Kalpana Kalpana (Editor)

Hyperelastic material

Updated on
Edit
Like
Comment
Share on FacebookTweet on TwitterShare on LinkedInShare on Reddit
Hyperelastic material

A hyperelastic or Green elastic material is a type of constitutive model for ideally elastic material for which the stress-strain relationship derives from a strain energy density function. The hyperelastic material is a special case of a Cauchy elastic material.

Contents

For many materials, linear elastic models do not accurately describe the observed material behaviour. The most common example of this kind of material is rubber, whose stress-strain relationship can be defined as non-linearly elastic, isotropic, incompressible and generally independent of strain rate. Hyperelasticity provides a means of modeling the stress-strain behavior of such materials. The behavior of unfilled, vulcanized elastomers often conforms closely to the hyperelastic ideal. Filled elastomers and biological tissues are also often modeled via the hyperelastic idealization.

Ronald Rivlin and Melvin Mooney developed the first hyperelastic models, the Neo-Hookean and Mooney–Rivlin solids. Many other hyperelastic models have since been developed. Other widely used hyperelastic material models include the Ogden model and the Arruda–Boyce model.

Saint Venant–Kirchhoff model

The simplest hyperelastic material model is the Saint Venant–Kirchhoff model which is just an extension of the linear elastic material model to the nonlinear regime. This model has the form

S = λ   tr ( E ) 1 + 2 μ E

where S is the second Piola–Kirchhoff stress and E is the Lagrangian Green strain, λ and μ are the Lamé constants, and 1 is the second order unit tensor.

The strain-energy density function for the St. Venant–Kirchhoff model is

W ( E ) = λ 2 [ tr ( E ) ] 2 + μ tr ( E 2 )

and the second Piola–Kirchhoff stress can be derived from the relation

S = W E   .

Classification of hyperelastic material models

Hyperelastic material models can be classified as:

1) phenomenological descriptions of observed behavior

  • Fung
  • Mooney–Rivlin
  • Ogden
  • Polynomial
  • Saint Venant–Kirchhoff
  • Yeoh
  • Marlow
  • 2) mechanistic models deriving from arguments about underlying structure of the material

  • Arruda–Boyce model
  • Neo-Hookean
  • 3) hybrids of phenomenological and mechanistic models

  • Gent
  • Van der Waals
  • Generally, a hyperelastic model should satisfy the Drucker stability criterion. Some hyperelastic models satisfy the Valanis-Landel hypothesis which states that the strain energy function can be separated into the sum of separate functions of the principal stretches ( λ 1 , λ 2 , λ 3 ) :

    W = f ( λ 1 ) + f ( λ 2 ) + f ( λ 3 ) .

    First Piola–Kirchhoff stress

    If W ( F ) is the strain energy density function, the 1st Piola–Kirchhoff stress tensor can be calculated for a hyperelastic material as

    P = W F or P i K = W F i K .

    where F is the deformation gradient. In terms of the Lagrangian Green strain ( E )

    P = F W E or P i K = F i L   W E L K   .

    In terms of the right Cauchy–Green deformation tensor ( C )

    P = 2   F W C or P i K = 2   F i L   W C L K   .

    Second Piola–Kirchhoff stress

    If S is the second Piola–Kirchhoff stress tensor then

    S = F 1 W F or S I J = F I k 1 W F k J   .

    In terms of the Lagrangian Green strain

    S = W E or S I J = W E I J   .

    In terms of the right Cauchy–Green deformation tensor

    S = 2   W C or S I J = 2   W C I J   .

    The above relation is also known as the Doyle-Ericksen formula in the material configuration.

    Cauchy stress

    Similarly, the Cauchy stress is given by

    σ = 1 J   W F F T   ;     J := det F or σ i j = 1 J   W F i K   F j K   .

    In terms of the Lagrangian Green strain

    σ = 1 J   F W E F T or σ i j = 1 J   F i K   W E K L   F j L   .

    In terms of the right Cauchy–Green deformation tensor

    σ = 2 J   F W C F T or σ i j = 2 J   F i K   W C K L   F j L   .

    The above expressions are valid even for anisotropic media (in which case, the potential function is understood to depend implicitly on reference directional quantities such as initial fiber orientations). In the special case of isotropy, the Cauchy stress can be expressed in terms of the left Cauchy-Green deformation tensor as follows:

    σ = 2 J   B W B or σ i j = 2 J   B i k   W B k j   .

    Incompressible hyperelastic materials

    For an incompressible material J := det F = 1 . The incompressibility constraint is therefore J 1 = 0 . To ensure incompressibility of a hyperelastic material, the strain-energy function can be written in form:

    W = W ( F ) p   ( J 1 )

    where the hydrostatic pressure p functions as a Lagrangian multiplier to enforce the incompressibility constraint. The 1st Piola–Kirchhoff stress now becomes

    P = p   J F T + W F = p   F T + F W E = p   F T + 2   F W C   .

    This stress tensor can subsequently be converted into any of the other conventional stress tensors, such as the Cauchy Stress tensor which is given by

    σ = P F T = p   1 + W F F T = p   1 + F W E F T = p   1 + 2   F W C F T   .

    Compressible isotropic hyperelastic materials

    For isotropic hyperelastic materials, the Cauchy stress can be expressed in terms of the invariants of the left Cauchy–Green deformation tensor (or right Cauchy–Green deformation tensor). If the strain energy density function is W ( F ) = W ^ ( I 1 , I 2 , I 3 ) = W ¯ ( I ¯ 1 , I ¯ 2 , J ) = W ~ ( λ 1 , λ 2 , λ 3 ) , then

    σ = 2 I 3 [ ( W ^ I 1 + I 1   W ^ I 2 ) B W ^ I 2   B B ] + 2 I 3   W ^ I 3   1 = 2 J [ 1 J 2 / 3 ( W ¯ I ¯ 1 + I ¯ 1   W ¯ I ¯ 2 ) B 1 J 4 / 3   W ¯ I ¯ 2   B B ] + [ W ¯ J 2 3 J ( I ¯ 1   W ¯ I ¯ 1 + 2   I ¯ 2   W ¯ I ¯ 2 ) ]   1 = 2 J [ ( W ¯ I ¯ 1 + I ¯ 1   W ¯ I ¯ 2 ) B ¯ W ¯ I ¯ 2   B ¯ B ¯ ] + [ W ¯ J 2 3 J ( I ¯ 1   W ¯ I ¯ 1 + 2   I ¯ 2   W ¯ I ¯ 2 ) ]   1 = λ 1 λ 1 λ 2 λ 3   W ~ λ 1   n 1 n 1 + λ 2 λ 1 λ 2 λ 3   W ~ λ 2   n 2 n 2 + λ 3 λ 1 λ 2 λ 3   W ~ λ 3   n 3 n 3

    (See the page on the left Cauchy–Green deformation tensor for the definitions of these symbols).

    Incompressible isotropic hyperelastic materials

    For incompressible isotropic hyperelastic materials, the strain energy density function is W ( F ) = W ^ ( I 1 , I 2 ) . The Cauchy stress is then given by

    σ = p   1 + 2 [ ( W ^ I 1 + I 1   W ^ I 2 ) B W ^ I 2   B B ] = p   1 + 2 [ ( W I ¯ 1 + I 1   W I ¯ 2 )   B ¯ W I ¯ 2   B ¯ B ¯ ] = p   1 + λ 1   W λ 1   n 1 n 1 + λ 2   W λ 2   n 2 n 2 + λ 3   W λ 3   n 3 n 3

    where p is an undetermined pressure. In terms of stress differences

    σ 11 σ 33 = λ 1   W λ 1 λ 3   W λ 3   ;     σ 22 σ 33 = λ 2   W λ 2 λ 3   W λ 3

    If in addition I 1 = I 2 , then

    σ = 2 W I 1   B p   1   .

    If λ 1 = λ 2 , then

    σ 11 σ 33 = σ 22 σ 33 = λ 1   W λ 1 λ 3   W λ 3

    Consistency with linear elasticity

    Consistency with linear elasticity is often used to determine some of the parameters of hyperelastic material models. These consistency conditions can be found by comparing Hooke's law with linearized hyperelasticity at small strains.

    Consistency conditions for isotropic hyperelastic models

    For isotropic hyperelastic materials to be consistent with isotropic linear elasticity, the stress-strain relation should have the following form in the infinitesimal strain limit:

    σ = λ   t r ( ε )   1 + 2 μ ε

    where λ , μ are the Lame constants. The strain energy density function that corresponds to the above relation is

    W = 1 2 λ   [ t r ( ε ) ] 2 + μ   t r ( ε 2 )

    For an incompressible material t r ( ε ) = 0 and we have

    W = μ   t r ( ε 2 )

    For any strain energy density function W ( λ 1 , λ 2 , λ 3 ) to reduce to the above forms for small strains the following conditions have to be met

    W ( 1 , 1 , 1 ) = 0   ;     W λ i ( 1 , 1 , 1 ) = 0 2 W λ i λ j ( 1 , 1 , 1 ) = λ + 2 μ δ i j

    If the material is incompressible, then the above conditions may be expressed in the following form.

    W ( 1 , 1 , 1 ) = 0 W λ i ( 1 , 1 , 1 ) = W λ j ( 1 , 1 , 1 )   ;     2 W λ i 2 ( 1 , 1 , 1 ) = 2 W λ j 2 ( 1 , 1 , 1 ) 2 W λ i λ j ( 1 , 1 , 1 ) = i n d e p e n d e n t o f   i , j i 2 W λ i 2 ( 1 , 1 , 1 ) 2 W λ i λ j ( 1 , 1 , 1 ) + W λ i ( 1 , 1 , 1 ) = 2 μ     ( i j )

    These conditions can be used to find relations between the parameters of a given hyperelastic model and shear and bulk moduli.

    Consistency conditions for incompressible I 1 {displaystyle I_{1}} based rubber materials

    Many elastomers are modeled adequately by a strain energy density function that depends only on I 1 . For such materials we have W = W ( I 1 ) . The consistency conditions for incompressible materials for I 1 = 3 , λ i = λ j = 1 may then be expressed as

    W ( I 1 ) | I 1 = 3 = 0 and W I 1 | I 1 = 3 = μ 2 .

    The second consistency condition above can be derived by noting that

    W λ i = W I 1 I 1 λ i = 2 λ i W I 1 and 2 W λ i λ j = 2 δ i j W I 1 + 4 λ i λ j 2 W I 1 2 .

    These relations can then be substituted into the consistency condition for isotropic incompressible hyperelastic materials.

    References

    Hyperelastic material Wikipedia