分享

Hyperelastic material

 小白人哈 2017-05-16
Continuum mechanics
Laws[show]
Scientists[show]
Stress-strain curves for various hyperelastic material models.

A hyperelastic or Green elastic material[1] 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.

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.[2] The behavior of unfilled, vulcanized elastomers often conforms closely to the hyperelastic ideal. Filled elastomers and biological tissues[3][4] 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.

Hyperelastic material models[edit]

Saint Venant–Kirchhoff model[edit]

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

{\displaystyle {\boldsymbol {S}}=\lambda ~{\text{tr}}({\boldsymbol {E}}){\boldsymbol {\mathit {1}}}+2\mu {\boldsymbol {E}}}

where {\displaystyle {\boldsymbol {S}}} is the second Piola–Kirchhoff stress and {\displaystyle {\boldsymbol {E}}} is the Lagrangian Green strain, {\displaystyle \lambda } and {\displaystyle \mu } are the Lamé constants, and {\displaystyle {\boldsymbol {\mathit {1}}}} is the second order unit tensor.

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

{\displaystyle W({\boldsymbol {E}})={\frac {\lambda }{2}}[{\text{tr}}({\boldsymbol {E}})]^{2}+\mu {\text{tr}}({\boldsymbol {E}}^{2})}

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

{\displaystyle {\boldsymbol {S}}={\cfrac {\partial W}{\partial {\boldsymbol {E}}}}~.}

Classification of hyperelastic material models[edit]

Hyperelastic material models can be classified as:

1) phenomenological descriptions of observed behavior

2) mechanistic models deriving from arguments about underlying structure of the material

3) hybrids of phenomenological and mechanistic models

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 {\displaystyle (\lambda _{1},\lambda _{2},\lambda _{3})} :

{\displaystyle W=f(\lambda _{1})+f(\lambda _{2})+f(\lambda _{3})\,.}

Stress-strain relations[edit]

Compressible hyperelastic materials[edit]

First Piola–Kirchhoff stress[edit]

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

{\displaystyle {\boldsymbol {P}}={\frac {\partial W}{\partial {\boldsymbol {F}}}}\qquad {\text{or}}\qquad P_{iK}={\frac {\partial W}{\partial F_{iK}}}.}

where {\displaystyle {\boldsymbol {F}}} is the deformation gradient. In terms of the Lagrangian Green strain ( {\displaystyle {\boldsymbol {E}}} )

{\displaystyle {\boldsymbol {P}}={\boldsymbol {F}}\cdot {\frac {\partial W}{\partial {\boldsymbol {E}}}}\qquad {\text{or}}\qquad P_{iK}=F_{iL}~{\frac {\partial W}{\partial E_{LK}}}~.}

In terms of the right Cauchy–Green deformation tensor ( {\displaystyle {\boldsymbol {C}}} )

{\displaystyle {\boldsymbol {P}}=2~{\boldsymbol {F}}\cdot {\frac {\partial W}{\partial {\boldsymbol {C}}}}\qquad {\text{or}}\qquad P_{iK}=2~F_{iL}~{\frac {\partial W}{\partial C_{LK}}}~.}

Second Piola–Kirchhoff stress[edit]

If {\displaystyle {\boldsymbol {S}}} is the second Piola–Kirchhoff stress tensor then

{\displaystyle {\boldsymbol {S}}={\boldsymbol {F}}^{-1}\cdot {\frac {\partial W}{\partial {\boldsymbol {F}}}}\qquad {\text{or}}\qquad S_{IJ}=F_{Ik}^{-1}{\frac {\partial W}{\partial F_{kJ}}}~.}

In terms of the Lagrangian Green strain

{\displaystyle {\boldsymbol {S}}={\frac {\partial W}{\partial {\boldsymbol {E}}}}\qquad {\text{or}}\qquad S_{IJ}={\frac {\partial W}{\partial E_{IJ}}}~.}

In terms of the right Cauchy–Green deformation tensor

{\displaystyle {\boldsymbol {S}}=2~{\frac {\partial W}{\partial {\boldsymbol {C}}}}\qquad {\text{or}}\qquad S_{IJ}=2~{\frac {\partial W}{\partial C_{IJ}}}~.}

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

Cauchy stress[edit]

Similarly, the Cauchy stress is given by

{\displaystyle {\boldsymbol {\sigma }}={\cfrac {1}{J}}~{\cfrac {\partial W}{\partial {\boldsymbol {F}}}}\cdot {\boldsymbol {F}}^{T}~;~~J:=\det {\boldsymbol {F}}\qquad {\text{or}}\qquad \sigma _{ij}={\cfrac {1}{J}}~{\cfrac {\partial W}{\partial F_{iK}}}~F_{jK}~.}

In terms of the Lagrangian Green strain

{\displaystyle {\boldsymbol {\sigma }}={\cfrac {1}{J}}~{\boldsymbol {F}}\cdot {\cfrac {\partial W}{\partial {\boldsymbol {E}}}}\cdot {\boldsymbol {F}}^{T}\qquad {\text{or}}\qquad \sigma _{ij}={\cfrac {1}{J}}~F_{iK}~{\cfrac {\partial W}{\partial E_{KL}}}~F_{jL}~.}

In terms of the right Cauchy–Green deformation tensor

{\displaystyle {\boldsymbol {\sigma }}={\cfrac {2}{J}}~{\boldsymbol {F}}\cdot {\cfrac {\partial W}{\partial {\boldsymbol {C}}}}\cdot {\boldsymbol {F}}^{T}\qquad {\text{or}}\qquad \sigma _{ij}={\cfrac {2}{J}}~F_{iK}~{\cfrac {\partial W}{\partial C_{KL}}}~F_{jL}~.}

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:[5]

{\displaystyle {\boldsymbol {\sigma }}={\cfrac {2}{J}}~{\boldsymbol {B}}\cdot {\cfrac {\partial W}{\partial {\boldsymbol {B}}}}\qquad {\text{or}}\qquad \sigma _{ij}={\cfrac {2}{J}}~B_{ik}~{\cfrac {\partial W}{\partial B_{kj}}}~.}

Incompressible hyperelastic materials[edit]

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

{\displaystyle W=W({\boldsymbol {F}})-p~(J-1)}

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

{\displaystyle {\boldsymbol {P}}=-p~J{\boldsymbol {F}}^{-T}+{\frac {\partial W}{\partial {\boldsymbol {F}}}}=-p~{\boldsymbol {F}}^{-T}+{\boldsymbol {F}}\cdot {\frac {\partial W}{\partial {\boldsymbol {E}}}}=-p~{\boldsymbol {F}}^{-T}+2~{\boldsymbol {F}}\cdot {\frac {\partial W}{\partial {\boldsymbol {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

{\displaystyle {\boldsymbol {\sigma }}={\boldsymbol {P}}\cdot {\boldsymbol {F}}^{T}=-p~{\boldsymbol {\mathit {1}}}+{\frac {\partial W}{\partial {\boldsymbol {F}}}}\cdot {\boldsymbol {F}}^{T}=-p~{\boldsymbol {\mathit {1}}}+{\boldsymbol {F}}\cdot {\frac {\partial W}{\partial {\boldsymbol {E}}}}\cdot {\boldsymbol {F}}^{T}=-p~{\boldsymbol {\mathit {1}}}+2~{\boldsymbol {F}}\cdot {\frac {\partial W}{\partial {\boldsymbol {C}}}}\cdot {\boldsymbol {F}}^{T}~.}

Expressions for the Cauchy stress[edit]

Compressible isotropic hyperelastic materials[edit]

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 {\displaystyle W({\boldsymbol {F}})={\hat {W}}(I_{1},I_{2},I_{3})={\bar {W}}({\bar {I}}_{1},{\bar {I}}_{2},J)={\tilde {W}}(\lambda _{1},\lambda _{2},\lambda _{3})} , then

{\displaystyle {\begin{aligned}{\boldsymbol {\sigma }}&={\cfrac {2}{\sqrt {I_{3}}}}\left[\left({\cfrac {\partial {\hat {W}}}{\partial I_{1}}}+I_{1}~{\cfrac {\partial {\hat {W}}}{\partial I_{2}}}\right){\boldsymbol {B}}-{\cfrac {\partial {\hat {W}}}{\partial I_{2}}}~{\boldsymbol {B}}\cdot {\boldsymbol {B}}\right]+2{\sqrt {I_{3}}}~{\cfrac {\partial {\hat {W}}}{\partial I_{3}}}~{\boldsymbol {\mathit {1}}}\\&={\cfrac {2}{J}}\left[{\cfrac {1}{J^{2/3}}}\left({\cfrac {\partial {\bar {W}}}{\partial {\bar {I}}_{1}}}+{\bar {I}}_{1}~{\cfrac {\partial {\bar {W}}}{\partial {\bar {I}}_{2}}}\right){\boldsymbol {B}}-{\cfrac {1}{J^{4/3}}}~{\cfrac {\partial {\bar {W}}}{\partial {\bar {I}}_{2}}}~{\boldsymbol {B}}\cdot {\boldsymbol {B}}\right]\\&\qquad \qquad +\left[{\cfrac {\partial {\bar {W}}}{\partial J}}-{\cfrac {2}{3J}}\left({\bar {I}}_{1}~{\cfrac {\partial {\bar {W}}}{\partial {\bar {I}}_{1}}}+2~{\bar {I}}_{2}~{\cfrac {\partial {\bar {W}}}{\partial {\bar {I}}_{2}}}\right)\right]~{\boldsymbol {\mathit {1}}}\\&={\cfrac {2}{J}}\left[\left({\cfrac {\partial {\bar {W}}}{\partial {\bar {I}}_{1}}}+{\bar {I}}_{1}~{\cfrac {\partial {\bar {W}}}{\partial {\bar {I}}_{2}}}\right){\bar {\boldsymbol {B}}}-{\cfrac {\partial {\bar {W}}}{\partial {\bar {I}}_{2}}}~{\bar {\boldsymbol {B}}}\cdot {\bar {\boldsymbol {B}}}\right]+\left[{\cfrac {\partial {\bar {W}}}{\partial J}}-{\cfrac {2}{3J}}\left({\bar {I}}_{1}~{\cfrac {\partial {\bar {W}}}{\partial {\bar {I}}_{1}}}+2~{\bar {I}}_{2}~{\cfrac {\partial {\bar {W}}}{\partial {\bar {I}}_{2}}}\right)\right]~{\boldsymbol {\mathit {1}}}\\&={\cfrac {\lambda _{1}}{\lambda _{1}\lambda _{2}\lambda _{3}}}~{\cfrac {\partial {\tilde {W}}}{\partial \lambda _{1}}}~\mathbf {n} _{1}\otimes \mathbf {n} _{1}+{\cfrac {\lambda _{2}}{\lambda _{1}\lambda _{2}\lambda _{3}}}~{\cfrac {\partial {\tilde {W}}}{\partial \lambda _{2}}}~\mathbf {n} _{2}\otimes \mathbf {n} _{2}+{\cfrac {\lambda _{3}}{\lambda _{1}\lambda _{2}\lambda _{3}}}~{\cfrac {\partial {\tilde {W}}}{\partial \lambda _{3}}}~\mathbf {n} _{3}\otimes \mathbf {n} _{3}\end{aligned}}}

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

[show]Proof 1:
[show]Proof 2:
[show]Proof 3:

Incompressible isotropic hyperelastic materials[edit]

For incompressible isotropic hyperelastic materials, the strain energy density function is {\displaystyle W({\boldsymbol {F}})={\hat {W}}(I_{1},I_{2})} . The Cauchy stress is then given by

{\displaystyle {\begin{aligned}{\boldsymbol {\sigma }}&=-p~{\boldsymbol {\mathit {1}}}+2\left[\left({\cfrac {\partial {\hat {W}}}{\partial I_{1}}}+I_{1}~{\cfrac {\partial {\hat {W}}}{\partial I_{2}}}\right){\boldsymbol {B}}-{\cfrac {\partial {\hat {W}}}{\partial I_{2}}}~{\boldsymbol {B}}\cdot {\boldsymbol {B}}\right]\\&=-p~{\boldsymbol {\mathit {1}}}+2\left[\left({\cfrac {\partial W}{\partial {\bar {I}}_{1}}}+I_{1}~{\cfrac {\partial W}{\partial {\bar {I}}_{2}}}\right)~{\bar {\boldsymbol {B}}}-{\cfrac {\partial W}{\partial {\bar {I}}_{2}}}~{\bar {\boldsymbol {B}}}\cdot {\bar {\boldsymbol {B}}}\right]\\&=-p~{\boldsymbol {\mathit {1}}}+\lambda _{1}~{\cfrac {\partial W}{\partial \lambda _{1}}}~\mathbf {n} _{1}\otimes \mathbf {n} _{1}+\lambda _{2}~{\cfrac {\partial W}{\partial \lambda _{2}}}~\mathbf {n} _{2}\otimes \mathbf {n} _{2}+\lambda _{3}~{\cfrac {\partial W}{\partial \lambda _{3}}}~\mathbf {n} _{3}\otimes \mathbf {n} _{3}\end{aligned}}}

where {\displaystyle p} is an undetermined pressure. In terms of stress differences

{\displaystyle \sigma _{11}-\sigma _{33}=\lambda _{1}~{\cfrac {\partial W}{\partial \lambda _{1}}}-\lambda _{3}~{\cfrac {\partial W}{\partial \lambda _{3}}}~;~~\sigma _{22}-\sigma _{33}=\lambda _{2}~{\cfrac {\partial W}{\partial \lambda _{2}}}-\lambda _{3}~{\cfrac {\partial W}{\partial \lambda _{3}}}}

If in addition {\displaystyle I_{1}=I_{2}} , then

{\displaystyle {\boldsymbol {\sigma }}=2{\cfrac {\partial W}{\partial I_{1}}}~{\boldsymbol {B}}-p~{\boldsymbol {\mathit {1}}}~.}

If {\displaystyle \lambda _{1}=\lambda _{2}} , then

{\displaystyle \sigma _{11}-\sigma _{33}=\sigma _{22}-\sigma _{33}=\lambda _{1}~{\cfrac {\partial W}{\partial \lambda _{1}}}-\lambda _{3}~{\cfrac {\partial W}{\partial \lambda _{3}}}}

Consistency with linear elasticity[edit]

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[edit]

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:

{\displaystyle {\boldsymbol {\sigma }}=\lambda ~\mathrm {tr} ({\boldsymbol {\varepsilon }})~{\boldsymbol {\mathit {1}}}+2\mu {\boldsymbol {\varepsilon }}}

where {\displaystyle \lambda ,\mu } are the Lame constants. The strain energy density function that corresponds to the above relation is[1]

{\displaystyle W={\tfrac {1}{2}}\lambda ~[\mathrm {tr} ({\boldsymbol {\varepsilon }})]^{2}+\mu ~\mathrm {tr} ({\boldsymbol {\varepsilon }}^{2})}

For an incompressible material {\displaystyle \mathrm {tr} ({\boldsymbol {\varepsilon }})=0} and we have

{\displaystyle W=\mu ~\mathrm {tr} ({\boldsymbol {\varepsilon }}^{2})}

For any strain energy density function {\displaystyle W(\lambda _{1},\lambda _{2},\lambda _{3})} to reduce to the above forms for small strains the following conditions have to be met[1]

{\displaystyle {\begin{aligned}&W(1,1,1)=0~;~~{\cfrac {\partial W}{\partial \lambda _{i}}}(1,1,1)=0\\&{\cfrac {\partial ^{2}W}{\partial \lambda _{i}\partial \lambda _{j}}}(1,1,1)=\lambda +2\mu \delta _{ij}\end{aligned}}}

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

{\displaystyle {\begin{aligned}&W(1,1,1)=0\\&{\cfrac {\partial W}{\partial \lambda _{i}}}(1,1,1)={\cfrac {\partial W}{\partial \lambda _{j}}}(1,1,1)~;~~{\cfrac {\partial ^{2}W}{\partial \lambda _{i}^{2}}}(1,1,1)={\cfrac {\partial ^{2}W}{\partial \lambda _{j}^{2}}}(1,1,1)\\&{\cfrac {\partial ^{2}W}{\partial \lambda _{i}\partial \lambda _{j}}}(1,1,1)=\mathrm {independentof} ~i,j\neq i\\&{\cfrac {\partial ^{2}W}{\partial \lambda _{i}^{2}}}(1,1,1)-{\cfrac {\partial ^{2}W}{\partial \lambda _{i}\partial \lambda _{j}}}(1,1,1)+{\cfrac {\partial W}{\partial \lambda _{i}}}(1,1,1)=2\mu ~~(i\neq j)\end{aligned}}}

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 {\displaystyle I_{1}} based rubber materials[edit]

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

{\displaystyle W(I_{1}){\biggr |}_{I_{1}=3}=0\quad {\text{and}}\quad {\cfrac {\partial W}{\partial I_{1}}}{\biggr |}_{I_{1}=3}={\frac {\mu }{2}}\,.}

The second consistency condition above can be derived by noting that

{\displaystyle {\cfrac {\partial W}{\partial \lambda _{i}}}={\cfrac {\partial W}{\partial I_{1}}}{\cfrac {\partial I_{1}}{\partial \lambda _{i}}}=2\lambda _{i}{\cfrac {\partial W}{\partial I_{1}}}\quad {\text{and}}\quad {\cfrac {\partial ^{2}W}{\partial \lambda _{i}\partial \lambda _{j}}}=2\delta _{ij}{\cfrac {\partial W}{\partial I_{1}}}+4\lambda _{i}\lambda _{j}{\cfrac {\partial ^{2}W}{\partial I_{1}^{2}}}\,.}

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

References[edit]

  1. ^ Jump up to: a b c d R.W. Ogden, 1984, Non-Linear Elastic Deformations, ISBN 0-486-69648-0, Dover.
  2. Jump up ^ Muhr, A. H. (2005). Modeling the stress-strain behavior of rubber. Rubber chemistry and technology, 78(3), 391-425. [1]
  3. Jump up ^ Hao Gao et al., "A finite strain nonlinear human mitral valve model with fluid-structure interaction", Int J Numer Method Biomed Eng. 2014 Dec; 30(12): 1597–1613.
  4. Jump up ^ Fei Jia et al., "Morphoelasticity in the development of brown alga Ectocarpus siliculosus: from cell rounding to branching", J R Soc Interface. 2017 Feb; 14(127): 20160596.
  5. Jump up ^ Y. Basar, 2000, Nonlinear continuum mechanics of solids, Springer, p. 157.
  6. Jump up ^ Fox & Kapoor, Rates of change of eigenvalues and eigenvectors, AIAA Journal, 6 (12) 2426–2429 (1968)
  7. Jump up ^ Friswell MI. The derivatives of repeated eigenvalues and their associated eigenvectors. Journal of Vibration and Acoustics (ASME) 1996; 118:390–397.

See also[edit]

    本站是提供个人知识管理的网络存储空间,所有内容均由用户发布,不代表本站观点。请注意甄别内容中的联系方式、诱导购买等信息,谨防诈骗。如发现有害或侵权内容,请点击一键举报。
    转藏 分享 献花(0

    0条评论

    发表

    请遵守用户 评论公约

    类似文章 更多