Introduction

Vulcanized rubber is an isotropic material consisting of crosslinked polymers. This material is typically modeled as an incompressible, hyperelastic solid, where deformation is characterized by the left Cauchy–Green tensor B = FFT defined in terms of the deformation gradient F (Gurtin 1981 p. 46). This objective tensor is usually assumed to be consistent with the isochoric constraint det(B) = 1. However, applying tensile stress to vulcanized rubber causes volume to increase until strain-induced crystallization results in a decrease at large strains (Treloar 2009 pp. 295, 20–23). Filled and unfilled natural rubber also exhibit hysteresis at intermediate strains (Omnès et al. 2008, Treloar 2009 pp. 87, 89, 92), implying work in a closed cycle of deformation. While this behavior is inconsistent with predictions for hyperelastic solids (Gurtin 1981 p. 190), these models are still used to fit extensional data at large strains.

Constitutive equations for hyperelastic solids derive stress from an energy function. Rivlin (1948) developed models for rubber-like materials based on functions of the invariants tr(B) and tr(B−1). For example, the function

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

involving constants C1 and C2 determines Cauchy stress

$$ \mathbf{T}=-\mathrm{p}\mathbf{I}+2{\mathrm{C}}_1\mathbf{B}-2{\mathrm{C}}_2{\mathbf{B}}^{-1} $$
(2)

to within an unspecified pressure p associated with the constraint det(B) = 1. While many models use this approach (Hoss and Marczak 2010), most energy functions depend only on the invariant tr(B). For example, Treloar (1943) used network theory with a Gaussian distribution function to derive the neo-Hookean model

$$ \mathbf{T}=-\mathrm{p}\mathbf{I}+2{\mathrm{C}}_1\mathbf{B}, $$
(3)

where I is the identity tensor. Theories for non-Gaussian networks of flexible chains lead to similar models (Arruda and Boyce 1993; Horgan et al. 2004), where C1 becomes a function of tr(B).

Unfortunately, models based on energy functions involving tr(B) and tr(B−1) have generally failed to describe data for rubber-like materials as summarized by Destrade et al. (2017).

“Following World War II, a huge research effort was launched to find an explicit strain-energy function able to describe accurately the experimental data obtained from the testing of natural and synthetic rubbers. However, in spite of decades of intensive work in that area, to this day there is still no effective model able to perform this task in a satisfying and universal way. … Here, a satisfactory model is defined as a model able to describe the experimental data first of all from a qualitative point of view and then from a quantitative point with acceptable relative errors of prediction with respect to the data.”

Perhaps using models for hyperelastic materials outside their range of validity contributed to this lack of success.

Varga (1966) suggested an alternate approach to modeling rubber-like materials using the left stretch tensor V. This symmetric tensor and the material rotation R determine a polar decomposition F = VR of the deformation gradient (Richter 1952; Gurtin 1981 p. 46). Varga’s linear principal stress–strain model assumes stress is proportional to the objective strain tensor V − I, which is also used here to describe strain. This constitutive equation is derivable from an energy function involving tr(V). Varga’s approach, like the neo-Hookean model (3), accurately describes Treloar’s (1944) extensional data for vulcanized natural rubber out to 50–70% strain depending on the type of deformation (Ogden 1972a, Fig. 1). Even with this limited range, Varga (1966) obtains good results for several problems involving finite deformation.

Fig. 1
figure 1

A fit of extensional data (Treloar 1944) for |V| < 3.5 using Eq. (7), where coefficients m1 = 0.740 MPa and m2 = 0.0261 MPa determine all curves with a goodness of fit R2 = 0.9988, relative error E = 4.6%, and sum of squared error SSE = 0.010 (MPa)2. Residuals are differences between data and fitted values. These differences are normalized by the root mean squared error RMSE = 0.0168 MPa, which is about 1% of full scale for 39 data points shown in the figure. See Appendix 4 for additional information about determining model coefficients and statistics

Other models for hyperelastic materials specify the energy function in terms of positive principal stretches (λ1, λ2, λ3) associated with a spectral decomposition of V (Gurtin 1981 pp. 11–12). For example, Mooney (1940) assumed shear stress and strain are proportional in simple shear to develop the energy function (1) in terms of (λ1, λ2, λ3). Ogden (1972a, 1984 p. 494) later fit Treloar’s (1944) extensional data out to 600% strain as well as biaxial data obtained by Jones and Treloar (1975) using a function involving six constants

$$ w=\sum \limits_{\mathrm{i}=1}^3{\mu}_{\mathrm{i}}\left({\lambda_1}^{a_{\mathrm{i}}}+{\lambda_2}^{a_{\mathrm{i}}}+{\lambda_3}^{a_{\mathrm{i}}}-3\right), $$
(4)

where exponents (α1, α2, α3) can take non-integer values. Varga’s model is a special case of (4) with constants μ1 and α1 = 1, while the function (1) also has this form using expressions tr(B) = λ12 + λ22 + λ32 and tr(B−1) = λ1−2 + λ2−2 + λ3−2 for the invariants.

Varga (1966 p. 88), Ogden (1972a), and Treloar (2009 p. 215) all reject Rivlin’s (1948) requirement that strain energy depends only on squared principal stretches. Specifically, Treloar (2009 p. 233) states that “…the restriction of the strain energy function to even powers of extension ratios have no necessary basis in physical reality.” However, the function (1) can be expressed in the separable form

$$ w=W\left({\lambda}_1\right)+W\left({\lambda}_2\right)+W\left({\lambda}_3\right) $$
(5)

suggested by Valanis and Landel (1967). An accurate description of biaxial data apparently requires an energy function to be consistent with this form, like Ogden’s model (4). While this constraint is not sufficient, models based on non-separable energy functions generally fail to fit and superpose biaxial data. See Treloar (2009 pp. 236–251) and Ogden (1984 pp. 488–501) for further discussion of separable energy functions and their role in describing hyperelastic materials.

General constitutive equations for isotropic elastic solids use the deformation tensor B or the stretch tensor V to describe large strains (Gurtin 1981 pp. 170–171; Truesdell and Noll 1992 p. 140). Models based on B can easily be converted to a function of V since B = V2. For example, the Mooney–Rivlin model (2) is expressed in terms of the left stretch tensor in Appendix 1. However, converting a simple model like Varga’s to an isotropic function of B results in a very complicated expression requiring the positive root of a quartic equation (see Appendix 1). This asymmetry suggests that V may be a better variable for modeling elastic materials, especially since Varga’s model provides an accurate description of Treloar’s (1944) extensional data for relatively small strains. A model for rubber elasticity is developed in the next section using this approach.

Elastic solids

A constitutive equation for isotropic elastic solids (see Appendix 1) can be expressed as

$$ \mathbf{T}=-\mathrm{p}\mathbf{I}+{\mathrm{M}}_1\mathbf{V}+{\mathrm{M}}_2{\mathbf{V}}^2 $$
(6)

in terms of the left stretch tensor V, an unspecified pressure p and scalar coefficients M1 and M2. These moduli may depend on moment invariants I1 = tr(V) and I2 = tr(V2), while the remaining invariant J = det(V) = 1 is constant for incompressible materials. This solid has a fixed reference configuration, where V = I and the Cauchy stress T reduces to a pressure determined by boundary conditions. For small deformations from this reference state (F ≈ I), the left stretch tensor and moment invariants become

$$ \mathbf{V}\approx \mathbf{I}+\boldsymbol{\upvarepsilon}, \mathrm{tr}\left(\mathbf{V}\right)\approx 3+\mathrm{tr}\left(\boldsymbol{\upvarepsilon} \right),\mathrm{tr}\left({\mathbf{V}}^2\right)\approx 3+2\mathrm{tr}\left(\boldsymbol{\upvarepsilon} \right) $$

to within products of the non-objective Cauchy strain ε = [grad(u) + grad(u)T]/2, where the displacement gradient can be expressed as F − I (Gurtin 1981 p. 42). The resulting approximation

$$ \mathbf{T}\approx -\mathrm{p}\mathbf{I}+2\upmu \boldsymbol{\upvarepsilon} $$

determines stress for an incompressible linear elastic solid with a positive shear modulus μ = M1/2 + M2, where material parameters are evaluated at I1 = 3 and I2 = 3.

Coefficients M1 and M2 in Eq. (6) are usually specified by fitting extensional data like Treloar (1944). Attempts involving powers of I1/3 and I2/3 suggest the expressions M1 = m1 and M2 = m2I2/3, where constants m1 and m2 satisfy the constraint μ = m1/2 + m2 > 0 to ensure a positive shear modulus. The first term after pressure in the resulting constitutive equation

$$ \mathbf{T}=-\mathrm{p}\mathbf{I}+{\mathrm{m}}_1\mathbf{V}+{\mathrm{m}}_2{I}_2{\mathbf{V}}^2/3 $$
(7)

is essentially Varga’s model, while the last term with V2 = B and I2 = tr(B) is a variant of the neo-Hookean model (3). This term also appears in models derived from an energy function involving tr(B)2 as listed in Treloar (2009 p. 231) and Hoss and Marczak (2010). This result suggests that the constitutive eq. (7) can also be derived from an energy function.

An incompressible elastic solid (6) is hyperelastic (see Appendix 2) if Cauchy stress can be expressed as

$$ \mathbf{T}=-\mathrm{p}\mathbf{I}+\frac{\partial w}{{\partial I}_1}\mathbf{V}+2\frac{\partial w}{{\partial I}_2}{\mathbf{V}}^2 $$
(8)

in terms of a scalar function w(I1, I2) of invariants I1 = tr(V) and I2 = tr(V2). The constitutive model (7) is readily obtained using the function

$$ w\left({I}_1,{I}_2\right)={\mathrm{m}}_1\left({I}_1-3\right)+{\mathrm{m}}_2\left({I}_2^2-9\right)/12, $$
(9)

which is convex if m2 ≥ 0. The moment invariants become I1 = λ1 + λ2 + λ3 and I2 = λ12 + λ22 + λ32 in terms of principal stretches consistent with the constraint J = λ1λ2λ3 = 1. Substituting these results into (9) gives an alternate form of the energy function

$$ \hat{\mathrm{w}}\left({\lambda}_1,{\lambda}_2,{\lambda}_3\right)={\mathrm{m}}_1\left({\lambda}_1,{\lambda}_2,{\lambda}_3\hbox{--} 3\right)+{\mathrm{m}}_2\left({\lambda_1}^4+{\lambda_2}^4+{\lambda_3}^4+2{\lambda_1}^{\hbox{--} 2}+2{\lambda_2}^{\hbox{--} 2}+2{\lambda_3}^{\hbox{--} 2}\hbox{--} 9\right)/12 $$
(10)

in terms of stretches. This expression is a special case of the energy function (4) with constants (μ1 = m1, μ2 = m2/12, μ3 = m2/6, α1 = 1, α2 = 4, α3 =  – 2). Ogden (1984 pp. 495, 498) used similar exponent values (α1 = 1.3, α2 = 5, α3 =  – 2) and (α1 = 1.3, α2 = 4, α3 =  – 2) to fit extensional data by Treloar (1944) and Jones and Treloar (1975). The function

$$ W\left(\uplambda \right)={\mathrm{m}}_1\left(\uplambda -1\right)+{\mathrm{m}}_2\left({\uplambda}^4+2{\uplambda}^{-2}-3\right)/12 $$

can be used to express (10) in the separable form (5), where \( dW/ d\lambda ={\mathrm{m}}_1+{\mathrm{m}}_2\left({\lambda}^3-{\lambda}^3\right)/3=\hat{W}\left(\lambda \right) \) determines a shift factor λŴ(λ) for superposing biaxial data (Treloar 2009 pp. 241).

Treloar (1944) performed experiments on vulcanized natural rubber in equibiaxial, planar, and uniaxial extension. These results are usually expressed in terms of the nominal or first Piola–Kirchhoff stress (Gurtin 1981 p. 178)

$$ \mathbf{P}=\det \left(\mathbf{F}\right){\mathbf{TF}}^{-\mathbf{T}}, $$

where det(F) = 1 for incompressible materials. Expressions for the nominal stress component Pxx given in Appendix 3 are fit to Treloar’s data (see Steimann et al. 2012) using a method described in Appendix 4. These comparisons are limited to a range consistent with an upper bound on the stretch magnitude |V|, where |V|2 = I2/2 ≥ 3/2. For example, the term involving m1 fits data (R2 = 0.9961) for 1.225 ≤ |V| < 1.7, similar to comparisons by Ogden 1972a, Fig. 1). The term involving m2 becomes necessary outside this range, resulting in a better fit (R2 = 0.9988) out to |V| = 3.5 as shown in Fig. 1. This figure also shows the distribution of error using normalized residuals (see Appendix 4). While this approach can fit data (R2 = 0.9968) out to |V| = 5, the root mean squared error increases significantly (0.0168 → 0.0583 MPa). This increased error and a poorer fit suggest different material behavior for |V| > 3.5. Nah et al. (2010) question the value of modeling vulcanized natural rubber as hyperelastic for strains greater than 300%.

Simple shear results in Appendix 3 accurately predict derived data shown in Fig. 2. Here, shear stress Txy is obtained as a function of shear strain γ from planar extension measurements in Fig. 1 using the equivalence relations

$$ \gamma =\uplambda -{\uplambda}^{-1},{\mathrm{T}}_{\mathrm{xy}}={\mathrm{P}}_{\mathrm{xx}}/\left(1+{\uplambda}^{-2}\right) $$
(11)

given by Treloar (2009 pp. 84, 93 with a sign correction on the last exponent). This agreement is not surprising since the constitutive model (7) fits planar extension data in Fig. 1 and identically satisfies (11). The ratio Txy/γ is relatively constant in Fig. 2, decreasing by less than 20% from an initial value μ = m1/2 + m2 = 0.396 MPa. Destrade et al. (2017) obtain 8% to 24% higher estimates for μ by fitting models only to Treloar’s uniaxial data (see Appendix 5).

Fig. 2
figure 2

Shear stress values derived from Treloar’s planar extension data shown in Fig. 1 using the equivalence relations (11). Curves are predictions of Eq. (7) using coefficients from Fig. 1 with correlation R2 = 0.9990 to derived data. The deviation from predicted values RMSE = 0.0167 MPa is about 1% of full scale for derived data shown in the figure

As expected, model (7) is consistent with the universal result (Gurtin 1981 p. 177)

$$ {\mathrm{T}}_{\mathrm{xx}}-{\mathrm{T}}_{\mathrm{yy}}=\upgamma {\mathrm{T}}_{\mathrm{xy}} $$

relating primary normal stress difference N1 = Txx − Tyy to shear stress Txy for all elastic solids. This result and (11) suggest another equivalence relation N1 = Pxx(λ − λ−1)/(1 + λ−2) with planar extension data. While N1 ≥ 0, the secondary normal stress difference N2 = Tyy − Tzz is typically negative with − m1 ≤ N2 ≤ 0 for m1 > 0. In this case, the ratio N2/N1 increases from an initial value − m1/8μ = − 0.234 to zero like steady shear data for polystyrene solutions (Hua et al. 1999). Tractions associated with Tyy are negative if m1 > 0, which implies axial elongation in torsion observed by Poynting (1913) for twisted rubber cords. Similar results are obtained for silicone rubber in Appendix 5, while a negative Poynting effect involving axial contraction (m1 < 0) is predicted for porcine liver tissue. Models derived from an energy function depending only on the invariant tr(B) = tr(V2) predict no Poynting effect (Horgan and Smayda 2012).

In Treloar’s various extensional tests, hysteresis appears during unloading (Treloar 2009 pp. 87, 89, 92) at some point in the range 3.1 < |V| < 3.6. While the constitutive Eq. (7) does not predict this behavior, correlation with Treloar’s data for |V| > 3.5 can be improved by simply replacing m2 with m2/[1 − β(I1 − 3)]. Gent (1996) suggested a similar model for hyperelastic solids using a dimensionless parameter β−1 and the invariant tr(B) = tr(V2) instead of I1 = tr(V). Here, the resulting constitutive equation

$$ \mathbf{T}=-\mathrm{p}\mathbf{I}={\mathrm{m}}_1\mathbf{V}+{\mathrm{m}}_2{I}_2{\mathbf{V}}^2/3\left[1-\upbeta \left({I}_1-3\right)\right] $$
(12)

reduces to (7) for β = 0. If β > 0, stress becomes unbounded for I1 = 3 + β−1. This limited extensibility is associated with crosslinked polymers being stretched to their full extent. This limit should decrease at larger strain as more chains become fully extended. In this case, the parameter β increases with tensile strain for rubber-like materials described by the model (12).

Expressions for the nominal stress component Pxx predicted by (12) are obtained from results in Appendix 3 with m2 replaced by m2/[1 − β(I1 − 3)]. As shown in Fig. 3, these expressions accurately fit Treloar’s loading data to |V| = 5 with R2 similar to the value in Fig. 1. The parameter β decreases by 62% (0.045 → 0.017) as the limiting magnitude goes from 5 to 3.5, becoming negative for |V| = 3.4. This sign change suggests an upper bound on the hyperelastic range, where a similar fit is obtained out to |V| = 3.5 using model (7) or (12). Strain-induced crystallization may dominate uniaxial response for λ > 7, leading to hardening shown in Fig. 3. The constitutive Eq. (12) underestimates this response at large tensile strains similar to Ogden’s model (1972a, Fig. 4). This deviation suggests a hardening mechanism is active at very large strains.

Fig. 3
figure 3

A fit of extensional data (Treloar 1944) to |V| = 5 using Eq. (12), where coefficients m1 = 0.762 MPa, m2 = 0.0216 MPa, and β = 0.0446 determine curves with correlation R2 = 0.9990, relative error E = 5.2%, and SSE = 0.045 (MPa)2. The distribution of error is normalized by RMSE = 0.0309 MPa, which is about 0.8% of full scale for 50 data points (λ < 7). The uniaxial curve is extrapolated beyond fitted data to show model deviation at large strains. This fit is visually similar to results by Ogden 1972a, Fig. 4) using the six-constant energy function (4)

Fig. 4
figure 4

A fit of extensional data (Treloar 1944) for a vulcanized natural rubber using the model (14), where coefficients m1 = 0.824 MPa, m2 = 0.0191 MPa, β = 0.0390, and α = 4.04E−5 determine curves with correlation R2 = 0.9989, relative error E = 11.9%, and SSE = 0.15 (MPa)2. The distribution of error is normalized by RMSE = 0.0530 MPa, which is about 0.8% of full scale for 56 data points shown in the figure

The constitutive Eq. (12) is not derivable from an energy function, which allows work in closed processes (Gurtin 1981 p. 185). Such materials can unload along a different path, where the area between curves is the work per volume associated with a cycle. Hyperelastic solids must traverse the loading curve during this cycle, resulting in no hysteresis. However, at least part of the stress (12) is derivable from an energy function, since the alternate form

$$ \mathbf{T}=\hbox{--} \mathrm{p}\mathbf{I}+{\mathrm{m}}_1\mathbf{V}+{\mathrm{m}}_2{I}_2{\mathbf{V}}^2/3+{\mathrm{m}}_2\upbeta \left({I}_1\hbox{--} 3\right){I}_2{\mathbf{V}}^2/3\left[1\hbox{--} \upbeta \left({I}_1\hbox{--} 3\right)\right], $$
(13)

splits this constitutive equation into model (7) and a term involving β that can dissipate energy (see Appendix 2). This term is about 9.2% of total stress for planar and 19% for uniaxial and equibiaxial extension at the limit of fitted data in Fig. 3. This additive decomposition is consistent with observations by Göritz (1992) on filled and unfilled rubber. This electron microscopy study suggests that highly oriented short chains can significantly influence stress for these materials. These chains limit extensibility leading to hysteresis at intermediate to large strains associated here with positive values of β.

Hardening of vulcanized natural rubber at large strain is associated with chains aligning in the direction of extension (Treloar 2009 p. 20–23). The constitutive Eq. (12) can be modified to predict effects of strain-induced crystallization by adding a hardening term. For example, the equation

$$ \mathbf{T}=\hbox{--} \mathrm{p}\mathbf{I}+{\mathrm{m}}_1\mathbf{V}+{\mathrm{m}}_2{I}_2{\mathbf{V}}^2/3\left\{1\hbox{--} \upbeta \left[\left({I}_1\hbox{--} 3\right)+\upalpha {\left({I}_2\hbox{--} 3\right)}^3\right]\right\} $$
(14)

adds a term associated with the dimensionless parameter α to fit Treloar’s extensional data shown in Fig. 4. This model can be split like (12)

$$ \mathbf{T}=\hbox{--} \mathrm{p}\mathbf{I}+{\mathrm{m}}_1\mathbf{V}+{\mathrm{m}}_2{I}_2{\mathbf{V}}^2/3+{\mathrm{m}}_2\upbeta \left[\left({I}_1\hbox{--} 3\right)+\upalpha {\left({I}_2\hbox{--} 3\right)}^3\right]{I}_2{\mathbf{V}}^2/3\left\{1\hbox{--} \upbeta \left[\left({I}_1\hbox{--} 3\right)+\upalpha {\left({I}_2\hbox{--} 3\right)}^3\right]\right\} $$

into the constitutive Eq. (7) and a term involving α and β that can dissipate energy. Like (13), this term is not derivable from an energy function for m1 constant. The parameter α decreases by 94% (4.04E−5 → 0.248E−5) as the limiting magnitude goes from 5.4 to 4.5. A similar fit is obtained out to |V| = 5 using model (12) or (14), which suggests a lower bound on the stretch magnitude associated with hardening. However, the additional term in (14) may only be necessary for filled rubber (see Appendix 5) or natural rubber at sufficiently large strains. For example, model (12), corresponding to α = 0, is sufficient to describe extensional and shear data shown in Appendix 5 for a variety of rubber-like materials.

Treloar’s uniaxial data shown in Fig. 5 displays extensive hysteresis during unloading, which is not predicted by constitutive equations for hyperelastic materials. The unloading curve requires a modified version of (14), where the hardening term exponent is increased from 3 to 5. This modified model accurately fits both loading and unloading curves shown in the figure, but does not do as well with data in Fig. 4. The work per volume associated with this closed cycle is the area 2.24 MPa between these curves. The shear modulus μ = m1/2 + m2 drops 16% (0.415 → 0.347 MPa) for the unloading curve. Parameters β and α decrease from positive values during tensile loading to negative values for the unloading curve. The drop in β is consistent with a dynamic constraint β \( {\dot{\mathrm{I}}}_2 \) ≥ 0 imposed by dissipation (see Appendix 2), where the invariant I2 decreases during the unloading phase. Plotting this invariant as a function of λ yields a convex curve with a minimum at λ = 1. Consequently, the parameter β should also be negative while unloading from compressive stresses outside the hyperelastic range. The dynamic constraint further suggests that dissipative stress associated with β can play a significant role in shear wave propagation and other unsteady motions.

Fig. 5
figure 5

Fits of uniaxial data (Treloar 2009 p. 87) for a vulcanized natural rubber using model (14), where points are obtained from curves in the reference. Coefficients m1 = 0.785 MPa, m2 = 0.0225 MPa, β = 0.0176, and α = 3.66E−8 fit loading data with correlation R2 = 0.9999 and shear modulus μ = 0.415 MPa. Coefficients m1 = 0.586 MPa, m2 = 0.0542 MPa, β = −0.599, and α = − 1.02E−8 fit unloading data with correlation R2 = 0.9989 and shear modulus μ = 0.347 MPa

While agreement with Treloar’s extensional data is encouraging, the parameter β associated with limited extensibility exhibits a complicated dependence on invariants of V. Assume β is an increasing function of J − 1, where the isochoric constraint J = det(V) = 1 becomes a less accurate approximation outside the hyperelastic range. Since volume increases for rubber subject to tensile stress (Penn 1970), the density ρ should decrease relative to the reference state value ρ0. In this case, the expression J = ρ0/ρ (Gurtin 1981 p. 88) implies a positive value for β that increases with tensile stress. The 13% drop in β between Figs. 3 and 4 is consistent with volume reduction during strain-induced crystallization (Treloar 2009 pp. 22–23). A similar argument suggests that β becomes negative and decreases with compressive stress. This parameter could also depend on the material rate \( \dot{J} \) as well as \( {\dot{I}}_2 \) to exhibit the decrease during unloading shown in Fig. 5. Pressure should be specified as a function of invariants (Ogden 1972b) to model materials as compressible.

The special case (12) obtained by setting α = 0 in the constitutive Eq. (14) is sufficient to fit Jones and Treloar’s biaxial data (see Ogden 1984 p. 494). The normal stress difference N1 = Txx − Tyy obtained from Appendix 3 is plotted as a function of stretches λ1 and λ2 in Fig. 6. As expected, the parameter β is negative for compressive stresses, increasing from − 0.36 to − 0.0092 as the minimum λ1 stretch goes from 0.18 to 0.5. For λ1 > 0.5, the special case (7) accurately fits biaxial data (R2 = 0.9997) in Fig. 6 since β ≈ 0. If additional data were available at larger strains, a separate fit with positive β might be required to describe tensile stress outside the hyperelastic range. Note that non-zero values of β imply hysteresis for a similar vulcanized natural rubber tested by Treloar (1944). Also note that uniaxial and equibiaxial extension bound biaxial response in the I1I2 plane similar to descriptions by Urayama (2006) and Treloar (2009 p. 218) using invariants tr(B) and tr(B−1).

Fig. 6
figure 6

A fit of biaxial data (Jones and Treloar 1975) for a vulcanized natural rubber using Eq. (12), where coefficients m1 = 0.722 MPa, m2 = 0.0901 MPa, and β = − 0.358 determine curves with correlation R2 = 0.9995, relative error E = 9.4%, and SSE = 0.11 (MPa)2. The distribution of error is normalized by RMSE = 0.0302 MPa, which is about 0.5% of full scale for 125 data points shown in the figure. This fit compares well to correlations based on the energy function (4) using twice as many constants (Ogden 1984 p. 494; Ogden et al. 2004, Fig. 11)

The exceptional fit of biaxial data in Fig. 6 is aided by separability of the associated energy function (9). The normal stress difference in biaxial extension predicted by model (7) has the alternate form

$$ {\mathrm{N}}_1={\uplambda}_1\hat{W}\left({\uplambda}_1\right)-{\uplambda}_2\hat{W}\left({\uplambda}_2\right),\kern0.5em \hat{W}\left(\uplambda \right)={\mathrm{m}}_1+{\mathrm{m}}_2\left({\uplambda}^3-{\uplambda}^{-3}\right)/3 $$
(15)

implying a vertical shift λ2Ŵ(λ2) between curves in the hyperelastic range. Extending this result by replacing m2 with m2/[1 − β(I1 − 3)] essentially skews the shift between curves. However, Treloar (2009 p. 243) successfully superposed all data in Fig. 6 to a single curve. A similar result is obtained in Fig. 7 by applying the shift factor λ2Ŵ(λ2) in Eq. (15), where constants m1 and m2 fit biaxial data for λ1 > 0.5. This simple shift does surprisingly well at aligning points outside the hyperelastic range (λ1 < 0.5). This result highlights the extended role separability plays in describing biaxial data in Fig. 6.

Fig. 7
figure 7

Biaxial data from Fig. 6 are shifted up to a single curve using Eq. (15) with coefficients m1 = 0.890 MPa and m2 = 0.0478 MPa. These moduli fit data for λ1 > 0.5 in Fig. 6 with correlation R2 = 0.9997 and deviation RMSE = 0.0202 MPa, which is about 0.4% of full scale for 103 points

Equibiaxial tension approximates the state of stress in a spherical balloon, where tangential stress components Tϕϕ ≈ Tθθ are nearly equal. The volume of material is about 4πt0R02 for a thin-walled balloon (t0/R0 < 0.1) with initial radius R0 and thickness t0. An inflated balloon with radius R and thickness t has the approximate volume 4πtR2. Deformation is isochoric if t/t0 ≈ (R/R0)−2, where the tangential stretch λ = R/R0 is essentially a ratio of arc lengths. A force balance on a hemispherical section implies 2πtRTϕϕ ≈ πR2pg in terms of the gage pressure pg. Applying the isochoric constraint leads to an approximation P ≈ Tϕϕ3 for the scaled pressure P = pgR0/2 t0. This approximation becomes a function of tangential stretch

$$ P\approx {\mathrm{m}}_1\left({\uplambda}^{-2}-{\uplambda}^{-5}\right)+{\mathrm{m}}_2\left({\uplambda}^{-1}-{\uplambda}^{-7}\right)\left(2{\uplambda}^2+{\uplambda}^{-4}\right)/3 $$
(16)

using the equibiaxial result for Txx in Appendix 3. Replacing m2 with m2/[1 − β(2λ + λ−2 − 3)] extends (16) beyond the hyperelastic range. The resulting expression exhibits limited extensibility, where P becomes unbounded at λ corresponding to the largest positive root of the polynomial 2λ3 − (3 + β−1)λ2 + 1 = 0. Both expressions for scaled pressure are plotted in Fig. 8 using material parameters for vulcanized natural rubber.

Fig. 8
figure 8

A plot of scaled pressure P as a function of tangential stretch λ for a spherical balloon using both forms of Eq. (16). The lower curve uses moduli m1 = 0.740 MPa and m2 = 0.0261 MPa from Fig. 1, while the upper curve uses parameters m1 = 0.762 MPa, m2 = 0.0216 MPa, and β = 0.0446 from Fig. 3. Both curves have a maximum P = 0.263 MPa at λ = 1.375 followed by a pressure minimum outside the hyperelastic range. The lower curve increases along a straight line after the minimum, while the upper curve becomes unbounded for λ ≈ 11.95. This point decreases to λ ≈ 6.2 using results for the model (14) with parameters from Fig. 4. This limit suggests a bound on the size at rupture as shown in the figure

Curves in Fig. 8 exhibit a pressure maximum at small strains first observed by Osborne (1909) for rubber balloons. This point is measurable in experiments that control the volume of air in a balloon. Varga’s model (1966 p. 154) predicts a pressure maximum at λ = (5/2)1/3 ≈ 1.357, while Ogden (1972a) calculated the value \( \lambda ={\left[\left(2{\alpha}_1+3\right)/\left(3-{\alpha}_1\right)\right]}^{1/3{\alpha}_1} \) for − 3/2 < α1 < 3 using the first term in (4). If m2 > 0 and 0 < m2/m1 < 0.18768, the maximum predicted by (16) occurs in the range 1.357 < λ < 1.888 followed by a pressure minimum shown in Fig. 8. For comparison, the Mooney–Rivlin model (2) exhibits a maximum in the range 1.383 < λ < 1.840 if C2 > 0 and 0 < C2/C1 < 0.21446 (Mangan and Destrade 2015). The scaled pressure (16) is a strictly increasing function of λ if m2 > 0 and m2/m1 > 0.18768. The constraint m2/m1 < − 0.5 implies a positive shear modulus μ = m1/2 + m2 for m1 < 0 and m2 > 0. In this case, the pressure (16) is also a strictly increasing function of λ like the predicted curve for porcine liver tissue in Appendix 5.

Discussion

The model (14) based on the left stretch tensor V describes rubber-like materials using four parameters linked to physical processes observed in specific ranges of deformation. This constitutive equation describes hysteresis at intermediate to large strains as a consequence of limited extensibility, while the effect of crystallization at very large strains is modeled using a simple hardening term. The parameter α associated with this term vanishes at smaller tensile strains, reducing (14) to the model (12). The parameter β associated with limited extensibility tends to decrease in compression and increase in tension, with a slight decrease at very large tensile strains. This behavior suggests a dependence on density as the isochoric constraint becomes less accurate at larger strains. An abrupt drop to negative values during uniaxial unloading is consistent with a dynamic constraint imposed by dissipation. The parameter β vanishes for intermediate to small strains, reducing (12) and (14) to the constitutive Eq. (7). This equation describes hyperelastic solids, where remaining parameters satisfy the constraint m1/2 + m2 > 0 ensuring a positive shear modulus. This model can be derived from an energy function (9), which is convex if m2 ≥ 0. This function can be expressed in the separable form (10), which is a special case of Ogden’s model (4). The modulus m1 determines a Poynting effect predicted by Eqs. (7), (12), and (14) for rubber-like materials.

Data correlation for a wide variety of materials, including natural and synthetic rubber as well as soft tissue, suggest the utility of simple constitutive equations based on the left stretch tensor V. Specifically, the model (14) and special cases (12) and (7) provide excellent results with R2 > 0.995 over a large range of deformation for biaxial, planar, and uniaxial extension. The Varga term involving m1 is essential to this capability as well as predictions in simple shear and the inflation of spherical membranes. An equivalent expression in terms of the left Cauchy–Green tensor B is significantly more complicated, requiring the positive root of a quartic equation. While guessing this expression seems unlikely, simpler models based on B have failed to describe the entire range of behavior shown here. Regarding the proposed description of hysteresis, Ericksen’s (1977) prescient advice seems appropriate.

“Gradually, it has become clear that elasticity theory can predict effects that we do not commonly think of as being associated with the adjective elastic. In such cases, we should, I think, let elasticity theory enter into free competition with other theories capable of describing the effect at hand.”

Regardless of the outcome, constitutive equations presented here provide a useful starting point for accurately describing rubber-like materials in terms of an objective measure of deformation.