Mathematical Foundations of

Similar documents
Existence of minimizers for the pure displacement problem in nonlinear elasticity

Progress and puzzles in nonlinear elasticity

Weak Convergence Methods for Energy Minimization

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

MAT 570 REAL ANALYSIS LECTURE NOTES. Contents. 1. Sets Functions Countability Axiom of choice Equivalence relations 9

3 2 6 Solve the initial value problem u ( t) 3. a- If A has eigenvalues λ =, λ = 1 and corresponding eigenvectors 1

2 BAISHENG YAN When L =,it is easily seen that the set K = coincides with the set of conformal matrices, that is, K = = fr j 0 R 2 SO(n)g: Weakly L-qu

3 (Due ). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

Weak convergence methods for nonlinear partial differential equations.

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

Riemann integral and volume are generalized to unbounded functions and sets. is an admissible set, and its volume is a Riemann integral, 1l E,

An introduction to Mathematical Theory of Control

2 (Bonus). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

PROBLEMS. (b) (Polarization Identity) Show that in any inner product space

1 Directional Derivatives and Differentiability

1 Topology Definition of a topology Basis (Base) of a topology The subspace topology & the product topology on X Y 3

2 A Model, Harmonic Map, Problem

Topological properties

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

Contents: 1. Minimization. 2. The theorem of Lions-Stampacchia for variational inequalities. 3. Γ -Convergence. 4. Duality mapping.

EXISTENCE OF SOLUTIONS TO ASYMPTOTICALLY PERIODIC SCHRÖDINGER EQUATIONS

Calculus of Variations. Final Examination

l(y j ) = 0 for all y j (1)

Lebesgue Integration on R n

THEOREMS, ETC., FOR MATH 515

Convex Analysis and Economic Theory Winter 2018

1/12/05: sec 3.1 and my article: How good is the Lebesgue measure?, Math. Intelligencer 11(2) (1989),

SYMMETRY RESULTS FOR PERTURBED PROBLEMS AND RELATED QUESTIONS. Massimo Grosi Filomena Pacella S. L. Yadava. 1. Introduction

REAL AND COMPLEX ANALYSIS

SPECTRAL PROPERTIES OF THE LAPLACIAN ON BOUNDED DOMAINS

Nonlinear elasticity and gels

Convex Analysis and Optimization Chapter 2 Solutions

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

Partial Differential Equations, 2nd Edition, L.C.Evans The Calculus of Variations

INVERSE FUNCTION THEOREM and SURFACES IN R n

(1) Consider the space S consisting of all continuous real-valued functions on the closed interval [0, 1]. For f, g S, define

Probability and Measure

Optimization and Optimal Control in Banach Spaces

Γ-convergence of functionals on divergence-free fields

Euler Equations: local existence

SHARP BOUNDARY TRACE INEQUALITIES. 1. Introduction

PEAT SEISMOLOGY Lecture 2: Continuum mechanics

MATH 51H Section 4. October 16, Recall what it means for a function between metric spaces to be continuous:

VISCOSITY SOLUTIONS. We follow Han and Lin, Elliptic Partial Differential Equations, 5.

Optimal Transportation. Nonlinear Partial Differential Equations

2) Let X be a compact space. Prove that the space C(X) of continuous real-valued functions is a complete metric space.

ENERGY-MINIMIZING INCOMPRESSIBLE NEMATIC ELASTOMERS

Coupled second order singular perturbations for phase transitions

A NUMERICAL ITERATIVE SCHEME FOR COMPUTING FINITE ORDER RANK-ONE CONVEX ENVELOPES

g 2 (x) (1/3)M 1 = (1/3)(2/3)M.

Analysis Qualifying Exam

Partial Differential Equations

We denote the derivative at x by DF (x) = L. With respect to the standard bases of R n and R m, DF (x) is simply the matrix of partial derivatives,

Regularity for Poisson Equation

Stanford Mathematics Department Math 205A Lecture Supplement #4 Borel Regular & Radon Measures

Topology, Math 581, Fall 2017 last updated: November 24, Topology 1, Math 581, Fall 2017: Notes and homework Krzysztof Chris Ciesielski

Finite-dimensional spaces. C n is the space of n-tuples x = (x 1,..., x n ) of complex numbers. It is a Hilbert space with the inner product

DEVELOPMENT OF MORSE THEORY

Lecture Notes in Advanced Calculus 1 (80315) Raz Kupferman Institute of Mathematics The Hebrew University

3. (a) What is a simple function? What is an integrable function? How is f dµ defined? Define it first

Some lecture notes for Math 6050E: PDEs, Fall 2016

REGULARITY FOR INFINITY HARMONIC FUNCTIONS IN TWO DIMENSIONS

Chapter 4. Inverse Function Theorem. 4.1 The Inverse Function Theorem

Everywhere differentiability of infinity harmonic functions

S chauder Theory. x 2. = log( x 1 + x 2 ) + 1 ( x 1 + x 2 ) 2. ( 5) x 1 + x 2 x 1 + x 2. 2 = 2 x 1. x 1 x 2. 1 x 1.

Global Maxwellians over All Space and Their Relation to Conserved Quantites of Classical Kinetic Equations

Sobolev Spaces. Chapter 10

If Y and Y 0 satisfy (1-2), then Y = Y 0 a.s.

Course Summary Math 211

Analysis-3 lecture schemes

Multivariable Calculus

L p Spaces and Convexity

Symmetric Matrices and Eigendecomposition

1 Continuity Classes C m (Ω)

Introduction to Functional Analysis

Chapter 4. Multiple Integrals. 1. Integrals on Rectangles

Math The Laplacian. 1 Green s Identities, Fundamental Solution

Global minimization. Chapter Upper and lower bounds

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

1 Math 241A-B Homework Problem List for F2015 and W2016

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

Laplace s Equation. Chapter Mean Value Formulas

Asymptotic behavior of infinity harmonic functions near an isolated singularity

Real Analysis Problems

MAT 257, Handout 13: December 5-7, 2011.

Cavitation and fracture in nonlinear elasticity

Relaxation methods and finite element schemes for the equations of visco-elastodynamics. Chiara Simeoni

Mathematical Tripos Part IA Lent Term Example Sheet 1. Calculate its tangent vector dr/du at each point and hence find its total length.

are Banach algebras. f(x)g(x) max Example 7.4. Similarly, A = L and A = l with the pointwise multiplication

Chapter 2. Metric Spaces. 2.1 Metric Spaces

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

Real Analysis Math 131AH Rudin, Chapter #1. Dominique Abdi

ON SELECTION OF SOLUTIONS TO VECTORIAL HAMILTON-JACOBI SYSTEM

Chapter 0. Preliminaries. 0.1 Things you should already know

Glimpses on functionals with general growth

Continuous Functions on Metric Spaces

van Rooij, Schikhof: A Second Course on Real Functions

Analysis Finite and Infinite Sets The Real Numbers The Cantor Set

GLOBAL LIPSCHITZ CONTINUITY FOR MINIMA OF DEGENERATE PROBLEMS

4 Constitutive Theory

Transcription:

Mathematical Foundations of Elasticity Theory John Ball Oxford Centre for Nonlinear PDE J.M.Ball

Reading Ciarlet Antman Silhavy

Lecture note 2/1998 Variationalmodels for microstructure and phase transitions Stefan Müller Prentice Hall 1983 http://www.mis.mpg.de/

J.M. Ball, Some open problems in elasticity. In Geometry, Mechanics, and Dynamics, pages 3--59, Springer, New York, 2002. You can download this from my webpage http://www.maths.ox.ac.uk/~ball

Elasticity Theory The central model of solid mechanics. Rubber, metals (and alloys), rock, wood, bone can all be modelled as elastic materials, even though their chemical compositions are very different. For example, metals and alloys are crystalline, with grains consisting of regular arrays of atoms. Polymers (such as rubber) consist of long chain molecules that are wriggling in thermal motion, often joined to each other by chemical bonds called crosslinks. Wood and bone have a cellular structure 6

A brief history 1678 Hooke's Law 1705 Jacob Bernoulli 1742 Daniel Bernoulli 1744 L. Euler elastica(elastic rod) 1821 Navier, special case of linear elasticity via molecular model (Dalton s atomic theory was 1807) 1822 Cauchy, stress, nonlinearand linear elasticity For a long time the nonlinear theory was ignored/forgotten. 1927 A.E.H. Love, Treatise on linear elasticity 1950's R. Rivlin, Exact solutions in incompressiblenonlinear elasticity (rubber) 1960 -- 80 Nonlinear theory clarified by J.L. Ericksen, C. Truesdell 1980 -- Mathematical developments, applications to materials, biology 7

Kinematics Label the material points of the body by the positions x Ω they occupy in the reference configuration. 8

Deformation gradient F =Dy(x,t), F iα = y i x α. 9

Invertibility To avoid interpenetration of matter, we require that for each t, y(,t) is invertible on Ω, with sufficiently smooth inverse x(, t). We also suppose that y(, t) is orientation preserving; hence J =detf(x,t)>0 for x Ω. (1) Bytheinversefunctiontheorem,ify(,t)isC 1, (1) implies that y(, t) is locally invertible. 10

11

Global inverse function theorem for C 1 deformations Let Ω R n be a bounded domain with Lipschitz boundary Ω (in particular Ω lies on one side of Ω locally). Let y C 1 ( Ω;R n ) with detdy(x)>0 for all x Ω and y Ω one-to-one. Then y is invertible on Ω. Proof uses degree theory. cf Meisters and Olech, Duke Math. J. 30 (1963) 63-80. 12

Notation M n n = {real n n matrices} M n n + = {F M n n :detf >0} SO(n) = {R M n n :R T R=1,detR=1} = {rotations}. If a R n, b R n, the tensor product a b is the matrix with the components (a b) ij =a i b j. [Thus (a b)c=(b c)a if c R n.]

Variationalformulation of nonlinear elastostatics We suppose for simplicity that the body is homogeneous, i.e. the material response is the sameateachpoint. Inthiscasethetotalelastic energy corresponding to the deformation y=y(x) is given by I(y)= W(Dy(x))dx, Ω wherew =W(F)isthestored-energyfunction ofthematerial. WesupposethatW :M+ 3 3 R is C 1 and bounded below, so that without loss of generality W 0.

We will study the existence/nonexistence of minimizers of I subject to suitable boundary conditions. Issues. 1. What function space should we seek a minimizer in? This controls the allowable singularities in deformations and is part of the mathematical model. 2. What boundary conditions should be specified? 3. What properties should we assume about W?

The Sobolevspace W 1,p W 1,p ={y:ω R 3 : y 1,p < }, where y 1,p = { ( Ω [ y(x) p + Dy(x) p ]dx) 1/p if 1 p< esssup x Ω ( y(x) + Dy(x) ) if p= Dy is interpreted in the weak(or distributional) sense, so that Ω for all ϕ C 0 (Ω). y i x α ϕdx= Ω y i ϕ i x α dx

WeassumethatybelongstothelargestSobolev space W 1,1 =W 1,1 (Ω;R 3 ), so that in particu- lar Dy(x) is well defined for a.e. x Ω, and I(y) [0, ].

y W 1, Lipschitz y W 1,1 (because every y W 1,1 is absolutely continuous on almost every line parallel to a given direction)

Cavitation y(x)= 1+ x x x y W 1,p for 1 p<3. Exercise: prove this.

Boundary conditions We suppose that Ω= Ω 1 Ω 2 N, where Ω 1, Ω 2 are disjoint relatively open subsets of Ω and N has two-dimensional Hausdorff measure H 2 (N)=0 (i.e. N has zero area). We suppose that y satisfies mixed boundary conditions of the form y Ω1 =ȳ( ), where ȳ : Ω 1 R 3 is a given boundary displacement.

By formally computing d dτ I(y+τϕ) τ=0= d dτ W(Dy+τDϕ)dx t=0=0, Ω we obtain the weak form of the Euler-Lagrange equation for I, that is D FW(Dy) Dϕdx=0 ( ) Ω for all smooth ϕ with ϕ Ω1 =0. T R (F)=D F W(F) is the Piola-Kirchhoff stress tensor. (A B=tr A T B)

If y, Ω 1 and Ω 2 are sufficiently regular then (*) is equivalent to the pointwise form of the equilibrium equations DivD F W(Dy)=0 in Ω, together with the natural boundary condition of zero applied traction D F W(Dy)N =0 on Ω 2, wheren =N(x)denotestheunitoutwardnormal to Ω at x.

Ω 1 Example: boundary conditions for buckling of a bar Ω 2 Ω2 Ω 1 1 ȳ(x)=x ȳ(x 1,x 2,x 3 )=(λx 1,x 2,x 3 ) y( Ω 2 ) traction free λ<1

Hence our variational problem becomes: Does there exist y minimizing in I(y)= Ω W(Dy)dx A={y W 1,1 :y invertible,y Ω1 =ȳ}? We will make the invertibility condition more precise later.

Properties of W To try to ensure that deformations are invertible, we suppose that (H1) W(F) as detf 0+. So as to also prevent orientation reversal we define W(F)= if detf 0. Then W :M 3 3 [0, ] is continuous.

Thus if I(y) < then detdy(x) > 0 a.e.. However this does not imply local invertibility (think of the map (r,θ) (r,2θ) in plane polar coordinates, which has constant positive Jacobian but is not invertible near the origin), nor is it clear that detdy(x) µ>0 a.e..

We suppose that W is frame-indifferent, i.e. (H2) W(RF)=W(F) for all R SO(3),F M 3 3. In order to analyze (H2) we need some linear algebra.

Square root theorem Let C be a positive symmetric n n matrix. Then there is a unique positive definite symmetric n n matrix U such that (we write U =C 1/2 ). C=U 2 28

Formula for the square root Since C is symmetric it has a spectral decomposition C= n i=1 λ i ê i ê i. Since C>0, it follows that λ i >0. Then U = n i=1 λ 1/2 i ê i ê i satisfies U 2 =C. 29

Polar decomposition theorem Let F M+ n n. Then there exist positive definite symmetric U, V and R SO(n) such that F =RU =VR. These representations (right and left respectively) are unique. 30

Proof. Suppose F =RU. Then U 2 =F T F := C. Thus if the right representation exists U must be the square root of C. But if a R n is nonzero, Ca a = Fa 2 > 0, since F is nonsingular. Hence C > 0. So by the square root theorem, U =C 1/2 exists and is unique. Let R=FU 1. Then R T R=U 1 F T FU 1 =1 and detr=detf(detu) 1 =+1. The representation F =VR 1 is obtained similarly using B:=FF T, and it remains to prove R=R 1. ButthisfollowsfromF =R 1 ( R T 1 VR 1 ), and the uniqueness of the right representation. 31

Exercise: simple shear y(x)=(x 1 +γx 2,x 2,x 3 ). F = cosψ sinψ 0 sinψ cosψ 0 0 0 1 cosψ sinψ 0 sinψ 1+sin2 ψ cosψ 0 0 0 1 tanψ = γ 2. As γ 0+ the eigenvectors of U and V tend to 1 (e 2 1 +e 2 ), 1 (e 2 1 e 2 ),e 3. 32,

Hence (H2) implies that W(F)=W(RU)=W(U)= W(C). Conversely if W(F)= W(U) or W(F)= W(C) then (H2) holds. Thus frame-indifference reduces the dependence ofw onthe9elementsoff tothe6elements of U or C.

Material Symmetry In addition, if the material has a nontrivial isotropy group S, W satisfies the material symmetry condition W(FQ)=W(F) for all Q S,F M 3 3 +. ThecaseS=SO(3)correspondstoanisotropic material.

The strictly positive eigenvalues v 1,v 2,v 3 of U (or V) are called the principal stretches. Proposition 1 W is isotropic iff W(F)=Φ(v 1,v 2,v 3 ), where Φ is symmetric with respect to permutations of the v i. Proof. Suppose W is isotropic. Then F = RDQforR,Q SO(3)andD=diag(v 1,v 2,v 3 ). Hence W = W(D). But for any permutation P of 1,2,3 there exists Q such that Qdiag(v 1,v 2,v 3 ) Q T =diag(v P1,v P2,v P3 ). The converse holds since Q T F T FQ has the same eigenvalues (namely v 2 i ) as FT F.

(H1) and (H2) are not sufficient to prove the existence of energy minimizers. We also need growth and convexity conditions on W. The growth condition will say something about how fastw growforlargevaluesoff. Theconvexity condition corresponds to a statement of the type stress increases with strain. We return tothequestionofwhatthecorrectformofthis convexity condition is later; for a summary of older thinking on this question see Truesdell& Noll.

Why minimize energy? This is the deep problem of the approach to equilibrium, having its origins in the Second Law of Thermodynamics. We will see how rather generally the balance of energy plus a statement of the Second Law lead to the existence of a Lyapunov function for the governing equations.

Balance of Energy ( 1 ) d dt E 2 ρ R y t 2 +U dx = b y tdx E + t R y t ds + E E rdx q R NdS, (1) E for all E Ω, where ρ R =ρ R (x) is the density in the reference configuration, U is the internal energy density, b is the body force, t R is the Piola-Kirchhoff stress vector, q R the reference heat flux vector and r the heat supply.

Second Law of Thermodynamics WeassumethisholdsintheformoftheClausius Duhem inequality d q R N r dt ηdx ds+ dx (2) E E θ Eθ for all E, where η is the entropy and θ the temperature.

The Ballistic Free Energy Suppose that the the mechanical boundary conditions are that y=y(x,t) satisfies y(,t) Ω1 = ȳ( ) and the condition that the applied traction on Ω 2 is zero, and that the thermal boundary condition is θ(,t) Ω3 =θ 0, q R N Ω\ Ω3 =0, where θ 0 >0 is a constant. Assume that the heat supply r is zero, and that the body force is given by b= grad y h(x,y),

Thusfrom(1),(2)withE=Ωandtheboundary conditions d dt Ω [ 1 Ω t R y t ds 2 ρ R y t 2 +U θ 0 η+h Ω Ω θ ( 1 θ 0 ) ] dx q R NdS = 0. So E = Ω[ 12 ρ R y t 2 +U θ 0 η+h ] dx is a Lyapunov function, and it is reasonable to suppose that typically (y t,y,θ) tends as t to a (local) minimizer of E.

For thermoelasticity, W(F, θ) can be identified with the Helmholtz free energy U(F,θ) θη(f, θ). Hence, if the dynamics and boundary conditions are such that as t we have y t 0 and θ θ 0, then this is close to saying that y tends to a local minimizer of I θ0 (y)= Ω [W(Dy,θ 0)+h(x,y)]dx. The calculation given follows work of Duhem, Ericksen and Coleman & Dill.

Of course a lot of work would be needed to justify this (we would need well-posedness of suitable dynamic equations plus information on asymptotic compactness of solutions and more; this is currently out of reach). Note that it is not the Helmoltz free energy that appears in the expression for E but U θ 0 η, where θ 0 is the boundary temperature. Forsomeremarksonthecasewhenθ 0 depends on x see J.M. Ball and G. Knowles, Lyapunov functions for thermoelasticity with spatially varying boundary temperatures. Arch. Rat. Mech. Anal., 92:193 204, 1986.

Existence in one dimension To make the problem nontrivial we consider an inhomogeneous one-dimensional elastic material with reference configuration Ω = (0, 1) and stored-energy function W(x, p), with corresponding total elastic energy 1 I(y)= [W(x,y x(x))+h(x,y(x))]dx, 0 where h is the potential energy of the body force.

We seek to minimize I in the set of admissible deformations A={y W 1,1 (0,1): y x (x)>0 a.e., y(0)=α,y(1)=β}, where α < β. (We could also consider mixed boundary conditions y(0) = α, y(1) free.) (Note the simple form taken by the invertibility condition.)

Hypotheses on W: We suppose for simplicity that W :[0,1] (0, ) [0, ) is continuous. (H1) W(x,p) as p 0+. As before we define W(x,p)= if p 0. (H2) is automatically satisfied in 1D. (H3) W(x,p) Ψ(p) for all p>0, x (0,1), where Ψ:(0, ) [0, ) is continuous with lim p Ψ(p) p =.

(H4) W(x,p) is convex in p, i.e. W(x,λp+(1 λ)q) λw(x,p)+(1 λ)w(x,q) for all p>0,q>0,λ (0,1),x (0,1). If W is C 1 in p then (H4) is equivalent to the stress W p (x,p) being nondecreasing in the strain p. (H5) h:[0,1] R [0, ) is continuous.

Theorem 1 Under the hypotheses (H1)-(H5) there exists y that minimizes I in A. Proof. A is nonempty since z(x)=α+(β α)x belongs to A. Since W 0,h 0, 0 l=inf y A I(y)<. Let y (j) be a minimizing sequence, i.e. y (j) A, I(y (j) ) l as j.

We may assume that lim j 10 W(x,y (j) x )dx=l 1, lim j 10 h(x,y (j) )dx=l 2, where l=l 1 +l 2. Since Ω Ψ(y(j) x )dx M <, by the de la Vallée Poussin criterion(see e.g. One-dimensional variational problems, G. Buttazzo, M. Giaquinta, S. Hildebrandt, OUP, 1998 p 77) there exists a subsequence, which we continue to call y (j) x converging weakly in L 1 (0,1) to some z.

Lety (x)=α+ x 0 z(s)ds,sothaty x=z. Then y (j) (x)=α+ x 0 y (j) x (s)ds y (x) for all x [0,1]. In particular y (0)=α, y (1)=β. By Mazur s theorem, there exists a sequence z (k) = j=k λ (k) j y x (j) of finite convex combinations of the y x (j) converging strongly to z, and so without loss of generality almost everywhere.

By convexity 1 0 W(x,z(k) )dx 1 0 sup j k j=k 1 λ (k) j W(x,y (j) x )dx 0 W(x,y(j) x )dx. Letting k, by Fatou s lemma 1 0 W(x,y x)dx= 1 0 W(x,z)dx l 1. But this implies that y x(x) > 0 a.e. and so y A.

Also by Fatou s lemma 1 0 h(x,y )dx liminf j 1 0 h(x,y(j) )dx=l 2. Hence l I(y ) l 1 +l 2 =l. 1 2 So I(y )=l and y is a minimizer.

Discussion of (H3) We interpret the superlinear growth condition (H3) for a homogeneous material. y(x)=px 1/p 1 Total stored-energy = W(p) p. So lim p W(p) p = says that you can t get a finite line segment from an infinitesimal one with finite energy.

Simplified model of atmosphere y(1)=α x=1 Assume constant density ρ R and pressure p R in the reference configuration. Assume gas deforms adiabatically so that the pressure p and density ρ satisfy pρ γ =p R ρ γ R, where γ>1 is a constant. x=0 y(0)=0

The potential energy of the column is I(y) = 1 0 = ρ R g [ p R (γ 1)y x (x) γ 1+ρ Rgy(x) 1 0 where k= p R ρ R g(γ 1). [ k y x (x) γ 1+(1 x)y x(x) ] dx ] dx, We seek to minimize I in A={y W 1,1 (0,1):y x >0 a.e., y(0)=0,y(1)=α}.

Then the minimum ofi(y) on A is attained iff α α crit, where α crit = γ ( ) pr 1/γ. γ 1 ρ R g y α crit can be interpreted as the finite height of the atmosphere predicted by this simplified model (cf Sommerfeld). α α crit If α>α crit then minimizing sequences y (j) for I converge to the minimizer for α=α crit plus a vertical portion. 1 x

For details of the calculation see J.M. Ball, Loss of the constraint in convex variational problems, in Analyse Mathématiques et Applications, Gauthier-Villars, Paris, 1988, where a general framework is presented for minimizing general framework is presented for minimizing a convex functional subject to a convex constraint, and is applied to other problems such as Thomas-Fermi and coagulation-fragmentation equations.

Discussion of (H4) For simplicity consider the case of a homogeneous material with stored-energy function W =W(p). The proof of existence of a minimizer used the direct method of the calculus of variations, the key point being that if W is convex then E(p)= 1 0 W(p)dx isweaklylowersemicontinuousinl 1 (0,1), i.e. p j pinl 1 (0,1)(that is 1 0 p (j) vdx 1 0 pvdx for all v L (0,1)) implies 1 W(p)dx liminf 0 j 1 0 W(p(j) )dx.

Proposition 2 (Tonelli) IfE isweaklylowersemicontinuousinl 1 (0,1) then W is convex. Proof. Define p (j) as shown. q p p (j) p (j) λp+(1 λ)q in L 1 (0,1) 10 W(p (j) )dx=λw(p)+(1 λ)w(q) W(λp+(1 λ)q) λ j 1 λ j 2 j 1 x

If W(x, ) is not convex then the minimum is in general not attained. For example consider the problem inf y(0)=0,y(1)= 3 2 1 0 [ (yx 1) 2 (y x 2) 2 y x +(y 3 2 x)2 ] dx. 3/2 Then the infimum is zero but is not attained. slopes 1 and 2 1

More generally we have the following result. Theorem 2 If W :R [0, ] is continuous and I(y)= 1 0 [W(y x )+h(x,y)]dx attains a minimum on A={y W 1,1 :y(0)=0,y(1)=β} forallcontinuoush:[0,1] R [0, ) andall β then W is convex.

Let l=inf y A 10 W(y x )dx, m=inf y A 10 [W(y x )+ y βx 2 ]dx. Then l m. Let ε > 0 and pick z A with 10 W(z x )dx l+ε. Define z (j) A by z (j) (x)=β k j +j 1 z(j(x k j )) for k j x k+1,k=0,...,j 1. j

Then So z (j) (x) βx = β( k j x)+j 1 z(j(x k j )) I(z (j) ) = j+1 k j k j 1 j+1 k=0 j 1 j 1 1 k=0 l+2ε Cj 1 for k j x k+1. j W(z x (j(x k j ))dx+ 1 0 W(z x)dx+cj 2 for j sufficiently large. Hence m l and so l=m. 0 z(j) βx 2 dx

But by assumption there exists y with I(y )= 1 0 W(y x)dx+ 1 0 y βx 2 dx=l. Hence y (x)=βx and thus 10 W(y x )dx W(β) for all y A. Taking in particular y(x)= { px if 0 x λ qx+λ(p q) if λ x 1, with β=λp+(1 λ)q we obtain W(λp+(1 λ)q) λw(p)+(1 λ)w(q) as required.

There are two curious special cases when the minimum is nevertheless attained when W is not convex. Suppose that (H1), (H3) hold and that either (i) I(y)= 1 0 W(x,y x )dx, or (ii) I(y)= 1 0 [W(y x )+h(y)]dx. Then I attains a minimum on A.

(i)canbefoundinaubert&tahraoui(j.differential Eqns 1979) and uses the fact that inf y A 10 W(x,y x )dx=inf y A 10 W (x,y x )dx, where W is the lower convex envelope of W. W(x,p) W (x,p) p

(ii) I(y)= 1 0 [W(y x)+h(y)]dx Exercise. Hint: Write x=x(y) and use (i).

The Euler-Lagrange equation in 1D Let y minimize I in A and let ϕ C 0 (0,1). Formally calculating d I(y+τϕ) =0 dτ I(y+τϕ) τ=0=0 we obtain the weak form of the Euler-Lagrange equation 1 0 [W p(x,y x )ϕ x +h y (x,y)ϕ]dx=0.

However, there is a serious problem in making this calculation rigorous, since we need to pass to the limit τ 0+ in the integral [ ] 1 W(x,yx +τϕ x ) W(x,y x ) dx. 0 τ and the only obvious information we have is that 1 0 W(x,y x)dx<.

But W p can be much bigger than W. For example, when p is small and W(p) = 1 p then W p (p) = 1 p2 is much bigger than W(p). Or if p is large and W(p)=expp 2 then W p (p) =2pexpp 2 is much bigger than W(p). It turns out that this is a real problem and not just a technicality. Even for smooth elliptic integrands satisfying a superlinear growth condition there is no general theorem of the one-dimensional calculus of variations saying that a minimizer satisfies the Euler-Lagrange equation.

Consider a general one-dimensional integral of the calculus of variations 1 I(u)= f(x,u,u x)dx, 0 where f is smooth and elliptic (regular), i.e. f pp µ>0 for all x,u,p. Suppose also that lim p f(x,u,p) p = for all x,u.

Supposeu W 1,1 (0,1)satisfiestheweakform of the Euler-Lagrange equation 1 0 [f pϕ x +f u ϕ]dx=0 for all ϕ C 0 (0,1), so that in particular f p,f u L 1 loc (0,1). Then x f p = f u+const a where a (0,1), and hence (Exercise) u x is bounded on compact subsets of (0, 1). It follows that u is a smooth solution of the Euler-Lagrange equation d dx f p=f u on (0,1).

For this problem one has that I(u +tϕ)= for t 0, if ϕ(0) 0. u u +tϕ Also if u (j) W 1, converges a.e. to u then I(u (j) ) (the repulsion property), explaining the numerical results.

Derivation of the EL equation for 1D elasticity Theorem 3 Let (H1) and (H5) hold and suppose further that W p exists and is continuous in (x,p) [0,1] (0, ),andthath y existsandiscontinuous in (x,y) [0,1] R. If y minimizes I in A then W p (,y x ( )) C 1 ([0,1]) and d dx W p(x,y x (x))=h y (x,y(x)) for all x [0,1].(EL)

Proof. Pick any representative of y x and let Ω j = {x [0,1] : 1 j y x(x) j}. Then Ω j is measurable, Ω j Ω j+1 and meas ( [0,1]\ j=1 Ω j) =0. Givenj, letz L (0,1)with Ω j zdx=0. For ε sufficiently small define y ε W 1,1 (0,1) by yx(x)=y ε x (x)+εχ j (x)z(x), y ε (0)=α, where χ j is the characteristic function of Ω j.

Then d dε I(yε ) ε=0 x ] = [W p (x,y x )z+h y χ jzds dx Ω[ 0 x ] = W p (x,y x ) h y(s,y(s))ds Ω j 0 =0. z(x) dx Hence x W p (x,y x ) h y(s,y(s))ds=c j in Ω j, 0 and clearly C j is independent of j.

Corollary 1 Let the hypotheses of Theorem 3 hold and assume further that lim max W p(x,p)= ( ). p 0+ x [0,1] Then there exists µ>0 such that y x (x) µ>0 a.e. x [0,1]. Proof. W p (x,y x (x)) C> by (EL).

Remark. (*) is satisfied if W(x,p) is convex in p for all x [0,1],0<p ε for some ε>0 and if Proof. (H1+) lim min W(x,p)=. p 0+ x [0,1] W p (x,p) W(x,ε) W(x,p). ε p

Corollary 2. If (*), (H3) and the hypotheses of Theorem 3 hold, and if W(x, ) is strictly convex then y has a representative in C 1 ([0,1]). Proof. W p (x,y x (x)) has a continuous representative. So there is a subset S [0,1] of full measure such thatw p (x,y x (x))is continuousons. We need to show that that if {x j },{ x j } S with x j x, x j x then the limits lim j y x (x j ), lim j y x ( x j ) exist and are finite and equal.

By (H3) and convexity, lim p min x [0,1] W p (x,p)=. Hencewemay assume that y x (x j ) z,y x ( x j ) z [µ, ) with z z. Hence W p (x,z)=w p (x, z), which by strict convexity implies z= z.

Existence of minimizersin 3D elastostatics

Minimize in I(y)= Ω W(Dy)dx A={y W 1,1 :detdy(x)>0 a.e., y Ω1 =ȳ}. (Notethatwehaveforthetimebeingreplaced the invertibility condition by the local conditiion detdy(x)>0 a.e., which is easier to handle.)

SofarwehaveassumedthatW :M 3 3 + [0, ) is continuous, and that (H1) W(F) as detf 0+, so that setting W(F) = if detf 0, we have that W : M 3 3 [0, ] is continuous, and that W is frame-indifferent, i.e. (H2) W(RF)=W(F) for all R SO(3),F M 3 3. (In fact (H2) plays no direct role in the existence theory.)

Growth condition y=fx 1 F lim F W(F) F 3 = says that you can t get a finite line segment from an infinitesimal cube with finite energy.

We will use growth conditions a little weaker than this. Note that if W(F) C(1+ F 3+ε ) then any deformation with finite elastic energy Ω W(Dy(x))dx is in W 1,3+ε and so is continuous.

Convexity conditions ThekeydifficultyisthatW isneverconvex,so that we can t use the same method to prove existence of minimizers as in 1D. Reasons 1. Convexity of W is inconsistent with (H1) because M+ 3 3 is not convex.

A=diag(1,1,1) 1 W( 1 2 (A+B))= > 1 2 W(A)+1 2 W(B) 2 (A+B)=diag(0,0,1) detf <0 detf >0 B=diag( 1, 1,1)

2. IfW isconvex,thenanyequilibriumsolution (solution of the EL equations) is an absolute minimizer of the elastic energy Proof. I(z)= I(y)= Ω W(Dz)dx Ω W(Dy)dx. Ω [W(Dy)+DW(Dy) (Dz Dy)]dx=I(y). This contradicts common experience of nonunique equlibria, e.g. buckling.

Examples of nonunique equilibrium solutions Pure zero traction problem.

For an isotropic material could we assume instead that Φ(v 1,v 2,v 3 ) is convex in the principal stretches v 1,v 2,v 3? This is a consequence of the Coleman-Noll inequality. While it is consistent with (H1) it is not in general satisfied for rubber-like materials, which are almost incompressible (v 1 v 2 v 3 =1), and so have nonconvex sublevel sets. v 2 v 1 v 2 =1 v 3 =1 nonconvex sublevel set Φ c v 1

Rank-one matrices and the Hadamard jump condition N y piecewise affine Dy=A, x N >k Dy=B, x N <k x N =k Let C = A B. Then Cx = 0 if x N = 0. Thus C(z (z N)N) = 0 for all z, and so Cz=(CN N)z. Hence A B=a N Hadamard jump condition

More generally this holds for y piecewise C 1, with Dy jumping across a C 1 surface. N Dy + (x 0 )=A x 0 Dy (x 0 )=B A B=a N Exercise: prove this by blowing up around x using y ε (x)=εy( x x 0 ε ).

Rank-one convexity W is rank-one convex if the map t W(F +ta N) is convex for each F M 3 3 and a R 3,N R 3. (Same definition for M m n.) Equivalently, W(λF +(1 λ)g) λw(f)+(1 λ)w(g) if F,G M 3 3 with F G = a N and λ (0,1).

F Rank-one cone Λ={a N :a,n R 3 } Rank-one convexity is consistent with(h1) becausedet(f+ta N)islinearint,sothatM 3 3 + is rank-one convex (i.e. if F,G M 3 3 + λf +(1 λ)g M 3 3 +.) with F G=a N then

A specific example of a rank-one convex W is an elastic fluid for which W(F)=h(detF), with h convex and lim δ 0+ h(δ)=. Such a W is rank-one convex because if F,G M+ 3 3 with F G=a N, and λ (0,1) then W(λF +(1 λ)g) = h(λdetf +(1 λ)detg) λh(detf)+(1 λ)h(detg) = λw(f)+(1 λ)w(g).

If W C 1 (M+ 3 3 ) then W is rank-one convex ifft DW(F+ta N) a N isnondecreasing. The linear map y(x)=(f +ta N)x=Fx+ta(x N) represents a shear relative to Fx parallel to a planeπwithnormaln inthereferenceconfiguration, in the direction a. The corresponding stress vector across the plane Π is t R =DW(F +ta N)N, and so rank-one convexity says that the component t R a in the direction of the shear is monotone in the magnitude of the shear.

If W C 2 (M+ 3 3 ) then W is rank-one convex iff d 2 dt 2W(F +ta N) t=0 0, for all F M 3 3 +,a,n R3, or equivalently D 2 W(F)(a N,a N)= 2 W(F) F iα F jβ a i N α a j N β 0, (Legendre-Hadamard condition).

The strengthened version D 2 W(F)(a N,a N)= 2 W(F) F iα F jβ a i N α a j N β µ a 2 N 2, forallf,a,n andsomeconstantµ>0iscalled strong ellipticity. One consequence of strong ellipticity is that it implies the reality of wave speeds for the equations of elastodynamics linearized around a uniform state y=fx.

Equation of motion of pure elastodynamics ρ R ÿ=divdw(dy), where ρ R is the (constant) density in the reference configuration. Linearized around the uniform state y = Fx the equations become ρ R ü i = 2 W(F) F iα F jβ u j,αβ.

The plane wave is a solution if where u=af(x N ct) C(F)a=c 2 ρ R a, C ij = 2 W(F) F iα F jβ N α N β. Since C(F)>0 by strong ellipticity, it follows that c 2 >0 as claimed.

Quasiconvexity(C.B. Morrey, 1952) Let W :M m n [0, ] be Borel measurable. W is said to be quasiconvex at F M m n if the inequality W(F +Dϕ(x))dx W(F)dx Ω Ω holds for any ϕ W 1, 0 (Ω;R m ), and is quasiconvexifitisquasiconvexateveryf M m n. Here Ω R n is any bounded open set whose boundary Ω has zero n-dimensional Lebesgue measure.

Remark W 1, 0 (Ω;R m ) is defined as the closure of C0 (Ω;Rm ) in the weak* topology of W 1, (Ω;R m ) (and not in the norm topology - why?). That is ϕ W 1, 0 (Ω;R m ) if there exists a sequence ϕ (j) C0 (Ω;Rm ) such that (j) (j) ϕ ϕ, Dϕ Dϕ in L. Sometimes the definition is given with C 0 replacing W 1, 0. This is the same if W is finite and continuous, but it is not clear (to me) if itisthesameifw iscontinuousandtakesthe value +.

Setting m=n=3 we see that W is quasiconvex if for any F M 3 3 the pure displacement problem to minimize I(y)= W(Dy(x))dx Ω subject to the linear boundary condition y(x)=fx, x Ω, has y(x)=fx as a minimizer.

Proposition 3 Quasiconvexity is independent of Ω. Proof. Suppose the definition holds for Ω, and let Ω 1 be another bounded open subset of R n suchthat Ω 1 hasn-dimensionalmeasurezero. BytheVitalicoveringtheoremwecanwriteΩ as a disjoint union Ω= i=1 (a i +ε i Ω 1 ) N, where N is of zero measure.

Ω Ω 1 Let ϕ W 1, 0 (Ω 1 ;R n ) and define ϕ(x)= { εi ϕ( x a i ε i ) for x a i +ε i Ω 1 0 otherwise.

Then ϕ W 1, 0 (Ω;R n ) and W(F +Dϕ(x))dx Ω = a i +ε i Ω 1 W(F +Dϕ i=1 = = i=1 ( measω measω 1 ( x ai ε i ) )dx ε n i Ω 1 W(F +Dϕ(x))dx ) Ω 1 W(F +Dϕ(x))dx (measω)w(f).

Another form of the definition that is equivalent for finite continuous W is that Q W(Dy)dx (measq)w(f) foranyy W 1, suchthatdyistherestriction to a cube Q (e.g. Q=(0,1) n ) of a Q-periodic map on R n with Q Dydx=F. One can even replace periodicity with almost periodicity(see J.M. Ball, J.C. Currie, and P.J. Olver. Null Lagrangians, weak continuity, and variational problems of arbitrary order. J. Functional Anal., 41:135 174, 1981).

Theorem 4 If W is continuous and quasiconvex then W is rank-one convex. Remark This is not true in general if W is not contin- uous. As an example, define for given nonzero a,n W(0)=W(a N)=0,W(F)= otherwise. Then W is clearly not rank-one convex, but it isquasiconvexbecausegivenf 0,a N there is no ϕ W 1, 0 with F +Dϕ(x) {0,a N}

Proof We prove that W(F) λw(f (1 λ)a N)+(1 λ)w(f+λa N) for any F M m n,a R m,n R m,λ (0,1). Without loss of generality we suppose that N =e 1. We follow an argument of Morrey. Let D=( (1 λ),λ) ( ρ,ρ) n 1 and let D ± j be the pyramid that is the convex hull of the origin and the face of D with normal ±e j.

1 λ x j Dϕ=λa e 1 (1 λ)a e 1 D 1 - ρ ρ 1 λ(1 λ)a e j ρ 1 λ(1 λ)a e j D 1 + ρ D j + D- j λ x 1 Let ϕ W 1, 0 (D;R m ) be affine in each D ± j with ϕ(0)=λ(1 λ)a. The values of Dϕ are shown.

By quasiconvexity (2ρ) n 1 W(F) (2ρ)n 1 λ W(F (1 λ)a e 1 ) n + (2ρ)n 1 (1 λ) W(F +λa e 1 ) n n (2ρ) n 1 + [W(F +ρ 1 λ(1 λ)a e j ) 2n j=2 +W(F ρ 1 λ(1 λ)a e j )] SupposeW(F)<. Thendividingby(2ρ) n 1, letting ρ 0+ and using the continuity of W, we obtain W(F) λw(f (1 λ)a e 1 )+(1 λ)w(f +λa e 1 ) as required.

Now suppose that W(F (1 λ)a e 1 ) and W(F+λa e 1 )arefinite. Theng(τ)=W(F+ τa e 1 ) lies below the chordjoining the points ( (1 λ),g( (1 λ))),(λ,g(λ)) whenever g(τ)<, and since g is continuous it follows that g(0)=w(f)<. g (1 λ) λ

Corollary 3 If m = 1 or n = 1 then a continuous W : M m n [0, ] is quasiconvex iff it is convex. Proof. If m = 1 or n = 1 then rank-one convexity is the same as convexity. If W isconvex (forany dimensions) then W is quasiconvex by Jensen s inequality: 1 measω W ( Ω 1 measω W(F +Dϕ)dx Ω (F +Dϕ)dx ) =W(F).

Theorem 5 Let W :M m n [0, ] be Borel measurable, and Ω R n a bounded open set. A necessary condition for I(y)= W(Dy)dx Ω to be sequentially weak* lower semicontinuous on W 1, (Ω;R m ) is that W is quasiconvex. Proof. Let F M m n,ϕ W 1, 0 (Q;R m ), where Q=(0,1) n.

Given k write Ω as the disjoint union Ω= (a (k) j +ε (k) j=1 j Q) N k, wherea (k) j R m, ε (k) j <1/k,measN k =0,and define y (k) (x)= Fx+ε (k) j ϕ F x x a(k) j ε (k) j for x a (k) j +ε (k) j Q otherwise.

Then y (k) Fx in W 1, as k and so (measω)w(f) liminf = j=1 = = j=1 a (k) j +ε (k) j Q ε (k)n j ( measω measq Q ) k I(y(k) ) W(F +Dϕ W(F +Dϕ)dx Q W(F +Dϕ)dx x a (k) j ε (k) j )dx

Theorem 6 (Morrey, Acerbi-Fusco, Marcellini) Let Ω R n be bounded and open. Let W : M m n [0, ) be quasiconvex and let 1 p. If p< assume that 0 W(F) c(1+ F p ) for all F M m n. Then I(y)= Ω W(Dy)dx is sequentially weakly lower semicontinuous(weak* if p= ) on W 1,p. Proof omitted. Unfortunately the growth conditionconflictswith (H1),so wecan tusethis to prove existence in 3D elasticity.

Theorem 7 (Ball & Murat) Let W :M m n [0, ] be Borel measurable, and let Ω R n be bounded open and have boundary of zero n-dimensional measure. If I(y)= [W(Dy)+h(x,y)]dx Ω attains an absolute minimum on A={y :y Fx W 1,1 0 (Ω;Rm )} for all F and all smooth nonnegative h, then W is quasiconvex. Proof (Exercise: use the same method as Theorem 2 and Vitali.)

Theorem 8 (van Hove) Let W(F)=c ijkl F ij F kl be quadratic. Then W is rank-one convex W is quasiconvex. Proof. Let W be rank-one convex. Since for any ϕ W 1, 0 Ω [W(F +Dϕ) W(F)]dx= Ω c ijklϕ i,j ϕ k,l dx we just need to show that the RHS is 0. Extend ϕ by zero to the whole of R n and take Fourier transforms.

By the Plancherel formula c ijklϕ i,j ϕ k,l dx = Ω R nc ijklϕ i,j ϕ k,l dx = 4π 2 R nre[c ijklˆϕ i ξ j ˆϕ k ξ l ]dξ 0 as required.

Null Lagrangians When does equality hold in the quasiconvexity condition? That is, for what L is Ω L(F +Dϕ(x))dx= Ω L(F)dx for all ϕ W 1, 0 (Ω;R m )? We call such L quasiaffine.

Theorem 9 (Landers, Morrey, Reshetnyak...) If L : M m n R is continuous then the following are equivalent: (i) L is quasiaffine. (ii) L is a (smooth) null Lagrangian, i.e. the Euler-Lagrange equations DivD F L(Du) = 0 hold for all smooth u. (iii) L(F)=constant + d(m,n) k=1 c k J k (F), where J(F) = (J 1 (F),...,J d(m,n) (F)) consists of all the minors of F. (e.g. m = n = 3: L(F)=const.+C F +D coff +edetf). (iv) u L(Du) is sequentially weakly continuous from W 1,p L 1 for sufficiently large p (p>min(m,n) will do).

Ideas of proofs. (i) (iii) use L rank-one affine (Theorem 4). (iii) (iv) Take e.g. J(Du) = u 1,1 u 2,2 u 1,2 u 2,1 if u is smooth. So if ϕ C 0 (Ω) Ω J(Du) ϕdx= = (u 1 u 2,2 ),1 (u 1 u 2,1 ),2 Ω [u 1u 2,1 ϕ,2 u 1 u 2,2 ϕ,1 )dx. True for u W 1,2 by approximation.

If u (j) u in W 1,p,p>2, then Ω J(Du(j) )ϕdx = Ω [u(j) 1 u(j) Ω J(Du)ϕdx 2,1 ϕ,2 u (j) 1 u(j) 2,2 ϕ,1]dx since u (j) 1 u 1 in L p, u (j) 2,1 u 2,1 in L p, and since J(Du (j) ) is bounded in L p/2 it follows that J(Du (j) ) J(Du) in L p/2. For the higher-order Jacobians we use induction based on the identity (u 1,...,u m ) m (x 1,...,x m ) = s=1 ( 1) s+1 x s ( u 1 (u 2,...,u m ) (x 1,...,ˆx s,...,x m ) )

For example, suppose m=n=3, p 2, u (j) u in W 1,p,cofDu (j) bounded in L p, detdu (j) χ in L 1. Then χ=detdu. (iv) (i) by Theorem 5. (i) (ii) d dt Ω L(F +Dϕ+tDψ)dx =0 t=0 Ω for all ϕ,ψ C 0 DL(F +Dϕ) Dψdx=0. }{{} Du

Polyconvexity Definition W is polyconvex if there exists a convex function g:r d(m,n) (, ] such that W(F)=g(J(F)) for all F M m n. e.g. W(F)=g(F,detF) if m=n=2, W(F)=g(F,cofF,detF) if m=n=3, with g convex.

Theorem 10 Let W : M m n [0, ] be Borel measurable and polyconvex, with g lower semicontinuous. Then W is quasiconvex. Proof. Writing 1 fdx= fdx, Ω measω Ω Ω W(F +Dϕ(x))dx = Jensen g Ω ( g(j(f +Dϕ(x)))dx Ω = g(j(f)) = W(F). ) J(F +Dϕ)dx

Remark If m 3,n 3 then there are quadratic rankone convex W that are not polyconvex. Such W cannot be written in the form W(F)=Q(F)+ N αl J (l) 2 (F), l=1 whereq 0isquadraticand thej (l) 2 are2 2 minors (Terpstra, D. Serre).

Examples and counterexamples We have shown that W convex W polyconvex W quasiconvex W rank-one convex. The reverse implications are all false if m > 1,n > 1, except that it is not known whether W rank-one convex W quasiconvex when n m=2. W polyconvex W convex since any minor is polyconvex.

Example (Dacorogna & Marcellini) m=n=2 W γ (F)= F 4 2γ F 2 detf, γ R, where F 2 =tr(f T F). It is not known for what γ the function W γ is quasiconvex.

W γ is convex γ 2 3 2 polyconvex γ 1 quasiconvex γ 1+ε for some (unknown) ε>0 rank-one convex γ 2 1.1547... 3 Numerically (Dacorogna-Haeberly) 1+ε=1.1547... In particular W quasiconvex W polyconvex (see also later).

Theorem 11 (Sverak 1992) If n 2,m 3 then W rank-one convex W quasiconvex. Sketch of proof. Itisenoughtoconsiderthecasen=2,m=3. Consider the periodic function u:r 2 R 3 u(x)= 1 2π sin2πx 1 sin2πx 2 sin2π(x 1 +x 2 ).

Then Du(x) = L:= cos2πx 1 0 0 cos2πx 2 cos2π(x 1 +x 2 ) cos2π(x 1 +x 2 ) r 0 0 s t t :r,s,t R a.e.

L is a 3-dimensional subspace of M 3 2 and the only rank-one lines in L are in the r,s,t 1 0 0 0 directions (i.e. parallel to 0 0, 0 1, 0 0 0 0 0 0 or 0 0 ). 1 1 Hence g(f)= rst is rank-one affine on L.

However (0,1) 2g(Du)dx= 1 4 <0=g violating quasiconvexity. For F M 3 2 define ( (0,1) 2Dudx ), f ε,k (F)=g(PF)+ε( F 2 + F 4 )+k F PF 2, where P :M 3 2 L is orthogonal projection. Can check that for each ε > 0 there exists k(ε)>0 such that f ε =f ε,k(ε) isrank-oneconvex, and we still get a contradiction.

So is there a tractable characterization of quasiconvexity? This is the main road-block of the subject. Theorem 12 (Kristensen 1999) For m 3,n 2 there is no local condition equivalent to quasiconvexity (for example, no condition involving W and any number of its derivatives at an arbitrary matrix F). Idea of proof. Sverak s W is locally quasiconvex, i.e. it coincides with a quasiconvex function in a neighbourhood of any F.

This might lead one to think that no characterization of quasiconvexity is possible. On the other hand Kristensen also proved Theorem 13 (Kristensen) For m 2,n 2 polyconvexity is not a local condition. For example, one might contemplate a characterization of the type W quasiconvex W is the supremum of a family of special quasiconvex functions(including null Lagrangians).

Existence based on polyconvexity Wewillshowthatitispossibletoprovetheexistence of minimizers for mixed boundary value problems if we assume W is polyconvex and problems if we assume W is polyconvex and satisfies (H1) and appropriate growth conditions. Furthermore the hypotheses are satisfied by various commonly used models of natural rubber and other materials.

Theorem 14 Suppose that W satisfies (H1) and the hypotheses (H3) W(F) c 0 ( F 2 + coff 3/2 ) c 1 for all F M 3 3, where c 0 >0, (H4)W ispolyconvex,i.e. W(F)=g(F,cofF,detF) for all F M 3 3 for some continuous convex g. Assume that there exists some y in A={y W 1,1 (Ω;R 3 ):y Ω1 =ȳ} with I(y)<, where H 2 ( Ω 1 )>0 and ȳ: Ω 1 R 3 ismeasurable. Thenthereexists a global minimizer y of I in A.

Theorem 14 is a refinement (weakening the growth conditions) of J.M. Ball, Convexity conditions and existence theorems in nonlinear elasticity. Arch. Rat. Mech. Anal., 63:337 403, 1977 (see also J.M. Ball. Constitutive inequalities and existence theorems in nonlinear elastostatics. In R.J. Knops, editor, Nonlinear Analysis and Mechanics, Heriot-Watt Symposium, Vol. 1. Pitman, 1977.) due to S. Müller, T. Qi, and B.S. Yan. On a new class of elastic deformations not allowing for cavitation. Ann. Inst. Henri Poincaré, Analyse Nonlinéaire, 11:217243, 1994.

Proof of Theorem 14 Togiveareasonablysimpleproofwewillcombine (H1), (H3), (H4) into the single hypothesis W(F)=g(F,cofF,detF) forsomecontinuousconvexfunctiong:m 3 3 3 3 M 3 3 R R + with g(f,h,δ) < iff δ>0 and g(f,h,δ) c 0 ( F p + H p )+h(δ), for all F M 3 3, where p 2, 1 p + 1 p = 1, c 0 > 0 and h : R [0, ] is continuous with h(δ)< iff δ>0 and lim δ h(δ) δ =.

Letl=inf y A I(y)andlety (j) beaminimizing sequence for I in A, so that lim j I(y(j) )=l. Then since by assumption l < we may as- sume that l+1 I(y (j) ) ( c0 [ Dy (j) p + cofdy (j) p ] for all j. Ω +h(detdy (j) ) ) dx

Lemma 1 There exists a constant d>0 such that ( Ω z p dx d Ω Dz p dx+ for all z W 1,p (Ω;R 3 ). Ω 1 zda p ) Proof. Suppose not. Then there exists z (j) such that 1= for all j. ( Ω z(j) p dx j Ω Dz(j) p dx+ Ω 1 z (j) da p )

Thus z (j) is bounded in W 1,p and so there is a subsequence z (j k) z in W 1,p. Since Ω is Lipschitz we have by the embedding and trace theorems that z (j k) z strongly in L p, z (j k) z in L 1 ( Ω). In particular Ω z p dx=1.

But since p is convex we have that Ω Dz p dx liminf k Ω Dz(j k) p dx=0. Hence Dz=0 in Ω, and since Ω is connected it follows that z = constant a.e. in Ω. But also we have that Ω 1 zda=0, and since H 2 ( Ω 1 )>0 it follows that z=0, contradicting Ω z p dx=1.

By Lemma 1 the minimizing sequence y (j) is bounded in W 1,p and so we may assume that y (j) y in W 1,p for some y. But also we have that cofdy (j) is bounded in L p and that Ω h(detdy(j) )dx is bounded. So we may assume that cofdy (j) H in L p and that detdy (j) δ in L 1. Bytheresultsontheweakcontinuityofminors we deduce that H=cofDy and δ=detdy.

Let u (j) =(Dy (j),cofdy (j),detdy (j) ), u=(dy,cofdy,detdy )). Then u (j) u in L 1 (Ω;R 19 ). Butg isconvex,andsousingmazur stheorem as in the proof of Theorem 1, I(y )= g(u)dx liminf Ω j Ω g(u(j) )dx = lim j I(y (j) )=l. But y (j) Ω1 =ȳ y Ω1 in L 1 ( Ω 1 ;R 3 ) and so y A and y is a minimizer.

Incompressible elasticity Rubber is almost incompressible. Thus very large forces, and a lot of energy, are required to change its volume significantly. Such materials are well modelled by the constrained theory of incompressible elasticity, in which the deformation gradient is required to satisfy the pointwise constraint detf =1.

Existence of minimizersin incompressible elasticity Theorem 15 Let U = {F M 3 3 : detf = 1}. Suppose W : U [0, ) is continuous and such that (H3) W(F) c 0 ( F 2 + coff 3 2) c 1 for all F U, (H4) W ispolyconvex,i.e. W(F)=g(F,cofF) for all F U for some continuous convex g. Assume that there exists some y in A={y W 1,1 (Ω;R 3 ):detdy(x)=1 a.e.,y Ω1 =ȳ} with I(y)<, where H 2 ( Ω 1 )>0 and ȳ: Ω 1 R 3 ismeasurable. Thenthereexists a global minimizer y of I in A.

Proof. For simplicity suppose that W(F)=g(F,cofF) for some continuous convex g:m 3 3 M 3 3 R, where g(f,h) c 0 ( F p + H q ) c 1 forallf,h,where p 2,q p and c 0 >0. Lettingy (j) beaminimizingsequencetheonly newpointistoshowthattheconstraintissatisfied. But since detdy (j) =1 we have by the weak continuity properties that for a subsequence detdy (j) detdy in L, so that the weak limit satisfies detdy=1.

Models of natural rubber 1. Modelled as an incompressible isotropic material. Constraint is detf =v 1 v 2 v 3 =1. Neo-Hookean material Φ=α(v 2 1 +v2 2 +v2 3 3), whereα>0isaconstant. Thiscanbederived from a simple statistical mechanics model of long-chain molecules.

Mooney-Rivlin material. Φ=α(v 2 1 +v2 2 +v2 3 3) +β((v 2 v 3 ) 2 +(v 3 v 1 ) 2 +(v 1 v 2 ) 2 3), where α > 0,β > 0 are constants. Gives a better fit to bi-axial experiments of Rivlin & Saunders. Ogden materials. Φ= N i=1 α i (v p i 1 +vp i 2 +vp i 3 3) + M i=1 where α i,β i,p i,q i are constants. β i ((v 2 v 3 ) q i+(v 3 v 1 ) q i+(v 1 v 2 ) q i 3),

e.g. for a certain vulcanised rubber a good fit is given by N =2,M =1,p 1 =5.0, p 2 =1.3,q 1 =2,α 1 =2.4 10 3,α 2 =4.8, β 1 =0.05kg/cm 2. The high power 5 allows a better modelling of the tautening of rubber as the long-chain molecules are highly stretched and the cross-links tend to prevent further stretchand the cross-links tend to prevent further stretching.

2. Modelled as compressible isotropic material Add h(v 1 v 2 v 3 ) to above Φ, where h is convex, h(δ) as δ 0+, and h has a steep minimum near δ=1.

Convexity properties of isotropic functions Let R n + ={x=(x 1,...,x n ) R n :x i 0}. Theorem 16 (Thompson & Freede) Let n 1 and for F M n n let W(F)=Φ(v 1,...,v n ), where Φ is a symmetric real-valued function of the singular values v i of F. Then W is convex on M n n iff Φ is convex on R n + and nondecreasing in each v i.

Proof. Necessity. If W is convex then clearly Φ is convex. Also for fixed nonnegative v 1,...,v k 1,v k+1,...,v n g(v)=w(diag(v 1,...,v k 1, v,v k+1,...,v n )) is convex and even in v. But any convex and even function of v is nondecreasingfor v>0. The sufficiency uses von Neumann s inequality - for the details see Ball (1977).

Lemma 2 (von Neumann) Let A,B M n n have singular values v 1 (A)... v n (A), v 1 (B)... v n (B) respectively. Then vi n max tr(qarb)= (A)v i (B). Q,R O(n) i=1

ApplyingTheorem16weseethatifp 1then Φ p (F)=v p 1 +vp 2 +vp 3 is a convex function of F. Since the singular values of coff are v 2 v 3,v 3 v 1,v 1 v 2 it also follows that if q 1 Ψ q (F)=(v 2 v 3 ) q +(v 3 v 1 ) q +(v 1 v 2 ) q is a convex function of coff.

Hence the incompressible Ogden material Φ= N i=1 α i (v p i 1 +vp i 2 +vp i 3 3) + M i=1 β i ((v 2 v 3 ) q i+(v 3 v 1 ) q i+(v 1 v 2 ) q i 3), is polyconvex if the α i 0,β i 0, p 1... p N 1,q 1... q M 1. And in the compressible case if we add a convex function h=h(detf) of detf then under the same conditions the stored-energy function is polyconvex.

It remains to check the growth condition of Theorems 14, 15, namely W(F) c 0 ( F 2 + coff 3/2 ) c 1 for all F M 3 3, where c 0 >0. This holds for the Ogden materials provided p 1 2,q 1 3 2 and α 1>0,β 1 >0. This includes the case of the Mooney-Rivlin material, but not the neo-hookean material. In the incompressible case, Theorem 15 covers the case of the stored-energy function if p 3. Φ=α(v p 1 +vp 2 +vp 3 3)

For the neo-hookean material (incompressible orwithh(detf)added)itisnotknownifthere exists an energy minimizer in A, but it seems unlikely because of the phenomenon of cavitation.

In particular the stored-energy function W(F)=α( F 2 3)+h(detF) is not W 1,2 quasiconvex (same definition but with the test functions ϕ W 1,2 0 instead of W 1, 0 ). To see this consider the radial deformation y: B(0,1) R 3 given by where R= x. y(x)= r(r) R x,

Since y i = r(r) R x i it follows that y i,α = r(r) R δ iα+ ( r r R R ) xi x α R, that is ( ) Dy(x)= r R 1+ r r x R R x R. In particular Dy(x) 2 =r 2 +2 ( r(r) R ) 2.

Set F =λ1 where λ>0. Then W(λ1)=3α(λ 2 1)+h(λ 3 ). On the other hand, if we choose r 3 (R)=R 3 +λ 3 1, then y(x)=λx for x =1 and ( detdy(x)=r r 2=1. R)

Then B(0,1) [α( Dy(x) 2 3)+h(detDy(x))]dx =4π 1 0 R2 α ( R 3 +λ 3 1 R 3 which is of order λ 2 for large λ. ) 4 3 ( R 3 +λ 3 )2 1 3 3 +2 +h(1) dr, R 3 Hence B(0,1) W(Dy(x))dx< for large λ. B(0,1) W(λ1)dx

SinceW 1,2 quasiconvexityisnecessaryforweak lower semicontinuity of I(y) in W 1,2 this suggests that the minimum is not attained. (For a further argument suggesting this see J.M. Ball, Progress and Puzzles in Nonlinear Elasticity, Proceedings of course on Poly-, Quasiand Rank-One Convexity in Applied Mechanics, CISM, Udine, 2010.

However there is an existence theory that covers the neo-hookean and other cases of polyconvex energies with slow growth, due to M. Giaquinta, G. Modica and J. Souček, Arch. Rational Mech. Anal. 106 (1989), no. 2, 97-159 using Cartesian currents. In the previous examusing Cartesian currents. In the previous examplethiswouldgiveastheminimizery(x)=λx, i.e. the function space setting does not allow cavitation. A simpler proof of this result is in S. Müller, Weak continuity of determinants and nonlinear elasticity, C. R. Acad. Sci. Paris Ser. I Math. 307 (1988), no. 9, 501-506.

The Euler-Lagrange equations Suppose that W C 1 (M+ 3 3 ). Can we show that the minimizer y in Theorem 14 satisfies the weak form of the Euler-Lagrange equa- tions? As we have seen the standard form of these are formally obtained by computing d dτ I(y+τϕ) τ=0= d dτ W(Dy+τDϕ)dx t=0=0, Ω for smooth ϕ with ϕ Ω1 =0.

This leads to the weak form D FW(Dy) Dϕdx=0 Ω for all smooth ϕ with ϕ Ω1 =0. As we have seen, the problem in deriving this weakformisthat DW(Dy) canbebiggerthan W(Dy), and that we do not know if detdy(x) µ>0 a.e. It is an open problem to give hypotheses under which this or the above weak form can be proved.

However, it turns out to be possible to derive other weak forms of the Euler-Lagrange equations by using variations involving compositions of maps. We consider the following conditions that may be satisfied by W: (C1) D F W(F)F T K(W(F)+1) for all F M 3 3 +, where K>0 is a constant, and (C2) F T D F W(F) K(W(F)+1) for all F M 3 3 +, where K>0 is a constant.

As usual, denotes the Euclidean norm on M 3 3,forwhichtheinequalities F G F G and FG F G hold. But of course the conditions are independent of the norm used up to a possible change in the constant K. Proposition 4 Let W satisfy (C2). Then W satisfies (C1).

Proof SinceW isframe-indifferentthematrixd F W(F)F T is symmetric (this is equivalent to the symmetry of the Cauchy stress tensor T = (detf) 1 T R (F)F T ). To prove this, note that d =D dt W(exp(Kt)F) t=0=d F W(F) (KF)=0 for all skew K. Hence D F W(F)F T 2 = [D F W(F)F T ] [F(D F W(F)) T ] = [F T D F W(F)] [F T D F W(F)] T F T D F W(F) 2, from which the result follows.

Example Let W(F)=(F T F) 11 + 1 detf. Then W is frame-indifferent and satisfies (C1) but not (C2). Exercise: check this. As before let A={y W 1,1 (Ω;R 3 ):y Ω1 =ȳ}.

We say that y is a W 1,p local minimizer of I(y)= W(Dy)dx Ω in A if I(y)< and I(z) I(y) for all z A with z y 1,p sufficiently small.

Theorem 17 For some 1 p < let y A W 1,p (Ω;R 3 ) be a W 1,p local minimizer of I in A. (i) Let W satisfy (C1). Then Ω [D FW(Dy)Dy T ] Dϕ(y)dx=0 for all ϕ C 1 (R 3 ;R 3 ) such that ϕ and Dϕ are uniformly bounded and satisfy ϕ(y) Ω1 =0 in the sense of trace. (ii) Let W satisfy (C2). Then Ω [W(Dy)1 DyT D F W(Dy)] Dϕdx=0 for all ϕ C 1 0 (Ω;R3 ).

We use the following simple lemma. Lemma 3 (a) If W satisfies (C1) then there exists γ>0 such that if C M+ 3 3 and C 1 <γ then D F W(CF)F T 3K(W(F)+1) for all F M+ 3 3. (b) If W satisfies (C2) then there exists γ>0 such that if C M+ 3 3 and C 1 <γ then F T D F W(FC) 3K(W(F)+1) for all F M 3 3 +.

Proof of Lemma 3 We prove (a); the proof of (b) is similar. We first show that there exists γ >0 such that if C 1 <γ then W(CF)+1 3 (W(F)+1) for all F M3 3 2 +. For t [0,1] let C(t)=tC+(1 t)1. Choose γ (0, 1 6K )sufficientlysmallsothat C 1 <γ implies that C(t) 1 2 for all t [0,1]. This is possible since 1 = 3<2.

For C 1 <γ we have that W(CF) W(F) 1 d = 0 dt W(C(t)F)dt 1 = D FW(C(t)F) [(C 1)F]dt 0 1 = D F W(C(t)F)(C(t)F) T ((C 1)C(t) 1 )dt 0 K 1 2Kγ 0 [W(C(t)F)+1] C 1 C(t) 1 dt 1 0 (W(C(t)F)+1)dt.

Let θ(f)=sup C 1 <γ W(CF). Then W(CF) W(F) θ(f) W(F) 2Kγ(θ(F)+1). Hence (θ(f)+1)(1 2Kγ) W(F)+1, from which W(CF)+1 3 2 (W(F)+1) follows.

Finally, if C 1 < γ we have from (C1) and the above that D F W(CF)F T = D F W(CF)(CF) T C T as required. K(W(CF)+1) C T 3K(W(F)+1),

Proof of Theorem 17 Given ϕ as in the theorem, define for τ sufficiently small Then y τ (x):=y(x)+τϕ(y(x)). Dy τ (x)=(1+τdϕ(y(x)))dy(x) a.e. x Ω. and so y τ A. Also detdy τ (x) > 0 for a.e. x Ω and lim τ 0 y τ y W 1,p=0.

Hence I(y τ ) I(y) for τ sufficiently small. But 1 τ (I(y τ) I(y)) 1 d ds W((1+sτDϕ(y(x)))Dy(x))dsdx = 1 τ Ω 0 1 = DW((1+sτDϕ(y(x)))Dy(x)) Ω 0 [Dϕ(y(x))Dy(x)] ds dx.

Since by Lemma 3 the integrand is bounded by the integrable function 3K(W(Dy(x))+1) sup z R 3 Dϕ(z), we may pass to the limit τ 0 using domi- nated convergence to obtain Ω [D FW(Dy)Dy T ] Dϕ(y)dx=0, as required.

(ii) This follows in a similar way to (i) from Lemma 3(b). We just sketch the idea. Let ϕ C 1 0 (Ω;R3 ). Forsufficientlysmall τ >0the mapping θ τ defined by θ τ (z):=z+τϕ(z) belongstoc 1 (Ω;R 3 ), satisfies detdθ τ (z)>0, and coincides with the identity on Ω. By the global inverse function theorem θ τ is a diffeomorphism of Ω to itself.

Thus the inner variation y τ (x):=y(z τ ), x=z τ +τϕ(z τ ) defines a mapping y τ A, and Dy τ (x)=dy(z τ )[1+τDϕ(z τ )] 1 a.e. x Ω. Since y W 1,p it follows easily that y τ y W 1,p 0 as τ 0.

Changing variables we obtain I(y τ )= from which Ω W(Dy(z)[1+τDϕ(z)] 1 ) det(1+τdϕ(z))dz, Ω [W(Dy)1 DyT D F W(Dy)] Dϕdx=0 follows from (C2) and Lemma 3 using dominated convergence.

Interpretations of the weak forms TointerpretTheorem17(i), wemakethefollowing Invertibility Hypothesis. y is a homeomorphism of Ω onto Ω :=y(ω), Ω is a bounded domain, and the change of variables formula Ω f(y(x))detdy(x)dx= Ω f(z)dz holds whenever f :R 3 R is measurable, provided that one of the integrals exists.

Theorem 18 Assume that the hypotheses of Theorem 17 and the Invertibility Hypothesis hold. Then Ω σ(z) Dϕ(z)dz=0 for all ϕ C 1 (R 3 ;R 3 ) such that ϕ y( Ω1 ) =0, wherethecauchy stress tensor σ is definedby σ(z):=t(y 1 (z)), z Ω andt(x)=(detdy(x)) 1 D F W(Dy(x))Dy(x) T. Proof. Since by assumption y( Ω) is bounded, we can assume that ϕ and Dϕ are uniformly bounded. The result then follows straighforwardly.

Thus Theorem 17 (i) asserts that y satisfies the spatial (Eulerian) form of the equilibrium equations. Theorem 17(ii), on the other hand, involves the so-called energy-momentum tensor E(F)=W(F)1 F T D F W(F), F and is a multi-dimensional version of the Du Bois Reymond or Erdmann equation of the one-dimensional calculus of variations, and is the weak form of the equation DivE(Dy)=0.

The hypotheses (C1) and (C2) imply that W has polynomial growth. Proposition 4 Suppose W satisfies (C1) or (C2). Then for some s>0 W(F) M( F s + F 1 s ) for all F M 3 3 +.

Proof Let V M 3 3 be symmetric. For t 0 d dt W(etV ) = (D F W(e tv )e tv ) V From this it follows that = (e tv D F W(e tv )) V K(W(e tv )+1) V. W(e V )+1 (W(1)+1)e K V. Now set V = lnu, where U = U T > 0, and denote by v i the eigenvalues of U.

Since lnu =( it follows that 3 i=1 (lnv i ) 2 ) 1/2 3 i=1 lnv i, e K lnu (v K 1 +v K 3 3 3 ( C( 1 )(vk 2 +v K 2 )(vk 3 +v K 3 ) 3 3 K K 3 i=1 3 i=1 v K i + v 3K i + i=1 3 i=1 v K i ) 3 v 3K i ) C 1 [ U 3K + U 1 3K ], where C>0,C 1 >0 are constants.

We thus obtain W(U) M( U 3K + U 1 3K ), where M = C 1 (W(1)+1). The result now follows from the polar decomposition F = RU of an arbitrary F M+ 3 3, where R SO(3), U =U T >0.

IfW =Φ(v 1,v 2,v 3 )isisotropicthenboth(c1) and(c2) are equivalent(exercise) to the condition that (v 1 Φ,1,v 2 Φ, 2,v 3 Φ,3 ) K(Φ(v 1,v 2,v 3 )+1) for all v i > 0 and some K > 0, where Φ,i = Φ/ v i. Now for p 0,q 0 3 i=1 v i (v p v 1 +vp 2 +vp 3 ) i =p(vp 1 +vp 2 +vp 3 ), 3 i=1 v i ((v 2 v 3 ) q +(v 3 v 1 ) q +(v 1 v 2 ) q ) v i =2q((v 2 v 3 ) q +(v 3 v 1 ) q +(v 1 v 2 ) q )

And 3 i=1 v i h(v 1 v 2 v 3 ) v i =3v 1v 2 v 3 h (v 1 v 2 v 3 ). Hence both (C1) and (C2) hold for compressible Ogden materials if p i 0,q i 0,α i 0,β i 0, and h 0, for all δ>0. δh (δ) K 1 (h(δ)+1)

Exercise. Work out a corresponding treatment of weak forms of the Euler-Lagrange equation in the incompressible case.

Existence of minimizerswith body and surface forces Mixed displacement-traction problems. Suppose that Ω= Ω 1 Ω 2, where Ω 1 Ω 2 =. Considerthe deadload boundary conditions: y Ω1 = ȳ t R Ω2 = t R, where t R (x)=d F W(Dy(x))N(x) is the Piola- Kirchhoff stress vector, N(x) is the unit outward normal to Ω and t R L 1 ( Ω 2 ;R 3 ).

Suppose the body force b is conservative, so that b(y)= grad y Ψ(y), where Ψ = Ψ(y) is a real-valued potential. The most important example is gravity, for which b = ρ e, where e = (0,0,1), where which b = ρ R e 3, where e 3 = (0,0,1), where thedensityinthereferenceconfigurationρ R > 0 is constant. In this case we can take Ψ(y)= gy 3.

Consider the functional I(y)= Ω [W(Dy)+Ψ(y)]dx R yda. Ω 2 t Then formally a local minimizer y satisfies [D FW(Dy) Dϕ b(y) ϕ]dx R ϕda=0 Ω Ω2 t for all smooth ϕ with ϕ Ω1 =0, and thus DivD F W(Dy)+b = 0 in Ω, t R Ω2 = t R.

Theorem 19 Suppose that W satisfies (H1) and (H3) W(F) c 0 ( F 2 + coff 3/2 ) c 1 for all F M 3 3, where c 0 >0, (H4)W ispolyconvex,i.e. W(F)=g(F,cofF,detF) for all F M 3 3 for some continuous convex g. Assume further that Ψ is continuous and such that Ψ(y) d 0 y s d 1 for constants d 0 > 0,d 1 > 0,1 s < 2, and that t R L 2 ( Ω 2 ;R 3 ).

Assume that there exists some y in A={y W 1,1 (Ω;R 3 ):y Ω1 =ȳ} with I(y)<, where H 2 ( Ω 1 )>0 and ȳ: Ω 1 R 3 ismeasurable. Thenthereexists a global minimizer y of I in A. Sketch of proof. We need to get a bound on a minimizing sequence y (j). By the trace theorem there is a constant c 2 >0 such that Ω [ Dy 2 + y 2 ]dx d 0 y 2 da Ω 2 for all y W 1,2.

Using Lemma 1 and the boundary condition on Ω 1, we thus have for any y A, I(y) c 0 2 Ω Dy 2 dx+c 0 +m Ω y 2 dx d 0 Ω cofdy 3 2dx Ω y s dx 1 [ε 1 t 2 2 R +ε y ]da+const. 2 Ω2 Thus choosing a small ε I(y) a 0 Ω [ Dy 2 + y 2 + cofdy 3 2]dx a 1 for all y A and constants a 0 > 0,a 1, giving the necessary bound on y (j).

Pure traction problems. Suppose Ω 1 = and that b=b 0 is constant. Then choosing ϕ= const. we find that a necessary condition for a local minimum is that Ω b 0dx+ Ω 2 t R da=0, sayingthatthetotalappliedforceonthebody is zero. If this condition holds then I is invariant to the addition of constants, and it is convenient to remove this indeterminacy by minimizing I subject to the constraint Ω ydx=0.

Wethengettheexistenceofaminimizerunder the same hypotheses as Theorem 19, but using the Poincaré inequality Ω y 2 dx C ( Ω Dy 2 dx+ ( Ω ydx ) 2 ). It is also possible to treat mixed displacement pressure boundary conditions (see Ball 1977), which are conservative.

Invertibility Recall the Global Inverse Function Theorem, that if Ω R n is a bounded domain with Lipschitz boundary Ω and if y C 1 ( Ω;R n ) with detdy(x)>0 for all x Ω and y Ω one-to-one,theny isinvertibleon Ω. Can we prove a similar theorem for mappings in a Sobolev space?

Before discussing this question let us note an amusing example showing that failure of y to be C 1 at just two points of the boundary can invalidate the theorem. y Rubber sheet stuck to rigid wires Inner wire twisted through π about vertical axis Yellow region double covered

A. Weinstein, A global invertibility theorem for manifolds with boundary, Proc. Royal Soc. Edinburgh, 99 (1985) 283 284. shows that a local homeomorphism from a compact, connected manifold with boundary to a pact, connected manifold with boundary to a simply connected manifold without boundary is invertible if it is one-to-one on each component of the boundary.

Results for y W 1,p,p>n, (so that y is continuous). J.M. Ball, Global invertibility of Sobolev functions and the interpenetration of matter, Proc. Royal Soc. Edinburgh 90a(1981)315-328.

Theorem 20 Let Ω R n be a bounded domain with Lipschitz boundary. Let ȳ : Ω R n be continuous in Ω and one-to-one in Ω. Let p > n and let y W 1,p (Ω;R n ) satisfy y Ω = ȳ Ω, detdy(x)>0 a.e. in Ω. Then (i) y( Ω)=ȳ( Ω), (ii) y maps measurable sets in Ω to measurable sets in ȳ( Ω), and the change of variables formula A f(y(x))detdy(x)dx= ȳ(a) f(v)dv

holds for any measurable A Ω and any measurable f :R n R, provided one of the integrals exist, (iii) y is one-to-one a.e., i.e. the set S={v ȳ( Ω):y 1 (v) contains more than one element} has measure zero, (iv) if v ȳ(ω) then y 1 (v) is a continuum contained in Ω, while if v ȳ(ω) then each connectedcomponentofy 1 (v)intersects Ω.

y y 1 (v) y 1 ( v) v v Note that ȳ(ω) is open by invariance of domain. Proof of theorem uses degree theory and change of variables formula of Marcus and Mizel. Examples with complicated inverse images y 1 (v) can be constructed.

Theorem 21 Let the hypotheses of Theorem 20 hold, let ȳ(ω) satisfy the cone condition, and suppose that for some q>n Ω (Dy(x)) 1 q detdy(x)dx<. Then y is a homeomorphism of Ω onto ȳ(ω), and the inverse function x(y) belongs to W 1,q (ȳ(ω);r n ). The matrix of weak derivatives x( ) is given by Dx(v)=Dy(x(v)) 1 a.e. in ȳ(ω). If further ȳ(ω) is Lipschitz then y is a homeomorphism of Ω onto ȳ( Ω).

Note that formally we have Ω (Dy(x)) 1 q detdy(x)dx = ȳ(ω) Dx(v) q dv. Idea of proof. Get the inverse as the limit of a sequence of mollified mappings. Suppose x( ) is the inverse of y. Let ρ ε be a mollifier, i.e. ρ ε 0, suppρ ε B(0,ε), R n ρ ε (v)dv=1,anddefine x ε (v)= ȳ(ω) ρ ε(v u)x(u)du.

Changing variables we have x ε (v)= Ω ρ ε(v y(z))zdetdy(z)dz. In this way the mollified inverse is expressible directly in terms of y, and one can show that for any smooth domain D ȳ(ω) we have Dx ε(v) q dv M <, D where M is independent of sufficiently small ε. Then we can extract a weakly convergent subsequenceinw 1,p (D;R n )foreveryd,giving a candidate inverse.

For the pure displacement boundary-value problem with boundary condition y Ω =ȳ Ω for which the existence of minimizers was proved intheorem14,wegetthatanyminimizerisa homeomorphism provided we strengthen (H3) by assuming that W(F) c 0 ( F p + coff q +(detf) s ) c 1, where p>3,q>3,s> 2q q 3, and that ȳ satisfies the hypotheses of Theorem 21.

With this assumption we have that for any y A with I(y)<, Ω Dy(x) 1 σ detdy(x)dx = Ω cofdy(x) σ (detdy(x)) 1 σ dx c ( cofdy q +(detdy) (1 σ) ( q σ ) )dx Ω <, where σ= q(1+s) q+s >3, since (1 σ) ( q σ ) = s.

An interesting approach to the problem of invertibility(i.e. non-interpenetration of matter) in mixed boundary-value problems is given in P.G. Ciarlet and J. Nečas, Unilateral problems in nonlinear three-dimensional elasticity, Arch. Rational Mech. Anal., 87:(1985) 319 338. They proposed minimizing I(y)= W(Dy)dx Ω subject to the boundary condition y Ω1 = ȳ and the global constraint detdy(x)dx volume(y(ω)), Ω

If y W 1,p (Ω;R 3 ) with p>3 then a result of Marcus & Mizel says that Ω detdy(x)dx= y(ω) cardy 1 (v)dv. sothat theconstraintimpliesthat y isone-toone almost everywhere. They showed that IF the minimizer y is sufficiently smooth then this constraint corresponds to smooth self-contact.

They then proved the existence of minimizers satisfying the constraint for mixed boundary conditions under the growth condition W(F) c 0 ( F p + coff q +(detf) s ) c 1, with p > 3,q p p 1,s > 0. (The point is to show that the constraint is weakly closed.)

Results in the space A + p,q(ω)={y:ω R n ;Dy L p (Ω;M n n ), cofdy L q (Ω;M n n ),detdy(x)>0 a.e. in Ω}, following V. Šverák, Regularity properties of deformations with finite energy, Arch. Rat. Mech. Anal. 100(1988)105-127. For the results we are interested in Šverák assumes p>n 1,q p p 1, but Qi, Müller, Yan show his results go through with p>n 1,q n n 1, which we assume. Notice that then,since (detf)1=f(coff) T,wehave detf coff n 1, n so that detdy L 1 (Ω).

Infact,ifDy L n 1,cofDy L n n 1 thendetdy belongs to the Hardy space H 1 (Ω) (Iwaniec & Onninen 2002). If y A + p,q(ω) with p>n 1,q n 1 n then it is possible to define for every a Ω a set-valued image F(a,y), and thus the image F(A) of a subset A of Ω by F(A)= a A F(a,y).

Furthermore Theorem 22 Assume p n. (i)yhasarepresentativeỹwhichiscontinuous outside a singular set S of Hausdorff dimension n p. (ii) H n 1 (F(a))=0 for all a Ω. (iii) For each measurable A Ω, F(A) is measurable and L n (F(A)) detdy(x)dx. A In particular L n (F(S))=0.

We suppose that Ω is C and for simplicity thatȳisadiffeomorphismofsomeopenneighbourhood Ω 0 of Ω onto ȳ(ω 0 ). Now suppose that y A + p,q(ω) with y Ω =ȳ Ω. Given v ȳ( Ω), let G(v)={x Ω:v F(x)}. Thus G(v) consists of all inverse images of v.

Theorem 23 (i)foreachv ȳ( Ω)thesetG(v)isanonempty continuum in Ω. (ii) For each measurable A ȳ( Ω) the set G(A)= v A G(v) is measurable and L n (A)= detdy(x)dx. dx. G(A) (iii) Let T ={v ȳ( Ω);diamG(v)>0}. Then H n 1 (T)=0. Thus we can define the inverse function x(v) for all v T, and Šverák proves that x( ) W 1,1 (ȳ(ω)).

Regularity of minimizers Open Problem: Decide whether or not the global minimizer y in Theorem 14 is smooth. Here smooth means C in Ω, and C up to Here smooth means C in Ω, and C up to the boundary (except in the neighbourhood of points x 0 Ω 1 Ω 2 where singularities can be expected).

Clearly additional hypotheses on W are needed for this to be true. One might assume, for example, that W :M 3 3 + R is C, and that W is strictly polyconvex (i.e. that g is strictly convex). Also for regularity up to the boundary we would need to assume both smoothness of the boundary (except perhaps at Ω 1 Ω 2 ) and that ȳ is smooth. The precise nature of these extra hypotheses is to be determined.

The regularity is unsolved even in the simplest special cases. In fact the only situation in which smoothness of y seems to have been proved is for the pure displacement problem with small boundary displacements from a stress-free state. For this case Zhang (1991), stress-free state. For this case Zhang (1991), following work of Sivaloganathan, gave hypotheses under which the smooth solution to the equilibrium equations delivered by the implicit function theorem was in fact the unique global minimizer y of I given by Theorem 14.

An even more ambitious target would be to somehow classify possible singularities in minimizers of I for generic stored-energy functions W. If at the same time one could associate with each such singularity a condition on W that prevented it, one would also, by imposing all such conditions simultaneously, possess a set of hypotheses implying regularity. Itispossibletogoalittlewayinthisdirection.

Jumps in the deformation gradient y piecewise affine N Dy=A, x N >k Dy=B, x N <k x N =k A B=a N. When can such a y with A B be a weak solution of the equilibrium equations?

Theorem 24 Suppose that W :M+ 3 3 R has a local minimizer F 0. Then every piecewise affine map y as above is C 1 (i.e. A = B) iff W is strictly rank-one convex, i.e. the map t W(F +ta N) is strictly convex for each F M 3 3 and a R 3,N R 3. J.M. Ball, Strict convexity, strong ellipticity, and regularity in the calculus of variations. Proc. Camb. Phil. Soc., 87(1980)501 513.

Sufficiency. Suppose y is a weak solution. Then (DW(A) DW(B))N =0. Let θ(t)=w(b+ta N). Then θ (1)>θ (0) and so (DW(A) DW(B)) a N =a (DW(A) DW(B))N >0, a contradiction. The necessity use a characterization of strictly convexc 1 functionsdefinedonanopenconvex subset U of R n.

Theorem 25 A function ϕ C 1 (U) is strictly convex iff (i) there exists some z U with ϕ(w) ϕ(z)+ ϕ(z) (w z) for w z sufficiently small, and (ii) ϕ is one-to-one. The sufficiency follows by applying this to ϕ(a)=w(b+a N). the assumption implying that ϕ is one-toone.

Proof of Theorem 25 for the case when U = R n and ϕ(z) as z. z Let z R n. Let w R n minimize h(v)=ϕ(v) ϕ(z) v. By the growth condition w exists and so h(w)= ϕ(w) ϕ(z)=0. Hence w=z, the minimum is unique, and so ϕ(v)>ϕ(z)+ ϕ(z) (v z) for all v z, giving the strict convexity.

Discontinuities in y An example is cavitation, which we know is prevented(together with other discontinuities) e.g. by W(F) c 0 F n c 1.

Counterexamples to regularity

Quasiconvexity and partial regularity

In view of the above and other counterexamples for elliptic systems, if minimizers are smooth itmustbeforspecialreasonsapplyingtoelasticity. Plausible such reasons are: (a) the integrand depends only on Dy (and perhaps x), and not y perhaps x), and not y (b) the dimensions m=n=3 are low, (c) the frame-indifference of W. (d) invertibility of y.

2D incompressible example. Minimize I(y)= B Dy 2 dx, where B =B(0,1) is the unit disc in R 2, and y:b R 2, in the set of admissible mappings A={y W 1,2 (B;R 2 ):detdy=1 a.e.,y B =ȳ}, whereinpolarcoordinatesȳ:(r,θ) ( 1 2 r,2θ). Then there exists a global minimizer y of I in A. (Note that A is nonempty since ȳ A.) But since by degree theory there are no C 1 maps y satisfying the boundary condition, it is immediate that y is not C 1.

Solid phase transformations Displacive phase transformations are characterizedbyachangeofshapeinthecrystallattice at a critical temperature. e.g. cubic tetragonal

Energy minimization problem for single crystal

Frame-indifference requires ψ(rf,θ)=ψ(f,θ) for all R SO(3). If the material has cubic symmetry then also ψ(fq,θ)=ψ(f,θ) for all Q P 24, where P 24 is the group of rotations of a cube.

Energy-well structure K(θ)={F M 3 3 + Assume :F minimizes ψ(,θ)} austenite martensite Assuming the austenite has cubic symmetry, andgiventhetransformationstrainu 1 say,the N variantsu i arethedistinctmatricesqu 1 Q T, where Q P 24.

Cubic to tetragonal (e.g. Ni 65 Al 35 ) N =3 U 1 =diag(η 2,η 1,η 1 ) U 2 =diag(η 1,η 2,η 1 ) U 3 =diag(η 1,η 1,η 2 )

Exchange of stability

Atomistically sharp interfaces for cubic to tetragonal transformation in NiMn Baele, van Tenderloo, Amelinckx

Macrotwins in Ni 65 Al 35 involving two tetragonal variants (Boullay/Schryvers) 252

Martensitic microstructures in CuAlNi (Chu/James) 253

There are no rank-one connections between matrices A,B in the same energy well. In general there is no rank-one connection between A SO(3) and B SO(3)U i.

Theorem 26 Theorem 26 Let D = U 2 V 2 have eigenvalues λ 1 λ 2 λ 3. Then SO(3)U and SO(3)V are rank-one connected iff λ 2 = 0. There are exactly two solutions up to rotation provided λ 1 < λ 2 = 0<λ 3, and the corresponding N s are orthogonal iff tru 2 =trv 2, i.e. λ 1 = λ 3.

Gradient Young measures y (j) :Ω R m 257

The Young measure encodes the information on weak limits of all continuous functions of Dy (j k). Thus f(dy (j k) ) ν x,f. In particular Dy (j k) νx =Dy(x).

Simple laminate A B=a N Dy (j) λa+(1 λ)b Young measure ν x =λδ A +(1 λ)δ B

260

(Classical) austenite-martensite interface in CuAlNi (C-H Chu and R.D. James)

Gives formulae of the crystallographic theory of martensite (Wechsler, Lieberman, Read) 24 habit planes for cubicto-tetragonal