Fact-checked by Grok 2 weeks ago

Hyperelastic material

A hyperelastic material is a type of nonlinear material whose stress- relationship is derived entirely from a scalar that depends on the current deformation state, enabling accurate modeling of large, reversible deformations without dissipation. These materials exhibit ideally behavior under finite strains, often up to several hundred percent, and are typically assumed to be isotropic and nearly incompressible, with the determined solely by the instantaneous rather than loading history or rate. Hyperelastic models are essential for simulating the mechanical response of soft solids like elastomers, where traditional fails due to geometric nonlinearities and material softening or stiffening. The theoretical foundation of hyperelasticity emerged in the 1940s as an extension of to finite deformations. Melvin Mooney introduced an early phenomenological model in 1940, postulating that the elastic potential of isotropic materials depends on two strain invariants for incompressible cases, laying the groundwork for large-deformation theory. Ronald Rivlin advanced this in 1948 by formalizing the use of strain invariants in the strain energy function for both compressible and incompressible isotropic materials, establishing the mathematical framework that ensures objectivity and thermodynamic consistency. Subsequent developments, such as the Ogden model in 1972, generalized the approach by expressing the strain energy in terms of principal stretches, providing greater flexibility to fit experimental data across various deformation modes like uniaxial tension, biaxial extension, and shear. Key hyperelastic models include the Neo-Hookean, which assumes a Gaussian chain network for rubbers and uses only the first ; the Mooney-Rivlin, incorporating the first two invariants to capture upturn in stress- curves at moderate ; and more advanced forms like the Yeoh or Arruda-Boyce models for broader ranges. These models are implemented in finite element analysis software to predict behaviors such as the of rubber bladders or the compression of seals, assuming hyperelasticity holds under quasi-static, isothermal conditions. Hyperelastic materials find widespread applications in and , including automotive tires, O-rings, biomedical implants, and , due to their , low relative to bulk , and ability to maintain constancy during deformation.

Fundamentals

Definition and Properties

Hyperelastic materials are idealized nonlinear solids capable of undergoing large, reversible deformations without , where the stress-strain relationship is derived from a scalar defined per unit reference volume. This framework, also known as Green elasticity, assumes the existence of a unique strain energy function W that depends on the deformation state, ensuring that the material response is conservative and fully recoverable upon unloading. Unlike viscoelastic materials, which exhibit time-dependent and loss, or materials that undergo permanent deformation, hyperelasticity features path-independent stress-strain behavior, making it a specialized case of nonlinear suitable for modeling purely responses under finite strains. Key properties of hyperelastic materials include their ability to sustain substantial strains—often up to 500% or more—while maintaining near-ideally elastic behavior, with stresses determined solely by the current deformation rather than loading history. These materials are typically characterized by high bulk moduli, rendering them nearly incompressible, and low shear moduli, which facilitate large shape changes without significant volume alteration. The response is often isotropic in the undeformed state, though anisotropic formulations exist for materials with directional properties, and homogeneity is assumed throughout the body. The concept originated in the 19th century with George Green's introduction of the strain energy function as a foundational element of elasticity theory, later formalized in the modern context of finite strain theory through contributions by researchers like Rivlin and Mooney in the mid-20th century. Real-world examples include rubber and elastomers, which can exhibit elastic strains exceeding 500%, as well as biological tissues such as arteries and skin, and soft polymers used in applications like seals and medical devices. These materials highlight the practical relevance of hyperelastic models in engineering and biomechanics, where large deformations occur without failure.

Kinematics and Strain Measures

In the of finite deformations for hyperelastic materials, the describe the mapping from the reference configuration to the current configuration without regard to forces. The primary quantity is the deformation gradient tensor \mathbf{F}, defined as \mathbf{F} = \frac{\partial \mathbf{x}}{\partial \mathbf{X}}, where \mathbf{x} is the position vector in the current configuration and \mathbf{X} is the position vector in the reference configuration. This second-order tensor captures both and , with the \det \mathbf{F} = J > 0 ensuring preservation of material orientation and distinguishability of inside from outside. The theorem provides a unique of \mathbf{F} into orthogonal and stretch components: \mathbf{F} = \mathbf{R} \mathbf{U} = \mathbf{V} \mathbf{R}, where \mathbf{R} is the proper orthogonal rotation tensor (\mathbf{R}^T \mathbf{R} = \mathbf{I}, \det \mathbf{R} = 1), \mathbf{U} is the right stretch tensor (symmetric, positive definite, referenced to the undeformed ), and \mathbf{V} is the left stretch tensor (symmetric, positive definite, referenced to the deformed ). This decomposition separates motion from pure deformation, facilitating the analysis of in hyperelastic models. Derived from \mathbf{F}, the right Cauchy-Green deformation tensor is \mathbf{C} = \mathbf{F}^T \mathbf{F}, which is symmetric and positive definite, measuring deformation relative to the . The left Cauchy-Green deformation tensor is \mathbf{b} = \mathbf{F} \mathbf{F}^T, symmetric and positive definite, measuring deformation relative to the . These tensors admit three principal invariants: I_1 = \tr \mathbf{C}, I_2 = \frac{1}{2} [(\tr \mathbf{C})^2 - \tr (\mathbf{C}^2)], and I_3 = \det \mathbf{C}, which are used to characterize isotropic hyperelastic responses due to their invariance under rotations. The Green-Lagrange strain tensor, a common finite strain measure in the reference configuration, is defined as \mathbf{E} = \frac{1}{2} (\mathbf{C} - \mathbf{I}), symmetric by construction and reducing to the strain tensor for small deformations. Complementarily, the Eulerian-Almansi strain tensor, an Eulerian finite strain measure in the current configuration, is \mathbf{e} = \frac{1}{2} (\mathbf{I} - \mathbf{b}^{-1}), also symmetric and suitable for describing strains post-deformation. The volumetric change is quantified by the Jacobian J = \det \mathbf{F} = \sqrt{I_3}, representing the local volume ratio between current and reference configurations; hyperelastic materials may be compressible (J \neq 1) or incompressible (J = 1). Principal stretches \lambda_i ( i=1,2,3 ) are the positive eigenvalues of \mathbf{U} or \mathbf{V}, with \lambda_i^2 being the eigenvalues of \mathbf{C} or \mathbf{b}; they enable spectral representations of deformation, such as \mathbf{C} = \sum_{i=1}^3 \lambda_i^2 \mathbf{N}_i \otimes \mathbf{N}_i, where \mathbf{N}_i are principal directions in the reference configuration.

Material Models

Strain Energy Potentials

In hyperelastic materials, the mechanical behavior is characterized by a W, defined as a scalar-valued of the deformation gradient tensor \mathbf{F}, such that W = W(\mathbf{F}). This represents the elastic stored per unit volume in the configuration. The total potential for a body is then given by the \int_V W(\mathbf{F}) \, dV over the volume V. For isotropic hyperelastic materials, W is typically expressed in terms of the principal invariants I_1, I_2, and I_3 of the right Cauchy-Green deformation tensor \mathbf{C} = \mathbf{F}^T \mathbf{F}, yielding W = W(I_1, I_2, I_3). The thermodynamic foundation of hyperelasticity derives from the Clausius-Duhem inequality, which enforces the second law of thermodynamics for continuum media under isothermal conditions. This inequality implies that the Helmholtz free energy density \psi (with W = \rho_0 \psi, where \rho_0 is the reference mass density) must satisfy \dot{W} = \mathbf{P} : \dot{\mathbf{F}}, where \mathbf{P} is the first Piola-Kirchhoff stress tensor, ensuring that the rate of change of stored energy equals the mechanical power input. Consequently, the stress follows as \mathbf{P} = \frac{\partial W}{\partial \mathbf{F}}. In the small-strain limit, this reduces to the second Piola-Kirchhoff stress being \mathbf{S} = \frac{\partial W}{\partial \mathbf{E}}, where \mathbf{E} is the Green-Lagrange strain tensor, and W increases monotonically with deformation to maintain thermodynamic consistency. To account for compressibility, the strain energy density is often decoupled into volumetric and isochoric contributions: W = W_\text{vol}(J) + W_\text{iso}(\bar{I}_1, \bar{I}_2), where J = \det \mathbf{F} is the Jacobian measuring change. Here, W_\text{vol} governs the resistance to volumetric deformation, while W_\text{iso} describes at constant . The modified invariants for the isochoric part are \bar{I}_1 = J^{-2/3} I_1 and \bar{I}_2 = J^{-4/3} I_2, which remove the volumetric effects from the original invariants to focus on shear-dominated responses. Extensions to anisotropic hyperelasticity incorporate material microstructure through additional dependencies in W, such as structural tensors derived from preferred directions (e.g., fiber orientations in biological tissues like arteries or tendons). For transversely isotropic cases with a single fiber family along direction \mathbf{a}, the strain energy becomes W = W(I_1, I_2, I_3, I_4, I_5), where I_4 = \mathbf{a} \cdot \mathbf{C} \mathbf{a} and I_5 = \mathbf{a} \cdot \mathbf{C}^2 \mathbf{a} are pseudo-invariants coupling deformation to fiber stretch and shear; the isotropic baseline remains foundational, with anisotropy superimposed via these tensors. For material stability, the strain energy function W must satisfy convexity conditions, ensuring positive definiteness of the elasticity tensors and preventing non-physical responses like material softening or phase separation. A key requirement is the Legendre-Hadamard (rank-one convexity) condition, which mandates that the second variation of W with respect to rank-one perturbations be non-negative: D^2 W(\mathbf{F})[( \mathbf{m} \otimes \mathbf{n}), (\mathbf{m} \otimes \mathbf{n})] \geq 0 for all nonzero vectors \mathbf{m}, \mathbf{n} and deformation gradients \mathbf{F}, guaranteeing positive wave speeds and ellipticity of the governing equations.

Classification and Specific Models

Hyperelastic material models are broadly classified into phenomenological and statistical (or micromechanical) categories. Phenomenological models, such as the Mooney-Rivlin and Ogden forms, are empirical constructs derived from macroscopic experimental observations, often expressed in terms of invariants to fit stress- without explicit to underlying microstructure. In contrast, statistical models, like the Neo-Hookean, originate from molecular theories of networks, modeling the of cross-linked chains under deformation to capture rubber-like behavior at the microscale. Additionally, mathematical criteria such as polyconvexity and quasiconvexity ensure the existence of solutions in value problems by imposing convexity conditions on the , promoting physical realism in numerical simulations. The Neo-Hookean model represents the simplest isotropic hyperelastic formulation, with strain energy density given by
W = \frac{\mu}{2} (\bar{I}_1 - 3) + f(J),
where \mu is the , \bar{I}_1 is the first deviatoric , and f(J) accounts for volumetric changes. Developed from of ideal rubber networks, it excels in describing small to moderate strains in rubbers, recovering for infinitesimal deformations, but underpredicts stiffening at large strains.
The Mooney-Rivlin model extends the Neo-Hookean by incorporating the second deviatoric invariant, with
W = C_1 (\bar{I}_1 - 3) + C_2 (\bar{I}_2 - 3) + f(J),
where C_1 and C_2 are material constants related to initial shear modulus ($2(C_1 + C_2)). This two-parameter phenomenological model better captures shear stiffening and upturn in stress-strain curves for elastomers under moderate deformations, though it requires careful parameter fitting to avoid unphysical responses at high strains.
The Ogden model adopts a spectral form in principal stretches, expressed as
W = \sum_{k=1}^N \frac{\mu_k}{\alpha_k} (\lambda_1^{\alpha_k} + \lambda_2^{\alpha_k} + \lambda_3^{\alpha_k} - 3) + f(J),
with material parameters \mu_k and \alpha_k. This versatile phenomenological approach, often using 1–3 terms, fits complex, non-monotonic stress-strain behaviors in rubbers and biological tissues, offering flexibility for uniaxial, biaxial, and data, but demands extensive experimental calibration.
The Saint Venant-Kirchhoff model serves as a direct extension of to finite strains, with
W = \frac{\lambda}{2} (\operatorname{tr} \mathbf{E})^2 + \mu \operatorname{tr}(\mathbf{E}^2),
where \lambda and \mu are Lamé constants, and \mathbf{E} is the Green-Lagrange strain tensor. It applies to moderate deformations in metals but exhibits non-physical behavior, such as loss of ellipticity and convexity failure, under large compressions or tensions, limiting its use to small finite strains.
The Yeoh model employs a polynomial expansion in the first deviatoric invariant,
W = \sum_{i=1}^N C_i (\bar{I}_1 - 3)^i + f(J),
commonly truncated at for practical use. This accurately represents the nonlinear response of filled rubbers in uniaxial , capturing initial stiffening and S-shaped curves, though higher-order terms may lead to without constraints.
The Arruda-Boyce model is a statistical approach based on an eight- representation of the network, with density
W = \mu \sum_{i=1}^{5} \frac{N_i}{i} (\bar{\lambda}_c^i - 1) + f(J),
where \mu is the , N_i are coefficients related to the number of Kuhn segments, and \bar{\lambda}_c is the effective chain stretch derived from \bar{I}_1. It provides accurate fitting for rubber-like materials up to large strains (around 500–700%), improving on the Neo-Hookean by accounting for limited chain extensibility.
Model selection typically involves fitting parameters to uniaxial , equibiaxial, and experimental data, prioritizing those ensuring polyconvexity for computational stability; for instance, the Venant-Kirchhoff often fails convexity at strains beyond 20–30%, while Neo-Hookean and Ogden maintain robustness up to 500% extension in rubbers.

Stress-Strain Relations

Compressible Materials

In compressible hyperelastic materials, the stress-strain relations are derived from the strain energy density function W(\mathbf{F}), where \mathbf{F} is the deformation gradient, using measures defined in the . The first Piola-Kirchhoff stress tensor \mathbf{P}, also known as the nominal stress, is obtained directly as \mathbf{P} = \frac{\partial W}{\partial \mathbf{F}}. This tensor relates the components of force in the current to the components of area in the reference , providing a nonsymmetric measure suitable for expressing the constitutive response without reference to the deformed . The second Piola-Kirchhoff stress tensor \mathbf{S} is related to \mathbf{P} by \mathbf{S} = \mathbf{F}^{-1} \mathbf{P} and, equivalently, \mathbf{S} = 2 \frac{\partial W}{\partial \mathbf{C}}, where \mathbf{C} = \mathbf{F}^T \mathbf{F} is the right Cauchy-Green deformation tensor. Unlike \mathbf{P}, \mathbf{S} is symmetric and serves as the work conjugate to the Green-Lagrange tensor \mathbf{E} = \frac{1}{2} (\mathbf{C} - \mathbf{I}), making it particularly useful for variational formulations in the reference configuration. The \boldsymbol{\sigma} in the current configuration can be obtained from \mathbf{S} via \boldsymbol{\sigma} = \frac{1}{J} \mathbf{F} \mathbf{S} \mathbf{F}^T, where J = \det \mathbf{F} is the ; however, the Piola-Kirchhoff tensors are emphasized here for their compatibility with reference-based analyses. For isotropic compressible hyperelastic materials, the strain energy density depends on the principal invariants I_1 = \tr(\mathbf{C}), I_2 = \frac{1}{2} [(\tr \mathbf{C})^2 - \tr(\mathbf{C}^2)], and I_3 = \det \mathbf{C} = J^2 of \mathbf{C}, so W = W(I_1, I_2, I_3). The second Piola-Kirchhoff stress then takes the explicit form \mathbf{S} = 2 \left( \frac{\partial W}{\partial I_1} + I_1 \frac{\partial W}{\partial I_2} \right) \mathbf{I} - 2 \frac{\partial W}{\partial I_2} \mathbf{C} + 2 I_3 \frac{\partial W}{\partial I_3} \mathbf{C}^{-1}, where \mathbf{I} is the identity tensor. This expression arises from the chain rule applied to the invariants and ensures the material response is objective and isotropic. The absence of incompressibility constraints in compressible models allows the Jacobian J to vary freely, capturing the full volumetric response and enabling the representation of materials with initial Poisson's ratios in the range from 0 to 0.5. Piola-Kirchhoff stresses are advantageous in finite element methods for hyperelastic simulations, as they are formulated in the fixed reference configuration, facilitating straightforward integration of the weak form and avoiding issues with mesh distortion in large-deformation problems.

Incompressible Materials

In hyperelastic materials exhibiting incompressibility, the deformation must satisfy the that the of the deformation gradient tensor J = \det(\mathbf{F}) = 1, ensuring no volume change occurs during deformation. This condition implies that the material response is purely deviatoric, with any volumetric contribution handled either by assuming an infinite volumetric or through a to approximate the . Such materials are characterized by a W that depends solely on the isochoric (volume-preserving) invariants of the deformation, focusing on and distortion without volumetric expansion or contraction. The second Piola-Kirchhoff stress tensor \mathbf{S} for incompressible hyperelastic materials is modified to incorporate the incompressibility constraint, given by \mathbf{S} = 2 \frac{\partial W}{\partial \mathbf{C}} - p \mathbf{C}^{-1}, where \mathbf{C} = \mathbf{F}^T \mathbf{F} is the right Cauchy-Green deformation tensor, and p is an indeterminate that enforces the incompressibility constraint J = 1 and is determined by solving the , such as through equations or boundary conditions. This pressure term adjusts the to enforce J = 1, as the direct derivative of W with respect to \mathbf{C} yields only the deviatoric response. The first Piola-Kirchhoff tensor \mathbf{P} follows similarly, expressed as \mathbf{P} = \frac{\partial W}{\partial \mathbf{F}} - p \mathbf{F}^{-T}, which is generally non-symmetric due to the incorporation of the through the term. This ensures that the reflects the directional dependence of the deformation while maintaining the incompressibility condition. For isotropic incompressible hyperelastic materials, the deviatoric component of the second Piola-Kirchhoff can be expressed in terms of the invariants I_1 and I_2 of \mathbf{C} as \mathbf{S}_{\mathrm{dev}} = 2 \left( \frac{\partial W}{\partial I_1} + I_1 \frac{\partial W}{\partial I_2} \right) \mathbf{I} - 2 \frac{\partial W}{\partial I_2} \mathbf{C}, with the full including the hydrostatic contribution -p \mathbf{C}^{-1}. This form captures the material's response to deformations through the dependence on the invariants, commonly used in models like the Mooney-Rivlin potential. The hydrostatic pressure p functions as a in the variational formulation, enforcing the incompressibility constraint and remaining indeterminate within the constitutive law itself. Its value is determined by solving the , typically through equilibrium equations or specified traction conditions on the boundary, allowing the model to adapt to external loads without altering the volume. Incompressible hyperelastic models are particularly suitable for modeling rubber-like materials, where the approaches 0.5, indicating near-perfect volume preservation under deformation. For instance, compounds exhibit a of approximately 0.4997, making these models ideal for capturing large-strain behavior in elastomers. Additionally, the constraint simplifies analyses in two-dimensional plane strain problems, where the out-of-plane stretch is , inherently satisfying J = 1 and reducing .

Cauchy Stress Expressions

Isotropic Compressible Cases

The Cauchy stress tensor \boldsymbol{\sigma} represents the true stress in the current configuration, defined as the force per unit current area, and for hyperelastic materials is given by \boldsymbol{\sigma} = \frac{1}{J} \mathbf{F} \frac{\partial W}{\partial \mathbf{F}} \mathbf{F}^T, where J = \det \mathbf{F} is the determinant of the deformation gradient \mathbf{F}, and W is the function. For isotropic compressible hyperelastic materials, where W = W(I_1, I_2, I_3) with I_1, I_2, and I_3 the principal invariants of the right Cauchy-Green tensor \mathbf{C} = \mathbf{F}^T \mathbf{F} (or equivalently of the left Cauchy-Green tensor \mathbf{b} = \mathbf{F} \mathbf{F}^T), the Cauchy admits an explicit representation in the current configuration as \boldsymbol{\sigma} = \frac{2}{J} \left[ \left( \frac{\partial W}{\partial I_1} + I_1 \frac{\partial W}{\partial I_2} \right) \mathbf{b} - \frac{\partial W}{\partial I_2} \mathbf{b}^2 + I_3 \frac{\partial W}{\partial I_3} \mathbf{b}^{-1} \right], with \mathbf{I} the identity tensor. This form arises from the assumption, which ensures the is a symmetric function of \mathbf{b}, and is obtained by chain-rule differentiation of W with respect to \mathbf{F}. A common approach decomposes the Cauchy into deviatoric and volumetric components, \boldsymbol{\sigma} = \boldsymbol{\sigma}_\mathrm{dev} + \boldsymbol{\sigma}_\mathrm{vol}, reflecting the separation of the strain energy as W = W_\mathrm{dev}(\bar{I}_1, \bar{I}_2) + W_\mathrm{vol}(J), where \bar{I}_1 = J^{-2/3} I_1 and \bar{I}_2 = J^{-4/3} I_2 are modified invariants. The volumetric part is \boldsymbol{\sigma}_\mathrm{vol} = \frac{\partial W_\mathrm{vol}}{\partial J} \mathbf{I}, and for a W_\mathrm{vol} = \frac{\kappa}{2} (J - 1)^2 (where \kappa is the initial ), this simplifies to \boldsymbol{\sigma}_\mathrm{vol} = \kappa (J - 1) \mathbf{I}. The deviatoric part captures response and is traceless, ensuring the decomposition aligns with the material's isochoric and dilatational behaviors. In uniaxial tension along the 1-direction with principal stretches \lambda_1 = \lambda and transverse stretches \lambda_2 = \lambda_3 = \sqrt{J / \lambda} (determined by zero transverse stresses), the axial Cauchy stress is \sigma_{11} = \frac{\lambda}{J} \frac{\partial W}{\partial \lambda}, where the derivative accounts for dependence on all stretches via W. These expressions are particularly suited for numerical solution of boundary value problems in the current configuration, as they directly provide forces on deformed surfaces; however, accurate computation of J is essential to handle volumetric changes without ill-conditioning near incompressibility. For specific models, consider the compressible Neo-Hookean material with W = \frac{\mu}{2} (\bar{I}_1 - 3) + \frac{\kappa}{2} (J - 1)^2, where \mu is the initial shear modulus. The corresponding is \boldsymbol{\sigma} = \frac{\mu}{J^{5/3}} \left( \mathbf{b} - \frac{1}{3} \operatorname{tr}(\mathbf{b}) \mathbf{I} \right) + \kappa (J - 1) \mathbf{I}, combining a deviatoric term scaled by J^{-5/3} with the volumetric contribution.

Isotropic Incompressible Cases

For isotropic incompressible hyperelastic materials, the Jacobian determinant satisfies J = 1, ensuring volume preservation throughout the deformation. The \boldsymbol{\sigma} takes the form \boldsymbol{\sigma} = -p \mathbf{I} + 2 \left( \frac{\partial W}{\partial I_1} + I_1 \frac{\partial W}{\partial I_2} \right) \mathbf{b} - 2 \frac{\partial W}{\partial I_2} \mathbf{b}^2, where p is the hydrostatic pressure, \mathbf{I} is the identity tensor, W = W(I_1, I_2) is the depending on the first two principal invariants of the left Cauchy-Green deformation tensor \mathbf{b} = \mathbf{F} \mathbf{F}^T , I_1 = \mathrm{tr} \mathbf{b}, and I_2 = \frac{1}{2} [ (\mathrm{tr} \mathbf{b})^2 - \mathrm{tr} (\mathbf{b}^2) ]. This expression decomposes the stress into a purely hydrostatic contribution -p \mathbf{I} and a deviatoric part that captures the material's response, with the deviatoric component being trace-free. The hydrostatic pressure p is not derived from the strain energy function W, as the incompressibility constraint renders the volumetric response indeterminate within the material model itself. Instead, p is determined by solving the equilibrium equation \nabla \cdot \boldsymbol{\sigma} = \mathbf{0} (in the absence of forces) or by enforcing specified tractions on the deformed . This Lagrange multiplier-like role of p enforces the constraint J = 1 while allowing the model to accommodate arbitrary hydrostatic stresses without volume change. In the principal frame of the deformation, where the principal stretches \lambda_i (with i = 1,2,3) satisfy the incompressibility condition \prod_{i=1}^3 \lambda_i = 1 , the principal components of the Cauchy stress simplify to \sigma_i = \lambda_i \frac{\partial W}{\partial \lambda_i} - p, \quad i = 1,2,3. This representation facilitates analytical solutions for deformations aligned with principal directions, such as uniaxial or equibiaxial extensions, by allowing p to be eliminated using conditions on two principal stresses. A representative application arises in the Mooney-Rivlin model, where W = C_1 (I_1 - 3) + C_2 (I_2 - 3) with material constants C_1 > 0 and C_2 \geq 0. For a deformation involving principal stretch \lambda in the \mathbf{e}_1 direction (with transverse stretches $1/\sqrt{\lambda}), the Cauchy stress component in that direction, after determining p from zero lateral tractions, is \boldsymbol{\sigma} = -p \mathbf{I} + 2 \left( C_1 + \frac{C_2}{\lambda^2} \right) (\lambda^2 - \lambda^{-1}) \mathbf{e}_1 \otimes \mathbf{e}_1. This expression highlights the model's ability to capture nonlinear stiffening in tension, commonly observed in rubber-like materials. The incompressible formulation offers analytical advantages in simulations of , where the trace-free deviatoric simplifies finite element implementations and enables efficient handling of large shear-dominated deformations without volumetric locking. However, a key limitation emerges in pure hydrostatic loading, where the deformation remains isochoric (), rendering p indeterminate from the constitutive relation alone and leading to singularities in the field that require careful boundary condition specification to resolve.

Consistency with Linear Elasticity

General Conditions for Isotropic Models

In the small strain limit, where the norm of the Green-Lagrange strain tensor E approaches zero (or equivalently, the infinitesimal strain tensor \varepsilon \to 0), the strain energy density function W of an isotropic hyperelastic material must reduce to the quadratic form characteristic of linear isotropic elasticity: W \approx \frac{\lambda}{2} (\operatorname{tr} \varepsilon)^2 + \mu \operatorname{tr}(\varepsilon^2), where \lambda and \mu are the Lamé constants representing the material's resistance to volumetric and deviatoric deformations, respectively. This ensures that the hyperelastic response seamlessly transitions to the classical Hookean behavior for infinitesimal deformations, a fundamental requirement for physical consistency across strain regimes. Consistency with linear elasticity is enforced through the second derivatives of W with respect to the right Cauchy-Green deformation tensor \mathbf{C} evaluated at the identity \mathbf{C} = \mathbf{I}. Specifically, the shear modulus arises from \partial^2 W / \partial \mathbf{C}^2 \big|_{\mathbf{C}=\mathbf{I}} = \mu \mathbf{I} for the deviatoric (shear) response, while the bulk modulus \kappa = \lambda + (2/3)\mu is determined from the volumetric contributions via higher-order second derivatives at the reference state. These conditions guarantee that the initial tangent stiffness tensor matches the isotropic linear elastic form, with positive \mu > 0 and \kappa > 0 required for material stability and ellipticity of the equilibrium equations. For isotropic models where W = W(I_1, I_2, I_3) and I_1, I_2, I_3 are the principal invariants of \mathbf{C}, the small strain behavior is analyzed via a Taylor series expansion of W around the reference values I_1 = 3, I_2 = 3, I_3 = 1 (corresponding to \mathbf{C} = \mathbf{I}). The partial derivatives of W with respect to the invariants satisfy the condition that ensures zero stress in the undeformed configuration, and the second-order terms yield the Lamé constants: \mu = 2 \left( \frac{\partial W}{\partial I_1} + \frac{\partial W}{\partial I_2} \right) \bigg|_{\mathbf{I}}, \quad \lambda = \kappa - \frac{2}{3} \mu, where the initial bulk modulus \kappa = 4 \frac{\partial^2 W}{\partial I_3^2} \big|_{I_3=1} for the volumetric response, ensuring positive definiteness for thermodynamic stability under hydrostatic loading. These relations allow calibration of model parameters to experimental small-strain data, such as Young's modulus and Poisson's ratio. Beyond , global mathematical well-posedness requires polyconvexity of W, a condition that W can be expressed as a convex function of the deformation gradient \mathbf{F} and its cofactors to ensure the of minimizers for boundary value problems. This convexity extends naturally to the small regime, where the quadratic form is inherently for positive moduli, preventing non-physical solutions like interpenetration of . Polyconvexity is a for proving theorems in nonlinear elasticity, particularly for isotropic materials. Representative examples illustrate these conditions. The Neo-Hookean model, with W = (\mu/2)(I_1 - 3) + U(I_3), directly satisfies the shear modulus requirement via its single parameter \mu = 2 \partial W / \partial I_1 \big|_{\mathbf{I}}, while the volumetric function U is chosen to match \kappa. In contrast, the Ogden model, W = \sum_{i=1}^N (\mu_i / \alpha_i) (\lambda_1^{\alpha_i} + \lambda_2^{\alpha_i} + \lambda_3^{\alpha_i} - 3) + U(J), requires tuning of multiple \mu_i and \alpha_i pairs to achieve the desired \mu and \lambda in the small strain limit, often via least-squares fitting to linear elastic data.

Conditions for I₁-Based Incompressible Models

In I₁-based incompressible hyperelastic models, the strain energy density function is expressed as W = f(\bar{I}_1 - 3), where \bar{I}_1 is the first invariant of the deviatoric right Cauchy-Green deformation tensor, and the second and third invariants are ignored for simplicity. A representative example is the Neo-Hookean model, where f(x) = \frac{\mu}{2} x and \mu is the . These models are particularly suited to rubber-like materials, deriving from of polymer chain networks. The small-strain shear modulus arises directly as \mu = 2 \frac{df}{dx} \big|_{x=0}, linking the nonlinear response to linear elasticity limits and aligning with chain density predictions \mu = NkT from rubber elasticity theory, where N is the number of chains per unit volume, k is Boltzmann's constant, and T is temperature. In the incompressible limit, the bulk modulus \kappa \to \infty, corresponding to Poisson's ratio \nu = 0.5 and an undefined Lamé constant \lambda, though the Lame constant λ becomes very large (λ → ∞) in the incompressible limit, corresponding to Poisson's ratio ν = 0.5, while the shear modulus μ remains finite. Consistency with linear elasticity is recovered for small deformations, where the stretch \lambda = 1 + \varepsilon and |\varepsilon| \ll 1. The deviatoric second Piola-Kirchhoff stress simplifies to S \approx 2\mu \varepsilon_{\text{dev}}, matching the deviatoric part of \sigma = 2\mu \varepsilon + \lambda (\operatorname{tr} \varepsilon) I under infinite , ensuring the Cauchy stress remains finite via . The Mooney-Rivlin model, with W = C_1 (\bar{I}_1 - [3](/page/3)) + C_2 (\bar{I}_2 - [3](/page/3)), reduces to the Neo-Hookean form when C_2 = 0, yielding small-strain \mu = 2(C_1 + C_2); this reduction holds for strains below 100%, but C_2 > 0 is required for accurate representation at larger strains, though it introduces complexities in handling the small-strain response due to the added . Experimental validation in uniaxial tests extracts \mu from the initial stress-strain , confirming the model's fidelity at low strains; however, pure I₁-based models like Neo-Hookean underpredict strain stiffening beyond moderate levels without I₂ contributions, as evidenced by poorer fits to rubber data compared to Mooney-Rivlin. This addresses a key aspect in incompressible hyperelasticity: explicit management of small-strain behavior under volume preservation.

References

  1. [1]
    Applied Mechanics of Solids (A.F. Bower) Section 3.5: Hyperelastic ...
    Hyperelastic constitutive laws are used to model materials that respond elastically when subjected to very large strains.
  2. [2]
    [PDF] Hyperelasticity 101 - OSTI.GOV
    o Hyperelastic materials can experience large strains (up to 500%) and most of it if not all is recoverable. o For example, rubber is a hyperelastic ...
  3. [3]
  4. [4]
    A Theory of Large Elastic Deformation | Journal of Applied Physics
    Mooney; A Theory of Large Elastic Deformation. J. Appl. Phys. 1 September 1940; 11 (9): 582–592. https://doi.org/10.1063/1.1712836. Download citation file ...
  5. [5]
    Large elastic deformations of isotropic materials. I. Fundamental ...
    The mathematical theory of small elastic deformations has been developed to a high degree of sophistication on certain fundamental assumptions.
  6. [6]
    Large deformation isotropic elasticity – on the correlation of theory ...
    The purpose of this paper is, while making full use of the inherent simplicity of isotropic elasticity, to construct a strain-energy function.
  7. [7]
  8. [8]
    Preface to the special issue on “Mechanics of Fibre-Reinforced ...
    Apr 8, 2015 · He thus gave rise to the part of solid mechanics which is now known as Green elasticity or hyperelasticity. An alternative route of historical ...<|control11|><|separator|>
  9. [9]
    A review on material models for isotropic hyperelasticity - Melly - 2021
    Nov 29, 2021 · The development of what is now robust research in the mechanics of hyperelastic materials began in the early 1940s when Mooney, through his ...1 Introduction · 2.1 Invariants Based · 3 Micromechanical Models
  10. [10]
    Polar Decomposition
    The polar decomposition theorem states that any deformation gradient tensor FiJ can be multiplicatively decomposed into the product of an orthogonal tensor ...
  11. [11]
    Green & Almansi Strains - Continuum Mechanics
    So the Almansi strain tensor is therefore defined as. e=12(I−F−T⋅F−1) This can also be written as.
  12. [12]
    [PDF] Chapter 6 Elasticity
    A material is said to be hyperelastic (or Green elastic) if there exists a strain energy function W such that the stress power per unit undeformed volume is the ...
  13. [13]
    Hyperelastic Materials - COMSOL Documentation
    A hyperelastic material is defined by its elastic strain energy density Ws, which is a function of the elastic strain state. It is often referred to as the ...
  14. [14]
    A General Approach to Derive Stress and Elasticity Tensors for ... - NIH
    In the material description or the reference configuration, the elasticity tensor ℂ is defined as the gradient of the PK2 stress to its work conjugate strain, ...
  15. [15]
    4.6. Hyperelasticity - BME-MM
    The hyperelastic constants in the strain-energy density function of a material determine its mechanical response. Therefore, in order to obtain successful ...<|control11|><|separator|>
  16. [16]
  17. [17]
    Limitations of the St. Venant–Kirchhoff material model in large strain ...
    We present a comprehensive study discussing the St. Venant–Kirchhoff law and its shortcomings in large stain regimes, focusing on compressive loading.
  18. [18]
    [PDF] Lecture 8 Hyperelastic solids
    Feb 8, 2022 · – Vulcanized rubber undergoes very small volume changes at very high hydrostatic pressures. It is very much easier to change its shape than to ...Missing: limitations | Show results with:limitations
  19. [19]
    Getting Started with Abaqus: Keywords Edition (6.11)
    Since typical unfilled elastomers have ratios in the range of 1,000 to 10,000 ( to ) and filled elastomers have ratios in the range of 50 to 200 ( to ), this ...
  20. [20]
    Plane stress finite element modelling of arbitrary compressible ... - NIH
    Abstract. Modelling the large deformation of hyperelastic solids under plane stress conditions for arbitrary compressible and nearly incompressible material ...Plane Stress Finite Element... · Numerical Results · Cook's Cantilever
  21. [21]
    [PDF] A transversely isotropic visco-hyperelastic constitutive model for soft ...
    is the viscous part of S. For incompressible materials, equation (52) can be modified to obtain. S=2. ∂We. ∂C. +2. ∂Wv. ∂C. pC. 1,. П54ч where p is a ...
  22. [22]
    [PDF] Modelling the behaviour of rubber-like materials to obtain correlation ...
    Material: First grade natural rubber compound Ref N° 19066. 68 Shore A ... Poisson's ratio ( v ) = 0.4997. Mooney-Rivlin constants CJQ = 0.916 MPa. CQI ...
  23. [23]
    Analysis of the compressible, isotropic, neo-Hookean hyperelastic ...
    Jan 3, 2023 · Consequently, the general form of the Cauchy stress tensor for the CNH model is given by. {\varvec{\sigma}} = GJ^{ - 5/3} {\varvec{b}} + K ...
  24. [24]
    Mooney-Rivlin Models - Continuum Mechanics
    Introduction. Mooney-Rivlin models are popular for modeling the large strain nonlinear behavior of incompressible materials, i.e., rubber.Missing: original | Show results with:original
  25. [25]
    [PDF] 4.4 Isotropic Hyperelasticity
    Principal Directions of the Cauchy Stress and Left Cauchy ... This relation holds for all compressible isotropic hyperelastic materials under the stress.
  26. [26]
    Constitutive laws - 3.5 Hyperelasticity - Applied Mechanics of Solids
    Usually stress-strain laws are given as equations relating Cauchy stress (`true' stress) to left Cauchy-Green deformation tensor. For some computations it may ...
  27. [27]
    [PDF] arXiv:0712.0227v1 [cond-mat.mtrl-sci] 3 Dec 2007
    Dec 3, 2007 · With the first choice, the constant Cauchy stress (“the pre- ... σ = −pI + 2(∂W/∂I1 + I1∂W/∂I2)B − 2(∂W/∂I2)B2,. (2.5) where p ...
  28. [28]
    [PDF] the constitutive law of incompressible bodies and existence in ...
    The existence and uniqueness of the"indeterminate" pressure is established in a general case. In the context of nonlinear incompressible elasticity, we then ...