain based on the current stress triaxiality.

The strain at which instability starts could be either input with a curve (blue curve in Figure 217) using table_ID2 or input as a constrain strain using the option Inst_start.

In a uniaxial tensile test \(\sigma^{*}=\frac{1}{3}\).

  • If material instability is not included then damage is calculated using the red failure curve in Figure 217

  • If material instability is included and modeled using the curve input (table_ID2) then:

    • Damage due to diffuse necking starts when the plastic strain defined by the blue curve in Figure 217 is reached.

    • Damage due to diffuse necking is linear if Fad_exp=1 is used. Increasing the Fad_exp leads to more energy dissipated during damage. Figure 217 shows the influence of Fad_exp from 1 to 10 in stress-strain curve. It is recommend to use a Fad_exp value of 5 to 10.

    • Once the strain reaches the red curve the element fails.

../_images/image9u3.jpg

Figure 217: Influence of Parameter Fad_exp and Material Instability Region

  • If only Inst_start is used without curve input in table_ID2, then the diffuse necking plastic strain is the constant value, Inst_start, for all stress triaxiality.

../_images/imauyge94.jpg

Figure 218: Constant Inst_start Describing Diffuse Necking

Currently diffuse necking (material instability) in /FAIL/TAB1 can only be used with material law numbers > 28.

  1. Wierzbicki, Tomasz, “Addendum to the Research Proposal on Fracture of Advanced High Strength Steels”, page 19, January 2007.

  2. Wierzbicki, Tomasz. “Fracture of AHSS Sheets–Addendum to the Research Proposal on Fracture of Advanced High Strength Steels.” Impact and Crashworthiness Laboratory, Massachusetts Institute of Technology (2007).

Composites

Composite materials consist of two or more materials combined each other. Most composites consist of two materials, binder (matrix) and reinforcement. Reinforcements come in three forms, particulate, discontinuous fiber, and continuous fiber.

Composite Material

In PRADIOS the following material laws are used to describe composite material:

  • LAW12 and LAW14

  • LAW15 (recommended to use LAW25+/FAIL/CHANG)

  • LAW25

  • LAW19 (For Fabric and with only /PROP/TYPE9)

  • LAW58 (For Fabric only with /PROP/TYPE16)

table 19: Composite material, propert, failure model, and element type compatibility

Shell Element

Brick Element

Failure Model

LAW12

/PROP/TYPE6 (SOL_ORTH)

/PROP/TYPE21 (TSH_ORTH)

/PROP/TYPE22 (TSH_COMP)

/FAIL/HASHIN

/FAIL/PUCK

/FAIL/LAD_DAMA

LAW14

/PROP/TYPE6 (SOL_ORTH)

/PROP/TYPE21 (TSH_ORTH)

/PROP/TYPE22 (TSH_COMP)

/FAIL/HASHIN

/FAIL/PUCK

/FAIL/LAD_DAMA

LAW15

/PROP/TYPE9 (SH_ORTH)

/PROP/TYPE10 (SH_ORTH)

/PROP/TYPE11 (SH_SANDW)

/PROP/TYPE17 (STACK)

/PROP/TYPE19 (PLY)

/FAIL/CHANG

LAW25

/PROP/TYPE10 (SH_ORTH)

/PROP/TYPE11 (SH_SANDW)

/PROP/TYPE17 (STACK)

/PROP/TYPE19 (PLY)

/PROP/TYPE51 /PROP/PCOMPP

/PROP/TYPE6 (SOL_ORTH)

/PROP/TYPE14 (SOLID)

/PROP/TYPE20 (TSHELL)

/PROP/TYPE21 (TSH_ORTH)

/PROP/TYPE22 (TSH_COMP)

/FAIL/CHANG

/FAIL/HASHIN

/FAIL/PUCK

/FAIL/LAD_DAMA

(for solid only)

LAW19

/PROP/TYPE9 (SH_ORTH)

LAW58

/PROP/TYPE16 (SH_FABR)

LAW12 and LAW14

Describes orthotropic solid material which use the Tsai-Wu formulation. The materials are 3D orthotropic-elastic, before the Tsai-Wu criterion is reached. LAW12 is a generalization and improvement of LAW14.

Elastic Phase

Both material laws require Young’s modulus, shear modulus and Poisson ratio (9 parameters) to describe the material orthotropic in elastic phase.

../_images/image9-5.jpg

Figure 219:

../_images/Untitle11d.png

Stress Damage

../_images/image9-6.jpg

Figure 220:

Stress limits \(\sigma_{t1}\) and \(\sigma_{t2}\) and \(\sigma_{t3}\) (in tensile/compression) are requested for damage. These stress limits could be observed from a tensile test in 3 related directions.

../_images/image9-7.jpg

Figure 221:

Once stress limit is reached, then damage to material begins (stress reduced with damage parameter \(\sigma\)). If Damage (\(D_{i}=D_{i}+\sigma)\) reaches D=1, then stress is reduced to 0.

../_images/image9-8.jpg

Figure 222:

Tsai-Wu Yield Criteria

In LAW12 (3D_COMP), the Tsai-Wu yield criteria is:

\[F(\sigma)=F_{1}\sigma_{1}+F_{2}\sigma_{2}+F_{3}\sigma_{3}+F_{1}\sigma_{1}^{2}+F_{2}\sigma_{2}^{2}+F_{3}\sigma_{3}^{2}+F_{4}\sigma_{12}^{2}+F_{5}\sigma_{23}^{2}+F_{6}\sigma_{3}^{2}+2F_{1}\sigma_{1}\sigma_{2}+2F_{2}\sigma_{2}\sigma_{3}+2F_{1}\sigma_{1}\sigma_{3} \tag{169}\]

The 12 coefficients of the Tsai-Wu criterion could be determined using the yield stress from the following tests:

Tensile/Compression Tests

  • Longitude tensile/compression (in direction 1):

../_images/image9-9.jpg

Figure 223:

\[F_{1}=-\frac{1}{\sigma_{\rm ly}^{c}}+\frac{1}{\sigma_{\rm ly}^{t}} \tag{170}\]
\[F_{1}=\frac{1}{\sigma_{\rm ly}^{c}\sigma_{\rm ly}^{t}}\tag{171}\]
  • Transverse tensile/compression (in direction 2):

../_images/image10-0.jpg

Figure 224:

\[F_{2}=-,\frac{1}{\sigma_{2y}^{c}}+\frac{1}{\sigma_{2y}^{t}}\tag{172}\]
\[F_{22}=\frac{1}{\sigma_{2y}^{c}\sigma_{2y}^{t}}\tag{173}\]
  • Transverse tensile/compression (in direction 3):

../_images/image10-1.jpg

Figure 225:

\[F_{3}=-,\frac{1}{\sigma_{3y}^{c}}+\frac{1}{\sigma_{3y}^{t}}\tag{174}\]
\[F_{33}=\frac{1}{\sigma_{3y}^{c}\sigma_{3y}^{t}}\tag{175}\]

Then the interaction coefficients can be calculated as:

\[F_{12}=-,\frac{1}{2}\sqrt{\left(F_{11}F_{22}\right)}\tag{176}\]
\[F_{23}=-,\frac{1}{2}\sqrt{\left(F_{22}F_{33}\right)}\tag{177}\]
\[F_{13}=-,\frac{1}{2}\sqrt{\left(F_{11}F_{33}\right)}\tag{178}\]

Shear Tests

  • Shear in plane 1-2 test:

../_images/image1-02.jpg

Figure 226:

\(\sigma^{t}_{12y}\), and \(\sigma^{c}_{12y}\) can result from the sample tests below:

../_images/image1031.jpg

Figure 227:

\[F_{44}=\frac{1}{\sigma^{c}_{12y}\sigma^{t}_{12y}}\tag{179}\]
  • Shear in plane 1-3

../_images/image1-04.jpg

Figure 228:

\(\sigma^{t}_{3,1y}\) and \(\sigma^{c}_{3,1y}\), can result from the sample tests below:

../_images/image10-5.jpg

Figure 229:

\[F_{66}=\frac{1}{\sigma_{31y}^{c}\sigma_{31y}^{t}}\tag{180}\]
  • Shear in plane 2-3:

../_images/image10-6.jpg

Figure 230:

\[F_{55}=\frac{1}{\sigma_{23y}^{c}\sigma_{23y}^{t}}\tag{181}\]

The parameters shown below in LAM12 and LAM14 are requested to calculate the Tsai-Wu criteria:

image16

Figure 231:

The yield surface for Tsai-Wu is \(F(\sigma)=1\). As long as \(\left(F(\sigma)\leq 1\right)\), the material is in the elastic phase. Once \(\left(F(\sigma)>1\right)\), the yield surface is exceeded and the material is in nonlinear phase.

In these two material laws, the following factors could also be considered for the yield surface.

  • Plastic work \(W_{p}\) with parameter B and n

  • Strain rate \(\dot{\varepsilon}\) with parameter \(\dot{\varepsilon}_{0}\) and c.

\[F\big{(}W_{p},\dot{\varepsilon}\big{)}=\big{(}1+BW_{p}^{n}\big{)}\Big{(}1+c\mathrm{ln}\frac{\dot{\varepsilon}}{\dot{\varepsilon}_{0}}\Big{)}\tag{182}\]

Then the yield surface will be \(F(\sigma)=F\big{(}W_{p},\dot{\varepsilon}\big{)}\).

  • Material will be in elastic phase, if \(F(\sigma)\leq F\big{(}W_{p},\dot{\varepsilon}\big{)}\)

  • Material will be in nonlinear phase, if \(F(\sigma)>F\big{(}W_{p},\dot{\varepsilon}\big{)}\)

This yield surface \(F\big{(}W_{p},\dot{\varepsilon}\big{)}\) will be limited with \(f_{\mathrm{max}}(F\big{(}W_{p},\dot{\varepsilon}\big{)}\leq f_{\mathrm{max}})\), where \(f_{\mathrm{max}}\) is the maximum value of the Tsai-Wu criterion limit.

\(f_{\mathrm{max}}=\Big{(}\frac{\sigma_{\mathrm{max}}}{\sigma_{y}}\Big{)}^{2}\)

Depending on parameter B, n, c and \(\dot{\varepsilon}_{0},\) the yield surface is between 1 and \(f_{\mathrm{max}}\).

../_images/image1-08.jpg

Figure 232: Tsai-WU yield criteria in 1-2 plane

LAW25 (Tsai-WU and CRASURV)

LAW25 is the most commonly used composite material in PRADIOS. It can be used with shell and solid elements. The two formulations available in LAW25 are the Tsai-Wu and CRASURV formulations.

Elastic Phase

In the elastic phase, Young’s modulus (3 parameters), shear modulus (3 parameters) and one parameter for Poisson ratio are required to describe the orthotropic material.

../_images/Untit2352led.png

Tsai-Wu Yield Criteria for \(\mathbf{I_{form}}\) =0 and =1

The Tsai-Wu yield surface in LAW25 is defined with 6 coefficient:

\(\mathbf{I_{form}}\) : Tsai-Wu \(\mathbf{I_{form}}\) : CRASURV \(\left(F(W^{*}_{\mathbf{p}},\varepsilon,\sigma)\leq 1\right)\)

\(\left(F(\sigma)\leq F(W^{*}_{\mathbf{p}},\varepsilon)\right)\)

To check if material in yield, in Tsai-Wu (\(I_{form}\) will be compared with \(F(W^{*}_{\mathbf{p}},\varepsilon)\) at each stress state and in CRASURV (\(I_{form}=1)\) \(F(W^{*}_{\mathbf{p}},\varepsilon)\) will be simply compared with 1 at each stress state. These 6 coefficients could be determined with yield stress from these tests:

  • Tensile/compression tests

Longitudinal tensile/compression tests (in direction 1 which is fiber direction):

../_images/image110.png

Here, \((W^{*}_{p}=\frac{W_{p}}{W^{ref}_{p}})\) in compression:

\(\sigma^{c}_{1y}\big{(}W^{*}_{p}\cdot\varepsilon\big{)}=\sigma^{c}_{1y}\big{(}1+b^{c}_{1}\big{(}W^{*}_{p}\big{)}^{n^{c}_{1}}\big{(}1+c^{c}_{1}\ln\Big{(}\frac{\varepsilon}{\varepsilon_{0}}\Big{)}\big{)})\)

Here \((W^{*}_{p}=\frac{W_{p}}{W^{ref}_{p}})\)

Transverse tensile/compression tests (in direction 2)

../_images/image11-1.png
  • Shear tests

Shear in plane 1-2

../_images/image1-1-2.jpg

\(I_{form}\) = 0: Tsai-WU \(\left(F(\sigma)\leq F(W^{*}_{\mathbf{p}},\varepsilon)\right)\)

\(I_{form}\) = 1: CRASURV \(\left(F(W^{*}_{\mathbf{p}},\varepsilon,\sigma)\leq 1\right)\)

\(F_{12}=-\frac{a}{2}\sqrt{F_{11}F_{22}}\)

The default reduction factor, a-1, is typically used.

\(F\big{(}W_{p}^{*},\dot{\varepsilon}\big{)}=\big{(}1+b \big{(}W_{p}^{*}\big{)}^{n}\big{)}\cdot\Big{(}1+\mathrm{cln}\big{(}\frac{\dot{\varepsilon}}{\dot{\varepsilon}_{0}}\big{)}\Big{)}\)

Here, \((W^{*}_{p}=\frac{W_{p}}{W^{ref}_{p}})\)

\(F_{12}\big{(}W_{p}^{*},\dot{\varepsilon}\big{)}=-\frac{a}{2}\sqrt{F_{11}\big{(}W_{p}^{*},\dot{\varepsilon}\big{)}F_{22}\big{(}W_{p}^{*},\dot{\varepsilon}\big{)}}\)

The default reduction factor, a-1, is typically used.

Note that the relative plastic work \(W_{p}^{*}\) is used in Tsai-WU to calculate the yield surface; whereas in CRASURV, the relative plastic work is used to calculate the yield stress.

../_images/Untitl3422ed.png

The yield stress limit \(F(W^{*}_{\mathbf{p}},\varepsilon)\) is range of 1 and \(f_{max}\)

In LAW25 (Tsai-Wu and CRASURV) damage is a function of the total strain and the maximum damage factor.

If the total strain \(\varepsilon>\varepsilon_{t}\) or out of plane strain \(\gamma_{ini}<\gamma<\gamma_{\max}\), then the material is softened using the following method:

\[\sigma^{reduce}=\sigma\cdot(1-d_{i})\tag{184}\]

with i =1,2,3

Where, \(d_{i}\) is the damage factor and is defined as:

\[d_{i}=\min\left(\frac{\varepsilon_{i}-\varepsilon_{ti}}{\varepsilon_{i}}\cdot\frac{\varepsilon_{ini}}{\varepsilon_{ini}}, d_{\max}\right)\tag{185}\]

with i =1,2

\[d_{3}=\min\left(\frac{\gamma-\gamma_{ini}}{\gamma_{\max}-\gamma_{ini}}\cdot\frac{\gamma_{\max}}{\gamma},d_{\max}\right)\tag{186}\]

in direction 3 (delamination)

  • If the total strain is between \(\varepsilon_{t}<\varepsilon_{t}<\varepsilon_{f}\), the material begins to soften, but this damage is reversible. Once \(\varepsilon>\varepsilon_{f}\), then the damage is irreversible and if \(\varepsilon\geq\varepsilon_{m})\), then stress in material is reduced to 0.

  • Damage could be in elastic phase or in plastic phase. It depends on which phase \(\varepsilon_{t}\) and \(\varepsilon_{f}\) are defined in.

  • Element deletion is controlled by \(I_{\mathit{off}}\). Select a different \(I_{\mathit{off}}\) option to control the criteria of element deletion. For additional information, refer to \(I_{\mathit{off}}\) in LAW25 in the Reference Guide.

LAW19 and LAW58 for Fabric

PRADIOS has two material laws for modeling fabrics LAW19 and LAW58. LAW19 is an elastic orthotropic material and must be used with /PROP/TYPE9. LAW58 is hyperelastic anisotropic fabric material and must be used with /PROP/TYPE16.

Coupling between warp and weft directions could be defined in this material law to reproduce physical interaction between fibers. Both material laws are often used for airbag modeling.

In LAW58, two methods are provided to define the stress-strain behavior.

  • Nonlinear function (\(fct\_ID_{i}\)) curve to define the warp, weft and shear behavior

  • Young’s modulus, soften coefficient (B), straightening strain \(S_{i}\) and fiber bending modulus reduction factor \(\mathit{Flex}_{i}\)

    In warp and weft direction:

    \(\sigma_{ii}=E_{i}\varepsilon_{ii}-\frac{(B_{i}\varepsilon_{ii})}{2}\quad(i=1,2)\text{when}\frac{d\sigma}{d\varepsilon}>0\)

    \(\sigma_{ii}=\max_{\varepsilon_{ii}}\left(E_{i}\varepsilon_{ii}-\frac{(B_{i}\varepsilon_{ii})}{2}\right)\quad(i=1,2)\text{when}\frac{d\sigma}{d\varepsilon}\leq 0\)

    For in-plane shear in initial state, use \(G_{0}\). Once \(\alpha\) (angle between wrap and weft) reaches \(\alpha_{T}\) (shear lock angle), then use \(G_{T}\) to describe the strengthening.

\(\tau=G_{0}\text{tan}(\alpha)-\tau_{0}\text{if}\alpha\leq\alpha_{T}\)

\(\tau =\frac{G_{T}}{1+\tan^{2}(\alpha_{T})}\text{tan}(\alpha)+\left(G_{0}-\frac{G_{T}}{1+\tan^{2}(\alpha_{T})}\right)\text{tan}(\alpha_{T})-\tau_{0}\text{if}\alpha>\alpha_{T}\)

../_images/image1-16.jpg

Figure 233:

For out-of-plane shear stress-stain is described with \(G_{sh.}\)

../_images/image1-17.jpg

Figure 234:

For fabric, there is straight process at the beginning of tension. At this phase, fiber shows very soft, due to not being tightened yet.

../_images/image1-18.jpg

Figure 235:

In LAN58, use \(Flex_{i}\) to describe this behavior:

\[E_{fi}=Flex_{i}*E_{i} \tag{187}\]

After the fabric is tight (strengthening strain \(S_{i}\) is reached), then normal fiber elasticity \(E_{i}\) could be sed.

../_images/image1191.jpg

Figure 236:

Composite Modeling

Composite modeling techniques used in PRADIOS.

Each Layer with at Least One Solid

The model is large which leads to high accuracy results but need more computation resource.

../_images/image12-0.jpg

Figure 237:

Mixed Approach (Middle Layer Thick)

Defines shells for top and bottom layer and solid for middle layer.

../_images/image1-21.jpg

Figure 238:

Sandwich Shell Approach

Use only one shell element through the thickness. Use different materials to define multiple layers.

../_images/image12-2.jpg

Figure 239:

In composite shell modeling there are two types of problems.

  • The first is composite shell with isotropic layers (orthotropic in z direction). For example, by Windshield modeling. Where, material LAW27(PLAS_BRIT) or LAW36 (PLAS_TAB) can be used.

  • The second is composite shell with orthotropic layers (orthotropic in x, y and z). For example, Sandwich shell with external fiber glass layers and internal foam layer. Where, material LAW25 (COMPSH) and all shell formulation can be used, except for QEPH.

Composite Properties

Composite could be modeled with solid or shell element. Depending on the element type, the following properties can be used in PRADIOS to model a composite.

Shell Elements

Modeling composite in PRADIOS could be defined by Layer based or Ply based with different property.

  • Layer based modeling with /PROP/TYPE10 (SH_COMP), /PROP/TYPE11 (SH_SANDW)

  • Ply based modeling with /PROP/TYPE17 (STACK), /PROP/TYPE51, /PROP/PCOMPP+/STACK, /PROP/ TYPE19 (PLY) and /PLY

../_images/image12-3.jpg

Figure 240:

For ply-based modeling, info (like material, thickness, anisotropic angle, anisotropic axis angle and number of integration points) for each ply defined in /PROP/TYPE19 (or /PLY) and assembled in /PROP/ TYPE17 or /PROP/TYPE51 (or /STACK) with option Pply_\(D_{i}\).

../_images/image1241.jpg

Figure 241:

For ply-based modeling, the way of assemble plies could be “by ply” or “by substack”. “By ply” is simply pile up plies one by one from bottom to top. The way “by substack” need first pile up plies one by one to create each substack, and then each substack could either be combined or could also be stacked with substack connection “INT”.

../_images/image1251.jpg

Figure 242:

../_images/image12-6.jpg

Figure 243: Stack in /PROP/TYPE51

For composite property following topics are:

  • Layer (ply) number and integration points each layer (ply)

  • Anisotropy in layer (ply)

  • Layer (ply) thickness and position

  • Composite material used for Layer (ply)

Layer-based Properties

/PROP/TYPE10 (SH_COMP)

/PROP/TYPE11

(SH_SANDW)

N

Layer Numbers

Or Pply_IDi

Ply Numbers

*N*=0~100

*N*=1~100

IP of each layer/ply

1 per Layer

1 per Layer

Iint

Integration formulation

\(\phi_{i}\) + V Anisotropic direction

\(\phi_{i}\) + skew

Anisotropic direction

\(\theta_{drape}\) Ply orientation change

\(a_{i}\) Angle between anisotropic axis

\(t_{i}\) Layer/Ply thickness

\(I_{pos} + Z_{i}\)

Layer/Ply position

\(I_{pos}\) =2,3,4

Layer/Ply offset

\(mat\_ID_{i}\)

Material for each layer/ply

Use material defined in /PART

Must use same material type for all layers.

Commonly used Composite

Material Law

15, 25 and user material

15, 25 and user material

XFEM compatibility (crack propagation)

With /FAIL/JOHNSON,

/FAIL/ TAB1 and /FAIL/TBUTCHER

Plyxfem

Delamination between layer/ply

Minterply

Material between layer/ply

Ply-based Properties

/PROP/TYPE17 + / PROP/TYPE19

/PROP/TYPE51 + / PROP/TYPE19

PROP/PCOMPP+/ STACK+/PLY

N

Layer Numbers

Or \(Pply\_ ID_{i}\)

Ply Numbers

\(Pply\_ ID_{i}\) =1~200

\(Pply\_ ID_{i}\) =1~200

\(Pply\_ ID_{i}\) 1~n

IP of each layer/ply

1 per Ply

Npt_ply=1 in /PROP/

TYPE19`

1~9 per Ply

Npt_ply=1~9 in /

PROP/TYPE19

1~9 per Ply

Npt_ply=1~9 in /PLY

Iint

Integration formulation

Uniformed or Gauss

Uniformed or Gauss

\(\phi_{i}\) + V

Anisotropic direction

\(\phi_{i}\) + skew

Anisotropic direction

\(\theta_ {drape}\)

Ply orientation change

Defined in /PROP/

TYPE19

Defined in /PROP/

TYPE19

Defined in /PLY

\(a_{i}\)

Angle between anisotropic axis

in /PROP/TYPE19

in /PROP/TYPE19

in /PLY

\(t_{i}\)

Layer/Ply thickness

Different in /PROP/

TYPE19

Different in /PROP/

TYPE19

Different in /PLY

\(I_{pos}\) + \(Z_{i}\)

Layer/Ply position

\(I_{pos}\) =2,3,4

Layer/Ply offset

\(mat\_ ID_{i}\)

Material for each layer/

ply

Must use same

material type for all plies.

Different material

type allows for each ply.

Different material

type allows for each ply.

Commonly used

Composite Material

Law

25, 27, 36, 60, 72, 93 and user material

25, ≥28 and user material

25 and user material

XFEM compatibility

(crack propagation)

With /FAIL/JOHNSON

and /FAIL /TBUTCHER

Plyxfem

Delamination between layer/ply

Minterply

Material between layer/ ply

Only with LAW1+/ FAIL/LAD_DAM

Layer (Ply) Number N (\(Nply\_ID_{i}\)) and Integration Points each Layer (Ply)

For layer-based modeling which use /PROP/TYPE10, /TYPE11. N is the number of layers through the shell thickness. For these properties, there is one integration point (IP) each layer.

For ply-based modeling, which use /PROP/TYPE17, /TYPE51 and /PCOMPP. Pply_ID is the number ofplies through the shell thickness. Plies could be combined until n plies for these properties.

For TYPE17 only one integration point is allowed while for TYPE51 and /STACKF until 9 integration points are allowed. Number of integration point defined with option “Npt_ply” in property TYPE19 or /PLY.

Example (Ply) (/PROP/TYPE51)

In this example, Npi_ply=3 defined in /PROP/TYPE19, means 3 integration points defined per ply and with option \(I_{int}\) =1 defined in /PROP/TYPE51, then these 3 integration points are uniformly distributed through each ply thickness. each ply thickness.

../_images/imag-ae127.jpg

Figure 244:

If \(I_{int}\) =1 in /PROP/TYPE51, then integration points are distributed follow Forces and Moments Calculation Gauss Integration Scheme through each ply thickness.

../_images/ima-ge128.jpg

Figure 245:

It is possible to print animation results (plastic strain, damage, stress and strain tensor) in each specific integration point with /ANIM/SHELL/IDPLY/Keyword4/I/J (or /ANIM/SHELL/Keyword3/N/NIP).

For instance, use /ANIM/SHELL/IDPLY/EPSP/2/3 (or /ANIM/SHELL/EPSP/2/3`) to print plastic strain in third integration point (red highlighted integration point in Figure 245) of second ply (ply name Ply12). For additional print info about Integration points through shell thickness for composite properties, refer to Shell stress tensor output in animation. in the FAQs.

Anisotrophy in Layer (Ply)

  • The first anisotropic direction of material could be defined with angle \(\phi\) and gloval vector \(\mathbf{V}(V_{x}),(V_{Y}),(V_{Z})\). It is also possible to use angle \(\phi\) and skew. In this case, x-axis of skew replaced the global vector \(\mathbf{V}\).

    Project the global vector \(\mathbf{V}\) in shell element and then rotate \(\phi\) degree is the first anisotropic direction (also called Material direction 1). The positive direction of \(\phi\) is coding to shell normal \(\mathbf{n}\). For example, in Figure 246 rotate counterclockwise \(\phi\) degree is material direction 1.

The material direction 1 of the local element reference is normally the fiber direction. Then the material character (E-Module, yield stress, and so on) of direction 1 which defined in material law could be then applied in the correct direction on local element reference.

../_images/ima-e129.jpg

Figure 246:

  • Composite material could be orthotropic or anisotropic. In Radios it is possible to describe this character with anisotropic axis angle \(\alpha_{i}\) in ply-based properties. In case of \(\alpha_{i}=90^{\circ}\), then it describes orthotropic material. For layer based properties (TYPE10 and TYPE11) which without this option \(\alpha_{i}\), so that only orthotropic material could be defined.

../_images/image130-.jpg

Figure 247:

Figure 248 shows an Example in /PROP/TYPE11 which use skew to define global vector V.

../_images/im-age131.jpg

Figure 248:

For property TYPE11 and TYPE51, the anisotropic direction for some shell could also be initialized with \(\phi\) in keyword /INISHE/ORTH_LOC or /INISH3/ORTH_LOC. For property TYPE51, anisotropic axis angle could also be initialized with these keywords.

  • The orientation of anisotropy for specific shell elements or shell element groups could be change again with option drape_ID and def_orth in

/PROP/TYPE19 and /PLY. drape_ID defined in /DRAPE. With this feature, angle of anisotropic direction could be changed with \(\theta_{drape}\).

  • If use def_orth=1: Angle:

\(\phi_{r}\) skew or global vector \(\mathbf{V}\) will be ignored. Take shell local x-axis as vector \(\mathbf{V}_{*}\) and then rotate \(\phi\) degree is the first anisotropic direction. \(\phi_{s}\) (defined in /SHELL or /SH3N) is taken into account by compute angle \(\phi\).

\[\phi=\phi_{s}+\Delta\phi+\theta_{drape}\tag{188}\]
  • If use def_orth=2 (Default)

Project the global vector \(\mathbf{V}\) in shell element to vector \(\mathbf{V}^{*}\) and then rotate \(\phi\) degree is the first anisotropic direction. The angle \(\phi\) is computed as:

\[\phi=\phi_{i}+\Delta\phi+\theta_{drape} \tag{189}\]
../_images/ima-ge132.jpg

Figure 249:

Layer (Ply) Thickness and Position

  • For /PROP/TYPE10, layer thickness is simply averaged by layer number

\[t_{i}=Thick/N \tag{190}\]

and layers are automatically overlying one by one from bottom to top.

../_images/im-age133.jpg

Figure 250:

  • For property TYPE11, TYPE17, TYPE51 and /STACK, layer (ply) position and thickness depend on option \(I_{pos}\)

    ◦ If \(I_{pos}\) =0

User layer (ply) thickness input will be taken, and layer (ply) position will be automatically overlying one by one from bottom till top; but if,

\[\sum_{i}t_{i}\neq Thick \tag{191}\]

Then, layer (ply) thickness will be adjusted to \(t_{i}^{new}\), so that

\[\sum_{i}t_{i}^{new}= Thick \tag{192}\]

Layer (ply) position will be then adjusted, as well.

◦ If :math:`I_{pos}`=1

User layer(ply) input of thickness \(t_{i}\) and position \(Z_{i}\) will be taken. Sum of layer thickness will not be checked with (Thick).

For additional information, refer to Layer thickness and position calculation. in the FAQs.

  • For property TYPE17, TYPE51 and /STACK, it is also possible to offset the plies with :math:`I_{pos}`=2, 3, 4

  • \(I_{pos}\) =2: the shell element mid-surface is at \(Z_{0}\) from the bottom of the ply layout

../_images/image1341.jpg

Figure 251:

  • \(I_{pos}\) =3: the top of the ply layout is coincident with the element mid-surface

../_images/image1351.jpg

Figure 252:

  • \(I_{pos}\) =4: the bottom of the ply layout is coincident with the element mid-surface

../_images/im-age136.jpg

Figure 253:

  • For /PROP/TYPE17, /PROP/TYPE51 and /STACK, ply thickness could be changed by option in /DRAPE (which is used in /PROP/TYPE19 or /PLY). Then updated ply thickness is:

\[t_{i}^{new}=t_{i}\cdot Thinning \tag{193}\]

Composite Material Used for Layer (Ply)

Composite material LAW15 and LAW25 could be used for shell element. Failure model /FAIL/HASHIN, /FAIL/PUCK and /FAIL/LAD_DAMA with LAW25 and /FAIL/CHANG with LAW15 could be used to describe composite behavior for shell element. For additional information, refer to Composite Material.

  • Material for layer (ply)

    ◦ For property TYPE10, composite used material defined in /PART

    ◦ For property TYPE11 and TYPE17, composite used material defined in option mat_ \(ID_{i}\). Different material ID could be defined for each layer (ply). But they must use same material type. If using LAW25, then several different LAW25 cards can be used for different layer (ply).

    ◦ For property TYPE51 and /STACK, composite also used material defined in option mat_ \(ID_{i}\) and different material type or ID could be used for each ply.

  • Material between ply

    For property TYPE17, it is possible to define delamination between plies or stacks (with Plyxfem=2). This is very useful for delamination is the main driven of composite failure. The material between plies defined with Minterply. For the moment, LAW1+ /FAIL/LAD_DAM could be used to describe three different type of ply delamination.

../_images/image1371.jpg

Figure 254:

And then in this case additional variable :math:`_{1},_{2},_{3} are added on each node of ply by computation to ply by computation to simulate the delamination failure between plies.

../_images/im-age138.jpg

Figure 255:

Solid Element

The new composite technology allows you to make parts production more and thicker, as modeling those parts with shell elements is not enough. Thick shell can solve this problem. Compare with shell element, thick shell could direct connect with other solid part.

For Solid element, for the moment only layer-based modeling with property /PROP/TYPE22 (TSH_COMP) is available. This solid property is similar to shell property /PROP/TYPE11 by define composite.

Layer-based Properties

/PROP/TYPE22 (TSH_COMP)

Layer Numbers

\(I_{solid}\) =14:

\(I_{int}\) =9~200

\(I_{solid}\) =15:

IP of each layer/ply

\(I_{npts}\) =ijk=2~9

\(I_{npts}\) =j=1~200

Integration formulation

\(\phi_{i}\) + V, Anisotropic direction

\(\phi_{i}\) + skew, Anisotropic direction

\(\theta_ {draper}\) , Ply orientation change

\(a_{i}\) , Angle between anisotropic axis

\(t_{i}\) , Layer/Ply thickness

defined with factor \(t_{i}/t\)

\(I_{pos}\) + \(Z_{i}\) Layer/Ply position

\(I_{pos}\) =2,3,4, Layer/Ply offset

\(mat\_ID_{i}\) , Material for each layer/ ply

Different material type allows for each layer.

Commonly used Composite

Material LAW

LAW12, LAW14, LAW25 and user material

Plyxfem, Delamination between layer/ply

Minterply, Material between layer/ply

  • Layer Number and Integration Points Each Layer

    The layer number defined using the option \(I_{int}\). \(I_{int}\) is only used for \(I_{solid}\) =14 when the number of layers > 9.

    In this case, the thickness direction integration point defined by \(I_{npts}\) should be zero.

    Example, \(I_{cstr}\) = 010; \(I_{npts}\) = 202; \(I_{int}\) = 100 for a number of 100 layers in “s” direction

  • Anisotrophy in Layer (Ply)

Similar to shell property /PROP/TYPE11, reference vector V and angle \(\phi\) are used to define the material direction 1. The reference vector V project to the middle surface of solid element and turn \(\phi\) degree is the material direction 1.

../_images/im0age139.jpg

Figure 256:

  • Layer Thickness and Position

For solid element thickness and position defined by element mesh.

  • Composite Material Used for Layer

    ◦ With option mat_IDi, it is possible to use different material type for each layer

    ◦ Composite material LAW12, LAW14 and LAW25 could be used with this property

    ◦ Failure model /FAIL/HASHIN, /FAIL/PUCK and /FAIL/LAD_DAMA with these composite material laws are also accounted for

    ◦ Material referred to in the corresponding /PART card is only used for time step and interface stiffness calculation

    ◦ For LAW25, it is assumed that (for solids and thick shells) the material is elastic in transverse direction (material direction 2 andv 1) and the E33 value must be specified in such cases

For additional information, refer to “Composite material and Composite failure”.

See Also
/DRAPE (Starter)
Property and Elements (FAQ)

Composite Failure Model

In PRADIOS the following composite failure models may be used to describe composite material failure.

  • /FAIL/HASHIN

  • /FAIL/PUCK

  • /FAIL/LAD_DAMA

  • /FAIL/CHANG

    1. Gornet, “Finite Element Damage Prediction of Composite Structures”

A composite material consists of two different materials (matrix and reinforcement fiber). Each material has a different failure behavior. In PRADIOS it is possible to use different failure models for matrix and fiber in one composite element (for elements with property TYPE11, TYPE16, TYPE17, TYPE51, PCOMPP or TYPE22). For example, you may use /FAIL/HASHIN for fiber failure, /FAIL/PUCK for matrix failure and /FAIL/LAD_DAMA for delamination between layers or plies (if there is more than one layer or plies defined for the composite).

Besides the above typical composite failure models, /FAIL/FLD (used for isotropic brittle composite materials in layers(plies) as in, glass), /FAIL/ENERGY, /FAIL/TBUTCHER and /FAIL/TENSSTRAIN may also be used to describe failure for composite layers(plies).

/FAIL/HASHIN

In HASHIN failure, two primary failure modes are considered.

  • Fiber mode: composite fails, due to fiber rupture in tension or fiber buckling in compression. So, in /FAIL/HASHIN, tensile/shear fiber mode, compression fiber mode and crush mode are the fiber modes. If direction 1 is the fiber direction, then plane 23 is the predominant failure plane for fiber mode.

  • Matrix mode: composite fails, due to matrix cracking from the fiber. Failure matrix mode (or shear failure matrix mode) and delamination mode are both matrix modes. The failure plane for matrix mode is parallel to the fiber, and stress \(\sigma\) will not be considered in this mode.

../_images/im-age140.jpg

Figure 257: Fiber modes and matrix modes for uni-directional lamina model

Fibers in uni-directional lamina model 17 are only in one direction and in fabric lamina model are in two directions.

Uni-directional Lamina Model

Fabric Lamina Model

Damage

criteria

If*D*=1, then failure.

If 0 \(\leq D<1\), D then no failure.

With \(D=Max(F_{1}, F_{2},F_{3},F_{4})\)

If*D*=1, then failure.

If 0 \(\leq D<1\), D then no failure.

With \(D=Max(F_{1}, F_{2},F_{3},F_{4})\)

Tensile/ shear fiber

mode

image24

image25

Compression

fiber mode

image26

image27

Crush mode

image28

image29

Shear failure

matrix mode

image30

Failure matrix

mode

image33

Delamination

mode

image34

image35

image36

image37

image1 Note:\(\left\langle a\right\rangle=\begin{cases}a\ if\ a>0\\ 0\ if\ a<0\end{cases}\)

In /FAIL/HASHIN, material strength \(\sigma_{1}^{t},\sigma_{2}^{t},\sigma_{3}^{t},\sigma_{1}^{c},\sigma_{2}^{c}\) are derived from tension/compression test for composite.

Crush strength \(\sigma_{c}\) and fiber shear strength \(\sigma_{12}^{f}\) may be obtained from a quasi-static punch shear test (QS-PST). [22] Crush strength \(\sigma_{c}\) from the span to punch ratio (SPR) =0 and fiber shear strength \(\sigma_{12}^{f}\) from SPR=1.1.

\(\phi\) is the Coulomb friction angle. It is observed that composite shear strength is higher if the composite is also under compression (rather than under tension). This is due to the friction between matrix and fiber.

The shear strength is assumed to be proportional to compression stress and is computed as:

\[S_{12}=\sigma_{12}^{m}+tan \phi\left\langle-\sigma_{22}\right\rangle \tag{194}\]

image22

Figure 258:

Function angle \(\phi\) may be fitted with an Off-Axis compression test with different angles \(\sigma\) (for example, 30 , 45 , 60 , …).

../_images/ima--ge142.jpg

Figure 259:

\(\sigma_{12}^{m},\sigma_{13}^{m},\sigma_{23}^{m}\) may be derived from a matrix shear test in three directions.

\(S_{del}\) is the scale factor for delamination criteria. It may be fitted with composite delamination experimental datain order for delamination failure to correlate with the damage area in experiment.

/FAIL/PUCK

In Puck failure, two types of failure are considered.

  • Fiber fracture: composite fails, due to the fiber reaching the tensile or compression strength limit.

  • Inter fiber failure (IFF): composite fails, due to the fiber matrix cracking.

Damage criteria

If D =1, then failure. If \(0\leq D<1 D\), then failure.

With \(D=Max\big{(}e_{f}(tensile),e_{f}(compression), e_{f}(ModeA),e_{f}(modeB),e_{f}(ModeC)\big{)}\)

Fiber fraction

failure

Tensile fiber failure mode: \(\sigma_{11}>0\)

\(e_{f}(tensile)=\frac{\sigma_{11}}{\sigma}_ {1}^{t}\)

Compressive fiber failure mode: \(\sigma_{11}>0\)

\(e_{f}(compression)=\frac{|\sigma_{11}|} {\sigma_{1}^{c}}\)

Inter fiber failure (IFF)

18

Mode A, if \(\sigma_{22}>0\)

image38

Figure 260:

\(e_{f}(ModeA)=\frac{1}{\sigma_{12}}\big{[} \sqrt{\big{(}\frac{\sigma_{12}}{\sigma_{2}^{t}} \big{)}^{2}\sigma_{22}2+\sigma_{12}2}+p_{12}^{+} \sigma_{22}\big{]}\)

Mode C, if \(\sigma_{22}<0\)

image23

Figure 261:

\(e_{f}(ModeС)=\big{[}\big{(}\frac{\sigma_{12}} {2(1+p_{22})\sigma_{12}}\big{)}^{2}+\big{(}\frac {\sigma_{22}}{\sigma_{2}^{c}}\big{)}^{2}\big{]} \big{(}\frac{\sigma_{2}^{c}}{-\sigma_{22}}\big{)}\)

Mode B:

image39

Figure 262:

\(e_{f}(ModeB)=\frac{1}{\sigma_{12}}\big{(} \sqrt{\sigma_{12}^{2}+\big{(}p_{12}\sigma_{22} \big{)}^{2}}+p_{12}\sigma_{22}\big{)}\)

In inter fiber failure, Mode A shows failure under tension in transverse fiber direction (90 degrees to fiber direction), and in this case, shear loading could reduce the failure limit.

If under compression in transverse fiber direction, at first increasing compression will increase composite shear loading (Mode B). If compression continues to increase, then shear loading will decrease (Mode C).

Input Parameters

For fiber fracture failure, you could obtain fiber strengths \(\sigma_{1}^{t},\sigma_{1}^{c}\) from tension and compression composite tests in the fiber direction.

For inter fiber failure, you could obtain strengths \(\sigma_{2}^{t},\sigma_{2}^{c}\) from tension and compression composite tests in the transverse fiber direction.

You could obtain shear strength \(\sigma_{12}\) with a pure shear test (\(\sigma_{2}=\sigma_{1}=0\)).

With \(\sigma_{2}^{t},\sigma_{2}^{c}\sigma_{12}\), then \(p_{22}\) and \(p_{12}\) for Mode C and Mode B may be determined.

With \(\sigma_{2}^{t},\sigma_{12}\) and additional tension-shear tests in the transverse fiber direction, \(p_{12}^{+}\) may be determined. The additional tension-shear test in transverse fiber direction could take equal tension-shear (by \(\sigma_{22}=\sigma_{12}\)) loading.

Then you may derive the fracture curve in \(\sigma_{22}-\sigma_{12}\) plane, as shown below.

../_images/image1461.jpg

Figure 263: IFF fracture curve in \(\sigma_{22}-\sigma_{12}\) plane

For \(p_{12}^{+},p_{12}^{-},p_{22}^{-}\) parameters 19. For carbon fiber composite, use \(p_{12}^{+}=0.35, p_{12}^{-}=0.3, p_{22}^{-}=0.2\) and for glass fiber composite, use \(p_{12}^{+}=0.3, p_{12}^{-}=0.25, p_{22}^{-}=0.2\).

/FAIL/LAD_DAMA

/FAIL/LAD_DAMA is used to describe delamination between composite layers (damage propagation in the matrix). Assume that layers are connected through a virtual interface (contact).

../_images/i-mage147.jpg

Figure 264:

For example, if composite is under loading, shown below, the traction \(\sigma\) and displacement in direction 3

../_images/im8age148.jpg

Figure 265:

The area under the traction versus displacement curve is the absorbed energy by delamination, which is also termed the strain energy of the damage interface. The failure is governed by this strain energy is described below, considering 3 Modes of delamination:

\[E_{D}=\frac{1}{2}\Bigg{[}\frac{\left<\sigma_{3}\right>^{2}}{K_{3}(1-d_{3})}+\frac{\left<-\,\sigma_{3}\right>^{2}}{K_{3}}+\frac{\sigma_{3}z^{2}}{K_{2}(1-d_{2})}+\frac{\sigma_{3}z^{2}}{K_{3}(1-d_{3})}\Bigg{]}\tag{195}\]

Where, \(\sigma_{3},\sigma_{32},\sigma_{31}\) are the stresses shown below in three modes of delamination behavior.

../_images/image1371.jpg

Figure 266:

With strain energy of delamination \(E_{D}\) , the thermodynamic force (contact force of virtual interface), also termed damage energy release rates may be calculated for these three modes:

  • Model I (DCB specimen:sup:21)

\(Y_{d_{3}}=\frac{\partial E_{D}}{\partial d_{3}}\bigg{|}_{\sigma=\text{csf}}=\frac{1}{2}\frac{\left<\sigma_{3}y\right>^{2}}{K_{\frac{3}{3}}(1-d_{3})^{2}}\)

  • Model II (ENF specimen:sup:21)

\(Y_{d_{2}}=\frac{\partial E_{D}}{\partial d_{2}}\bigg{|}_{\sigma=\text{csf}}=\frac{1}{2}\frac{\sigma_{3}z^{2}}{K_{\frac{3}{3}}(1-d_{2})^{2}}\)

  • Model III

\(Y_{d_{1}}=\frac{\partial E_{D}}{\partial d_{1}}\bigg{|}_{\sigma=\text{csf}}=\frac{1}{2}\frac{\sigma_{3}z^{2}}{K_{\frac{3}{3}}(1-d_{1})^{2}}\)