Graphical Representations of the Regions of Rank-One-Convexity of some Strain Energies

Similar documents
On the Rank 1 Convexity of Stored Energy Functions of Physically Linear Stress-Strain Relations

2.1 Strain energy functions for incompressible materials

Numerical Properties of Spherical and Cubical Representative Volume Elements with Different Boundary Conditions

In recent years anisotropic materials have been finding their way into aerospace applications traditionally

Existence of minimizers for the pure displacement problem in nonlinear elasticity

On the Development of Volumetric Strain Energy Functions

International Journal of Pure and Applied Mathematics Volume 58 No ,

Lectures on. Constitutive Modelling of Arteries. Ray Ogden

The Class of Simo & Pister-type Hyperelasticity Relations

Conditions for Strong Ellipticity of Anisotropic Elastic Materials

DEVELOPMENT OF A CONTINUUM PLASTICITY MODEL FOR THE COMMERCIAL FINITE ELEMENT CODE ABAQUS

Inverse Design (and a lightweight introduction to the Finite Element Method) Stelian Coros

Continuum Mechanics and the Finite Element Method

MODELING OF ELASTO-PLASTIC MATERIALS IN FINITE ELEMENT METHOD

Two-dimensional examples of rank-one convex functions that are not quasiconvex

Measurement of deformation. Measurement of elastic force. Constitutive law. Finite element method

AN ANISOTROPIC PSEUDO-ELASTIC MODEL FOR THE MULLINS EFFECT IN ARTERIAL TISSUE

Constitutive Equations

Mechanical Properties of Polymer Rubber Materials Based on a New Constitutive Model

An O(n) invariant rank 1 convex function that is not polyconvex

Constitutive models. Constitutive model: determines P in terms of deformation

Lecture #6: 3D Rate-independent Plasticity (cont.) Pressure-dependent plasticity

Mathematical Background

Natural States and Symmetry Properties of. Two-Dimensional Ciarlet-Mooney-Rivlin. Nonlinear Constitutive Models

Constitutive models: Incremental plasticity Drücker s postulate

This introductory chapter presents some basic concepts of continuum mechanics, symbols and notations for future reference.

The Finite Element Method for the Analysis of Non-Linear and Dynamic Systems. Prof. Dr. Eleni Chatzi Lecture ST1-19 November, 2015

Two-Phase Deformation Fields and Ellipticity of Nonlinear Elastic Materials

Continuum mechanics V. Constitutive equations. 1. Constitutive equation: definition and basic axioms

Elements of Continuum Elasticity. David M. Parks Mechanics and Materials II February 25, 2004

EXPERIMENTAL IDENTIFICATION OF HYPERELASTIC MATERIAL PARAMETERS FOR CALCULATIONS BY THE FINITE ELEMENT METHOD

Plane Strain Test for Metal Sheet Characterization

Classification of Prostate Cancer Grades and T-Stages based on Tissue Elasticity Using Medical Image Analysis. Supplementary Document

Duality method in limit analysis problem of non-linear elasticity

Nonlinear Problems of Elasticity

Symmetry Properties of Two-Dimensional Ciarlet-Mooney-Rivlin Constitutive Models in Nonlinear Elastodynamics

ELASTOPLASTICITY THEORY by V. A. Lubarda

Understand basic stress-strain response of engineering materials.

Course No: (1 st version: for graduate students) Course Name: Continuum Mechanics Offered by: Chyanbin Hwu

arxiv: v1 [math.ap] 20 Jun 2018

The Non-Linear Field Theories of Mechanics

Comparative Study of Variation of Mooney- Rivlin Hyperelastic Material Models under Uniaxial Tensile Loading

HANDBUCH DER PHYSIK HERAUSGEGEBEN VON S. FLÜGGE. BAND VIa/2 FESTKÖRPERMECHANIK II BANDHERAUSGEBER C.TRUESDELL MIT 25 FIGUREN

(MPa) compute (a) The traction vector acting on an internal material plane with normal n ( e1 e

Fundamentals of Linear Elasticity

DRIVING FORCE IN SIMULATION OF PHASE TRANSITION FRONT PROPAGATION

in this web service Cambridge University Press

Engineering Sciences 241 Advanced Elasticity, Spring Distributed Thursday 8 February.

Spline-Based Hyperelasticity for Transversely Isotropic Incompressible Materials

The elastic behavior of a rubber-like material for composite glass

KINEMATICS OF CONTINUA

Iranian Journal of Mathematical Sciences and Informatics Vol.2, No.2 (2007), pp 1-16

Mechanics of materials Lecture 4 Strain and deformation

Course Syllabus: Continuum Mechanics - ME 212A

The Finite Element Method II

On Constitutive Models for Limited Elastic, Molecular Based Materials

Exact Solutions in Finite Compressible Elasticity via the Complementary Energy Function

2 Kinematics of Deformation

Simulation of Thermomechanical Couplings of Viscoelastic Materials

Linear Cosserat elasticity, conformal curvature and bounded stiffness

XI. NANOMECHANICS OF GRAPHENE

MITOCW MITRES2_002S10nonlinear_lec15_300k-mp4

Constitutive Relations

A note on finite elastic deformations of fibre-reinforced non-linearly elastic tubes

A FAILURE CRITERION FOR POLYMERS AND SOFT BIOLOGICAL MATERIALS

Exercise: concepts from chapter 8

HIGHLY ADAPTABLE RUBBER ISOLATION SYSTEMS

Elastic Stress Ratchetting and Corotational Stress Rates

Other state variables include the temperature, θ, and the entropy, S, which are defined below.

POLITECNICO DI MILANO

Creasing Critical Strain Dependence on Surface Defect Geometry. EN234 Final Project

You may not start to read the questions printed on the subsequent pages until instructed to do so by the Invigilator.

circular voids On the interaction between two in a nonlinear elastic solid ACTA MECHANICA J. P. Zhang and R. C. Batra, Rolla, Missouri

ruhr.pad A polyconvex extension of the logarithmic Hencky strain energy R.J. Martin, I.D. Ghiba and P. Neff Preprint

Comparison of Models for Finite Plasticity

20. Rheology & Linear Elasticity

Regularity of Minimizers in Nonlinear Elasticity the Case of a One-Well Problem in Nonlinear Elasticity

Modelling Anisotropic, Hyperelastic Materials in ABAQUS

On some exponential decay estimates for porous elastic cylinders

The Exponentiated Hencky-Logarithmic Strain Energy. Part I: Constitutive Issues and Rank-One Convexity

NUMERICAL SIMULATION OF THE INELASTIC SEISMIC RESPONSE OF RC STRUCTURES WITH ENERGY DISSIPATORS

A new closed-form model for isotropic elastic sphere including new solutions for the free vibrations problem

An Exact Result for the Macroscopic Response of Particle-Reinforced Neo-Hookean Solids

CONSTITUTIVE RELATIONS FOR LINEAR ELASTIC SOLIDS

ANSYS Mechanical Basic Structural Nonlinearities

Microbuckling instability in elastomeric cellular solids

Combined Isotropic-Kinematic Hardening Laws with Anisotropic Back-stress Evolution for Orthotropic Fiber-Reinforced Composites

Stability of Thick Spherical Shells

Chapter 2. Rubber Elasticity:

Analytical formulation of Modified Upper Bound theorem

Slight compressibility and sensitivity to changes in Poisson s ratio

Nonlinear Structural Materials Module

On the Numerical Modelling of Orthotropic Large Strain Elastoplasticity

Linear Constitutive Relations in Isotropic Finite Viscoelasticity

Hyperelasticity and the Failure of Averages

Discrete Analysis for Plate Bending Problems by Using Hybrid-type Penalty Method

Upon Impact Numerical Modeling of Foam Materials

Stress, Strain, Mohr s Circle

The exponentiated Hencky-logarithmic strain energy. Part I: Constitutive issues and rank one convexity arxiv:submit/ [math.

Shock enhancement of cellular structures under impact loading: Part II Analysis

Transcription:

TECHNISCHE MECHANIK, 32, 2-5, (22), 227 237 submitted: November, 2 Graphical Representations of the Regions of Rank-One-Convexity of some Strain Energies R. Glüge, J. Kalisch Isotropic elastic energies which are quadratic in the strain measures of the Seth family are known not to be rankone-convex in the entire domain of invertible deformation gradients with positive determinant. Therefore, they are in principle capable of displaying a laminated microstructure. Nevertheless, they are commonly used for standard elastic solids. In general one does not observe a microstructure evolution due to the fact that the solution is not sought outside of the region of rank-one-convexity. Consequently, the question for the boundaries of the region of rank-one-convexity arises. We address this question by applying a set of necessary and sufficient conditions for rank-one-convexity to the mentioned elastic energies, and give graphical representations for the regions of rank-one-convexity. Introduction The modelling of elasticity of solids has reached a high level of sophistication, coming a long way from Hooke s observation that the force is proportional to the strain (66) and Green s existence theorem (839, Ferrers (87)) of an elastic energy w. The latter is the starting point for most of the approaches to elasticity. However, several restrictions have to be imposed on w, both from purely mathematical and physical considerations. Starting from the theory of hyperelastic simple materials, i.e.,w depends only on the deformation gradientf, it is imposed that w has to be quasiconvex inf (Morrey, 952), w has satisfy growth conditions, also known as coercivity, w should depend only through C = F T F = U 2 on F, in order to make w independent of rotations (material objectivity), w should be positive for any deformation. The first two conditions are necessary for the existence of a unique solution of the elastostatic boundary value problem. The third and fourth condition are imposed due to physical considerations. Unfortunately, it is very hard to change the order in which the constraints are imposed, since the linearity of F in the position vector x is an important property needed for the examination of the existence and uniqueness of a solution to the elastostatic boundary value problem. Since it is difficult to verify the quasiconvexity of an elastic energy, other notions of convexity have been examined, e.g., by Ball (977). Polyconvexity and rank-one-convexity proved to be more practical. Unlike quasiconvexity, a function w(f) can be checked pointwise for specific F for poly- and rank-one-convexity. Moreover, both appear to be very close to quasiconvexity, where the implications polyconvexity quasiconvexity rank-one-convexity hold. Rank-one-convexity is interesting for two reasons: It is related to the kinematic compatibility condition. The deformation gradient can only undergo rank-one jumps, i.e., F + F = ab, a,b R 3 \{}. () 227

This ensures that, at singular surfaces with the normal vectorb, the body does not open or overlap with itself. Poly- and quasiconvexity can also be linked to physical notions, namely to different stability criteria (Ball and Marsden, 984). However, the kinematic compatibility condition is more elementary. It appears to be very close to quasiconvexity. There are examples of elastic energies which are rank-oneconvex but not quasiconvex, firstly established by Sverak (992), but these strain energies are not rotationally invariant. It is still unknown whether a rotational invariant rank-one-convex strain energy which is not quasiconvex exists (Šilhavý, 22). The failure of polyconvexity has been demonstrated for the isotropic St. Venant-Kirchhoff elastic energy (Raoult, 986), and for isotropic linear stress strain relations based on the logarithmic strains by Bruhns et al. (2). Bruhns et al. also gave lower bounds for the limits of the region of rank-one-convexity and state that, for modelling phenomena with discontinuous deformation gradients, it may be important to know the exact locations of these limits. In Bertram et al. (27), all energy functions that are quadratic in Seth strain measures are shown to violate rank-one-convexity for one deformation state or another. However, since these states usually lie outside the domain within which a solution is sought, the functions are applied without problems in commercial finite element software. For example the FE system ABAQUS 6.7- employs an elastic law relating the Cauchy stresses linearly to the logarithmic strains in the large strain setting. (Although this information is not given in the documentation, it can be found easily in a uniaxial tension test.) Thus, before applying these laws to large deformations one should ask for the limit of the region of rank-one-convexity. Notation. Vectors are symbolized by lowercase bold letters, second order tensors by uppercase bold letters. The tensor product ab is defined by (ab) c = (b c)a. The dot represents a simple contraction, the number of dots corresponds to the number of contractions. Indices larger than 3 have to be taken modulo 3. The deformation gradient and its polar decomposition are F = R U, U = C, C = F T F. The eigenvalues of U, resp. the singular values of F, are denoted by,2,3. The strain energy is denoted by w, the indexing of w corresponds to the partial derivatives w.r.t.,2,3. 2 Necessary and Sufficient Conditions for Rank-One-Convexity Unfortunately, the full set of necessary and sufficient conditions for rank-one-convexity is known only for isotropic (Rosakis, 99; Šilhavý, 999; Dacorogna, 2) and incompressible isotropic (Zee and Sternberg, 983) strain energies. The starting point for the derivation is (ab) 2 w F 2 (ab) for all a R3,b R 3. (2) Further, in the case of incompressibility, a and b must obey b F a = (3) due to the fact that the jump of F at a singular surface must be volume preserving. Limiting the domain of a or b by the latter inequality, the restrictions for rank-one-convexity are less strict in the case of incompressibility. Due to the isotropy, w depends only on the singular values of F. Eliminating the quantifiers in eq. (2) is a laborious work. One way of doing this is normalize a and b determine 2 w/ F 2 at F = U (the substitutiona = R ã is always possible) carry out the contractions withb, which gives the acoustic tensor A assure positive definiteness of A The last step can be done by assuring the positivity of the principle minors. In doing so, one arrives at the following set of inequalities: The first principle minor gives the separate convexity conditions < w ii, (no sum) (4) 228

< 3 < 4 < /2 < < 2 / 3 < > / 2 > < 2 / 3 < Figure : Scheme for graphical representation. For later reference, some characteristic strain-driven tests are depicted, namely a simple shear test (dashed), isochoric extension/compression (dotted), uniaxial extension/compression without lateral straining (dash-dot), pure volume change (solid line). and the Baker-Ericksen-inequalities < w i i w j j i j, i j. (5) The second principle minor gives inequalities involving mixed derivatives <m ± ij + w ii w jj, i j (6) m ± ij =±w ij + w i ±w j i ± j. (7) The third principle minor gives inequalities involving mixed derivatives <m ± 2 w33 +m ± 3 w22 +m ± 23 w + w w 22 w 33, (8) which must be evaluated for the combinations {m + 2,m+ 3,m+ 23 }, {m 2,m+ 3,m 23 }, {m+ 2,m 3,m 23 }, and {m 2,m 3,m+ 23 }. These results can be found in Dacorogna (2). For the incompressible case, a similar procedure is possible. However, one has to take the derivative with respect tof on the subspace det(f) =, and incorporate eq. (3). 3 Graphical Representations of the Regions of Rank-One-Convexity Using the inequalities (4, 5, 6, 8), we are able to determine the regions of rank-one-convexity. w(, 2, 3 ) is symmetric w.r.t. index permutations. Therefore, the plotting range can be restricted by the ordering 2 3, which is achieved by introducing the parameters q = / 2 and q 2 = 2 / 3. The graphical representations are given in terms of q, q 2, 3 They have been constructed with the help of a computer algebra system. The figure layout is the same for all graphics. On the left, a 3D-representation is given, where the non-region of rank-one-convexity is filled. The lower back corner corresponds to the origin of the coordinates q, q 2, 3, and the upper front corner to q =, q 2 =, 3 = 4, see Fig.. Additionally, the region of relatively small strains i ( i) 2 < 2 is represented. On the right, a contour plot of 3 over 2 / 3 (the horizontal direction) and / 2 (the vertical direction) is given, which is more suitable for extracting data. It corresponds to a projection of the rank-one-convex boundary into theq q 2 plane. 3. Elastic Energies which are Quadratic in Seth Strain Measures In the case of a linear and isotropic stress strain relation, the elastic law and hence the elastic energy are completely determined by the choice of the strain measure and two elasticity constants. The elastic strain energy is then given 229

by w = ( µ+ ) (E 2 +E2 2 +E 2 2 3)+(E E 2 +E 2 E 3 +E E 3 ) (9) or normalized w = E 2 +E 2 2 +E 2 3 +β(e E 2 +E 2 E 3 +E E 3 ) β := µ+ 2 () with Lamé s constants,µ and the eigenvalues E i of the strain tensor. Lamé s constants are related to the bulkand shear modulus K and G by µ = G = K 2 3 G. () The Legendre-Hadamard condition is ensured by K >, G >, which allows to determine the range for valid β, < β < 2. (2) One can writeβ as a function of Poisson s ratioν β = 2ν ν (3) where < ν < 2 is admissible. Usually, the strains are defined as isotropic functions of the material stretching E = i E m ( i )u i u i. (4) The function E() defines the strain measure. Typically one uses a Seth strain, for which E m () := m (m ), m, (5) E () := ln(), (6) hold. For example, m = 2 corresponds to the Piola-Almansi strain. In the -dimensional homogeneous case, m = corresponds to ε = l/l, i.e., one relates the length change to the actual (true) length. The case m = corresponds to the logarithmic (Hencky) strain which is also referred to as the true strain. In the -dimensional homogeneous case, m = corresponds to ε = l/l, i.e., one relates the length change to the reference length. Thus, it is referred to as the nominal strain, engineering strain or Biot strain. The case m = 2 leads to the the Green, Lagrange or Finger strain. In Figs. 2 to 4, the regions of rank-one-convexity for m = 2,,,,2 and ν =,.,,.3,,.5 are plotted. It is found in general that for ν.5, the limit of the region of rank-one-convexity approaches the unstretched state = 2 = 3 =, independent of m. Further, for m, one can impose arbitrarily large compressive strains without violating rank-one-convexity, while for < m, one can impose an arbitrarily large volume increase without violating rank-one-convexity. The cases m =, m = and m = 2 are most interesting, and examined in detail. It is found that for m =, eq. 5 imposes no restrictions (the right hand side is positive for all strains and approaches as ν.5). eq. 8 is globally weaker than eq. 4 and eq. 6 together. for m =, eq. 4 imposes no restrictions (the right hand side is just 2). Eq. eq. 5 and eq. 6 restrict the very same domain, which is given by < β( i 2)+2( i+ + i+2 ). The latter approaches eq. 8 as ν.5. Eq. 8 is globally weaker than eq. 5 (resp. eq. 6). for m = 2, none of eq. 4, eq. 5 or eq. 6 includes the other one globally. However, they include together eq. 6. 23

.29.5.7..99.3 8.77.4.9 6.55.78 5 4.33 2.52.39 6..3 (a) m = 2, ν =..7.4.5.26.58.35.2.2.98.5.9 4.7.75.56 2 5 8.3.4.5 (c) m = 2, ν = 2.38.89 2.49.98 2. 2.2.68.47.76.54.32.26.5. 4 3 2 8 6 4 (d) m =, ν = (b) m =, ν =. (e) m = 2, ν = 2 (f) m =, ν = Figure 2: Regions of rank-one-convexity for m = 2, m = and ν =., ν =, ν =. 23

2.72 3.4 2.4 2.6 2.8 2.52.92.68 2.24.96.44.2.68.4.2.96.72 8 4.56 4 8 (a) m =, ν = (b) m =, ν = 3.49 3. 2.79 3.6 3.24 2.48 2.7 2.88 2.52.86.55 2.6.8.24.93 2.44.8.72.3 (c) m =, ν =.3.36 (d) m =, ν = Figure 3: Regions of rank-one-convexity for m = and ν =., ν =, ν =.3, ν =. 3.2 Ciarlet-Geymonat Strain Energy The Ciarlet-Geymonat strain energy (Ciarlet and Geymonat, 982) is conveniently denoted in terms of principal invariants of B, wcg = µ (IIIB lniiib ) + (IB lniiib 3) 4 2 (7) with IIIB = ( 2 3 )2 and IB = 2 + 22 + 23 and Lam e s constants (Simo and Hughes, 998;?). The original function contains six material parameters (Ciarlet (988), Exercise 23), which are reduced by approaching a linearly elastic isotropic material with and µ as F I. It is derived from the family of Ogden materials, but with the restriction to be polyconvex for positive values of and µ (Ciarlet (988), page 85). Thus, wcg is rank-one-convex in the entire domain of positive, 2, 3 if ν. We can normalize wcg by multiplying with 2/µ, where we are able to replace ν =. 2µ 2ν (8) This allows, similar to the preceding section, to parametrize the range of valid material parameters by ν on the normalized strain energy. For negative ν, the region of rank-one-convexity is bounded. It is smallest for ν =, for which a graphical representation is given in Fig. 5(a). 3.3 Blatz-Ko Strain Energy The strain energy of Blatz and Ko is of the form µ 2µ 2µ µ( f ) µf µ IB 3 + IB 3 + (IIIB ) + (IIIB ), wblatz Ko = 2 2 232 (9)

3.48 3.63 3.9 2.97 2.9 2.64 2.6 3.3 2.32 2.3.98 2.3.74.65.32.99.45.6 7 6.5 (a) m =, ν =. 2.45 3.6 3.3 2.24 2. 3 2.7.96.82 2.4 2..68.54.8.4.5.26.2.2.9 7.98 (c) m =, ν = 3.5.87 3.22 2.99.785.7.65 2.76 2.53.53.445 2.3 2.7.36.275.84.6.38.9.5.2.5 2 (d) m = 2, ν =.7 (b) m = 2, ν =. 7 (e) m =, ν = (f) m = 2, ν = Figure 4: Regions of rank-one-convexity for m =, m = 2 and ν =., ν =, ν =. 233.94

3.68 3.45 3.22 2.99 2.76 2.53 2.3 2.7.84.6.45 3.6 3.24 2.88 2.52 2.6.8.44.8.72.36 (a) Ciarlet-Geymonat strain energy at ν = (b) Reduced Blatz-Ko-strain energy (c) Figure 5: Region of rank-one-convexity of the Ciarlet-Geymonat strain energy at ν = (a), the reduced Blatz-Ko strain energy (b) and Attard s strain energy (c) for the material parameters of eq. (22). with I B = 2 + 2 2 + 2 3. Compared to its original form (Blatz and Ko (962), eq. 5) the expression ( 2ν)/ν is replaced by 2µ/ (eq. 8) and J 3 is replaced by III B. By fixing f = and /µ = (ν = 5) it is reduced to wblatz Ko = µ 2 (I B +2 III B 5) = µ 2 ( 2 + 2 2 + 2 3 +2 2 3 5) (2) see Blatz and Ko (962) eq. (67). The strain energywblatz Ko is used to model polymeric foams (Horgan, 996). However, Horgan noted that strong ellipticity is lost in case of f <, i.e., wblatz Ko displays loss of rankone-convexity. In wblatz Ko, there is no material parameter which affects the region of rank-one-convexity, i.e., a single graph is sufficient (Fig.5(b)). It is found that wblatz Ko does not fail to be rank-one-convex under purely volumetric deformations. To be more specific, rank-one-convexity is lost only if i / j < 688. 3.4 Attard s Strain Energy Attard (23) proposed a strain energy of the form w Attard = N n= ( An 2n (I C n 3)+ B ) n 2n (I n C 3) + M m= C m 2m (lnj)2m lnj N (A n B n ). (2) The first sum represents a shape-change energy contribution, the second and third sum incorporate the dependence of w Attard on volume changes. Usually, nonlinear strain energies employ more than two material parameters, i.e., it is a tedious task to make a parameter-study of the regions of rank-one-convexity. Therefore, we examine the rank-one-convexity for only one of the resultant strain energies, obtained by adopting Attard s strain energy to a rubber material. The nonzero material parameters (Attard (23) eq. 69) are A =.36MPa B = 2MPa A 2 =.MPa C = 2MPa. (22) The region of rank-one-convexity is, unlike the other examples presented here, not representable by a single projection of the rank-one-convex boundary into the 2 / 3 / 2 plane (Fig. 5(c)). Characteristic points for the loss of rank-one-convexity are eigenvalues very close to zero and,2,3. 4 Failure of Rank-One-Convexity in Characteristic Tests n= For completeness, we have examined five characteristic tests, namely simple shear: = +γ 2 /2 γ 4+γ 2 /2, 2 =, 3 = +γ 2 /2+γ 4+γ 2 /2, uniaxial stress state: = u, 2 and 3 fromw 2 = w 3 =, isochoric tension/compression: = u, 2 = 3 = / u, simple tension/compression: = u, 2 = 3 =, dilatation: = 2 = 3 = u, 234

m ν Simple Shear Uniax. Stress Isochor. T/C Simple T/C Dilatation. γ < 96 36 < u <.29 57 < u <.289 36 < u <.29 < u <.29. γ <.55.53 < u <.296 26 < u <.33.394 < u <.29 < u <.243-2 γ < 4.569 < u <.3.578 < u <.43.35 < u <.286 < u <.2.3 γ < 6 2 < u <.332.56 < u <.675.3 < u <.265 < u <.43 γ <.29 46 < u <.364 32 < u < 2.337 7 < u <.85 < u <.69. γ <.768.33 < u <.5.548 < u <.494.33 < u <.5 < u <.5. γ < 27.39 < u <.5.529 < u <.559.32 < u <.5 < u <.49 - γ <.99 39 < u <.544.53 < u <.655 72 < u <.498 < u <.333.3 γ <.43 8 < u <.595 62 < u <.837 39 < u <.478 < u <.269 γ <.366.54 < u <.67.39 < u < 2.434 3 < u <.368 < u <.42. γ <.849.4 < u < 2.78.33 < u < 2.638.4 < u < 2.78 < u < 2.78. γ <.94.65 < u < 2.78 94 < u < 2.829.33 < u < 2.78 < u < 2.266 γ < 2.6.87 < u < 3.3 83 < u < 3.74.24 < u < 2.78 < u <.947.3 γ < 2.226 4 < u < 3.469 68 < u < 3.4.4 < u < 2.78 < u <.73 γ < 2.485 5 < u < 8 45 < u < 3.932.3 < u < 2.78 < u <.535. γ < < u < < u < < u <.5 < u <. γ < < u <. < u <.2 < u <.579 < u < γ < < u < 5.999 < u < 5 < u < 67 < u <.3 γ < < u < 333 < u < 29 < u <.765 < u < γ < < u < 3.5 < u < 67 < u < 76 < u <. γ <.2.586 < u <.578 < u < 2..586 < u <.78 < u <. γ <.592 < u <.53 < u < 3 < u <.76 < u < 2 γ <.594 < u < < u < 39 < u < 7 < u <.3 γ <.593 < u < < u <.76 < u < 75 < u < γ <.588 < u < < u < 25 < u <.936 < u < Blatz-Ko γ < 2.349 < u < 2.867 6 < u < 2.46 68 < u < 3.732 < u < Table : Rank-one-convex ranges of characteristic tests for the Hookean strain energies and the reduced Blatz-Ko strain energy. whereuandγ are process parameters. Note that the eigenvalues in the list above are not ordered. Table contains the rank-one-convex ranges for the characteristic tests for the Hookean strain energies and Blatz-Ko s reduced strain energy. 5 Summary We examined the strain energies which are quadratic in the generalized Seth strain measures. It has been found that in all cases, in the incompressible limit, the rank-one-convex boundary approaches the undeformed state, for Seth strains with negative exponent and the logarithmic strains, a compressive volume change does not lead to the loss of rank-one-convexity, for Seth strains with positive exponent, a volume expansion does not lead to the loss of rank-one-convexity. From the Hookean strain energies which have been examined, the one which is quadratic in the nominal strain (m = ) displayed the largest rank-one-convex ranges in characteristic tests. We also examined three compressible non-hookean strain energies, namely Ciarlet-Geymonat s, Blatz-Ko s and Attard s strain energy. We confirmed that the two-parameter Ciarlet-Geymonat strain energy has no rank-one-convex limit for positive Poisson s ratios. It is worth noting that Ciarlet-Geymonat s strain energy and the corresponding Cauchy stresses are given by σ = (µ(b I)+(J 2 )/2I)/J, i.e. no explicit calculation of U or V is required. 235

The reduced Blatz-Ko strain energy (with ν = 5) fails to be rank-one-convex if and only if i / j < 688, i.e., no rank-one-convex failure under purely dilatorical deformations is observed. In the characteristic tests, the rank-one-convex range of the Blatz-Ko strain energy turns out to be similar to those Hookean strain energies. No specific conclusions could be drawn for Attard s strain energy, except for that for more complicated strain energies the regions of rank-one-convexity may have a complex shape in the space of the singular values off. We observed that, out of the set of necessary and sufficient conditions that we used here, the eq. (8) are always weaker than the restrictions eq. (4, 5, 6) together, and thus had no impact for the strain energies under consideration. At most, this restriction coincided with one of the others. However, this is due to the specific strain energies under consideration, and does not hold in general. Acknowledgement We like to express our gratitude to Miroslav Šilhavý and Patrizio Neff for providing valuable remarks. References Attard, M.: Finite strain isotropic hyperelasticity. Int. J. Solid. Struct., 4, (23), 4353 4378. Ball, J.: Convexity conditions and existence theorems in nonlinear elasticity. Arch. Ration. Mech. An., 63, (977), 337 43. Ball, J.; Marsden, J.: Quasiconvexity at the boundary, positivity of the second variation and elastic stability. Arch. Ration. Mech. An., 86, 3, (984), 25 277. Bertram, A.: Elasticity and Plasticity of Large Deformations - an Introduction, Third Edition. Springer Verlag, Berlin (22). Bertram, A.; Böhlke, T.; Šilhavý, M.: On the rank convexity of stored energy functions of physically linear stress-strain relations. J. Elasticity, 86, (27), 235 243. Blatz, P.; Ko, W.: Application of Finite Elastic Theory to the Deformation of Rubbery Materials. Transaction of the Society of Rheology, 6, (962), 223 25. Bruhns, O.; Xiao, H.; Meyers, A.: Constitutive inequalities for an isotropic elastic strain-energy function based on Henckys logarithmic strain tensor. P. Roy. Soc. Lond. A Mat., 457, (2), 227 2226. Ciarlet, P.: Mathematical Elasticity Volume : Three-Dimensional Elasticity. Studies in Mathematics and its Applications, Volume 2, Elsevier Science Publishers B.V. (988). Ciarlet, P.; Geymonat, G.: Sur les lois de comportement en élasticité non linéaire compressible. C. R. Acad. Sci. Paris Sér. II Méc. Phys. Chim. Sci. Univers Sci. Terre, 295, 4, (982), 423 426. Dacorogna, B.: Necessary and sufficient conditions for strong ellipticity of isotropic functions in any dimensions. Discrete and Continuous Dynamical Systems Series B,, 2, (2), 257 263. Ferrers, M.: Mathematical Papers of the late George Green. Machmillan and Co., London (87). Horgan, C.: Remarks on Ellipticity for the Generalized Blatz-Ko Constitutive Model for a Compressible Nonlinearly Elastic Solid. J. Elasticity, 42, (996), 65 76. Morrey, C.: Quasi-convexity and the lower semicontinuity of multiple integrals. Pac. J. Math., 2,, (952), 25 53. Raoult, A.: Non-polyconvexity of the stored energy function of a Saint Venant-Kirchhoff material. Applications of Mathematics, 3, 6, (986), 47 49. Rosakis, P.: Ellipticity and deformations with discontinuous gradients in finite elastostatics. Arch. Ration. Mech. An., 9, (99), 37. Simo, J.; Hughes, T.: Computational Inelasticity. Springer-Verlag, New York (998). Sverak, V.: Rank once convexity does not imply quasiconvexity. Proceedings of the Royal Society of Edinburgh A, 2, (992), 85 89. 236

Šilhavý, M.: On isotropic rank convex functions. Proceedings of the Royal Society of Edinburgh A, 29, (999), 8 5. Šilhavý, M.: Monotonicity of rotationally invariant convex and rank convex functions. Proceedings of the Royal Society of Edinburgh A, 32, (22), 49 435. Zee, L.; Sternberg, E.: Ordinary and strong ellipticity in the equilibrium theory of incompressible hyperelastic solids. Arch. Ration. Mech. An., 83,, (983), 53 9. Address: Otto-von-Guericke-Universität Magdeburg, Universitätsplatz 2, D-396 email: gluege@ovgu.de, jan.kalisch@ovgu.de 237