Fact-checked by Grok 2 weeks ago

Strain energy density function

The strain energy density function, often denoted as W or U, is a scalar-valued function in that quantifies the stored per unit reference volume in a material subjected to deformation. It serves as the foundational constitutive relation for hyperelastic materials, enabling the derivation of stress tensors—such as the second Piola-Kirchhoff stress S = 2 \frac{\partial W}{\partial \mathbf{C}}—directly from its partial derivatives with respect to kinematic measures like the right Cauchy-Green deformation tensor \mathbf{C} = \mathbf{F}^T \mathbf{F}, where \mathbf{F} is the deformation gradient. In hyperelasticity, the function W is typically formulated to depend on the principal invariants I_1, I_2, I_3 of \mathbf{C} for isotropic materials, ensuring path-independent loading-unloading behavior and thermodynamic consistency under isothermal conditions, as dictated by the Clausius-Duhem inequality. For nearly incompressible materials, volumetric terms are incorporated, such as W = \tilde{W}(\mathbf{C}) + U(J), where J = \det \mathbf{F} and U(J) penalizes deviations from incompressibility. This framework extends , where W reduces to a W = \frac{1}{2} \boldsymbol{\varepsilon} : \mathbb{C} : \boldsymbol{\varepsilon} in terms of the infinitesimal strain tensor \boldsymbol{\varepsilon} and the \mathbb{C}. Prominent models include the Neo-Hookean function W = \frac{\mu}{2} (I_1 - 3) + \frac{1}{2} K (J - 1)^2, which captures Gaussian chain statistics in networks like rubber, and the Mooney-Rivlin extension W = C_1 (I_1 - 3) + C_2 (I_2 - 3) + \frac{1}{2} K (J - 1)^2, accounting for additional nonlinearities in moderate strains. These functions are widely applied in simulating large deformations in biological tissues, elastomers, and engineering components, with parameters calibrated via experimental stress-strain data.

Fundamentals

Definition

The strain energy density , often denoted as W, is a scalar-valued that quantifies the elastic strain energy stored per unit reference volume of the material body. It relates the stored energy to measures of deformation, typically expressed as W = W(\varepsilon) for infinitesimal strains, where \varepsilon is the strain tensor, or W = W(\mathbf{F}) for finite deformations, where \mathbf{F} is the deformation gradient tensor. In finite deformations, this is per unit reference volume to account for the deformation gradient. This concept was introduced by George Green in 1839 as part of his foundational work on the mathematical theory of elasticity, where he proposed the as a of components to describe the in solids. Green's approach utilized the principle of to express the internal forces as the differential of a function dependent on the strains, revolutionizing the formulation of constitutive relations. In SI units, the has dimensions of per unit , measured in joules per cubic meter (J/m³). The total elastic U stored in a material body is obtained by integrating the over the body's reference : U = \int_V W \, dV. This provides the overall associated with the deformation, serving as a basis for variational principles in elasticity.

Physical Interpretation

The strain energy density function, often denoted as W, quantifies the potential stored per unit reference volume within a deformable as a result of deformation. It embodies the work expended by internal forces to achieve reversible straining, which is subsequently released as the material returns to its undeformed configuration in purely responses. This stored arises from the reconfiguration of or molecular bonds under load, maintaining without permanent alteration to the material's structure. In hyperelastic materials, W functions as a , contingent exclusively on the instantaneous deformation tensor and independent of the deformation history or loading trajectory. This path independence ensures that the total stored energy remains consistent for any sequence of deformations leading to the same final , underpinning the conservative nature of hyperelastic behavior. Such a property facilitates the modeling of materials that exhibit fully recoverable deformations under arbitrary loading paths. Conceptually, [W](/page/W) behaves like the of a coiled for strains, where accumulates quadratically with increasing deformation, reflecting a linear stress-strain relationship. However, under finite strains typical of large deformations, this accumulation turns nonlinear, allowing [W](/page/W) to capture phenomena such as stiffening or softening that deviate from simple storage. This transition highlights the function's adaptability to both modest distortions and extreme material excursions. Distinct from , which stems from the motion of material particles, W is exclusively potential in character, tied to static configurational changes in solids and inherently free of dissipative losses like those in plastic flow or viscous . This exclusivity positions W as a for analyzing reversible cycles in materials devoid of .

Mathematical Formulation

Infinitesimal Strain Case

In the infinitesimal strain case, the strain energy density function is developed under the assumptions of small deformations in theory, where the magnitude of the strain tensor satisfies |ε| << 1, allowing the neglect of higher-order terms in the expansion of the deformation gradient. This approximation simplifies the kinematics and constitutive relations, enabling a direct connection to classical mechanics principles. The appropriate strain measure is the infinitesimal (or Cauchy) strain tensor, defined as \boldsymbol{\varepsilon} = \frac{1}{2} \left( \nabla \mathbf{u} + (\nabla \mathbf{u})^T \right), where \mathbf{u} is the displacement field and \nabla denotes the gradient operator. This symmetric second-order tensor captures the linear approximation of the deformation, focusing on changes in length and angle without considering rotations explicitly. For hyperelastic materials in this regime, the strain energy density function takes a quadratic form W(\boldsymbol{\varepsilon}) = \frac{1}{2} \boldsymbol{\varepsilon} : \mathbf{C} : \boldsymbol{\varepsilon}, where \mathbf{C} is the fourth-order elasticity (stiffness) tensor, and : denotes the double contraction (a scalar product between tensors). This expression represents the elastic potential energy stored per unit volume due to deformation, assuming path-independent loading and reversible behavior. The corresponding Cauchy stress tensor \boldsymbol{\sigma} is derived from the potential as the partial derivative \boldsymbol{\sigma} = \frac{\partial W}{\partial \boldsymbol{\varepsilon}} = \mathbf{C} : \boldsymbol{\varepsilon}, which establishes the linear relation central to Hooke's law, linking stress directly to strain through the material's elastic properties encoded in \mathbf{C}. This formulation ensures that the stress vanishes when the strain is zero and supports the superposition principle for small perturbations.

Finite Strain Case

In the finite strain case, which addresses large deformations in nonlinear elasticity, the strain energy density function W is defined per unit volume in the reference configuration and serves as the core of hyperelastic constitutive models. This formulation accounts for geometric nonlinearities where traditional small-strain approximations fail, using objective measures that remain invariant under rigid body motions. The deformation gradient tensor \mathbf{F} = \frac{\partial \mathbf{x}}{\partial \mathbf{X}} describes the mapping from the reference configuration position \mathbf{X} to the current configuration position \mathbf{x}, with the assumption \det(\mathbf{F}) > 0 ensuring local invertibility and preservation of material orientation. Common objective strain measures derived from \mathbf{F} include the right Cauchy-Green deformation tensor \mathbf{C} = \mathbf{F}^T \mathbf{F} and the Green-Lagrange tensor \mathbf{E} = \frac{1}{2} (\mathbf{C} - \mathbf{I}), where \mathbf{I} is the identity tensor. The energy density W is typically expressed as a scalar of these measures, such as W = W(\mathbf{F}) or W = W(\mathbf{E}), capturing the stored without dependence on the deformation path for reversible processes. The stresses conjugate to these kinematic quantities are obtained through hyperelasticity relations: the first Piola-Kirchhoff stress tensor \mathbf{P} = \frac{\partial W}{\partial \mathbf{F}} relates nominal stress in the reference frame, while the second Piola-Kirchhoff stress tensor \mathbf{S} = \frac{\partial W}{\partial \mathbf{E}} = \mathbf{F}^{-1} \mathbf{P} provides a symmetric, objective measure pulled back to the reference configuration. A fundamental requirement is the principle of material objectivity (or frame-indifference), stipulating that W(\mathbf{Q} \mathbf{F}) = W(\mathbf{F}) for any proper orthogonal tensor \mathbf{Q} (i.e., \mathbf{Q}^T \mathbf{Q} = \mathbf{I}, \det(\mathbf{Q}) = 1), ensuring the energy is unaffected by superimposed rigid rotations. In the limit of small deformations where \|\mathbf{F} - \mathbf{I}\| \to 0, this finite strain framework reduces to the infinitesimal strain case.

Specific Models

Linear Elastic Materials

In linear elastic materials, the strain energy density function describes the stored per unit under small deformations, assuming a relationship between and strain. For isotropic materials, which exhibit uniform properties in all directions, the strain energy density W is expressed as W = \frac{\lambda}{2} (\operatorname{tr} \boldsymbol{\varepsilon})^2 + \mu \boldsymbol{\varepsilon} : \boldsymbol{\varepsilon}, where \boldsymbol{\varepsilon} is the infinitesimal strain tensor, \operatorname{tr} \boldsymbol{\varepsilon} is its (volumetric strain), : denotes the double contraction, \lambda is the first Lamé constant (related to ), and \mu is the second Lamé constant (). These Lamé constants connect to more commonly used parameters: E (stiffness under uniaxial tension) and \nu (lateral contraction ratio), via \lambda = \frac{E \nu}{(1 + \nu)(1 - 2\nu)}, \quad \mu = \frac{E}{2(1 + \nu)}. This form arises from the assumption of hyperelasticity in the linear regime, ensuring path-independent energy storage and symmetry in the -strain response. For anisotropic linear elastic materials, where properties vary with direction (as in or composites), the strain energy density takes a more general : W = \frac{1}{2} \boldsymbol{\varepsilon} : \mathbb{C} : \boldsymbol{\varepsilon}, with \mathbb{C} the fourth-order tensor relating \boldsymbol{\sigma} = \mathbb{C} : \boldsymbol{\varepsilon}. The tensor \mathbb{C} has 81 components in general, but symmetries of and (major and minor) reduce the number of independent constants to 21 for triclinic symmetry (lowest crystal symmetry, no rotational invariance). Higher symmetries further reduce this: for example, monoclinic symmetry (one mirror plane) yields 13 independent constants. To simplify computations, contracts the tensor into a 6×6 \mathbf{C}, mapping the six unique components (three normal, three ) to a , facilitating numerical implementation in analyses. The total strain energy U = \int_V W \, dV over the volume V governs through the principle of minimum , where the true field minimizes U subject to boundary conditions, subject to external loads. Stationarity condition \delta U = 0 (first variation) yields the Navier equations of in form for isotropic cases: \mu \nabla^2 \mathbf{u} + (\lambda + \mu) \nabla (\nabla \cdot \mathbf{u}) + \mathbf{f} = 0, with \mathbf{u} the and \mathbf{f} forces; analogous forms hold for anisotropic media via component-wise . This linear framework assumes strains and instantaneous elastic recovery, but it fails to capture behaviors under large deformations (where geometric nonlinearities dominate) or in viscoelastic materials (exhibiting time-dependent recovery).

Hyperelastic Materials

Hyperelastic materials are characterized by a constitutive response where the is derived from the of a scalar density function W with respect to the appropriate strain measure, ensuring a path-independent and conservative deformation process. In the context of finite strains, this framework posits that the material stores all work done during deformation as recoverable , with the obtained from partial derivatives of W with respect to the invariants of the deformation tensor, as developed in the general theory of large elastic deformations. For many hyperelastic materials, such as rubbers, an incompressible assumption is often adopted, where the strain energy density function depends on the first two principal invariants of the right Cauchy-Green deformation tensor \mathbf{C}, denoted W = W(I_1, I_2), with I_1 = \operatorname{tr} \mathbf{C} and I_2 = \frac{1}{2} \left( (\operatorname{tr} \mathbf{C})^2 - \operatorname{tr} (\mathbf{C}^2) \right). Volume preservation is enforced by the condition \det \mathbf{F} = 1, where \mathbf{F} is the deformation gradient, leading to an additional hydrostatic pressure term p in the expression, such that the Cauchy is \boldsymbol{\sigma} = -p \mathbf{I} + 2 \frac{\partial W}{\partial I_1} \mathbf{B} - 2 \frac{\partial W}{\partial I_2} \mathbf{B}^{-1}, with \mathbf{B} = \mathbf{F} \mathbf{F}^T the left Cauchy-Green tensor. A foundational phenomenological model is the Neo-Hookean form, which for compressible materials takes the strain energy density as W = \frac{\mu}{2} (I_1 - 3) + f(J), where \mu > 0 is the shear modulus, J = \det \mathbf{F}, and f(J) accounts for volumetric changes, often chosen as f(J) = \frac{\kappa}{2} (J - 1)^2 with bulk modulus \kappa \gg \mu to approximate near-incompressibility. This model extends the linear elastic response to moderate finite strains and is widely used for soft tissues and elastomers due to its simplicity and grounding in statistical mechanics of polymer networks. The Mooney-Rivlin model generalizes the Neo-Hookean by incorporating the second invariant, with the incompressible density given by W = C_{10} (I_1 - 3) + C_{01} (I_2 - 3), where C_{10} and C_{01} are material constants related to the initial by \mu = 2 (C_{10} + C_{01}). This form captures the stiffening behavior observed in rubbers under biaxial loading better than the Neo-Hookean, as validated through experiments on vulcanized rubber. For broader applicability to large deformations, the Ogden model expresses the strain energy in terms of the principal stretches \lambda_i (eigenvalues of \mathbf{V} = \sqrt{\mathbf{B}}) 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), where N is the number of terms (typically 1–3), and \mu_k, \alpha_k are fitted parameters ensuring \sum_k \mu_k \alpha_k = 2\mu > 0. This flexible polynomial form excels in reproducing complex stress-strain curves for materials like across multiaxial loading paths. These models are calibrated by fitting parameters to experimental data from controlled tests, such as uniaxial tension for initial modulus, equibiaxial extension for intermediate response, and simple shear for nonlinearity, minimizing errors in nominal versus stretch curves. For instance, Ogden parameters are often determined via least-squares optimization against combined uniaxial and biaxial datasets to ensure predictive accuracy beyond the calibration regime.

Properties and Constraints

Thermodynamic Requirements

The strain energy density function W must adhere to the to ensure consistency with the second law of thermodynamics in . For isothermal processes in solids, where \theta is constant, the takes the form \boldsymbol{\sigma} : \mathbf{D} - \rho \dot{\psi} \geq 0, with \psi as the specific (often identified with W / \rho for hyperelastic materials), \boldsymbol{\sigma} the , \mathbf{D} the rate-of-deformation tensor, and \rho the mass density. This reduces to zero for reversible , yielding the constitutive \boldsymbol{\sigma} = \rho \frac{\partial \psi}{\partial \mathbf{F}} \mathbf{F}^T, where \mathbf{F} is the deformation gradient, thereby linking directly to derivatives of W. Fundamental growth conditions on W guarantee that it represents a valid stored energy potential. Specifically, W(\mathbf{F}) \geq 0 for all physically admissible deformation gradients \mathbf{F} with J = \det \mathbf{F} > 0, and W = 0 only in the (undeformed) state corresponding to rigid-body motions. Additionally, the second of W at the state must form a tensor, ensuring local convexity and the of the material's tangent , which prevents unphysical negative responses under small deformations. In thermoelastic contexts, the strain energy density generalizes to W(\mathbf{F}, \theta) to account for effects, maintaining thermodynamic consistency via the extended Clausius-Duhem that includes and conduction terms. Seminal formulations, such as those by Chadwick and Seet, posit that W at a reference \theta_0 governs isothermal elasticity, with dependence introduced through coupled contributions; for instance, in finite thermoelasticity for isotropic materials, W(\mathbf{C}, \theta) = \hat{W}(\mathbf{C}, \theta_0) + -modulated terms involving the tensor, ensuring the s = -\partial \psi / \partial \theta \geq 0. A for small strains and changes is W(\boldsymbol{\varepsilon}, \theta) = W_{\text{elastic}}(\boldsymbol{\varepsilon}) - 3 K \alpha (\theta - \theta_0) \operatorname{tr}(\boldsymbol{\varepsilon}) + \frac{9}{2} K \alpha^2 (\theta - \theta_0)^2, where K is the and \alpha is the coefficient, capturing the reduction in stored energy due to thermal strains. The formulation of W typically corresponds to isothermal conditions, as it derives from the under constant temperature, which is appropriate for quasi-static loading where heat exchange maintains . In contrast, adiabatic processes (no heat exchange) yield higher effective moduli, with the adiabatic exceeding the isothermal one by the factor \gamma = C_p / C_v ( of specific heats at constant pressure and volume), reflecting faster wave propagation and stiffer response due to suppressed . This distinction arises because adiabatic elasticity involves conservation, altering the effective energy storage compared to the isothermal case used in standard definitions.

Material Stability Conditions

Material stability conditions impose mathematical restrictions on the strain energy density function W to ensure that the material response remains physically realistic, mechanically stable, and free from pathological behaviors such as spontaneous softening or loss of in solutions to boundary value problems. These conditions are essential for guaranteeing the well-posedness of the governing equations in both and finite strain theories, preventing unphysical phenomena that could arise from non-convex or otherwise ill-behaved energy functions. In particular, they address global and local , ensuring that the material resists arbitrary perturbations and supports proper wave propagation. In the infinitesimal strain regime, convexity of the strain energy density W(\varepsilon) with respect to the strain tensor \varepsilon is a fundamental requirement for global . Specifically, the \frac{\partial^2 W}{\partial \varepsilon \partial \varepsilon} must be positive semi-definite, i.e., \frac{\partial^2 W}{\partial \varepsilon_{ij} \partial \varepsilon_{kl}} \geq 0 for all components, which corresponds to the being positive definite. This condition ensures that the energy increases monotonically with deformation, preventing non-physical softening and guaranteeing unique minimizers for problems. Violation of convexity can lead to multiple states, undermining the material's under small perturbations. For linear elastic materials, this reduces to the standard of the tensor, a direct consequence of the of W. For finite strain hyperelasticity, polyconvexity provides a stronger and more appropriate stability criterion, as simple convexity in the deformation gradient \mathbf{F} is often insufficient due to the nonlinear geometry. A strain energy function W(\mathbf{F}) is polyconvex if it can be written as W(\mathbf{F}) = g(\mathbf{F}, \cof \mathbf{F}, \det \mathbf{F}), where g is a convex function of its arguments and satisfies appropriate growth conditions. This ensures the existence of global minimizers for the total energy functional in boundary value problems, even under large deformations, and implies rank-one convexity, which is necessary for local stability. Polyconvexity was established as a key condition for the mathematical well-posedness of nonlinear elasticity problems, particularly for compressible materials. The Baker-Ericksen inequalities represent another critical stability condition for isotropic hyperelastic materials, linking the strain energy derivatives to the monotonicity of principal stresses with respect to principal stretches. For materials where W = W(I_1, I_2, I_3) with invariants I_1, I_2, I_3 of the right Cauchy-Green tensor, the inequality \frac{\partial W}{\partial I_1} + I_1 \frac{\partial W}{\partial I_2} > 0 must hold, ensuring that greater principal stretches correspond to greater principal stresses in . This prevents anomalous behaviors where occurs in directions of extension, promoting realistic stress-stretch alignment and overall material integrity under multiaxial loading. Strong ellipticity is a local stability that ensures the hyperbolic nature of the dynamic equations and the ellipticity of the static equations, crucial for unique solutions and proper wave propagation. For a , it requires that the acoustic tensor be positive definite, specifically \mathbf{n} \cdot \left( \frac{\partial \mathbf{P}}{\partial \mathbf{F}} \right) \mathbf{n} > 0 for all unit vectors \mathbf{n} \neq \mathbf{0}, where \mathbf{P} is the first Piola-Kirchhoff tensor derived from W(\mathbf{F}). This maintains the material's ability to support incremental deformations without loss of hyperbolicity, particularly important in compressible isotropic models. Loss of strong ellipticity signals the onset of non-unique solutions and potential discontinuities in the deformation field. Failure to satisfy these stability conditions can precipitate critical material instabilities, such as under compression or strain localization in shear bands. In , loss of convexity or ellipticity allows to lower-energy deformed states, leading to sudden structural collapse even below limits. Similarly, violation of polyconvexity or strong ellipticity promotes localization, where deformation concentrates in narrow regions, resulting in shear banding or necking that accelerates failure in hyperelastic solids. These modes underscore the practical importance of enforcing stability conditions in constitutive modeling to predict and avoid catastrophic responses.

References

  1. [1]
    [PDF] Hyperelasticity 101 - OSTI.GOV
    A strain energy density function is used to define a hyperelastic material by postulating that the stress in the material can be obtained by taking the ...
  2. [2]
    [PDF] Introduction to Continuum Mechanics
    Another useful constitutive law in non-linear elasticity is defined by the strain energy function. ρ0ˇΨ(IC, IIC, IIIC) = µ. 2. (IC − 3) − µ ln J +. 1. 2λ(J ...
  3. [3]
    Applied Mechanics of Solids (A.F. Bower) Section 3.5: Hyperelastic ...
    If the strain energy density is a function of the Left Cauchy-Green deformation tensor B = F ⁡ F T the constitutive equation is automatically isotropic. To see ...
  4. [4]
    [PDF] A treatise on the mathematical theory of elasticity - HAL
    Apr 28, 2016 · The present volume contains the general mathematical theory of the elastic properties of the first class of bodies, and I· propose to treat the ...
  5. [5]
    EngArc - L - Strain-Energy Density - Engineering Archives
    Thus, if SI metric units are used, the strain-energy density is expressed in J/m3 or its multiples kJ/m3 and MJ/m3; if U.S. customary units are used, it is ...
  6. [6]
    Strain Energy in a Continuum - Engineering at Alberta Courses
    The increment in the deformation energy stored in a continuum arises naturally from the product of the components of the stress tensor and the components of ...Deformation Energy of a... · Energy Conjugates · Illustrative Example 3...
  7. [7]
    [PDF] Strain Energy in Linear Elastic Solids - Duke People
    Strain Energy in Linear Elastic Solids. 11. Summary. Strain energy is a kind of potential energy arising from the deformation of elastic solids. For structural ...
  8. [8]
    3.2 Linear Elastic Material Behavior - Applied Mechanics of Solids
    Based on these observations, we define the strain energy density of a solid as the work done per unit volume to deform a material from a stress free reference ...
  9. [9]
    Systematic Fitting and Comparison of Hyperelastic Continuum ...
    Jan 9, 2023 · Thus, a path independent mechanical work is guaranteed and the ... hyperelastic strain energy functions for near-incompressibility. Int ...
  10. [10]
    The Feynman Lectures on Physics Vol. II Ch. 39: Elastic Materials
    When a crystal is strained, its volume or its shape is changed. Such changes result in an increase in the potential energy of the crystal. To calculate the ...
  11. [11]
    [PDF] 2D/3D Elasticity - The deformation map - cs.wisc.edu
    Deformation Energy ( E ) [also known as strain energy] : Potential energy stored in elastic body, as a result of deformation. Energy density ( Ψ ) : Ratio ...
  12. [12]
    [PDF] Introduction to Linear Elasticity - UCI Mathematics
    Mar 25, 2024 · This notes introduces the theory of linear elasticity, which studies the de- formation of elastic solid bodies under external forces. The ...
  13. [13]
    Theory of Elasticity - Stephen Timoshenko, James Norman Goodier
    Theory of Elasticity, Front Cover, Stephen Timoshenko, James Norman Goodier, McGraw-Hill, 1969 - Science - 567 pages
  14. [14]
    Large elastic deformations of isotropic materials. I. Fundamental ...
    Rivlin R (1997) Large Elastic Deformations Collected Papers of R.S. Rivlin ... Rigbi Z (1969) A strain energy function for swollen crosslinked elastomers ...
  15. [15]
    [PDF] Linear elasticity
    Here λ and µ are material constants, called elastic moduli or Lamé ... must be interpreted as the elastic energy density in a deformed isotropic material.
  16. [16]
    Elastic Constants | Materials Project Documentation
    Mar 7, 2023 · The elastic tensor in Voigt notation is a 6 × 6 6\times6 6×6 symmetric matrix, indicating that the elastic tensor has 21 independent components.
  17. [17]
    Large elastic deformations of isotropic materials IV. further ... - Journals
    The equations of motion, boundary conditions and stress-strain relations for a highly elastic material can be expressed in terms of the stored-energy function.
  18. [18]
    Large elastic deformations of isotropic materials VII. Experiments on ...
    Large elastic deformations of isotropic materials VII. Experiments on the deformation of rubber. R. S. Rivlin.
  19. [19]
    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 ...
  20. [20]
    Large deformation isotropic elasticity: on the correlation of theory ...
    Large deformation isotropic elasticity: on the correlation of theory and experiment for compressible rubberlike solids. Raymond William Ogden.
  21. [21]
    [PDF] Improved Methods for Hyperelastic Material Characterizations ...
    In general practice, material constants determined from uniaxial tensile tests are used for establishing the constitutive equations for hyperelastic materials.