Skip to main content
Log in

A model for rubber elasticity

  • Original Contribution
  • Published:
Rheologica Acta Aims and scope Submit manuscript

Abstract

A constitutive equation for rubber-like materials is developed using the left stretch tensor. This process starts with a model for hyperelastic solids based on a separable energy function. This model accurately fits extensional data for vulcanized natural rubber until the onset of hysteresis at intermediate strains. Better predictions outside the hyperelastic range are obtained by directly modifying this constitutive equation to describe limited extensibility. The resulting model accurately fits biaxial, planar, and uniaxial extension data for a variety of rubber-like materials using three constants. This model also predicts simple shear results derived from planar extension data and characterizes inflation of spherical membranes for elastomers and soft tissue. A final modification accurately describes hardening associated with crystallization at large tensile strains.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

References

  • Alexander H (1968) A constitutive relation for rubber-like materials. Int J Eng Sci 6(9):549–563

    Google Scholar 

  • Arruda EM, Boyce MC (1993) A three-dimensional constitutive model for the large stretch behavior of rubber elastic materials. J Mech Phys Solids 41(2):389–412

    CAS  Google Scholar 

  • Bechir H, Chevalier L, Chaouche M, Boufala K (2006) Hyperelastic constitutive model for rubber-like materials based on the first Seth strain measures invariant. Eur J Mech A Solids 25(1):110–124

    Google Scholar 

  • Curran-Everett D (2011) Explorations in statistics: regression. Adv Physiol Educ 35:347–352

    Google Scholar 

  • Destrade M, Saccomandi G, Sgura I (2017) Methodical fitting for mathematical models of rubber-like materials. Proc R Soc A 473:20160811

    Google Scholar 

  • Dobrynin AV, Carrillo J-MY (2011) Universality in nonlinear elasticity of biological and polymeric networks and gels. Macromolecules 44:140–146

    CAS  Google Scholar 

  • Ericksen JL (1977) Special topics in elastostatics. Adv Appl Mech 17:189–244

    Google Scholar 

  • Fox JW, Goulbourne NC (2008) On the dynamic electromechanical loading of dielectric elastomer membranes. J Mech Phys Solids 56(8):2669–2686

    CAS  Google Scholar 

  • Gao Z, Lister K, Desai JP (2010) Constitutive modeling of liver tissue: experiment and theory. Ann Biomed Eng 38(2):505–516

    Google Scholar 

  • Gent AN (1996) A new constitutive relation for rubber. Rubber Chem Technol 69:59–61

    CAS  Google Scholar 

  • Göritz D (1992) Properties of rubber elastic systems at large strains. Angew Makromol Chem 202:309–329

    Google Scholar 

  • Gurtin ME (1981) An introduction to continuum mechanics. Academic Press, New York

    Google Scholar 

  • Hoger A, Carlson DE (1984) Determination of the stretch and rotation in the polar decomposition of the deformation gradient. Q J Appl Math 42:113–117 PDF

    Google Scholar 

  • Horgan CO, Ogden RW, Saccomandi S (2004) A theory of stress softening of elastomers based on finite chain extensibility. Proc R Soc Lond A 460:1737–1754

    CAS  Google Scholar 

  • Horgan CO, Smayda MG (2012) The importance of the second strain invariant in the constitutive modeling of elastomers and soft biomaterials. Mech Mater 51:43–52

    Google Scholar 

  • Hoss L, Marczak RJ (2010) A new constitutive model for rubber-like materials. Mecánica Computacional XXIX:2759–2773 PDF

    Google Scholar 

  • Hua CC, Schieber JD, Venerus DC (1999) Segment connectivity, chain-length breathing, segmental stretch, and constraint release in reptation models. III. Shear flows. Soc Rheol 43(3):701–717 PDF

    CAS  Google Scholar 

  • Jones DF, Treloar LRG (1975) The properties of rubber in pure homogeneous strain. J Phys D Appl Phys 8:1285–1304

    Google Scholar 

  • Lari DR, Schultz DS, Wang AS, On-Tat L, Stewart JM (2012) Scleral mechanics: comparing whole globe inflation and uniaxial testing. Exp Eye Res 94:128–135 PDF

    CAS  Google Scholar 

  • Mangan R, Destrade M (2015) Gent models for the inflation of spherical balloons. Int J Non-Linear Mech 68:52–58 PDF

    Google Scholar 

  • Mansouri MR, Darijani H (2014) Constitutive modeling of isotropic hyperelastic materials in an exponential framework using a self-contained approach. Int J Solids Struct 51:4316–4326

    CAS  Google Scholar 

  • Meunier L, Chagnon G, Favier D, Orgéas L, Vacher P (2008) Mechanical experimental characterisation and numerical modelling of an unfilled silicone rubber. Polym Test 27(6):765–777

    CAS  Google Scholar 

  • Miehe C, Lulei F (2001) A physically-based constitutive model for the finite viscoelastic deformations in rubbery polymers based on a directly evaluated micro–macro-transition. In: Besdo D, Schuster RH, Ihlemann J eds. Constitutive Models for Rubber II, Taylor & Francis, 117–128

  • Mooney M (1940) A theory of large elastic deformation. J Appl Phys 11:582–591

    Google Scholar 

  • Nah C, Lee GB, Lim JY, SenGupta R, Gent AN (2010) Problems in determining the elastic strain energy function for rubber. Int J Non-Linear Mech 45:232–235

    Google Scholar 

  • Ogden RW (1972a) Large deformation isotropic elasticity—on the correlation of theory and experiment for incompressible rubberlike solids. Proc R Soc Lond A 326:565-584

  • Ogden RW (1972b) Large deformation isotropic elasticity—on the correlation of theory and experiment for compressible rubberlike solids. Proc R Soc Lond A 328:567-583

  • Ogden RW (1984) Non-linear elastic deformations. John Wiley and Sons, New York

    Google Scholar 

  • Ogden RW, Saccomandi G, Sgura I (2004) Fitting hyperelastic models to experimental data. Comput Mech 34:484–502

    Google Scholar 

  • Omnès B, Thuillier S, Pilvin P, Grohens Y, Gillet S (2008) Effective properties of carbon black filled natural rubber: experiments and modeling. Composites Part A: App Sci Manufacturing 39(7):1141–1149

    Google Scholar 

  • Osborne WA (1909) The elasticity of rubber balloons and hollow viscera. Proc R Soc Lond B 81:485–499

    Google Scholar 

  • Penn RW (1970) Volume changes accompanying the extension of rubber. Trans Soc Rheol 14(4):509–517

    CAS  Google Scholar 

  • Poynting JH (1913) The changes in the length and volume of an India-rubber cord when twisted. India-Rubber J (4):6-7 (cited in Truesdell and Noll 1992, pp. 176, 235, 237 and 546)

  • Richter H (1952) Zur elnstizitätstheorie endlicher verformungen. Math Nachr 8:65–73

    Google Scholar 

  • Rivlin RS (1948) Large elastic deformations of isotropic materials IV. Further developments of the general theory. Phil Trans R Soc Lond A 241:379–397

    Google Scholar 

  • Sawyers K (1986) Comments on the paper Determination of the stretch and rotation in the polar decomposition of the deformation gradient by A. Hoger and D. E. Carlson. Q J Appl Math 44:309–311 PDF

    Google Scholar 

  • Steimann P, Hossain M, Possart G (2012) Hyperelastic models for rubber-like materials: consistent tangent operators and suitability for Treloar’s data. Arch Appl Mech 82(9):1183–1217 PDF

    Google Scholar 

  • Ting TCT (1985) Determination of C1/2, C–1/2 and more general isotropic tensor functions of C. J Elast 15:319–323

    Google Scholar 

  • Treloar LRG (1943) The elasticity of a network of long-chain molecules—II. Trans Faraday Soc 39:241–246

    CAS  Google Scholar 

  • Treloar LRG (1944) Stress-strain data for vulcanised rubber under various types of deformation. Trans Faraday Soc 40:59–70

    CAS  Google Scholar 

  • Treloar L (2009) The physics of rubber elasticity, 3rd edn. Clarendon Press, Oxford

    Google Scholar 

  • Truesdell C (ed) (1965) Continuum mechanics III, foundations of elasticity theory. Gordon and Breach, New York a parenthetial comment of the reference Richter H (1952)

  • Truesdell C (1984) Rational thermodynamics. Springer, New York

    Google Scholar 

  • Truesdell C, Noll W (1992) The nonlinear field theories of mechanics, 2nd edn. Springer, New York

    Google Scholar 

  • Urayama K (2006) An experimentalist’s view of the physics of rubber elasticity. J Polym Sci B Polym Phys 44(24):3440–3444

    CAS  Google Scholar 

  • Valanis KC, Landel RF (1967) The strain-energy function of a hyperelastic material in terms of the extension ratios. J Appl Phys 38:2997–3002

    CAS  Google Scholar 

  • VanArsdale WE (2003) Objective spin and the Rivlin–Ericksen model. Acta Mech 162:111–124

    Google Scholar 

  • Varga OH (1966) Stress–strain behavior of elastic materials. John Wiley and Sons, New York

    Google Scholar 

  • Yeoh OH (1990) Characterization of elastic properties of carbon-black-filled rubber vulcanizates. Rubber Chem Technol 63(5):792–805

    CAS  Google Scholar 

  • Yeoh OH, Fleming PD (1997) A new attempt to reconcile the statistical and phenomenological theories of rubber elasticity. J Polym Sci B Polym Phys Ed 35(12):1919–1932

    CAS  Google Scholar 

Download references

Acknowledgments

The application WebPlotDigitizer 4.1 was used to obtain numerical data from plot images. This exceptional program is copyrighted by the developer Ankit Rohatgi. Plot images were obtained directly from PDF files using Snagit 2019 or scanned from books (Figs. 5 and 6) using an Epson Perfection V700 Photo flatbed scanner. All graphs were generated using software developed in LabVIEW 2018. This software is available upon request as a LabVIEW project or a compiled Windows application requiring the LabVIEW 2018 32-bit Runtime.

Author information

Authors and Affiliations

Authors

Additional information

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendices

Model conversion

For isochoric deformation, the left stretch tensor V satisfies the Cayley–Hamilton equation (Gurtin 1981 p. 16)

$$ {\mathbf{V}}^3-{\upiota}_1{\mathbf{V}}^2+{\upiota}_2\mathbf{V}=\mathbf{I},{\upiota}_1=\mathrm{tr}\left(\mathbf{V}\right)={\boldsymbol{I}}_1,{\upiota}_2=\left[\mathrm{tr}{\left(\mathbf{V}\right)}^2-\mathrm{tr}\left({\mathbf{V}}^2\right)\right]/2=\left({I}_1^2-{I}_2\right)/2, $$

where the principal invariants1, ι2) are expressed in terms of the moment invariants (I1, I2). Multiplying this equation by V−2 and taking the trace implies the second result

$$ \mathrm{tr}\left(\mathbf{B}\right)=\mathrm{tr}\left({\mathbf{V}}^2\right)={I}_2,\mathrm{tr}\left({\mathbf{B}}^{-1}\right)=\mathrm{tr}\left({\mathbf{V}}^{-2}\right)={\left({I}_1^2-{I}_2\right)}^2/4-2{I}_1 $$

relating tr(B−1) to moment invariants of V, where B = V2. These results are used to convert function (1) into an expression

$$ w={\mathrm{C}}_1\left[\mathrm{tr}\left(\mathbf{B}\right)-3\right]+{\mathrm{C}}_2\left[\mathrm{tr}\left({\mathbf{B}}^{-1}\right)-3\right]={\mathrm{C}}_1\left({I}_2-3\right)+{\mathrm{C}}_2\left[{\left({I}_1^2-{I}_2\right)}^2/4-2{I}_1-3\right] $$

involving stretch tensor invariants, where C1 and C2 are constants. Using this energy function in (8) leads to an alternate form of the constitutive Eq. (2)

$$ \mathbf{T}+\mathrm{p}\mathbf{I}={\mathrm{C}}_2\left({I}_1^3-{I}_1{I}_2-2\right)\mathbf{V}+\left[2{\mathrm{C}}_1+{\mathrm{C}}_2\left({I}_2-{I}_1^2\right)\right]{\mathbf{V}}^2, $$

where the shear modulus μ = 2(C1 + C2) is positive. See Treloar (2009 pp. 211–229) for further discussion of the Mooney–Rivlin model.

The left stretch tensor can be expressed in terms of B using the Cayley–Hamilton equation for V. Multiplying this equation by V gives an alternate form

$$ {\mathbf{V}}^4-{\upiota}_1{\mathbf{V}}^3+{\upiota}_2{\mathbf{V}}^2=\mathbf{V} $$

in terms of principal invariants (ι1, ι2). Substituting for V3 using the Cayley–Hamilton equation results in the expression

$$ {\mathbf{B}}^2+\left({\upiota}_2-{\upiota_1}^2\right)\mathbf{B}-{\upiota}_1\mathbf{I}=\left(1-{\upiota}_1{\upiota}_2\right)\mathbf{V}, $$

where B = V2. Solving for V leads to an expression

$$ \mathbf{V}=\left[{\mathbf{B}}^2-\left({I}_2+{I_1}^2\right)\mathbf{B}/2-{I}_1\mathbf{I}\right]/\left[1-{I}_1\left({I_1}^2-{I}_2\right)/2\right], $$

involving moment invariants I1 = tr(V) and I2 = tr(V2) = tr(B). This expression can be written in terms of B−1 and B using a result

$$ {\mathbf{B}}^2={\mathbf{B}}^{-1}+\mathrm{tr}\left(\mathbf{B}\right)\mathbf{B}-\mathrm{tr}\left({\mathbf{B}}^{-1}\right)\mathbf{I} $$

obtained from the Cayley–Hamilton equation for B. A real root of the quartic equation

$$ {\left[{I}_1^2-\mathrm{tr}\left(\mathbf{B}\right)\right]}^2/4-2{I}_1=\mathrm{tr}\left({\mathbf{B}}^{-1}\right) $$

determines I1 ≥ 3 as a function of invariants tr(B) and tr(B−1). Ogden (1972a) alluded to the complexity of this isotropic function in the context of Varga’s (1966) model. See Hoger and Carlson (1984), Ting (1985), and Sawyers (1986) for a generalization where det(V) ≠ 1.

Constitutive equations for elastic solids determine Cauchy stress T = S(F) as a function of the deformation gradient F. This function satisfies the constraint S(F) = S(FH), where the symmetry transformation H is a constant rotation for isotropic materials. The polar decomposition F = VR leads to a necessary condition S(F) = S(VRH) = S(V) associated with the material rotation R = HT. The constitutive function also satisfies the constraint S(QVQT) = QS(V)QT, where Q(t) is a time-dependent rotation associated with a change of frame. This constraint implies the general model (6) for incompressible elastic solids. Substituting a complicated expression for V in terms of B, which is implicit in Gurtin’s derivation (1981 p. 167), leads to an alternate form of the general model. This result suggests that basic constitutive equations involving the natural variable V might provide a better description of isotropic elastic solids than simple models using the substitute variable B.

Hyperelastic solids

An incompressible elastic solid described by the constitutive eq. T = –pI + S(V) is hyperelastic if the symmetric tensor S(V) can be derived from a scalar energy function of the left stretch tensor V. Assume the Cauchy stress T satisfies an isothermal form of Planck’s inequality (Truesdell 1984 pp. 112–115)

$$ \mathbf{T}\cdotp \mathbf{L}-\uprho \dot{\psi}\ge 0 $$

involving the spatial velocity gradient L, density ρ, and a material derivative of the specific free energy function ψ(V), where the scalar product A·B = tr(ATB) is defined for any tensors A and B in terms of the trace function (Gurtin 1981 p. 5). For incompressible solids, this constraint reduces to

$$ \mathbf{S}\cdotp \mathbf{L}-\dot{w}\ge 0, $$

where w(V) = ρψ is the energy density function. The chain rule implies

$$ \dot{w}=\frac{\partial w}{\partial \mathbf{V}}\cdot \dot{\mathbf{V}}, $$

which can be expressed as

$$ \dot{w}=\frac{\partial w}{\partial \mathbf{V}}\mathbf{V}\cdotp \mathbf{L} $$

using an identity (VanArsdale 2003)

$$ \dot{\mathbf{V}}+{\mathbf{V}\mathbf{W}}_{\mathrm{R}}-{\mathbf{W}}_{\mathrm{R}}\mathbf{V}={\mathbf{L}}_{\mathrm{D}}\mathbf{V},\kern0.5em \mathbf{L}={\mathbf{W}}_{\mathrm{R}}+{\mathbf{L}}_{\mathrm{D}}=\dot{\mathbf{R}}{\mathbf{R}}^{\mathrm{T}}+\mathbf{R}\dot{\mathbf{U}}{\mathbf{U}}^{-1}{\mathbf{R}}^{\mathrm{T}} $$

associated with a polar decomposition F = RU = VR of the deformation gradient in terms of a material rotation R and symmetric stretch tensors U and V with positive principle values. Since the stress S and energy density w are independent of L, the isothermal form of Planck’s inequality is satisfied only if

$$ \mathbf{S}=\frac{\partial w}{\partial \mathbf{V}}\mathbf{V}, $$

where this material does not dissipate energy in any isochoric motion. Since w is an isotropic function of V, the energy density can be expressed as ŵ(I1, I2) in terms of the moment invariants I1 = tr(V) and I2 = tr(V2) (Gurtin 1981 p. 230). The resulting partial derivative

$$ \frac{\partial w}{\partial \mathbf{V}}=\frac{\partial \hat{w}}{\partial {I}_1}\mathbf{I}+2\frac{\partial \hat{w}}{\partial {I}_2}\mathbf{V} $$

leads to the expression (8) for stress in terms of the moment invariants.

However, only some models described by the general constitutive Eq. (6) have this property. For example, model (12) cannot be derived from an energy function but can be expressed in the alternate form (13). The preceding derivation suggests that one part is determined by an energy function, while the remaining stress can dissipate energy. The dissipation rate associated with this part

$$ \frac{{\mathrm{m}}_2\beta \left({I}_1-3\right){I}_2{\mathbf{V}}^2\cdotp \mathbf{L}}{3\left[1-\beta \left({I}_1-3\right)\right]}=\frac{{\mathrm{m}}_2\beta \left({I}_1-3\right){I}_2{\mathbf{V}}^2\cdotp \dot{\mathbf{V}}{\mathbf{V}}^{-1}}{3\left[1-\beta \left({I}_1-3\right)\right]}=\frac{{\mathrm{m}}_2\beta \left({I}_1-3\right){I}_2{\dot{I}}_2}{6\left[1-\beta \left({I}_1-3\right)\right]}\ge 0 $$

is non-negative for m2 > 0 if β İ2 ≥ 0 in dynamic processes. This constraint, which also applies to the constitutive Eq. (14), is consistent with the drop in β during uniaxial unloading from a large tensile strain shown in Fig. 5.

Model predictions

Predictions of model (7) with material constants m1 and m2 are listed below, where Txx, Txy, Tyy, Tzz, and Pxx denote Cauchy and nominal stress components associated with the orthonormal basis vectors \( {\overrightarrow{\mathrm{e}}}_{\mathrm{x}},{\overrightarrow{\mathrm{e}}}_{\mathrm{y}},{\overrightarrow{\mathrm{e}}}_{\mathrm{z}} \). Deformation is characterized by the left stretch tensor V and the deformation gradient F, where symbols γ, λ, λ1, λ2 denote the shear strain and principal stretches in coordinate directions. Tensors are expressed as linear combinations of basis tensors defined using a tensor product (Gurtin 1981 p. 4). Predictions of the constitutive Eqs. (12) and (14) are obtained by replacing m2 with m2/[1 − β(I1 − 3)] or m2/{1 – β[(I1 – 3) + α(I2 – 3)3]} in the expressions below.

  • Uniaxial extension:

$$ \mathbf{V}=\lambda {\overrightarrow{\mathbf{e}}}_x\otimes {\overrightarrow{\mathbf{e}}}_x+{\lambda}^{-1/2}\left({\overrightarrow{\mathbf{e}}}_y\otimes {\overrightarrow{\mathbf{e}}}_y+{\overrightarrow{\mathbf{e}}}_z\otimes {\overrightarrow{\mathbf{e}}}_z\right),\mathrm{tr}\left(\mathbf{V}\right)=\lambda +2{\lambda}^{\hbox{--} 1/2},\mathrm{tr}\left({\mathbf{V}}^2\right)={\lambda}^2+2{\lambda}^{\hbox{--} 1} $$
$$ {\mathrm{T}}_{\mathrm{yy}}={\mathrm{T}}_{\mathrm{zz}}=0\Rightarrow \mathrm{p}={\mathrm{m}}_1{\uplambda}^{\hbox{--} 1/2}+{\mathrm{m}}_2{\uplambda}^{\hbox{--} 1}\mathrm{tr}\left({\mathbf{V}}^2\right)/3\Rightarrow {\mathrm{T}}_{\mathrm{xx}}={\mathrm{m}}_1\left(\uplambda \hbox{--} {\uplambda}^{\hbox{--} 1/2}\right)+{\mathrm{m}}_2\left({\uplambda}^2\hbox{--} {\uplambda}^{\hbox{--} 1}\right)\mathrm{tr}\left({\mathbf{V}}^2\right)/3 $$
$$ \mathbf{F}=\lambda {\overrightarrow{\mathbf{e}}}_x\otimes {\overrightarrow{\mathbf{e}}}_x+{\lambda}^{-1/2}\left({\overrightarrow{\mathbf{e}}}_y\otimes {\overrightarrow{\mathbf{e}}}_y+{\overrightarrow{\mathbf{e}}}_z\otimes {\overrightarrow{\mathbf{e}}}_z\right)\Rightarrow {\mathrm{P}}_{\mathrm{xx}}={\mathrm{m}}_1\left(1\hbox{--} {\lambda}^{\hbox{--} 3/2}\right)+{\mathrm{m}}_2\left(\lambda \hbox{--} {\lambda}^{\hbox{--} 2}\right)\left({\lambda}^2+2{\lambda}^{\hbox{--} 1}\right)/3 $$
  • Planar extension:

$$ \mathbf{V}=\lambda {\overrightarrow{\mathbf{e}}}_x\otimes {\overrightarrow{\mathbf{e}}}_x+{\overrightarrow{\mathbf{e}}}_y\otimes {\overrightarrow{\mathbf{e}}}_y+{\lambda}^{-1}{\overrightarrow{\mathbf{e}}}_z\otimes {\overrightarrow{\mathbf{e}}}_z,\mathrm{tr}\left(\mathbf{V}\right)=\lambda +1+{\lambda}^{\hbox{--} 1},\mathrm{tr}\left({\mathbf{V}}^2\right)={\lambda}^2+1+{\lambda}^{\hbox{--} 2} $$
$$ {\mathrm{T}}_{\mathrm{zz}}=0\Rightarrow \mathrm{p}={\mathrm{m}}_1{\lambda}^{\hbox{--} 1}+{\mathrm{m}}_2{\lambda}^{\hbox{--} 2}\mathrm{tr}\left({\mathbf{V}}^2\right)/3\Rightarrow {\mathrm{T}}_{\mathrm{xx}}={\mathrm{m}}_1\left(\lambda \hbox{--} {\lambda}^{\hbox{--} 1}\right)+{\mathrm{m}}_2\left({\lambda}^2\hbox{--} {\lambda}^{\hbox{--} 2}\right)\mathrm{tr}\left({\mathbf{V}}^2\right)/3 $$
$$ \mathbf{F}=\lambda {\overrightarrow{\mathbf{e}}}_x\otimes {\overrightarrow{\mathbf{e}}}_x+{\overrightarrow{\mathbf{e}}}_y\otimes {\overrightarrow{\mathbf{e}}}_y+{\lambda}^{-1}{\overrightarrow{\mathbf{e}}}_z\otimes {\overrightarrow{\mathbf{e}}}_z\Rightarrow {\mathrm{P}}_{\mathrm{xx}}={\mathrm{m}}_1\left(1\hbox{--} {\lambda}^{\hbox{--} 2}\right)+{\mathrm{m}}_2\left(\lambda \hbox{--} {\lambda}^{\hbox{--} 3}\right)\left({\lambda}^2+1+{\lambda}^{\hbox{--} 2}\right)/3 $$
  • Equibiaxial extension:

$$ \mathbf{V}=\lambda \left({\overrightarrow{\mathbf{e}}}_x\otimes {\overrightarrow{\mathbf{e}}}_x+{\overrightarrow{\mathbf{e}}}_y\otimes {\overrightarrow{\mathbf{e}}}_y\right)+{\lambda}^{-2}{\overrightarrow{\mathbf{e}}}_z\otimes {\overrightarrow{\mathbf{e}}}_z,\mathrm{tr}\left(\mathbf{V}\right)=2\lambda +{\lambda}^{\hbox{--} 2},\mathrm{tr}\left({\mathbf{V}}^2\right)=2{\lambda}^2+{\lambda}^{\hbox{--} 4} $$
$$ {\mathrm{T}}_{\mathrm{zz}}=0\Rightarrow \mathrm{p}={\mathrm{m}}_1{\lambda}^{\hbox{--} 2}+{\mathrm{m}}_2{\lambda}^{\hbox{--} 4}\mathrm{tr}\left({\mathbf{V}}^2\right)/3\Rightarrow {\mathrm{T}}_{\mathrm{xx}}={\mathrm{m}}_1\left(\lambda \hbox{--} {\lambda}^{\hbox{--} 2}\right)+{\mathrm{m}}_2\left({\lambda}^2\hbox{--} {\lambda}^{\hbox{--} 4}\right)\mathrm{tr}\left({\mathbf{V}}^2\right)/3 $$
$$ \mathbf{F}=\lambda \left({\overrightarrow{\mathbf{e}}}_x\otimes {\overrightarrow{\mathbf{e}}}_x+{\overrightarrow{\mathbf{e}}}_y\otimes {\overrightarrow{\mathbf{e}}}_y\right)+{\lambda}^{-2}{\overrightarrow{\mathbf{e}}}_z\otimes {\overrightarrow{\mathbf{e}}}_z\Rightarrow {\mathrm{P}}_{\mathrm{xx}}={\mathrm{m}}_1\left(1\hbox{--} {\lambda}^{\hbox{--} 3}\right)+{\mathrm{m}}_2\left(\lambda \hbox{--} {\lambda}^{\hbox{--} 5}\right)\left(\ 2{\lambda}^2+{\lambda}^{\hbox{--} 4}\right)/3 $$
  • Biaxial extension:

$$ \mathbf{V}={\lambda}_1{\overrightarrow{\mathbf{e}}}_x\otimes {\overrightarrow{\mathbf{e}}}_x+{\lambda}_2{\overrightarrow{\mathbf{e}}}_y\otimes {\overrightarrow{\mathbf{e}}}_y+{\left({\lambda}_1{\lambda}_2\right)}^{-1}{\overrightarrow{\mathbf{e}}}_z\otimes {\overrightarrow{\mathbf{e}}}_z,\mathrm{tr}\left(\mathbf{V}\right)={\lambda}_1+{\lambda}_2+{\left({\lambda}_1{\lambda}_2\right)}^{\hbox{--} 1}, $$
$$ \mathrm{tr}\left({\mathbf{V}}^2\right)={\lambda_1}^2+{\lambda_2}^2+{\left({\lambda}_1{\lambda}_2\right)}^{\hbox{--} 2};{\mathrm{T}}_{\mathrm{xx}}\hbox{--} {\mathrm{T}}_{\mathrm{yy}}={\mathrm{m}}_1\left({\lambda}_1\hbox{--} {\lambda}_2\right)+{\mathrm{m}}_2\left({\lambda_1}^2\hbox{--} {\lambda_2}^2\right)\mathrm{tr}\left({\mathbf{V}}^2\right)/3, $$
  • Simple shear:

$$ \mathbf{V}={\left(4+{\upgamma}^2\right)}^{-1/2}\left[\left(2+{\upgamma}^2\right){\overrightarrow{\mathbf{e}}}_x\otimes {\overrightarrow{\mathbf{e}}}_x+\upgamma \left({\overrightarrow{\mathbf{e}}}_x\otimes {\overrightarrow{\mathbf{e}}}_y+{\overrightarrow{\mathbf{e}}}_y\otimes {\overrightarrow{\mathbf{e}}}_x\right)+2{\overrightarrow{\mathbf{e}}}_y\otimes {\overrightarrow{\mathbf{e}}}_y\right]+{\overrightarrow{\mathbf{e}}}_z\otimes {\overrightarrow{\mathbf{e}}}_z, $$
$$ \mathrm{tr}\left(\mathbf{V}\right)=1+{\left(4+{\upgamma}^2\right)}^{1/2},\mathrm{tr}\left({\mathbf{V}}^2\right)=3+{\upgamma}^2;{\mathrm{T}}_{\mathrm{xy}}={\mathrm{m}}_1\upgamma {\left(4+{\upgamma}^2\right)}^{\hbox{--} 1/2}+{\mathrm{m}}_2\upgamma \left(3+{\upgamma}^2\right)/3, $$
$$ {\mathrm{T}}_{\mathrm{xx}}={\mathrm{m}}_1\left[\left(2+{\upgamma}^2\right){\left(4+{\upgamma}^2\right)}^{\hbox{--} 1/2}\hbox{--} 1\right]+{\mathrm{m}}_2{\upgamma}^2\left(3+{\upgamma}^2\right)/3,{\mathrm{T}}_{\mathrm{yy}}={\mathrm{m}}_1\left[2{\left(4+{\upgamma}^2\right)}^{\hbox{--} 1/2}\hbox{--} 1\right],{\mathrm{T}}_{\mathrm{zz}}=0 $$

Fitting algorithm

The method used here searches for m model parameters that improve goodness of fit (also known as coefficient of determination)

$$ {\mathrm{R}}^2=1-\sum \limits_{\mathrm{i}=0}^{\mathrm{n}-1}{\left({y}_{\mathrm{i}}-{f}_{\mathrm{i}}\right)}^2/\sum \limits_{\mathrm{i}=0}^{\mathrm{n}-1}{\left({y}_{\mathrm{i}}-\overline{y}\right)}^2, $$

where predicted values fi are compared to n measurements yi with mean ͞y = Σyi/n. Residuals are errors yi − fi associated with this correlation. The root mean squared error

$$ \mathrm{RMSE}=\sqrt{\frac{1}{n-m}\sum \limits_{\mathrm{i}=0}^{n-1}{\left({y}_{\mathrm{i}}-{f}_{\mathrm{i}}\right)}^2} $$

is essentially the standard deviation of data about a fit with n − m degrees of freedom. This statistic normalizes residuals plotted in most figures, suggesting a confidence interval for predicted values. Root mean squared error is compared with the entire range of measurements ymax − ymin to obtain percent full scale. Relative errors Ei are defined by Ogden et al. (2004) as absolute residuals |yi − fi| normalized by the maximum of 0.5 or |yi|. The largest relative error E and the sum of squared error SSE = Σ(yi − fi)2 are provided with two significant digits for comparison with results in other papers. The goodness of fit is expressed with four significant digits, while model coefficients and the root mean squared error appear with three digits. A fit with R2 > 0.999 and RSME < 1% full scale is considered excellent, while correlations with R2 < 0.99 and RSME > 2% full scale are somewhat questionable for data shown in the figures.

The resulting algorithm randomly searches for new coefficients near the initial guess, where the change is usually less than 5% of the value. If a better fit is obtained with R2 closer to one, the search resumes from the new set of coefficients. This process stops after a specified number of attempts (typically 1000) fails to improve the fit. Like most methods, this approach is sensitive to an initial guess. However, this guess can usually be refined in real time for m < 4 since the search only involves model evaluations. The resulting coefficients listed in figure captions concurrently fit all data displayed within the specified range except for Figs. 5 and 18. Shear stress predictions in Figs. 2, 10, and 12 do not involve this algorithm, where statistics are provided to characterize correlation with derived data.

An alternate approach is used by Ogden et al. (2004), where fitted values are determined using a least-squares analysis based on the function SSE = Σ(fi − yi)2. These authors randomly vary their initial guess to improve this solution as determined by SSE and relative errors Ei. Plotting relative errors masks the trend of residuals (Curran-Everett 2011) by mapping all values to the positive quadrant. The significance of relative error is difficult to assess without additional information about experimental uncertainty. Since most test equipment has a fixed resolution, this uncertainty can be a large percentage of small data values. For example, Ogden et al. (2004) obtain relative errors up to 40% for small strains, where models seem to accurately describe extensional data. Smaller errors around 10% can occur at larger strains, where models visually deviate from data. Treloar’s uniaxial and equibiaxial measurements are fit simultaneously by Ogden et al. (2004) to obtain a set of coefficients, while excluding planar extension data. This paper uses an energy function to describe Treloar’s data well outside the hyperelastic range.

The approach in Ogden et al. (2004) is modified by Destrade et al. (2017), where fitted values are determined by minimizing the function Σ[(fi − yi)/yi]2 involving relative residuals. These authors must exclude measurements near zero to use this approach, which weights smaller values as more significant in determining a fit compared to a least-squares analysis by Ogden et al. (2004). This weighting is further exacerbated by an apparent choice of units, which stretches the y-axis by a factor of ten. At small strains, Destrade et al. (2017) obtain relative residuals around 15 to 20% for the neo-Hookean model even though fitted values appear to accurately describe data deemed most reliable. They determine model coefficients by fitting just uniaxial data. See referee Martin Kroon’s concern in Supplemental Material as well as observations by Urayama (2006) and Treloar (2009 p. 218) that uniaxial results are insufficient to distinguish between theories. Destrade et al. 2017, Fig. 9) limit predictions in simple shear to γ ≤ 1 and do not compare with values derived from Treloar’s planar extension data. They also use an energy function to describe Treloar’s uniaxial data well outside the hyperelastic range (λ < 5). Neither this paper nor Ogden et al. (2004) provides guidance concerning the significance of any particular relative error.

Fig. 9
figure 9

A fit of planar and uniaxial data (Meunier et al. 2008) for unfilled silicone rubber using Eq. (12), where coefficients m1 = 0.262 MPa, m2 = 0.190 MPa, and β = −0.404 determine curves with correlation R2 = 0.9989, relative error E = 10.4%, and SSE = 0.024 (MPa)2. Distribution of error is normalized by RMSE = 0.0275 MPa, which is about 0.8% of full scale for 35 data points shown in the figure

Data comparisons

The constitutive Eq. (12) is used to fit extensional data obtained from Mansouri and Darijani (2014) for different rubber-like materials. Correlations in Figs. 9, 11, 14, 15, 16 and 17 have a goodness of fit R2 > 0.995 and a root mean squared error RMSE < 2.2% of full scale. Predictions by (12) are similar to the best models evaluated by Mansouri and Darijani (2014, table 4) based on the sum of squared error SSE. In Figs. 10 and 12, simple shear predictions are compared with values derived from planar extension data using the equivalence relations (11). These predictions suggest axial elongation in torsion for silicone rubber, while the porcine liver tissue should exhibit a negative Poynting effect associated with axial contraction. A prediction for inflation of a spherical membrane is shown in Fig. 13 using parameters for porcine liver tissue. This plot is qualitatively similar to measurements on soft tissue by Osborne (1909) and Lari et al. (2012).

Fig. 10
figure 10

Shear stress values derived from planar extension data for an unfilled silicone rubber in Fig. 9 using the equivalence relations (11). An absolute value is used to map all data to this quadrant. Curves are predictions of Eq. (12) using coefficients from Fig. 9 with correlation R2 = 0.9907 to derived data. Deviation from predicted values is RMSE = 0.0209 MPa or about 1.8% of full scale over the entire range of shear strain (− 1.07 → 1.62). The ratio Txy/γ increases from an initial value μ = m1/2 + m2 = 0.321 MPa, while the ratio (Tyy − Tzz)/(Txx − Tyy) increases from − m1/8μ = − 0.102 to zero. This material should exhibit axial elongation in torsion since m1 > 0

Fig. 11
figure 11

A fit of planar extension data (Gao et al. 2010) for porcine liver tissue using Eq. (12), where coefficients m1 = − 0.946 kPa, m2 = 0.479 kPa, and β = 2.14 determine curves with correlation R2 = 0.9996, relative error E = 4.4%, and SSE = 0.0012 (kPa)2. Distribution of error is normalized by RMSE = 0.00866 kPa, which is about 0.7% of full scale for 29 data points shown in the figure. Note that the shear modulus μ = m1/2 + m2 = 0.006 kPa is positive for this soft tissue

Fig. 12
figure 12

Shear stress values derived from planar extension data for porcine liver tissue in Fig. 11 using the equivalence relations (11). Curves are predictions of Eq. (12) using coefficients from Fig. 11 with correlation R2 = 0.9997 to derived data. Deviation from predicted values is RMSE = 0.00545 kPa or about 0.6% of full scale for data shown in the figure. Note that the secondary normal stress difference Tyy − Tzz is positive. This material should exhibit axial contraction in torsion since m1 < 0

Fig. 13
figure 13

Scaled pressure P as a function of tangential stretch λ for a spherical membrane using the extended version of Eq. (16) with coefficients from Fig. 11. This curve becomes unbounded for λ > 1.516, suggesting a limit on the size at rupture. This trend is qualitatively similar to measurements by Osborne (1909) on a monkey bladder as well as porcine sclera data obtained by Lari et al. (2012). These authors also observe hysteresis during sclera deflation and uniaxial unloading, which is consistent with β ≠ 0

Fig. 14
figure 14

A fit of equibiaxial and uniaxial data (Alexander 1968) for a synthetic rubber neoprene using Eq. (12), where coefficients m1 = 1.90 MPa, m2 = 0.0597 MPa, and β = 0.0887 determine curves with correlation R2 = 0.9958, relative error E = 54%, and SSE = 8.7 (MPa)2. The distribution of error is normalized by RMSE = 0.628 MPa, which is about 2.2% of full scale for 25 data points shown in the figure

Fig. 15
figure 15

A fit of uniaxial data (Fox and Goulbourne 2008) for polyacrylate rubber VHB 4905 using Eq. (12), where coefficients m1 = 0.144 MPa, m2 = 0.00135 MPa, and β = 0.0872 determine a curve with correlation R2 = 0.9994, relative error E = 1.1%, and SSE = 0.00013 (MPa)2. The distribution of error is normalized by RMSE = 0.00308 MPa, which is about 0.7% of full scale for 17 data points shown in the figure. The value of β drops 79% to 0.0183 as the maximum stretch decreases to 6. A fit with β = 0 for λ < 6 has coefficients m1 = 0.137 MPa and m2 = 0.00204 MPa with correlation R2 = 0.9995, which suggests an upper bound on the hyperelastic range for this material

Fig. 16
figure 16

A fit of uniaxial data (Yeoh and Fleming 1997) for a vulcanized rubber using Eq. (12), where coefficients m1 = 0.168 MPa, m2 = 0.500 MPa, and β = −2.11 determine a curve with correlation R2 = 0.9992, relative error E = 11.9%, and SSE = 0.081 (MPa)2. The distribution of error is normalized by RMSE = 0.0949 MPa, which is about 1% of full scale for 12 data points shown in the figure

Fig. 17
figure 17

A fit of uniaxial data (Miehe and Lulei 2001) for carbon black–filled rubber b186 using Eq. (12), where coefficients m1 = 3.28 MPa, m2 = 0.0282, and β = 2.37 determine a curve with correlation R2 = 0.9979, relative error E = 13.5%, and SSE = 0.11 (MPa)2. The distribution of error is normalized by RMSE = 0.0745 MPa, which is about 1.4% of full scale for 23 data points shown in the figure. A better result can be obtained by fitting compressive and tensile stresses separately similar to Fig. 18

Fig. 18
figure 18

Uniaxial data for carbon black–filled rubber NR 70 obtained by Bechir et al. (2006), where compressive ( ) and tensile ( ) stresses are fit separately using models (7) and (14). Coefficients m1 = − 0.134 MPa, m2 = 3.00 MPa determine the compressive stress curve with shear modulus μ = 2.94 MPa and correlation R2 = 0.9985 RMSE = 0.140 MPa. Coefficients m1 = 2.04 MPa, m2 = 0.0906 MPa, β = 2.27, and α = −0.00713 determine the tensile stress curve with shear modulus μ = 1.11 MPa and correlation R2 = 0.9981, RMSE = 0.0681 MPa

Fig. 19
figure 19

A fit of uniaxial data (Dobrynin and Carrillo 2011) for rubber using Eq. (7), where coefficients m1 = 1.45 MPa and m2 = 0.0506 MPa determine a curve with correlation R2 = 0.9993, relative error E = 10.1%, and SSE = 0.024 (MPa)2. The distribution of error is normalized by RMSE = 0.0311 MPa, which is about 0.7% of full scale for 27 data points shown in the figure. This material should be hyperelastic over the entire range of deformation shown in the figure

Uniaxial data shown in Fig. 18 are obtained from a plot in Bechir et al. (2006, Fig. 8b) for carbon black–filled vulcanized natural rubber. Separate curves for compressive and tensile stress are fit using models (7) and (14). Such an approach appears to be necessary for carbon black–filled rubber due to amplification (Yeoh 1990) at sufficiently large tensile strains, where particle interactions can lead to increased stiffness, tensile strength, and hysteresis (Omnès et al. 2008). While correlation is good, the predicted shear modulus μ = m1/2 + m2 is not continuous at λ = 1. An unusual negative value for α in (14) essentially keeps the tensile stress curve from becoming unbounded at large strains. While using six constants to fit uniaxial data seems excessive, Bechir et al. (2006, Fig. 9) obtain worse correlation, especially for compressive stress, with the same number of constants.

Values shown in Fig. 19 are obtained from a plot in Destrade et al. (2017, Fig. 2a), where stresses are about 9.8 times too large in the stated units N/mm2 (Sec. 3a). These authors exclude the first 14 data points (λ < 1.1) from their analysis due to a Mooney plot discrepancy that they attribute to possible issues with the experiment. While some points are excluded here for λ < 1.5 due to large, overlapping plot symbols, scanned data do not exhibit the downturn shown in Destrade et al. (2017, Fig. 2c). This material is accurately described by model (7), which suggests hyperelastic behavior over the entire range of deformation shown in Fig. 19. The shear modulus μ = m1/2 + m2 = 0.777 MPa predicted for this material is within ± 2% of values (0.761 → 0.791 MPa) obtained by Destrade et al. (2017, tables 2, 3, 5) using 12 different models for hyperelastic solids. However, these authors obtain a 14% variation in μ (0.430 → 0.492 MPa) by fitting the same models to Treloar’s (1944) uniaxial data (Destrade et al. 2017, tables 2, 3, 4), where some correlations involve data outside the hyperelastic range. These values are typically larger with greater variation than the shear modulus (0.396 → 0.431 MPa) obtained for Treloar’s (1944) loading data shown in Figs. 1, 3, 4, and 5.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

VanArsdale, W.E. A model for rubber elasticity. Rheol Acta 59, 905–920 (2020). https://doi.org/10.1007/s00397-020-01229-1

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00397-020-01229-1

Keywords

Navigation