Skip to main content
Log in

Marginal Material Stability

  • Published:
Journal of Nonlinear Science Aims and scope Submit manuscript

Abstract

Marginal stability plays an important role in nonlinear elasticity because the associated minimally stable states usually delineate failure thresholds. In this paper we study the local (material) aspect of marginal stability. The weak notion of marginal stability at a point, associated with the loss of strong ellipticity, is classical. States that are marginally stable in the strong sense are located at the boundary of the quasi-convexity domain and their characterization is the main goal of this paper. We formulate a set of bounds for such states in terms of solvability conditions for an auxiliary nucleation problem formulated in the whole space and present nontrivial examples where the obtained bounds are tight.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

Notes

  1. Traditionally a strong local minimizer is associated with L topology. Our abuse of terminology should not cause problems, since in this paper we discuss only the necessary conditions. Clearly, all necessary conditions for a W 1,∞ sequential weak-* local minimizer will also be necessary for a L local minimizer.

  2. Variations can either be sequences as in the Definitions 2.1 and 2.3 or continuum families, such as G ϵ , where the limit as n→∞ is replaced by the limit as ϵ→0.

  3. The indicator function of a set equals zero on the set and +∞ on its complement.

  4. When d≥3 there is a canonical choice of the constant c, such that \(\phi - \boldsymbol{c}\in L^{2d/d-2}(\mathbb{R} ^{d};\mathbb{R} ^{m})\) (see Theorem A.1). When d=1 the choice of the constant is irrelevant. However, when d=2 there is no canonical choice of the constant c, which cannot be chosen arbitrarily.

  5. The functions in the much larger space \({\mathcal{S}}_{*}\) would possess infinitely many configurational degrees of freedom at infinity corresponding to the infinite variety of possible asymptotic behaviors of \(\phi \in{\mathcal{S}}_{*}\).

  6. In elasticity theory the tensors P(F) and P (F) are called the Piola–Kirchhoff tensor and the Eshelby tensor, respectively.

  7. By the regularity of the quasiconvex envelope theorem (Ball et al. 2000) the optimal fields must stay strictly away from the singularities of W(F). However, our results cannot guarantee that the field ε +e(ϕ) is indeed optimal.

References

  • Abeyaratne, R., Guo-Hua, J.: Dilatationally nonlinear elastic materials. I. Some theory. Int. J. Solids Struct. 25(10), 1201–1219 (1989)

    MATH  Google Scholar 

  • Agmon, S., Douglis, A., Nirenberg, L.: Estimates near the boundary for solutions of elliptic partial differential equations satisfying general boundary conditions. I. Commun. Pure Appl. Math. 12, 623–727 (1959)

    MathSciNet  MATH  Google Scholar 

  • Agmon, S., Douglis, A., Nirenberg, L.: Estimates near the boundary for solutions of elliptic partial differential equations satisfying general boundary conditions. II. Commun. Pure Appl. Math. 17, 35–92 (1964)

    MathSciNet  MATH  Google Scholar 

  • Alava, M., Nukala, P., Zapperi, S.: Statistical models of fracture. Adv. Phys. 55(3–4), 349–476 (2006)

    Google Scholar 

  • Allaire, G.: Shape Optimization by the Homogenization Method. Applied Mathematical Sciences, vol. 146. Springer, New York (2002)

    MATH  Google Scholar 

  • Allaire, G., Kohn, R.V.: Explicit optimal bounds on the elastic energy of a two-phase composite in two space dimensions. Q. Appl. Math. LI(4), 675–699 (1993a)

    MathSciNet  Google Scholar 

  • Allaire, G., Kohn, R.V.: Optimal bounds on the effective behavior of a mixture of two well-ordered elastic materials. Q. Appl. Math. LI(4), 643–674 (1993b)

    MathSciNet  Google Scholar 

  • Antman, S.S.: Nonlinear Problems of Elasticity, 2nd edn. Applied Mathematical Sciences, vol. 107. Springer, New York (2005)

    MATH  Google Scholar 

  • Baker, M., Ericksen, J.L.: Inequalities restricting the form of the stress-deformation relations for isotropic elastic solids and Reiner–Rivlin fluids. J. Wash. Acad. Sci. 44, 33–35 (1954)

    MathSciNet  Google Scholar 

  • Ball, J.M.: Convexity conditions and existence theorems in nonlinear elasticity. Arch. Ration. Mech. Anal. 63(4), 337–403 (1976/1977)

    Google Scholar 

  • Ball, J.: Strict convexity, strong ellipticity, and regularity in the calculus of variations. In: Mathematical Proceedings of the Cambridge Philosophical Society, vol. 87, pp. 501–513. Cambridge University Press, Cambridge (1980)

    Google Scholar 

  • Ball, J.M.: Discontinuous equilibrium solutions and cavitation in nonlinear elasticity. Philos. Trans. R. Soc. Lond. Ser. A 306(1496), 557–611 (1982)

    MATH  Google Scholar 

  • Ball, J.M.: A version of the fundamental theorem for Young measures. In: PDEs and Continuum Models of Phase Transitions, Nice, 1988. Lecture Notes in Physics, vol. 344, pp. 207–215. Springer, Berlin (1989)

    Google Scholar 

  • Ball, J.M.: Some open problems in elasticity. In: Geometry, Mechanics, and Dynamics, pp. 3–59. Springer, New York (2002)

    Google Scholar 

  • Ball, J.M., Marsden, J.E.: Quasiconvexity at the boundary, positivity of the second variation and elastic stability. Arch. Ration. Mech. Anal. 86(3), 251–277 (1984)

    MathSciNet  MATH  Google Scholar 

  • Ball, J.M., Murat, F.: Remarks on Chacon’s biting lemma. Proc. Am. Math. Soc. 107(3), 655–663 (1989)

    MathSciNet  MATH  Google Scholar 

  • Ball, J.M., Kirchheim, B., Kristensen, J.: Regularity of quasiconvex envelopes. Calc. Var. Partial Differ. Equ. 11(4), 333–359 (2000)

    MathSciNet  MATH  Google Scholar 

  • Ball, J., Koumatos, K., Seiner, H.: Nucleation of austenite in mechanically stabilized martensite by localized heating. J. Alloys Compd. (2011). doi:10.1016/j.jallcom.2011.11.070

    Google Scholar 

  • Bates, P.W., Fife, P.C.: The dynamics of nucleation for the Cahn–Hilliard equation. SIAM J. Appl. Math. 53(4), 990–1008 (1993)

    MathSciNet  MATH  Google Scholar 

  • Berdichevsky, V.L.: Seed of a melt in a solid. Dokl. Akad. Nauk SSSR 27, 80–84 (1983)

    Google Scholar 

  • Berdichevsky, V.L.: Variational Principles of Continuum Mechanics. I. Fundamentals. Springer, Berlin (2009a)

    MATH  Google Scholar 

  • Berdichevsky, V.L.: Variational Principles of Continuum Mechanics. II. Applications. Springer, Berlin (2009b)

    MATH  Google Scholar 

  • Biot, M.A.: Mechanics of Incremental Deformations. Wiley, New York (1965)

    Google Scholar 

  • Budiansky, B., Hutchinson, J., Lambropoulos, J.: Continuum theory of dilatant transformation toughening in ceramics. Int. J. Solids Struct. 19(4), 337–355 (1983)

    MATH  Google Scholar 

  • Chenchiah, I., Bhattacharya, K.: The relaxation of two-well energies with possibly unequal moduli. Arch. Ration. Mech. Anal. 187(3), 409–479 (2008)

    MathSciNet  MATH  Google Scholar 

  • Cherepanov, G.P.: Inverse problems of the plane theory of elasticity. J. Appl. Math. Mech. 38(6), 963–979 (1974)

    MathSciNet  Google Scholar 

  • Cherkaev, A.: Variational Methods for Structural Optimization. Springer, New York (2000)

    MATH  Google Scholar 

  • Cherkaev, A., Kucuk, I.: Detecting stress fields in an optimal structure. I. Two-dimensional case and analyzer. Struct. Multidiscip. Optim. 26(1–2), 1–15 (2004a)

    MathSciNet  MATH  Google Scholar 

  • Cherkaev, A., Kucuk, I.: Detecting stress fields in an optimal structure. II. Three-dimensional case. Struct. Multidiscip. Optim. 26(1–2), 16–27 (2004b)

    MathSciNet  MATH  Google Scholar 

  • Cherkaev, A., Zhang, Y.: Optimal anisotropic three-phase conducting composites: plane problem. Int. J. Solids Struct. 48(20), 2800–2813 (2011)

    Google Scholar 

  • Chipot, M., Kinderlehrer, D.: Equilibrium configurations of crystals. Arch. Ration. Mech. Anal. 103(3), 237–277 (1988)

    MathSciNet  MATH  Google Scholar 

  • Ciarlet, P.G.: Mathematical Elasticity. I. Three-Dimensional Elasticity. Studies in Mathematics and its Applications., vol. 20. North-Holland, Amsterdam (1988)

    Google Scholar 

  • Crandall, M.G., Rabinowitz, P.H.: Bifurcation from simple eigenvalues. J. Funct. Anal. 8, 321–340 (1971)

    MathSciNet  MATH  Google Scholar 

  • Dacorogna, B.: Quasiconvexity and relaxation of nonconvex problems in the calculus of variations. J. Funct. Anal. 46(1), 102–118 (1982)

    MathSciNet  MATH  Google Scholar 

  • Dacorogna, B.: Necessary and sufficient conditions for strong ellipticity of isotropic functions in any dimension. Discrete Contin. Dyn. Syst., Ser. B 1(2), 257–263 (2001)

    MathSciNet  MATH  Google Scholar 

  • Dolzmann, G.: Variational Methods for Crystalline Microstructure: Analysis and Computation. Lecture Notes in Mathematics, vol. 1803. Springer, Berlin (2003)

    Google Scholar 

  • Edelstein, W.S., Fosdick, R.L.: A note on non-uniqueness in linear elasticity theory. Z. Angew. Math. Phys. 19, 906–912 (1968). doi:10.1007/BF01602270

    MathSciNet  MATH  Google Scholar 

  • Erdmann, G.: Über die unstetige Lösungen in der Variationsrechnung. J. Reine Angew. Math. 82, 21–30 (1877)

    Google Scholar 

  • Ericksen, J.L., Toupin, R.A.: Implications of Hadamard’s conditions for elastic stability with respect to uniqueness theorems. Can. J. Math. 8, 432–436 (1956)

    MathSciNet  MATH  Google Scholar 

  • Eshelby, J.D.: The determination of the elastic field of an ellipsoidal inclusion, and related problems. Proc. R. Soc. Lond. Ser. A, Math. Phys. Sci. 241, 376–396 (1957)

    MathSciNet  MATH  Google Scholar 

  • Eshelby, J.D.: The elastic field outside an ellipsoidal inclusion. Proc. R. Soc. Lond. Ser. A, Math. Phys. Sci. 252, 561–569 (1959)

    MathSciNet  MATH  Google Scholar 

  • Eshelby, J.D.: Energy relations and energy momentum tensor in continuum mechanics. In: Kanninen, M., Adler, W., Rosenfeld, A., Jaffee, R. (eds.) Inelastic Behavior of Solids, pp. 77–114. McGraw-Hill, New York (1970)

    Google Scholar 

  • Eshelby, J.D.: The elastic energy-momentum tensor. J. Elast. 5(3–4), 321–335 (1975)

    MathSciNet  MATH  Google Scholar 

  • Freidin, A.B.: On new phase inclusions in elastic solids. Z. Angew. Math. Mech. 87(2), 102–116 (2007)

    MathSciNet  MATH  Google Scholar 

  • Freidin, A.B., Chiskis, A.M.: Zones of phase transitions in non-linear elastic isotropic materials. Part 1. Basic relations. Izv. Akad. Nauk SSSR, Meh. Tverd. Tela 4, 91–109 (1994a) (in Russian)

    Google Scholar 

  • Freidin, A.B., Chiskis, A.M.: Zones of phase transitions in non-linear elastic isotropic materials. Part 2. Incompressible materials with a potentials solely dependent on one of the invariants of the strain tensor. Izv. Akad. Nauk SSSR, Meh. Tverd. Tela 5, 49–61 (1994b) (in Russian)

    Google Scholar 

  • Gelfand, I.M., Fomin, S.V.: Calculus of Variations. Prentice Hall, New York (1963)

    Google Scholar 

  • Geymonat, G., Müller, S., Triantafyllidis, N.: Homogenization of non-linearly elastic materials, microscopic bifurcation and macroscopic loss of rank-one convexity. Arch. Ration. Mech. Anal. 122(3), 231–290 (1993)

    MATH  Google Scholar 

  • Giaquinta, M., Hildebrandt, S.: Calculus of Variations. I. The Lagrangian Formalism. Grundlehren der Mathematischen Wissenschaften, vol. 310. Springer, Berlin (1996)

    Google Scholar 

  • Gibiansky, L.V., Cherkaev, A.V.: Design of composite plates of extremal rigidity. Report 914, Ioffe Physicotechnical Institute, Leningrad, USSR (1984)

  • Gibiansky, L.V., Cherkaev, A.V.: Microstructures of composites of extremal rigidity and exact estimates of the associated energy density. Report 1115, Ioffe Physicotechnical Institute, Leningrad, USSR (1987)

  • Grabovsky, Y.: Nonsmooth analysis and quasi-convexification in elastic energy minimization problems. Struct. Optim. 10(3/4), 217–221 (1995)

    Google Scholar 

  • Grabovsky, Y.: Bounds and extremal microstructures for two-component composites: a unified treatment based on the translation method. Proc. R. Soc. Lond. Ser. A, Math. Phys. Sci. 452(1947), 945–952 (1996)

    Google Scholar 

  • Grabovsky, Y., Kohn, R.V.: Anisotropy of the Vigdergauz microstructure. J. Appl. Mech. 62(4), 1063–1065 (1995a)

    MATH  Google Scholar 

  • Grabovsky, Y., Kohn, R.V.: Microstructures minimizing the energy of a two phase elastic composite in two space dimensions. I. The confocal ellipse construction. J. Mech. Phys. Solids 43(6), 933–947 (1995b)

    MathSciNet  MATH  Google Scholar 

  • Grabovsky, Y., Truskinovsky, L.: The flip side of buckling. Contin. Mech. Thermodyn. 19(3–4), 211–243 (2007)

    MathSciNet  MATH  Google Scholar 

  • Grabovsky, Y., Truskinovsky, L.: Roughening instability of broken extremals. Arch. Ration. Mech. Anal. 200(1), 183–202 (2011)

    MathSciNet  MATH  Google Scholar 

  • Grabovsky, Y., Truskinovsky, L.: Generalized Clapeyron’s theorem (2013a, in preparation)

  • Grabovsky, Y., Truskinovsky, L.: Metastability in an elastic material with incompatible energy wells: an example. (2013b, in preparation)

  • Grabovsky, Y., Truskinovsky, L.: Metastability in the presence of phase boundaries. (2013c, in preparation)

  • Grabovsky, Y., Kucher, V., Truskinovsky, L.: Probing the limits of rank-one convexity (2013, in preparation)

  • Gurtin, M.E.: Configurational Forces as Basic Concepts of Continuum Physics. Applied Mathematical Sciences, vol. 137. Springer, New York (2000)

    Google Scholar 

  • Gutiérrez, S.: Laminations in linearized elasticity: the isotropic non-very strongly elliptic case. J. Elast. 53(3), 215–256 (1998/1999)

    Google Scholar 

  • Hill, R.: On uniqueness and stability in the theory of finite elastic strain. J. Mech. Phys. Solids 5, 229–241 (1957)

    MathSciNet  MATH  Google Scholar 

  • Hill, R.: On the elasticity and stability of perfect crystals at finite strain. Math. Proc. Camb. Philos. Soc. 77, 225–240 (1975)

    MATH  Google Scholar 

  • Hohlfeld, E., Mahadevan, L.: Unfolding the sulcus. Phys. Rev. Lett. 106, 105702 (2011)

    Google Scholar 

  • Kaganova, I.M., Roytburd, A.L.: Equilibrium between elastically-interacting phases. Sov. Phys. JETP 67(6), 1173–1183 (1988)

    Google Scholar 

  • Khachaturyan, A.G.: Theory of Structural Transformation in Solids. Wiley, New York (1983)

    Google Scholar 

  • Kinderlehrer, D.: Remarks about equilibrium configurations of crystals. In: Material Instabilities in Continuum Mechanics, Edinburgh, 1985–1986. Oxford Science Publications, pp. 217–241. Oxford University Press, New York (1988)

    Google Scholar 

  • Kinderlehrer, D., Pedregal, P.: Characterizations of Young measures generated by gradients. Arch. Ration. Mech. Anal. 115(4), 329–365 (1991)

    MathSciNet  MATH  Google Scholar 

  • Knowles, J.K., Sternberg, E.: On the ellipticity of the equations of nonlinear elastostatics for a special material. J. Elast. 5, 341–361 (1975)

    MathSciNet  MATH  Google Scholar 

  • Knowles, J.K., Sternberg, E.: On the failure of ellipticity and the emergence of discontinuous deformation gradients in plane finite elastostatics. J. Elast. 8(4), 329–379 (1978)

    MathSciNet  MATH  Google Scholar 

  • Knüpfer, H., Kohn, R.V.: Minimal energy for elastic inclusions. Proc. R. Soc. A, Math. Phys. Eng. Sci. 467(2127), 695–717 (2011)

    MATH  Google Scholar 

  • Knüpfer, H., Kohn, R., Otto, F.: Nucleation barriers for the cubic-to-tetragonal transformation. Commun. Pure Appl. Math. 66(6), 867–904 (2013)

    MATH  Google Scholar 

  • Kohn, R.V.: The relaxation of a double-well energy. Contin. Mech. Thermodyn. 3, 193–236 (1991)

    MathSciNet  MATH  Google Scholar 

  • Kohn, R.V., Strang, G.: Optimal design and relaxation of variational problems. Commun. Pure Appl. Math. 39(113–137), 139–182 (1986). 353–377

    MathSciNet  MATH  Google Scholar 

  • Kristensen, J.: On the non-locality of quasiconvexity. Ann. Inst. Henri Poincaré, Anal. Non Linéaire 16(1), 1–13 (1999)

    MathSciNet  MATH  Google Scholar 

  • Kublanov, L.B., Freidin, A.B.: Nuclei of a solid phase in a deformable material. Prikl. Mat. Meh. 52(3), 493–501 (1988)

    MathSciNet  Google Scholar 

  • Kunin, I., Sosnina, É.: Ellipsoidal inhomogeneity in an elastic medium. Sov. Phys. Dokl. 16, 534 (1971)

    MATH  Google Scholar 

  • Kunin, I., Sosnina, E.: Stress concentration on an ellipsoidal inhomogeneity in an anisotropic elastic medium. Prikl. Mat. Meh. 37, 306–315 (1973)

    Google Scholar 

  • Kunin, I., Mirenkova, G., Sosnina, E.: An ellipsoidal crack and needle in an anisotropic elastic medium. Prikl. Mat. Meh. 37(3), 524–531 (1973). See also in J. Appl. Math. Mech. 37(3), 501–508 (1973)

    Google Scholar 

  • Langer, J.: Metastable states. Physica 73(1), 61–72 (1974)

    Google Scholar 

  • Le Dret, H.: An example of H 1-unboundedness of solutions to strongly elliptic systems of partial differential equations in a laminated geometry. Proc. R. Soc. Edinb. A 105, 77–82 (1987)

    MATH  Google Scholar 

  • Li, J., Zhu, T., Yip, S., Vliet, K.J.V., Suresh, S.: Elastic criterion for dislocation nucleation. Mater. Sci. Eng. A, Struct. Mater.: Prop. Microstruct. Process. 365(1–2), 25–30 (2004)

    Google Scholar 

  • Lifshits, I.M., Gulida, L.S.: On nucleation under local melting. Dokl. Akad. Nauk SSSR 87(4), 523–526 (1952a) (in Russian, English version available)

    Google Scholar 

  • Lifshits, I.M., Gulida, L.S.: On the theory of local melting. Dokl. Akad. Nauk SSSR 87(3), 377–380 (1952b) (in Russian, English version available)

    Google Scholar 

  • Lu, J.: Elastic energy minimization and the shape of coherent precipitates. Ph.D. Thesis, New York University, New York, NY (1993)

  • Lurie, K.A.: Optimum control of conductivity of a fluid moving in a channel in a magnetic field. J. Appl. Math. Mech. 28(2), 316–327 (1964)

    MathSciNet  Google Scholar 

  • Lurie, K.A.: Applied Optimal Control Theory of Distributed Systems. Plenum, New York (1993)

    MATH  Google Scholar 

  • Maloney, C., Lemaître, A.: Universal breakdown of elasticity at the onset of material failure. Phys. Rev. Lett. 93(19), 195501 (2004)

    Google Scholar 

  • Maugin, G.A.: Material Inhomogeneities in Elasticity. Chapman & Hall, London (1993)

    MATH  Google Scholar 

  • Michel, J.C., Lopez-Pamies, O., Ponte Castañeda, P., Triantafyllidis, N.: Microscopic and macroscopic instabilities in finitely strained porous elastomers. J. Mech. Phys. Solids 55(5), 900–938 (2007)

    MathSciNet  MATH  Google Scholar 

  • Mielke, A., Sprenger, P.: Quasiconvexity at the boundary and a simple variational formulation of Agmon’s condition. J. Elast. 51(1), 23–41 (1998)

    MathSciNet  MATH  Google Scholar 

  • Morrey, C.B. Jr.: Quasi-convexity and the lower semicontinuity of multiple integrals. Pac. J. Math. 2, 25–53 (1952)

    MathSciNet  MATH  Google Scholar 

  • Morse, M.: The Calculus of Variations in the Large vol. 18. Am. Math. Soc., Providence (1934)

    MATH  Google Scholar 

  • Müller, S., Šverák, V.: Convex integration for Lipschitz mappings and counterexamples to regularity. Ann. Math. 157(3), 715–742 (2003)

    MATH  Google Scholar 

  • Müller, S., Sivaloganathan, J., Spector, S.J.: An isoperimetric estimate and W 1,p-quasiconvexity in nonlinear elasticity. Calc. Var. Partial Differ. Equ. 8(2), 159–176 (1999)

    MATH  Google Scholar 

  • Mura, T.: Micromechanics of Defects in Solids. Springer, Berlin (1987)

    Google Scholar 

  • Muskhelishvili, N.I.: Some Basic Problems of the Mathematical Theory of Elasticity. Noordhoff, Groningen (1953)

    MATH  Google Scholar 

  • Noether, E.: Invariante Variationsprobleme. In: Nachr. v. d. Ges. d. Wiss. zu Göttingen, vol. 1, pp. 235–257 (1918). English translation in Transport Theory and Statistical Mechanics, pp. 183–207 (1971)

    Google Scholar 

  • Ogden, R.: Non-linear Elastic Deformations. Dover, New York (1997)

    Google Scholar 

  • Pedregal, P.: Parametrized Measures and Variational Principles. Progress in Nonlinear Differential Equations and their Applications, vol. 30. Birkhäuser, Basel (1997)

    MATH  Google Scholar 

  • Pedregal, P.: Fully explicit quasiconvexification of the mean-square deviation of the gradient of the state in optimal design. Electron. Res. Announc. Am. Math. Soc. 7, 72–78 (2001) (electronic)

    MathSciNet  MATH  Google Scholar 

  • Pericak-Spector, K.A., Sivaloganathan, J., Spector, S.J.: An explicit radial cavitation solution in nonlinear elasticity. Math. Mech. Solids 7(1), 87–93 (2002)

    MathSciNet  MATH  Google Scholar 

  • Puglisi, G., Truskinovsky, L.: Thermodynamics of rate-independent plasticity. J. Mech. Phys. Solids 53(3), 655–679 (2005)

    MathSciNet  MATH  Google Scholar 

  • Roytburd, A., Slutsker, J.: Deformation of adaptive materials. Part I: Constrained deformation of polydomain crystals. J. Mech. Phys. Solids 47, 2299–2329 (1999a)

    MathSciNet  MATH  Google Scholar 

  • Roytburd, A., Slutsker, J.: Deformation of adaptive materials. Part II: Adaptive composite. J. Mech. Phys. Solids 47, 2331–2349 (1999b)

    MathSciNet  MATH  Google Scholar 

  • Roytburd, A., Slutsker, J.: Deformation of adaptive materials. Part III: Deformation of crystals with polytwin product phases. J. Mech. Phys. Solids 49(8), 1795–1822 (2001)

    MATH  Google Scholar 

  • Salman, O., Truskinovsky, L.: On the critical nature of plastic flow: one and two dimensional models. Int. J. Eng. Sci. 59, 219–254 (2012)

    Google Scholar 

  • Sethna, J.: Crackling noise and avalanches: scaling, critical phenomena, and the renormalization group. Les Houches 85, 257–288 (2007)

    Google Scholar 

  • Šilhavý, M.: The Mechanics and Thermodynamics of Continuous Media. Springer, Berlin (1997)

    MATH  Google Scholar 

  • Simpson, H.C., Spector, S.J.: On barrelling instabilities in finite elasticity. J. Elast. 14(2), 103–125 (1984)

    MathSciNet  MATH  Google Scholar 

  • Simpson, H.C., Spector, S.J.: On the positivity of the second variation in finite elasticity. Arch. Ration. Mech. Anal. 98(1), 1–30 (1987)

    MathSciNet  MATH  Google Scholar 

  • Simpson, H.C., Spector, S.J.: Necessary conditions at the boundary for minimizers in finite elasticity. Arch. Ration. Mech. Anal. 107(2), 105–125 (1989)

    MathSciNet  MATH  Google Scholar 

  • Simpson, H.C., Spector, S.J.: On bifurcation in finite elasticity: buckling of a rectangular rod. J. Elast. 92(3), 277–326 (2008)

    MathSciNet  MATH  Google Scholar 

  • Sokolowski, J., Zolesio, J.-P.: Introduction to Shape Optimization. Shape Sensitivity Analysis. Springer, Berlin (1992)

    MATH  Google Scholar 

  • Stein, E.M.: Singular Integrals and Differentiability Properties of Functions. Princeton Mathematical Series, vol. 30. Princeton University Press, Princeton (1970)

    MATH  Google Scholar 

  • Stuart, C.A.: Radially symmetric cavitation for hyperelastic materials. Ann. Inst. Henri Poincaré, Anal. Non Linéaire 2(1), 33–66 (1985)

    MathSciNet  MATH  Google Scholar 

  • Tanaka, T., Sun, S.-T., Hirokawa, Y., Katayama, S., Kucera, J., Hirose, Y., Amiya, T.: Mechanical instability of gels at the phase transition. Nature 325(6107), 796–798 (1987). 02

    Google Scholar 

  • Tartar, L.: A personalized introduction. In: The General Theory of Homogenization. Lecture Notes of the Unione Matematica Italiana, vol. 7. Springer, Berlin (2009)

    Google Scholar 

  • Truesdell, C., Noll, W.: The Non-linear Field Theories of Mechanics, 3rd edn. Springer, Berlin (2004)

    Google Scholar 

  • Truskinovsky, L.M.: Dynamics of nonequilibrium phase boundaries in a heat conducting non-linearly elastic medium. Prikl. Mat. Meh. 51(6), 1009–1019 (1987)

    MathSciNet  Google Scholar 

  • Truskinovsky, L., Zanzotto, G.: Ericksen’s bar revisited: energy wiggles. J. Mech. Phys. Solids 44(8), 1371–1408 (1996)

    MathSciNet  Google Scholar 

  • van der Waals, J.: The equilibrium between a solid body and a fluid phase, especially in the neighbourhood of the critical state. In: KNAW Proceedings, vol. 6, pp. 1903–1904 (1903)

    Google Scholar 

  • Van Hove, L.: Sur l’extension de la condition de Legendre du calcul des variations aux intégrales multiples à plusieurs fonctions inconnues. Proc. K. Ned. Akad. Wet. 50, 18–23 (1947). See also in Indag. Math. 9, 3–8 (1947)

    MATH  Google Scholar 

  • Vigdergauz, S.B.: Integral equation of the inverse problem of the plane theory of elasticity. Prikl. Mat. Meh. 40(3), 518–521 (1976)

    MATH  Google Scholar 

  • Vigdergauz, S.B.: On a case of the inverse problem of two-dimensional theory of elasticity. Prikl. Mat. Meh. 41(5), 927–933 (1977)

    MathSciNet  Google Scholar 

  • Vigdergauz, S.B.: Regular structures with extremal elastic properties. Mech. Solids 24(3), 57–63 (1989)

    Google Scholar 

  • Vigdergauz, S.B.: Two-dimensional grained composites of extreme rigidity. J. Appl. Mech. 61(2), 390–394 (1994)

    MATH  Google Scholar 

  • Young, L.C.: Generalized curves and the existence of an attained absolute minimum in the calculus of variations. C. R. Soc. Sci. Varsovie, Cl. III 30, 212–234 (1937)

    MATH  Google Scholar 

  • Young, L.C.: Generalized surfaces in the calculus of variations. Ann. Math. 43(1), 84–103 (1942)

    Google Scholar 

  • Young, L.C.: Lectures on the Calculus of Variations and Optimal Control Theory. Saunders, Philadelphia (1969). Foreword by Wendell H. Fleming

    MATH  Google Scholar 

  • Zaiser, M.: Scale invariance in plastic flow of crystalline solids. Adv. Phys. 55(1–2), 185–245 (2006)

    Google Scholar 

  • Zee, L., Sternberg, E.: Ordinary and strong ellipticity in the equilibrium theory of incompressible hyperelastic solids. Arch. Ration. Mech. Anal. 83, 53–90 (1983). doi:10.1007/BF00281087

    MathSciNet  MATH  Google Scholar 

  • Zhang, Z., James, R.D., Müller, S.: Energy barriers and hysteresis in martensitic phase transformations. Acta Mater. 57(15), 4332–4352 (2009)

    Google Scholar 

Download references

Acknowledgements

We thank R. Fosdick, R. Kohn, and anonymous reviewers for helpful comments. This material is based upon work supported by the National Science Foundation under Grant No. 1008092 and the French ANR grant EVOCRIT (2008–2012).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yury Grabovsky.

Additional information

Communicated by Paul Newton.

Appendices

Appendix A: Proof of Lemma 3.3

We note that \({\mathcal{S}}=W^{1,\infty}(\mathbb{R} ^{d};\mathbb{R} ^{m})\cap {\mathcal{S}} _{0}\). It will be important to use the following embedding theorem for the space \({\mathcal{S}}_{0}\).

Theorem A.1

Assume that \(\phi \in{\mathcal{S}}_{0}\) and d≥3. Then there exists a unique constant \(\boldsymbol{c}\in\mathbb{R} ^{m}\) such that \(\phi - \boldsymbol{c}\in L^{\frac{2d}{d-2}}(\mathbb{R} ^{d};\mathbb{R} ^{m})\).

Proof

First, we remark that without loss of generality we may take m=1. Now recall the well-known potential theory operators. The Riesz transforms R j are defined by

$${\mathcal{F}}(R_{j}f) (\xi )=i\frac{\xi_{j}}{|\xi |}\widehat{f}(\xi ), $$

where

$${\mathcal{F}}(f) (\xi )=\widehat{f}(\xi )=\int_{\mathbb{R} ^{d}}f( \boldsymbol{x} )\mathrm{e}^{2\pi i( \boldsymbol{x} ,\xi )}\,\mathrm{d}\boldsymbol{x} $$

is the Fourier transform. The operators R j map \(L^{2}(\mathbb{R} ^{d})\) into \(L^{2}(\mathbb{R} ^{d})\). The Riesz potential I 1 is defined by

$${\mathcal{F}}(I_{1}f) (\xi )=\frac{\widehat{f}(\xi )}{2\pi|\xi |}. $$

It maps \(L^{2}(\mathbb{R} ^{d})\) into \(L^{\frac{2d}{d-2}}(\mathbb{R} ^{d})\), (Stein 1970).

If ϕ were smooth and compactly supported, we would have

$$\phi=I_{1} \Biggl(\sum_{j=1}^{d}R_{j} \biggl(\frac{\partial\phi }{\partial x_{j}} \biggr) \Biggr). $$

Hence, we define

$$\psi( \boldsymbol{x})=I_{1} \Biggl(\sum_{j=1}^{d}R_{j}(g_{j}) \Biggr), $$

where \(\boldsymbol{g}=\nabla\phi\in L^{2}(\mathbb{R} ^{d};\mathbb{R} ^{d})\).

Let η(x) be an arbitrary smooth compactly supported function. By definition of the distributional derivative we have

$$\int_{\mathbb{R} ^{d}} \biggl\{g_{k}\frac{\partial\eta}{\partial x_{j}}-g_{j} \frac{\partial\eta}{\partial x_{k}} \biggr\}\,\mathrm{d}\boldsymbol{x}= -\biggl\langle\phi,\frac{\partial^2 \eta}{\partial x_{k}\partial x_{j}} \biggr\rangle+\biggl\langle\phi,\frac{\partial^2 \eta}{\partial x_{j}\partial x_{k}} \biggr\rangle=0. $$

By Plancherel’s identity

$$\int_{\mathbb{R} ^{d}}(\widehat{g}_{k}\xi_{j}- \widehat{g}_{j}\xi _{k})\overline{\widehat{\eta}}\, \mathrm{d}\xi =0. $$

We conclude that

$$ \widehat{g}_{k}(\xi )\xi_{j}= \widehat{g}_{j}(\xi )\xi_{k} $$
(A.1)

for a.e. \(\xi \in\mathbb{R} ^{d}\). Thus,

$$-2\pi i\xi_{k}\widehat{\psi}(\xi )=\sum_{j=1}^{d} \frac{\xi _{k}\xi _{j}\widehat{g}_{j}(\xi )}{|\xi |^{2}} =\widehat{g}_{k}(\xi ), $$

due to (A.1). By Plancherel’s identity

$$\int_{\mathbb{R} ^{d}}\psi( \boldsymbol{x})\frac{\partial\eta}{\partial x_{k}}\,\mathrm{d}\boldsymbol{x}= 2\pi i \int_{\mathbb{R} ^{d}}\xi_{k}\widehat{\psi}\,\overline{\widehat {\eta}}\,\mathrm{d}\xi = -\int_{\mathbb{R} ^{d}}\widehat{g}_{k}( \xi )\overline{\widehat{\eta }}\,\mathrm{d}\xi = -\int_{\mathbb{R} ^{d}}g_{k}( \boldsymbol{x})\eta( \boldsymbol{x})\,\mathrm{d}\boldsymbol{x}. $$

Therefore, ∇ψ=g=∇ϕ as distributions. The theorem is proved. □

The proof of Lemma 3.3 proceeds in different ways depending on the dimension d. If d=1, then

Let d=2. We estimate

$$\frac{1}{R}\int_{A_{R}}|\phi ||\nabla\phi |\, \mathrm{d}\boldsymbol{x}\le\|\phi \| _{\infty }\sqrt{3\pi} \biggl(\int _{A_{R}}|\nabla\phi |^{2}\,\mathrm{d}\boldsymbol{x} \biggr)^{1/2}\to0,\quad\text{as }R\to\infty. $$

By the Poincaré inequality

$$\int_{A_{R}}\big|\phi ( \boldsymbol{x})-\langle\phi \rangle_{A_{R}}\big|^{2} \,\mathrm{d}\boldsymbol{x}\le C_{0}R^{2}\int_{A_{R}}| \nabla\phi |^{2}\,\mathrm{d}\boldsymbol{x}, $$

where C 0 is the Poincaré constant for A 1. The boundedness of ϕ implies that there exists a sequence R=R k such that

$$\lim_{k\to\infty}\langle\phi \rangle_{A_{R_{k}}}= \boldsymbol{c}. $$

Hence, by the triangle inequality

$$\biggl(\int_{A_{R}}|\phi - \boldsymbol{c}|^{2}\,\mathrm{d}\boldsymbol{x} \biggr)^{1/2}\le \biggl(\int_{A_{R}}\big|\phi -\langle\phi \rangle_{A_{R}}\big|^{2}\,\mathrm{d}\boldsymbol{x} \biggr)^{1/2}+ |A_{R}|^{1/2}\big|\langle\phi \rangle_{A_{R}}- \boldsymbol{c}\big|. $$

Then,

Now assume that d≥3. By Theorem A.1 there exists a unique constant c, such that \(\phi - \boldsymbol{c}\in L^{\frac{2d}{d-2}}(\mathbb{R} ^{d};\mathbb{R} ^{m})\). Using the inequality ab≤(a 2+b 2)/2 we get

$$\frac{1}{R}\int_{A_{R}}|\phi - \boldsymbol{c}||\nabla\phi |\, \mathrm{d}\boldsymbol{x}+\frac {1}{R^{2}}\int_{A_{R}}|\phi - \boldsymbol{c}|^{2}\,\mathrm{d}\boldsymbol{x}\le \frac{1}{2}\int_{A_{R}}| \nabla\phi |^{2}\,\mathrm{d}\boldsymbol{x}+\frac{3}{2R^{2}}\int _{A_{R}}|\phi - \boldsymbol{c}|^{2}\,\mathrm{d}\boldsymbol{x}. $$

By Hölder inequality

$$\frac{1}{R^{2}}\int_{A_{R}}|\phi - \boldsymbol{c}|^{2}\, \mathrm{d}\boldsymbol{x}\le \biggl(\int_{A_{R}}|\phi - \boldsymbol{c}|^{\frac{2d}{d-2}}\, \mathrm{d}\boldsymbol{x} \biggr)^{\frac {d-2}{d}}\to 0,\quad\text{as }R\to\infty. $$

The lemma is proved.

Appendix B: Proof of Lemma 3.7

Let us begin with a technical lemma.

Lemma B.1

Suppose α(R)>0 is such that α(R)→0, as R→∞. Then there exists a monotonically increasing function h(R) with h(R)/R→0, as R→∞, such that

$$ \lim_{R\to\infty} \biggl(\frac{R}{h(R)} \biggr) \alpha(R)=0. $$
(B.1)

Proof

We define

$$h(R)=\max_{r<R} \bigl(r\sqrt{\alpha(r)} \bigr). $$

Then h(R) is monotonically increasing and h(R)/R→0, as R→∞. Indeed, for any ϵ>0

$$h(R)\le\max_{r<\epsilon R} \bigl(r\sqrt{\alpha(r)} \bigr)+ \max _{\epsilon R<r<R} \bigl(r\sqrt{\alpha(r)} \bigr)\le\epsilon R\sqrt{ \alpha(0)}+R\sqrt{\alpha(\epsilon R)}. $$

Therefore,

$$\mathop{\overline{\lim}}_{R\to\infty}\frac{h(R)}{R}\le\epsilon \sqrt{ \alpha(0)}. $$

Hence, h(R)/R→0, as R→∞. By definition of h(R) we have \(h(R)\ge R\sqrt{\alpha(R)}\). Therefore,

$$\frac{R}{h(R)}\le\frac{1}{\sqrt{\alpha(R)}}. $$

Thus,

$$\biggl(\frac{R}{h(R)} \biggr)\alpha(R)\le\sqrt{\alpha(R)}\to 0,\quad\text{as }R\to\infty. $$

 □

Now let us prove Lemma 3.7. First observe that for any \(\phi \in{\mathcal{S}}_{0}\)

$$\lim_{R\to\infty}\int_{B_{R}}|\nabla\phi |^{2}\,\mathrm{d}\boldsymbol{x}=\|\nabla\phi \|_{2}^{2}, $$

while

$$\lim_{R\to\infty}\int_{A_{R}(h(R))}|\nabla\phi |^{2}\,\mathrm{d}\boldsymbol{x}=0. $$

Hence, we only need to prove that there exist \(\boldsymbol{c}\in\mathbb{R} ^{m}\) and a monotonically increasing function h(R)=o(R) such that

$$ \lim_{R\to\infty}\frac{1}{h(R)^{2}}\int _{A_{R}(h(R))}|\phi - \boldsymbol{c} |^{2}\,\mathrm{d}\boldsymbol{x}=0. $$
(B.2)

Remark 3.6 implies that we need to prove (B.2) for d≥3. In that case, the constant \(\boldsymbol{c}\in\mathbb{R} ^{m}\) is chosen so that \(\phi - \boldsymbol{c}\in L^{\frac{2d}{d-2}}(\mathbb{R} ^{d};\mathbb{R} ^{m})\), which is possible by Theorem A.1. The Hölder inequality gives

$$\frac{1}{h(R)^{2}}\int_{A_{R}(h(R))}|\phi - \boldsymbol{c}|^{2}\, \mathrm{d}\boldsymbol{x}\le C \biggl(\frac{R}{h(R)} \biggr)^{\frac{2(d-1)}{d}} \biggl(\int _{A_{R}(h(R))}|\phi - \boldsymbol{c}|^{\frac{2d}{d-2}}\,\mathrm{d}\boldsymbol{x} \biggr)^{\frac{d-2}{d}}. $$

By Theorem A.1

$$\alpha(R)= \biggl(\int_{| \boldsymbol{x}|\ge R/2}|\phi - \boldsymbol{c}|^{\frac {2d}{d-2}}\, \mathrm{d}\boldsymbol{x} \biggr)^{\frac{d-2}{2(d-1)}}\to0,\quad\text{as }R\to\infty. $$

We see that in each of the three cases we have a function α(R)→0, as R→∞, which is independent of h(R). The application of Lemma B.1 concludes the proof of Lemma 3.7.

Appendix C: Proof of Theorem 3.9

Step 1: Asymptotics of \(\int_{B_{R}}|\nabla\phi |^{2}\,\mathrm{d}\boldsymbol{x}\).

We write x=p+R T t and |x|2=|p|2+|t|2. Therefore,

If we make the change of variables p=R u we obtain

$$\frac{1}{R^{d-k}}\int_{V_{R}}|\nabla\phi |^{2}\, \mathrm{d}\boldsymbol{x}=\int_{\{| \boldsymbol{t} |<R\} }\int_{\{| \boldsymbol{u}|<1\}} \bigl\{\big| \psi _{ \boldsymbol{t}}( \boldsymbol{t},R \boldsymbol{u})\big|^{2}+\big|\psi _{ \boldsymbol{p}}( \boldsymbol{t},R \boldsymbol{u})\big|^{2}\bigr\}\,\mathrm{d}\boldsymbol{u} \,\mathrm{d}\boldsymbol{t}. $$

Hence,

$$\frac{1}{R^{d-k}}\int_{V_{R}}|\nabla\phi |^{2}\, \mathrm{d}\boldsymbol{x}\le\int_{\{| \boldsymbol{u} |<1\} }\int_{{\mathbb{R} ^{k}}} \bigl \{\big|\psi _{ \boldsymbol{t}}( \boldsymbol{t},R \boldsymbol{u})\big|^{2}+\big|\psi _{ \boldsymbol{p}}( \boldsymbol{t},R \boldsymbol{u})\big|^{2}\bigr\}\,\mathrm{d}\boldsymbol{t} \,\mathrm{d}\boldsymbol{u}. $$

By the Riemann–Lebesgue lemma we get

$$\mathop{\overline{\lim}}_{R\to\infty}\frac{1}{R^{d-k}}\int _{V_{R}}|\nabla\phi |^{2}\,\mathrm{d}\boldsymbol{x} \le \omega_{d-k}\hspace{-8pt}\skew {33}-{\int _{Q_{d-k}}}\int_{{\mathbb{R} ^{k}}} \bigl\{\big| \psi _{ \boldsymbol{t}}( \boldsymbol{t}, \boldsymbol{p})\big|^{2}+\big|\psi _{ \boldsymbol{p}}( \boldsymbol{t}, \boldsymbol{p})\big|^{2}\bigr\}\,\mathrm{d}\boldsymbol{t} \,\mathrm{d}\boldsymbol{p}, $$

where ω n is the volume of the n-dimensional unit ball. Thus,

$$\mathop{\overline{\lim}}_{R\to\infty}\frac{1}{R^{d-k}}\int _{B_{R}}|\nabla\phi |^{2}\,\mathrm{d}\boldsymbol{x} \le \omega_{d-k}\hspace{-8pt}\skew {33}-{\int _{Q_{d-k}}}\int_{{\mathbb{R} ^{k}}} \bigl\{\big| \psi _{ \boldsymbol{t}}( \boldsymbol{t}, \boldsymbol{p})\big|^{2}+\big|\psi _{ \boldsymbol{p}}( \boldsymbol{t}, \boldsymbol{p})\big|^{2}\bigr\}\,\mathrm{d}\boldsymbol{t} \,\mathrm{d}\boldsymbol{p}. $$

To get the reverse inequality we write

$$\int_{B_{R}}|\nabla\phi |^{2}\,\mathrm{d}\boldsymbol{x}=\int _{\{| \boldsymbol{t}|<R\}}\int_{\{| \boldsymbol{p} |<r(R, \boldsymbol{t})\}} \bigl\{|\psi _{ \boldsymbol{t}}|^{2}+|\psi _{ \boldsymbol{p}}|^{2}\bigr\}\, \mathrm{d} \boldsymbol{p}\,\mathrm{d} \boldsymbol{t}, $$

where \(r(R, \boldsymbol{t})=\sqrt{R^{2}-| \boldsymbol{t}|^{2}}\). If we make a change of variables p=r(R,t)u we obtain

By the Riemann–Lebesgue lemma we get

for a.e. \(\boldsymbol{t}\in\mathbb{R} ^{k}\). By Fatous’s lemma we get

$$\mathop{\underline{\lim}}_{R\to\infty}\frac{1}{R^{d-k}}\int _{B_{R}}|\nabla\phi |^{2}\,\mathrm{d}\boldsymbol{x} \ge \omega_{d-k}\int_{\mathbb{R} ^{k}} \hspace{-8pt}\skew {33}-{\int _{Q_{d-k}}}\bigl\{\big| \psi _{ \boldsymbol{t}}( \boldsymbol{t}, \boldsymbol{p})\big|^{2}+\big|\psi _{ \boldsymbol{p}}( \boldsymbol{t}, \boldsymbol{p} )\big|^{2}\bigr\}\,\mathrm{d} \boldsymbol{p}\,\mathrm{d} \boldsymbol{t}. $$

Hence, we obtain the asymptotics of \(\int_{B_{R}}|\nabla\phi |^{2}\,\mathrm{d}\boldsymbol{x}\):

$$ \lim_{R\to\infty}\frac{1}{R^{d-k}}\int _{B_{R}}|\nabla\phi |^{2}\,\mathrm{d}\boldsymbol{x} = \omega_{d-k}\int_{\mathbb{R} ^{k}} \hspace{-8pt}\skew {33}-{\int _{Q_{d-k}}}\bigl\{\big| \psi _{ \boldsymbol{t}}( \boldsymbol{t}, \boldsymbol{p})\big|^{2}+\big|\psi _{ \boldsymbol{p}}( \boldsymbol{t}, \boldsymbol{p} )\big|^{2}\bigr\}\,\mathrm{d} \boldsymbol{p}\,\mathrm{d} \boldsymbol{t}. $$
(C.1)

In particular, we get

$$\mathop{\overline{\lim}}_{R\to\infty}\hspace{-8pt}\skew {33}-{\int _{B_{R}}}|\nabla\phi |^{2}\,\mathrm{d}\boldsymbol{x}\le\mathop{\overline{\lim}}_{R\to \infty} \frac{\omega_{d}}{R^{d}}\int_{V_{R}} |\nabla\phi |^{2}\, \mathrm{d}\boldsymbol{x}=0, $$

establishing (3.19).

Step 2: Proof of (3.17).

For any h(R)=o(R) we have, using (C.1),

$$\frac{\int_{A_{R}(h(R))}|\nabla\phi |^{2}\,\mathrm{d}\boldsymbol{x}}{\int_{B_{R}}|\nabla\phi |^{2}\,\mathrm{d}\boldsymbol{x}}= \frac{\int_{B_{R}}|\nabla\phi |^{2}\,\mathrm{d}\boldsymbol{x}-\int_{B_{R-h(R)}}|\nabla\phi |^{2}\,\mathrm{d}\boldsymbol{x}}{ \int_{B_{R}}|\nabla\phi |^{2}\,\mathrm{d}\boldsymbol{x}}=1- \biggl(1-\frac{h(R)}{R} \biggr)^{d-k}u_{R}, $$

where u R →1, as R→∞. Thus, (3.17) is proved.

Step 3: Proof of (3.18). We have

$$\int_{A_{R}(h(R))}|\phi |^{2}\,\mathrm{d}\boldsymbol{x}\le\int _{| \boldsymbol{t}|<R}\int_{| \boldsymbol{p} |<R}\big|\psi ( \boldsymbol{t}, \boldsymbol{p})\big|^{2}\,\mathrm{d}\boldsymbol{p} \,\mathrm{d}\boldsymbol{t}. $$

The periodicity in the p variable implies that for any domain \(\varOmega \subset\mathbb{R} ^{d-k}\)

$$\int_{\varOmega}\big|\psi ( \boldsymbol{t}, \boldsymbol{p})\big|^{2}\,\mathrm{d}\boldsymbol{p}\le v( \varOmega)\hspace{-8pt}\skew {33}-{\int _{Q_{d-k}}}\big|\psi ( \boldsymbol{t}, \boldsymbol{p})\big|^{2}\,\mathrm{d}\boldsymbol{p}, $$

where v(Ω) is the (dk)-volume of all period cells intersecting Ω. We further estimate that

$$v(\varOmega)\le|\varOmega+B_{M}|, $$

where M is the diameter of the period cell Q dk . When R>M we obtain

$$v\bigl(\bigl\{| \boldsymbol{p}|\le R+M\bigr\}\bigr)\le\omega_{d-k}(2R)^{d-k}. $$

Hence, we get the estimate

(C.2)

For convenience we introduce the truncated L 2 norm

$$\| \boldsymbol{f}\|_{2,R}^{2}=\int_{B_{R}}\int _{Q_{d-k}}| \boldsymbol{f}( \boldsymbol{t}, \boldsymbol{p} )|^{2}\,\mathrm{d} \boldsymbol{p}\,\mathrm{d} \boldsymbol{t}. $$

Lemma C.1

For every \(\psi \in{\mathcal{S}}_{k}(Q_{d-k})\) there exists a constant \(\boldsymbol{c}\in \mathbb{R} ^{m}\) such that

$$\mathop{\underline{\lim}}_{R\to\infty}\frac{\|\psi - \boldsymbol{c}\| ^{2}_{2,R}}{R^{2}}=0. $$

Proof

The proof of the lemma is different depending on whether k=1, k=2 or k≥3.

If k=1 we can use the assumption of uniform boundedness of ψ and conclude that

$$\frac{\|\psi \|^{2}_{2,R}}{R^{2}}\le\frac{2\|\psi \|_{\infty }^{2}}{R}\to0,\quad\text{as }R\to\infty. $$

If k≥3, then, according to Theorem A.1, for a.e. pQ dk there exists a unique vector c(p) such that

$$\int_{\mathbb{R} ^{k}}\big|\psi ( \boldsymbol{t}, \boldsymbol{p})- \boldsymbol{c}( \boldsymbol{p})\big|^{\frac {2k}{k-2}}<\infty. $$

However, we need a sharper statement

Lemma C.2

There exists \(\boldsymbol{c}\in\mathbb{R} ^{m}\) such that c(p)=c for a.e. pQ dk .

Proof

Let

$$\langle\psi \rangle_{Q_{d-k}}( \boldsymbol{t})=\frac{1}{|Q_{d-k}|}\int _{Q_{d-k}}\psi ( \boldsymbol{t}, \boldsymbol{p})\,\mathrm{d}\boldsymbol{p}. $$

The Poincaré inequality implies

$$\big\|\psi -\langle\psi \rangle_{Q_{d-k}}( \boldsymbol{t})\big\|_{2,R}^{2} \le C\|\psi _{ \boldsymbol{p}}\|_{2,R}^{2}. $$

Therefore, \(\|\psi -\langle\psi \rangle_{Q_{d-k}}( \boldsymbol{t})\|_{2,R}\) is bounded as R→∞. Next observe that \(\langle\psi \rangle_{Q_{d-k}}( \boldsymbol{t} )\in{\mathcal{S}}\) as a function of t. Hence, there exists \(\boldsymbol{c}\in\mathbb{R} ^{m}\) such that \(\langle\psi \rangle_{Q_{d-k}}( \boldsymbol{t})- \boldsymbol{c}\in L^{\frac{2k}{k-2}}(\mathbb{R} ^{k})\). It follows that

$$\big\| \boldsymbol{c}( \boldsymbol{p})- \boldsymbol{c}\big\|_{2,R}\le\big\| \boldsymbol{c}( \boldsymbol{p})-\psi \big\|_{2,R}+\big\|\psi -\langle \psi \rangle_{Q_{d-k}}( \boldsymbol{t})\big\|_{2,R} +\big\|\langle\psi \rangle_{Q_{d-k}}( \boldsymbol{t})- \boldsymbol{c}\big\|_{2,R}. $$

Let us apply the Hölder inequality to the first and third term on the left-hand side of the above inequality:

We conclude that

$$\mathop{\overline{\lim}}_{R\to\infty}\frac{1}{R^{2}}\big\| \boldsymbol{c}( \boldsymbol{p} )- \boldsymbol{c} \big\|_{2,R}^{2}<+\infty. $$

However, this would contradict

$$\big\| \boldsymbol{c}( \boldsymbol{p})- \boldsymbol{c}\big\|_{2,R}^{2}=|B_{R}|\int _{Q_{d-k}}\big| \boldsymbol{c}( \boldsymbol{p})- \boldsymbol{c} \big|^{2}\,\mathrm{d}\boldsymbol{p}, $$

unless c(p)=c for a.e. pQ dk . □

We will now establish Lemma C.1, in which the constant vector c is coming from Lemma C.2. For simplicity of notation ψ(t,p) will now stand for ψc. In order to prove Lemma C.1 we split the t-integral in the definition of ∥ψ2,R into the integral over the ball {|t|<ϵR} and the annulus {ϵR<|t|<R}. Then we apply the same Hölder inequality to both integrals and obtain the estimate

Lemma C.2 then implies that for a.e. pQ dk

$$\mathop{\overline{\lim}}_{R\to\infty}\frac{1}{R^{2}}\int _{| \boldsymbol{t} |<R}\big|\psi ( \boldsymbol{t}, \boldsymbol{p} )\big|^{2}\,\mathrm{d}\boldsymbol{t}\le \omega_{k}^{\frac{k}{2}}\epsilon^{2} \biggl(\int _{\mathbb{R} ^{k}}\big|\psi ( \boldsymbol{t}, \boldsymbol{p})\big|^{\frac{2k}{k-2}}\,\mathrm{d}\boldsymbol{t} \biggr)^{\frac{k-2}{k}}. $$

Thus, for a.e. pQ dk

$$ \lim_{R\to\infty}\frac{1}{R^{2}}\int _{| \boldsymbol{t}|<R}\big|\psi ( \boldsymbol{t}, \boldsymbol{p} )\big|^{2}\,\mathrm{d}\boldsymbol{t}=0. $$
(C.3)

By Hölder inequality and Theorem A.1

$$\frac{1}{R^{2}}\int_{| \boldsymbol{t}|<R}\big|\psi ( \boldsymbol{t}, \boldsymbol{p})\big|^{2}\, \mathrm{d}\boldsymbol{t}\le \omega_{k}^{\frac{2}{k}} \biggl(\int _{| \boldsymbol{t}|<R}\big|\psi ( \boldsymbol{t}, \boldsymbol{p} )\big|^{\frac{2k}{k-2}}\,\mathrm{d}\boldsymbol{t} \biggr)^{\frac{k-2}{k}}\le C\int_{\mathbb{R} ^{k}}\big|\psi _{ \boldsymbol{t}}( \boldsymbol{t}, \boldsymbol{p})\big|^{2}\,\mathrm{d}\boldsymbol{t}. $$

By the Lebesgue dominated convergence theorem \(\|\psi \|_{2,R}^{2}/R^{2}\to0\), as R→∞, since the function

$$\varPhi( \boldsymbol{p})=\int_{\mathbb{R} ^{k}}\big|\psi _{ \boldsymbol{t}}( \boldsymbol{t}, \boldsymbol{p})\big|^{2}\,\mathrm{d}\boldsymbol{t} $$

is integrable over Q dk .

The case k=2 is the most delicate. Let us define

$$\boldsymbol{c}_{R}( \boldsymbol{p})=\hspace{-8pt}\skew {33}-{\int _{| \boldsymbol{t}|<R}}\psi ( \boldsymbol{t}, \boldsymbol{p})\,\mathrm{d}\boldsymbol{t}. $$

Let R n →∞ be a strictly monotonic sequence such that

$$\lim_{n\to\infty}\langle \boldsymbol{c}_{R_{n}} \rangle_{Q_{d-2}}= \boldsymbol{c} $$

for some vector \(\boldsymbol{c}\in\mathbb{R} ^{m}\). We claim that

$$\lim_{R\to\infty}\big\| \boldsymbol{c}_{R}( \boldsymbol{p})-\langle \boldsymbol{c}_{R} \rangle _{Q_{d-2}}\big\|_{2}\,\mathrm{d}\boldsymbol{p}=0. $$

Indeed,

$$\big\| \boldsymbol{c}_{R}( \boldsymbol{p})-\langle \boldsymbol{c}_{R} \rangle_{Q_{d-2}} \big\|_{2}^{2}\le C\hspace{-8pt}\skew {33}-{\int _{| \boldsymbol{t}|<R}}\int_{Q_{d-2}}\big|\psi ( \boldsymbol{t}, \boldsymbol{p})-\langle\psi \rangle _{Q_{d-2}}( \boldsymbol{t})\big|^{2}\,\mathrm{d} \boldsymbol{p}\, \mathrm{d} \boldsymbol{t}. $$

Applying the Poincaré inequality for the inner integral we get

$$\big\| \boldsymbol{c}_{R}-\langle \boldsymbol{c}_{R} \rangle_{Q_{d-2}} \big\|_{2}^{2}\le C\hspace{-8pt}\skew {33}-{\int _{| \boldsymbol{t}|<R}}\int_{Q_{d-2}}\big|\psi _{ \boldsymbol{p}}( \boldsymbol{t}, \boldsymbol{p})\big|^{2}\,\mathrm{d} \boldsymbol{p}\,\mathrm{d} \boldsymbol{t} \to0,\quad \text{as } R\to\infty. $$

We now prove that

$$\lim_{n\to\infty}\frac{1}{R^{2}}\big\|\psi ( \boldsymbol{t}, \boldsymbol{p})- \boldsymbol{c} \big\|_{2,R}^{2}=0. $$

By triangle inequality we have

$$\big\|\psi ( \boldsymbol{t}, \boldsymbol{p})- \boldsymbol{c}\big\|_{2,R}\le\big\|\psi - \boldsymbol{c}_{R}( \boldsymbol{p}) \big\|_{2,R} +\big\| \boldsymbol{c}_{R}( \boldsymbol{p})-\langle \boldsymbol{c}_{R} \rangle_{Q_{d-2}}\big\|_{2,R} +\big\|\langle \boldsymbol{c}_{R} \rangle_{Q_{d-2}}- \boldsymbol{c}\big\|_{2,R}. $$

We compute

$$\big\|\langle \boldsymbol{c}_{R} \rangle_{Q_{d-2}}- \boldsymbol{c}\big\|_{2,R}^{2}=|Q_{d-2}| \pi R^{2}\big|\langle \boldsymbol{c}_{R} \rangle_{Q_{d-2}}- \boldsymbol{c}\big|^{2}. $$

Hence,

Therefore,

$$\lim_{R\to\infty}\frac{1}{R^{2}}\big\| \boldsymbol{c}_{R}( \boldsymbol{p})-\langle \boldsymbol{c}_{R} \rangle_{Q_{d-2}}\big\|_{2,R}^{2}=0. $$

Finally, we have

Using the uniform boundedness of ψ for the first term and the Poincaré inequality for the second term, we get

$$\big\|\psi - \boldsymbol{c}_{R}( \boldsymbol{p})\big\|_{2,R}^{2}\le C \epsilon^{2}R^{2}\|\psi - \boldsymbol{c}\| _{\infty}^{2} +R^{2}C_{\epsilon}\int_{Q_{d-2}}\int _{\{\epsilon R<| \boldsymbol{t}|<R\}}\big|\psi _{ \boldsymbol{t} }( \boldsymbol{t}, \boldsymbol{p})\big|^{2}\,\mathrm{d} \boldsymbol{t}\,\mathrm{d} \boldsymbol{p}. $$

Thus,

$$\mathop{\overline{\lim}}_{R\to\infty}\frac{1}{R^{2}}\big\|\psi - \boldsymbol{c} _{R}( \boldsymbol{p})\big\|_{2,R}^{2}\le C\epsilon^{2}\| \psi - \boldsymbol{c}\|_{\infty}^{2}. $$

The arbitrariness of ϵ>0 implies that

$$\lim_{R\to\infty}\frac{1}{R^{2}}\big\|\psi - \boldsymbol{c}_{R}( \boldsymbol{p}) \big\|_{2,R}^{2}=0. $$

Lemma C.1 is proved now.  □

Let R n →∞ be the monotonically increasing sequence for which \(\|\psi - \boldsymbol{c}\|_{2,R_{n}}/R_{n}\to0\), as n→∞. Let

$$\alpha(R)=\frac{\|\psi - \boldsymbol{c}\|_{2,R_{n}}}{R_{n}},\quad R_{n}\le R<R_{n+1}. $$

Then α(R)→0, as R→∞. By Lemma B.1 there exists a monotonically increasing function h(R) such that h(R)/R→0, as R→∞ and

$$\lim_{R\to\infty} \biggl(\frac{R}{h(R)} \biggr)\alpha(R)=0. $$

Hence,

$$\mathop{\underline{\lim}}_{R\to\infty} \biggl(\frac{\|\psi - \boldsymbol{c}\| _{2,R}}{h(R)} \biggr)^{2}\le \lim_{n\to\infty} \biggl(\frac{R_{n}}{h(R_{n})} \alpha(R_{n}) \biggr)^{2}=0. $$

The estimate (C.2) together with (3.17) now implies (3.18). Thus, we have proved that \({\mathcal{C}}_{k}\subset{\mathcal{S}}_{*}\) for any 1≤kd.

Step 4: Proof of the formula (3.24). We have

By the Riemann–Lebesgue lemma

Thus, in order to finish the proof of the theorem we need to show that

$$ \rho=\lim_{R\to\infty}\int_{| \boldsymbol{u}|\le1} \int_{| \boldsymbol{t}|\ge R\sqrt {1-| \boldsymbol{u}|^{2}}} W^{\circ}\bigl( \boldsymbol{F},\psi _{ \boldsymbol{t}}( \boldsymbol{t},R \boldsymbol{u}) \boldsymbol{R}+\psi _{ \boldsymbol{p}}( \boldsymbol{t},R \boldsymbol{u} ) \boldsymbol{Q}\bigr)\,\mathrm{d} \boldsymbol{t}\,\mathrm{d} \boldsymbol{u}=0. $$
(C.4)

Recall that \(\phi \in W^{1,\infty}(\mathbb{R} ^{d};\mathbb{R} ^{m})\). Hence, there exists a number C>0 depending on ∥ϕ1,∞, but independent of R such that

$$\rho\le C\mathop{\overline{\lim}}_{R\to\infty}\int_{| \boldsymbol{u}|\le 1} \int_{| \boldsymbol{t}|\ge R\sqrt{1-| \boldsymbol{u}|^{2}}} \bigl\{\big|\psi _{ \boldsymbol{t}}( \boldsymbol{t},R \boldsymbol{u})\big|^{2}+\big|\psi _{ \boldsymbol{p}}( \boldsymbol{t},R \boldsymbol{u})\big|^{2}\bigr\}\, \mathrm{d}\boldsymbol{t} \,\mathrm{d}\boldsymbol{u}. $$

For any ϵ∈(0,1) we have

$$\int_{| \boldsymbol{u}|\le1}\int_{| \boldsymbol{t}|\ge R\sqrt{1-| \boldsymbol{u}|^{2}}} \bigl\{\big|\psi _{ \boldsymbol{t}}( \boldsymbol{t},R \boldsymbol{u})\big|^{2}+\big|\psi _{ \boldsymbol{p}}( \boldsymbol{t},R \boldsymbol{u})\big|^{2}\bigr\}\,\mathrm{d}\boldsymbol{t} \,\mathrm{d}\boldsymbol{u}=T_{1}(R, \epsilon)+T_{2}(R,\epsilon), $$

where

If |u|≤1−ϵ and \(| \boldsymbol{t}|\ge R\sqrt{1-| \boldsymbol{u}|^{2}}\) then \(| \boldsymbol{t} |\ge R\sqrt{\epsilon(2-\epsilon)}\). In particular, \(| \boldsymbol{t}|\ge\sqrt {(2-\epsilon)/\epsilon }\), if R>1/ϵ. Therefore, by the Riemann–Lebesgue lemma

Also, by the Riemann–Lebesgue lemma

We conclude that ρ=0, since

$$\lim_{\epsilon\to0}T_{1}^{\infty}(\epsilon)=\lim _{\epsilon\to 0}T_{2}^{\infty }(\epsilon)=0. $$

Appendix D: Proof of Theorem 3.14

We may assume, without loss of generality, that n=e 1. By Lemma 3.2 in Müller and Šverák (2003), applied to the bounded domain Q=[0,1]d, there exists a sequence of functions u n (x) converging uniformly in Q to u 0(x)=x 1 a and such that ∥∇u n is a bounded sequence and for all 1≤jr

$$\lim_{n\to\infty}\big|\bigl\{ \boldsymbol{x}\in Q:\operatorname{dist}\bigl(\nabla \boldsymbol{u}_{n}( \boldsymbol{x}), \boldsymbol{H} _{j}\bigr)<1/n\bigr\}\big|=\lambda_{j}. $$

Let p n (x) denote the function defined in the layer 0<x 1<1, which is periodic with periods e 2,…,e d and equal to u n (x) on Q. Finally, we let

$$\boldsymbol{v}_{n}( \boldsymbol{x})= \begin{cases} \boldsymbol{a},&\text{if }x_{1}\ge1,\\[4pt] \boldsymbol{0},&\text{if }x_{1}\le0,\\[4pt] \boldsymbol{p}_{n}( \boldsymbol{x}),&\text{if }0<x_{1}<1. \end{cases} $$

Clearly, v n (x)→ϕ 0(x) uniformly in \(\mathbb{R} ^{d}\). However, the functions v n (x) have jump discontinuities across the surfaces Γ j,k ={x j =k,0<x 1<1}, j=2,…,d, \(k\in\mathbb{Z}\), as well as the planes Π 0={x 1=0} and Π 1={x 1=1}. Let ϵ n =∥v n ϕ 0. Then ϵ n →0 as n→∞. Let

$$\varGamma=\varPi_{0}\cup\varPi_{1}\cup \Biggl(\bigcup _{j=1}^{d} \biggl(\bigcup _{k\in\mathbb{Z}}\varGamma_{j,k} \biggr) \Biggr) $$

be the entire singular set. When n is sufficiently large there is a \(C^{\infty}(\mathbb{R} ^{d})\) function η n (x) that is periodic with periods e 2,…,e d , which is equal to 0 on Γ and 1 on \(\{ \boldsymbol{x}\in\mathbb{R} ^{d}:\operatorname{dist}( \boldsymbol{x},\varGamma)>\sqrt{\epsilon _{n}}\}\), and such that \(\|\nabla\eta_{n}\|_{\infty}\le C/\sqrt{\epsilon_{n}}\). It follows that the function

$$\phi _{n}( \boldsymbol{x})=\bigl(1-\eta_{n}( \boldsymbol{x})\bigr)\phi _{0}( \boldsymbol{x})+\eta_{n}( \boldsymbol{x}) \boldsymbol{v} _{n}( \boldsymbol{x}) $$

is Lipschitz continuous with

$$\|\nabla\phi _{n}\|_{\infty}\le\|\nabla\phi _{0} \|_{\infty}+\| \nabla \boldsymbol{u}_{n}\|_{\infty}+C\sqrt{ \epsilon_{n}}. $$

Obviously, ϕ n (x) converges uniformly to ϕ 0(x). In addition ∇ϕ n (x)=0 whenever \(x_{1}<-\sqrt{\epsilon_{n}}\) or \(x_{1}>1+\sqrt{\epsilon_{n}}\). It follows that (3.19) holds. Observe that ϕ n (x) has periods e 2,…,e d , since both v n (x) and η n (x) do. Thus, \(\psi _{n}\in{\mathcal{S}}_{1}^{0}\), where

$$\psi _{n}(t, \boldsymbol{p})=\phi _{n} \Biggl(t \boldsymbol{e}_{1}+\sum _{j=2}^{d}{p}_{j} \boldsymbol{e} _{j} \Biggr). $$

Hence, \(\phi _{n}\in{\mathcal{C}}_{1}\). To finish the proof of the theorem we need to establish (3.29). This is a consequence of the formula (3.24) and the relation

Appendix E: Proof of Lemma 4.4

The lemma is best proved in the (t,p) variables, where instead of the [0,1]dk periodic field ψ we use a Q dk periodic field, which we denote ψ as well, so that

$$\phi ( \boldsymbol{x})=\psi ( \boldsymbol{R} \boldsymbol{x}, \boldsymbol{Q} \boldsymbol{x}),\qquad\psi ( \boldsymbol{t}, \boldsymbol{p})=\phi \bigl( \boldsymbol{R} ^{T} \boldsymbol{t}+ \boldsymbol{Q}^{T} \boldsymbol{p}\bigr). $$

In terms of ψ Eq. (4.12) becomes

$$ \begin{cases} \nabla_{ \boldsymbol{t}}\cdot( \boldsymbol{P}( \boldsymbol{F}+\psi _{ \boldsymbol{t}} \boldsymbol{R}+\psi _{ \boldsymbol{p}} \boldsymbol{Q}) \boldsymbol{R}^{T})+ \nabla_{ \boldsymbol{p}}\cdot( \boldsymbol{P}( \boldsymbol{F}+\psi _{ \boldsymbol{t}} \boldsymbol{R}+\psi _{ \boldsymbol{p}} \boldsymbol{Q}) \boldsymbol{Q} ^{T})= \boldsymbol{0},\\[4pt] \nabla_{ \boldsymbol{t}}\cdot( \boldsymbol{R} \boldsymbol{P}^{*}( \boldsymbol{F}+\psi _{ \boldsymbol{t}} \boldsymbol{R}+\psi _{ \boldsymbol{p}} \boldsymbol{Q} ) \boldsymbol{R}^{T})= \nabla_{ \boldsymbol{p}}\cdot(\psi _{ \boldsymbol{t}}^{T} \boldsymbol{P}( \boldsymbol{F}+\psi _{ \boldsymbol{t}} \boldsymbol{R}+\psi _{ \boldsymbol{p}} \boldsymbol{Q}) \boldsymbol{Q}^{T}),\\[4pt] \nabla_{ \boldsymbol{p}}\cdot( \boldsymbol{Q} \boldsymbol{P}^{*}( \boldsymbol{F}+\psi _{ \boldsymbol{t}} \boldsymbol{R}+\psi _{ \boldsymbol{p}} \boldsymbol{Q} ) \boldsymbol{Q}^{T})= \nabla_{ \boldsymbol{t}}\cdot(\psi _{ \boldsymbol{p}}^{T} \boldsymbol{P}( \boldsymbol{F}+\psi _{ \boldsymbol{t}} \boldsymbol{R}+\psi _{ \boldsymbol{p}} \boldsymbol{Q}) \boldsymbol{R}^{T}), \end{cases} $$
(E.1)

while relation (4.18) reads

$$ \begin{cases} \int_{\mathbb{R} ^{k}}\int_{Q_{d-k}} \boldsymbol{R}\widehat{ \boldsymbol{P} }^{*} \boldsymbol{R} ^{T}\,\mathrm{d} \boldsymbol{p}\,\mathrm{d} \boldsymbol{t}= \boldsymbol{0},\\[4pt] \int_{\mathbb{R} ^{k}}\int_{Q_{d-k}}\psi _{ \boldsymbol{p} }^{T}\widehat{ \boldsymbol{P}} \boldsymbol{R}^{T}\,\mathrm{d} \boldsymbol{p}\,\mathrm{d} \boldsymbol{t}= \boldsymbol{0}, \end{cases} $$
(E.2)

where

$$\widehat{ \boldsymbol{P}}=\widehat{ \boldsymbol{P}}(\psi _{ \boldsymbol{t}} \boldsymbol{R}+\psi _{ \boldsymbol{p}} \boldsymbol{Q})= \boldsymbol{P} ( \boldsymbol{F} +\psi _{ \boldsymbol{t}} \boldsymbol{R}+\psi _{ \boldsymbol{p}} \boldsymbol{Q})- \boldsymbol{P}( \boldsymbol{F}). $$

Replacing \(\widehat{ \boldsymbol{P}}^{*}\) by its expression from (4.15) and using Eq. (E.1) we obtain

$$ \nabla_{ \boldsymbol{t}}\cdot\bigl( \boldsymbol{R}\widehat{ \boldsymbol{P}}^{*} \boldsymbol{R}^{T}\bigr)= \nabla_{ \boldsymbol{p}}\cdot\bigl(\psi _{ \boldsymbol{t}}^{T}\widehat{ \boldsymbol{P}} \boldsymbol{Q}^{T}\bigr) - \nabla_{ \boldsymbol{p}}\cdot\bigl( \boldsymbol{R} \boldsymbol{F}^{T} \boldsymbol{P}( \boldsymbol{F}+\psi _{ \boldsymbol{t}} \boldsymbol{R}+ \psi _{ \boldsymbol{p} } \boldsymbol{Q}) \boldsymbol{Q}^{T}\bigr), $$
(E.3)

since

$$\nabla_{ \boldsymbol{t}}\cdot\bigl( \boldsymbol{R} \boldsymbol{N}(\psi _{ \boldsymbol{t}} \boldsymbol{R}+\psi _{ \boldsymbol{p}} \boldsymbol{Q}) \boldsymbol{R}^{T}\bigr)= -\nabla_{ \boldsymbol{t}}\cdot\bigl( \psi _{ \boldsymbol{t}}^{T} \boldsymbol{P}( \boldsymbol{F})\bigr). $$

Let

Integrating (E.3) over Q dk and using the periodicity we obtain ∇ t f 1(t)=0. Similarly, integrating the third equation in (E.1) over Q dk we conclude that ∇ t f 2(t)=0. We estimate

$$\big|\widehat{ \boldsymbol{P}}^{*}\big|\le C\bigl(|\psi _{ \boldsymbol{t}}|^{2}+| \psi _{ \boldsymbol{p} }|^{2}\bigr),\qquad \big|\psi _{ \boldsymbol{p}}^{T} \widehat{ \boldsymbol{P}}\big|\le C\bigl(|\psi _{ \boldsymbol{t}}|^{2}+|\psi _{ \boldsymbol{p}}|^{2}\bigr), $$

since ψ t and ψ p are assumed to be uniformly bounded. Then \(\phi \in{\mathcal{C}}_{k}\) implies that \(\{ \boldsymbol{f}_{1}, \boldsymbol{f}_{2}\}\subset L^{1}(\mathbb{R} ^{k})\). The statement of the lemma follows from

$$\int_{\mathbb{R} ^{k}} \boldsymbol{f}_{1}( \boldsymbol{t})\,\mathrm{d}\boldsymbol{t}= \boldsymbol{0},\qquad\int _{\mathbb{R} ^{k}} \boldsymbol{f}_{2}( \boldsymbol{t})\,\mathrm{d}\boldsymbol{t}= \boldsymbol{0}, $$

which is a consequence of a simple observation that any L 1 divergence-free vector field f(t) on \(\mathbb{R} ^{k}\) must satisfy \(\int_{\mathbb{R} ^{k}} \boldsymbol{f}\,\mathrm{d} \boldsymbol{t}= \boldsymbol{0}\). Indeed, fL 1 implies that its Fourier transform \(\widehat{ \boldsymbol{f}}(\omega )\) is continuous. ∇⋅f=0 implies that \(\omega \cdot\widehat{ \boldsymbol{f}}(\omega )=0\) for any \(\omega \in\mathbb{R} ^{k}\). Fixing ω0 we obtain

$$\frac{\omega \cdot\widehat{ \boldsymbol{f}}(\epsilon\omega )}{|\omega |}=\frac {\epsilon\omega \cdot\widehat{ \boldsymbol{f}}(\epsilon\omega )}{|\epsilon\omega |}=0. $$

Passing to the limit as ϵ→0 and using continuity of \(\widehat{ \boldsymbol{f}}(\omega )\) we obtain \(\omega \cdot\widehat{ \boldsymbol{f}}( \boldsymbol{0})=0\). Thus \(\widehat{ \boldsymbol{f} }( \boldsymbol{0} )= \boldsymbol{0}\), since \(\omega \in\mathbb{R} ^{k}\setminus\{ \boldsymbol{0}\}\) was arbitrary.

Appendix F: Proof of Theorem 4.6

When the subspace \({\mathcal{L}}\) described by R is fixed we can simplify our notation by regarding the first k components of x as t and the remaining components as p. Then ∇ϕ=[ψ t ,ψ p ]. We then have the corresponding splitting of P=[P 1,P 2] and

$$\boldsymbol{P}^{*}=W\bigl( \boldsymbol{F}+[\psi _{ \boldsymbol{t}},\psi _{ \boldsymbol{p}}]\bigr) \left [ \begin{array}{c@{\quad}c} { \boldsymbol{I}_{k}}&{ \boldsymbol{0}}\\[4pt] { \boldsymbol{0}}&{ \boldsymbol{I}_{d-k}} \end{array} \right ]- \left [ \begin{array}{c} \psi _{ \boldsymbol{t}}^{T}\\[4pt] \psi _{ \boldsymbol{p}}^{T} \end{array} \right ] [ \boldsymbol{P}_{1}, \boldsymbol{P}_{2}]=\left [ \begin{array}{c@{\quad}c} { \boldsymbol{P}_{1}^{*}}&{-\psi _{ \boldsymbol{t}}^{T} \boldsymbol{P}_{2}}\\[4pt] {-\psi _{ \boldsymbol{p}}^{T} \boldsymbol{P}_{1}}&{ \boldsymbol{P}_{2}^{*}} \end{array} \right ]. $$

Similarly, splitting the t and p components we have

where

$$\widehat{ \boldsymbol{P}}^{*}_{1}=W^{\circ}\bigl( \boldsymbol{F},[\psi _{ \boldsymbol{t}},\psi _{ \boldsymbol{p}}]\bigr) I _{k}-\psi _{ \boldsymbol{t}}^{T}\widehat{ \boldsymbol{P}}_{1}, \qquad \widehat{ \boldsymbol{P}}^{*}_{2}=W^{\circ}\bigl( \boldsymbol{F},[\psi _{ \boldsymbol{t}}, \psi _{ \boldsymbol{p}}]\bigr) \boldsymbol{I} _{d-k}-\psi _{ \boldsymbol{p}}^{T} \widehat{ \boldsymbol{P}}_{2}. $$

Next we use the generalized Clapeyron theorem (Grabovsky and Truskinovsky 2013a) for \(\widehat{W}( \boldsymbol{H})\):

$$ \int_{| \boldsymbol{t}|\le R}\int_{Q_{d-k}} \widehat{W}(\nabla\phi )\,\mathrm{d} \boldsymbol{p}\,\mathrm{d} \boldsymbol{t} =\frac{1}{d}(T_{1}+T_{2}), $$
(F.1)

where

Next we observe that

$$\int_{\partial Q_{d-k}}\psi _{ \boldsymbol{t}}^{T}\widehat{ \boldsymbol{P}}_{2} \boldsymbol{n}_{ \boldsymbol{p} }\,\mathrm{d}S( \boldsymbol{p} )= \boldsymbol{0},\qquad \int _{\partial Q_{d-k}}(\widehat{ \boldsymbol{P}}_{2} \boldsymbol{n}_{ \boldsymbol{p}},\psi )\, \mathrm{d}S( \boldsymbol{p})=0, $$

since \(\psi _{ \boldsymbol{t}}^{T}\widehat{ \boldsymbol{P}}_{2}\) and \((\widehat{ \boldsymbol{P} }_{2})^{T}\psi \) are Q dk -periodic. By the divergence theorem we obtain

$$\int_{\partial Q_{d-k}}\bigl(\widehat{ \boldsymbol{P}}_{2}^{*} \boldsymbol{n}_{ \boldsymbol{p}}, \boldsymbol{p}\bigr)\,\mathrm{d}S( \boldsymbol{p} )=\int_{Q_{d-k}}\bigl \{ \bigl(\nabla_{ \boldsymbol{p}}\cdot\widehat{ \boldsymbol{P}}_{2}^{*}, \boldsymbol{p} \bigr)+\operatorname {Tr}\widehat{ \boldsymbol{P}}_{2}^{*}\bigr\} \, \mathrm{d}\boldsymbol{p}. $$

The Noether–Eshelby equation gives

$$\nabla_{ \boldsymbol{p}}\cdot\widehat{ \boldsymbol{P}}_{2}^{*}= \nabla_{ \boldsymbol{t}}\cdot\bigl(\psi _{ \boldsymbol{p} }^{T}\widehat{ \boldsymbol{P}}_{1}\bigr). $$

We also compute

$$\operatorname{Tr}\widehat{ \boldsymbol{P}}_{2}^{*}=(d-k)\widehat {W}- \operatorname{Tr}\bigl(\psi _{ \boldsymbol{p}}^{T}\widehat{ \boldsymbol{P}}_{2}\bigr). $$

Hence, we obtain

$$T_{2}=(d-k)\int_{| \boldsymbol{t}|\le R}\int_{Q_{d-k}} \widehat{W}(\nabla\phi )\,\mathrm{d}\boldsymbol{p} \,\mathrm{d}\boldsymbol{t}+T_{2}', $$

where

$$T_{2}'=\int_{| \boldsymbol{t}|=R}\int _{Q_{d-k}}\bigl(\psi _{ \boldsymbol{p}}^{T}\widehat{ \boldsymbol{P} }_{1} \boldsymbol{n}_{ \boldsymbol{t}}, \boldsymbol{p}\bigr)\,\mathrm{d}\boldsymbol{p}\,\mathrm{d}S( \boldsymbol{t}) -\int _{| \boldsymbol{t}|\le R}\int_{Q_{d-k}}\operatorname{Tr}\bigl( \psi _{ \boldsymbol{p} }^{T}\widehat{ \boldsymbol{P}}_{2}\bigr)\,\mathrm{d} \boldsymbol{p} \,\mathrm{d} \boldsymbol{t}. $$

Substituting this back into (F.1) we obtain

$$\int_{| \boldsymbol{t}|\le R}\int_{Q_{d-k}}\widehat{W}(\nabla \phi )\,\mathrm{d}\boldsymbol{p} \,\mathrm{d}\boldsymbol{t}=\frac{1}{k}\bigl(\widehat{T}_{1}(R)+ \widehat{T}_{2}(R)\bigr), $$

where

We observe that due to (4.17)

$$\lim_{R\to\infty}\widehat{T}_{2}(R)=-\operatorname{Tr} \biggl(\int_{Y_{k}}\psi _{ \boldsymbol{p}}^{T}\widehat{ \boldsymbol{P}}_{2}\,\mathrm{d} \boldsymbol{p}\,\mathrm{d} \boldsymbol{t} \biggr)=0. $$

To finish the proof of the theorem we need to show that \(\widehat {T}_{1}(R)\to0\), as R→∞.

We have \(|\widehat{ \boldsymbol{P}}|\le C|\nabla\phi |\) and \(|\widehat{ \boldsymbol{P} }^{*}|\le C|\nabla\phi |^{2}\), due to the uniform boundedness of ∇ϕ, where the constant C depends on ϕ, but is independent of R. Thus, \(|\widehat {T}_{1}(R)|\le CK(R)\) for a.e. R>1, where

$$K(R)=\int_{| \boldsymbol{t}|=R}\int_{Q_{d-k}}\bigl\{R|\nabla \phi |^{2}+|\phi - \boldsymbol{c} ||\nabla\phi |\bigr\}\,\mathrm{d}\boldsymbol{p}\,\mathrm{d}S( \boldsymbol{t}), $$

where \(\boldsymbol{c}\in\mathbb{R} ^{m}\) can be chosen arbitrarily. We have, after an application of the Cauchy–Schwartz inequality,

If k=1 or k=2 then the boundedness of ϕ implies that

$$ \lim_{R\to\infty}\frac{1}{R}\int _{R}^{2R}K(r)\,\mathrm{d}r=0. $$
(F.2)

If k≥3 then Lemma C.1 guarantees the choice of the constant \(\boldsymbol{c}\in\mathbb{R} ^{m}\) such that (F.2) holds. Therefore,

$$\mathop{\underline{\lim}}_{R\to\infty}K(R)=0. $$

Hence,

$$\biggl \vert \int_{Y}\widehat{W}(\nabla\phi )\, \mathrm{d}\boldsymbol{x}\biggr \vert \le C\mathop{\underline{\lim}}_{R\to\infty}K(R)=0. $$

The theorem is proved.

Appendix G: Proof of Lemma 5.5

We will prove that \({\mathcal{G}}={\mathcal{G}}_{0}\), where

$${\mathcal{G}}_{0}=\bigl\{ \boldsymbol{A}=\operatorname{diag}(A_{1}, \ldots,A_{k}),\ A_{i}>0,\ i=1,\ldots,k,\ \operatorname{Tr} \boldsymbol{A}=1\bigr\}. $$

If we write 〈Γ(n)〉 a in components,

$$\bigl(\bigl\langle\varGamma( \boldsymbol{n}) \bigr\rangle_{ \boldsymbol{a}} \bigr)_{ij}=\hspace{-8pt}\skew {33}-{\int _{\mathbb{S}^{k-1}}}\frac {{n}_{i}{n}_{j}}{ a_{i}a_{j}\sum_{s=1}^{k}a_{s}^{-2}{n}_{s}^{2}}\, \mathrm{d}S( \boldsymbol{n}), $$

we immediately see that \({\mathcal{G}}\subset{\mathcal{G}}_{0}\).

To each \(\boldsymbol{a}=(a_{1},\ldots,a_{k})\in\mathbb{R} ^{k}\) we associate (without relabeling) the diagonal matrix \(\boldsymbol{a}=\operatorname{diag}(a_{1},\ldots,a_{k})\). Let \(\Delta=\{(a_{1},\ldots,a_{k}):a_{i}> 0,\ \sum_{i=1}^{k}a_{i}=1\}\). Then the smooth map F(a)=〈Γ(n)〉 a maps Δ into itself. To prove the reverse inclusion \({\mathcal{G}}_{0}\subset{\mathcal{G}}\) we need to show that the map F:Δ→Δ is surjective. We first show that the differential of the map F is non-degenerate. This implies, via the inverse function theorem, that F(Δ) is an open subset of Δ.

In order to simplify the calculation we first change variables \(\boldsymbol{b}= \boldsymbol{a}^{-1}/\operatorname{Tr}( \boldsymbol{a}^{-1})\). Then

$$F( \boldsymbol{a})=G \biggl(\frac{ \boldsymbol{a}^{-1}}{\operatorname{Tr} \boldsymbol{a}^{-1}} \biggr),\qquad G( \boldsymbol{b})=\hspace{-8pt}\skew {33}-{\int _{ \mathbb{S}^{k-1}}}\frac{ \boldsymbol{b} \boldsymbol{n}\otimes \boldsymbol{b} \boldsymbol{n}}{| \boldsymbol{b} \boldsymbol{n} |^{2}}\,\mathrm{d}S( \boldsymbol{n}). $$

We compute

$$dF( \boldsymbol{a})\eta =-dG \biggl(\frac{ \boldsymbol{a}^{-1}}{\operatorname{Tr} \boldsymbol{a} ^{-1}} \biggr) \frac{ \boldsymbol{a}^{-1}\operatorname{Tr}( \boldsymbol{a}^{-1}\eta \boldsymbol{a}^{-1})- \boldsymbol{a} ^{-1}\eta \boldsymbol{a} ^{-1}\operatorname{Tr} \boldsymbol{a}^{-1}}{(\operatorname{Tr} \boldsymbol{a}^{-1})^{2}}, $$

where η is a diagonal trace-free matrix. If

$$\boldsymbol{a}^{-1}\operatorname{Tr}\bigl( \boldsymbol{a}^{-1}\eta \boldsymbol{a}^{-1} \bigr)- \boldsymbol{a}^{-1}\eta \boldsymbol{a} ^{-1}\operatorname{Tr} \boldsymbol{a} ^{-1}= \boldsymbol{0} $$

then η=λ a for some scalar λ. Taking traces we conclude that λ=0. Hence, the map

$$\eta \mapsto\frac{ \boldsymbol{a}^{-1}\operatorname{Tr}( \boldsymbol{a}^{-1}\eta \boldsymbol{a} ^{-1})- \boldsymbol{a} ^{-1}\eta \boldsymbol{a}^{-1}\operatorname{Tr} \boldsymbol{a}^{-1}}{(\operatorname{Tr} \boldsymbol{a} ^{-1})^{2}} $$

is a non-degenerate linear transformation on the space of diagonal trace-free matrices. Hence, dF is non-degenerate if and only if dG is non-degenerate. We compute explicitly

$$dG\eta =2\hspace{-8pt}\skew {33}-{\int _{\mathbb{S}^{k-1}}} \biggl\{\frac{\eta \boldsymbol{n}\odot \boldsymbol{b} \boldsymbol{n} }{| \boldsymbol{b} \boldsymbol{n}|^{2}}- \frac{ \boldsymbol{b} \boldsymbol{n}\otimes \boldsymbol{b} \boldsymbol{n}}{| \boldsymbol{b} \boldsymbol{n}|^{4}}( \boldsymbol{b} \boldsymbol{n},\eta \boldsymbol{n} ) \biggr\} \,\mathrm{d}S( \boldsymbol{n}), $$

where η is diagonal and \(\operatorname{Tr}\eta =0\). Suppose dGη=0 for some non-zero η. The lemma will be proved if we show that only for η=0. If this is not the case then we have \(\operatorname{Tr}(\eta \boldsymbol{b}^{-1}dG\eta )=0\). We compute (using commutativity of the diagonal matrix multiplication)

$$\operatorname{Tr}\bigl(\eta \boldsymbol{b}^{-1}dG\eta \bigr)=2\hspace{-8pt}\skew {33}-{\int _{ \mathbb{S}^{k-1}}}\frac {|\eta \boldsymbol{n} |^{2}| \boldsymbol{b} \boldsymbol{n}|^{2}-( \boldsymbol{b} \boldsymbol{n},\eta \boldsymbol{n})^{2}}{| \boldsymbol{b} \boldsymbol{n}|^{4}}\,\mathrm{d}S( \boldsymbol{n}). $$

The Cauchy–Schwartz inequality implies that the integrand is non-negative. For it to be zero we would need η n=α(n)bn for almost all \(\boldsymbol{n}\in\mathbb{S}^{k-1}\). The equivalent relation b −1 η n=α(n)n means that every unit vector is an eigenvector of b −1 η. Hence, there is a constant α 0 such that η=α 0 b. Taking the trace, we obtain α 0=0 and the non-degeneracy of dG is proved.

The lemma will follow, if we show that if a n a Δ and F(a n )→f , as n→∞ then f Δ. Let 1≤i,jk be a pair of indices such that \(a^{\circ}_{i}=0\) and \(a^{\circ}_{j}\neq0\). Such a pair exists, since a Δ. We claim that \(f^{\circ }_{j}=0\), finishing the proof of the lemma. We estimate

$$F_{j}( \boldsymbol{a})\le\hspace{-8pt}\skew {33}-{\int _{\mathbb{S}^{k-1}}}\frac {a_{j}^{-2}n_{j}^{2}}{a_{j}^{-2}n_{j}^{2}+a_{i}^{-2}n_{i}^{2}}\, \mathrm{d}S( \boldsymbol{n})= \hspace{-8pt}\skew {33}-{\int _{\mathbb{S}^{k-1}}}\frac {a_{i}^{2}n_{j}^{2}}{a_{i}^{2}n_{j}^{2}+a_{j}^{2}n_{i}^{2}}\,\mathrm{d}S( \boldsymbol{n}). $$

Now, by the Lebesgue bounded convergence theorem,

$$f^{\circ}_{j}=\lim_{n\to\infty}F_{j}( \boldsymbol{a}_{n})=0. $$

Lemma 5.5 is now proved.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Grabovsky, Y., Truskinovsky, L. Marginal Material Stability. J Nonlinear Sci 23, 891–969 (2013). https://doi.org/10.1007/s00332-013-9173-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00332-013-9173-6

Keywords

Mathematics Subject Classification

Navigation