Skip to main content
Log in

Legendre-Hadamard Conditions for Two-Phase Configurations

  • Published:
Journal of Elasticity Aims and scope Submit manuscript

Abstract

We generalize the classical Legendre-Hadamard conditions by using quadratic extensions of the energy around a set of two configurations and obtain new algebraic necessary conditions for nonsmooth strong local minimizers. The implied bounds of stability are easily accessible as we illustrate on a nontrivial example where quasiconvexification is unknown.

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

Similar content being viewed by others

Notes

  1. Microstructure based metastable states would also be disallowed, if the microstructure contains phase boundaries like a finite rank laminate.

  2. In this inequality \(QW_{\boldsymbol {F}\boldsymbol{F}}(\boldsymbol{F}_{\pm})\) is understood as a limiting value of \(QW_{\boldsymbol{F}\boldsymbol {F}}(\boldsymbol{F})\), \(\boldsymbol{F}\in\mathfrak{O}\).

  3. It holds in every nontrivial example of which we are aware.

References

  1. Ball, J.M.: Convexity conditions and existence theorems in nonlinear elasticity. Arch. Ration. Mech. Anal. 63(4), 337–403 (1976/77)

  2. Ball, J.M., James, R.: Fine phase mixtures as minimizers of energy. Arch. Ration. Mech. Anal. 100, 13–52 (1987)

    Article  MathSciNet  MATH  Google Scholar 

  3. 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)

    Article  MathSciNet  MATH  Google Scholar 

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

    Book  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  6. Dunn, J., Fosdick, R.: The Weierstrass condition for a special class of elastic materials. J. Elast. 34(2), 167–184 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  7. Fosdick, R., Royer-Carfagni, G.: Multiple natural states for an elastic isotropic material with polyconvex stored energy. J. Elast. 60(3), 223–231 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  8. Fu, Y.B., Freidin, A.B.: Characterization and stability of two-phase piecewise-homogeneous deformations. Proc. R. Soc., Math. Phys. Eng. Sci. 460(2051), 3065–3094 (2004)

    Article  ADS  MathSciNet  MATH  Google Scholar 

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

    MATH  Google Scholar 

  10. Grabovsky, Y., Kucher, V.A., Truskinovsky, L.: Weak variations of Lipschitz graphs and stability of phase boundaries. Contin. Mech. Thermodyn. 23(2), 87–123 (2011)

    Article  ADS  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  12. Grabovsky, Y., Truskinovsky, L.: Marginal material stability. J. Nonlinear Sci. 23(5), 891–969 (2013)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  13. Grinfeld, M.A.: Stability of interphase boundaries in solid elastic media. Prikl. Mat. Meh. 51(4), 628–637 (1987)

    MathSciNet  Google Scholar 

  14. Hadamard, J.: Sur quelques questions de Calcul des variations. Bull. Soc. Math. Fr. 30, 253–256 (1902)

    Google Scholar 

  15. James, R.D.: Finite deformation by mechanical twinning. Arch. Ration. Mech. Anal. 77(2), 143–176 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  16. Jensen, H.M., Christoffersen, J.: Kink band formation in fiber reinforced materials. J. Mech. Phys. Solids 45(7), 1121–1136 (1997)

    Article  ADS  Google Scholar 

  17. 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)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  ADS  MathSciNet  MATH  Google Scholar 

  19. McShane, E.J.: On the necessary condition of Weierstrass in the multiple integral problem of the calculus of variations. Ann. Math. 32(3), 578–590 (1931)

    Article  MathSciNet  MATH  Google Scholar 

  20. Merodio, J., Pence, T.: Kink surfaces in a directionally reinforced neo-Hookean material under plane deformation: I. Mechanical equilibrium. J. Elast. 62(2), 119–144 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  21. Merodio, J., Pence, T.: Kink surfaces in a directionally reinforced neo-Hookean material under plane deformation: II. Kink band stability and maximally dissipative band broadening. J. Elast. 62(2), 145–170 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  22. Ogden, R.W.: Large deformation isotropic elasticity: on the correlation of theory and experiment for compressible rubberlike solids. Proc. R. Soc. Lond. Ser. A, Math. Phys. Sci. 328(1575), 567–583 (1972)

    Article  ADS  MATH  Google Scholar 

  23. Simpson, H.C., Spector, S.J.: Some necessary conditions at an internal boundary for minimizers in finite elasticity. J. Elast. 26(3), 203–222 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  24. Taheri, A.: On critical points of functionals with polyconvex integrands. J. Convex Anal. 9(1), 55–72 (2002)

    MathSciNet  MATH  Google Scholar 

  25. Truesdell, C.A.: A First Course in Rational Continuum Mechanics, vol. 1, 2nd edn. Academic Press, New York (1991). General concepts

    MATH  Google Scholar 

Download references

Acknowledgements

This material is based upon work supported by the National Science Foundation under Grant No. DMS-1412058. LT was also supported by the French ANR contract EVOCRIT.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yury Grabovsky.

Appendices

Appendix A: Geometry of the Jump Set

1.1 A.1 Normal to the Jump Set

Suppose \(\boldsymbol{F}_{-}(s)\in\mathfrak{J}_{-}\) is a smooth curve passing through \(\boldsymbol{F}_{-}\) at \(s=0\). We would like to show by the implicit function theorem that there are smooth curves \(\boldsymbol{a}(s)\), \(\boldsymbol{n}(s)\) passing through \(\boldsymbol{a}\) and \(\boldsymbol{n}\) at \(s=0\) and satisfying

$$ \textstyle\begin{cases} (W_{\boldsymbol{F}}(\boldsymbol{F}_{-}(s)+\boldsymbol{a}(s)\otimes \boldsymbol{n}(s))-W_{\boldsymbol{F}}(\boldsymbol {F}_{-}(s)))\boldsymbol{n}(s)=\mathbf{0},\\ (W_{\boldsymbol{F}}(\boldsymbol{F}_{-}(s)+\boldsymbol{a}(s)\otimes \boldsymbol{n}(s))-W_{\boldsymbol{F}}(\boldsymbol {F}_{-}))^{T}\boldsymbol{a}(s)=\mathbf{0},\\ W(\boldsymbol{F}_{-}(s)+\boldsymbol{a}(s)\otimes\boldsymbol {n}(s))-W(\boldsymbol{F}_{-}(s))-W_{\boldsymbol{F}}(\boldsymbol {F}_{-}(s))\boldsymbol{n}(s)\cdot\boldsymbol{a}(s)=0. \end{cases} $$
(A.1)

Differentiating (A.1) with respect to \(s\) at \(s=0\) we obtain

$$ \textstyle\begin{cases} \boldsymbol{A}_{+}\dot{\boldsymbol{a}}+(\boldsymbol{B}_{+}+{}[ \!{}[\boldsymbol{P} ]\!])\dot{\boldsymbol {n}}+({}[\!{}[\mathsf{L} ]\!]\dot{\boldsymbol {F}}_{-})\boldsymbol{n}=\mathbf{0},\\ (\boldsymbol{B}_{+}+{}[\!{}[\boldsymbol{P} ]\! ])^{T}\dot{\boldsymbol{a}}+\boldsymbol{A}_{+}^{*}\dot {\boldsymbol{n}}+({}[\!{}[\mathsf{L} ]\!]\dot {\boldsymbol{F}}_{-})^{T}\boldsymbol{a}=\mathbf{0},\\ -\frac{1}{2}(\boldsymbol{A}_{+}\boldsymbol{a},\dot{\boldsymbol {a}})-\frac{1}{2}(\boldsymbol{A}_{+}^{*}\boldsymbol{n},\dot {\boldsymbol{n}})+({}[\!{}[\boldsymbol{P} ]\!] -\lbrace\!\!\lbrace\mathsf{L}\rbrace\!\!\rbrace{}[\!{}[ \boldsymbol{F} ]\!],\dot{\boldsymbol{F}}_{-})=0. \end{cases} $$
(A.2)

We now solve the first two equations in (A.2) for \(\dot {\boldsymbol{a}}\), \(\dot{\boldsymbol{n}}\). We obtain the system

$$ \mathbb{ C}_{+}\left [ \textstyle\begin{array}{c} {\dot{\boldsymbol{a}}}\\ {\dot{\boldsymbol{n}}} \end{array}\displaystyle \right ]=-\left [ \textstyle\begin{array}{c} {({}[\!{}[\mathsf{L} ]\!]\dot{\boldsymbol {F}}_{-})\boldsymbol{n}}\\ {({}[\!{}[\mathsf{L} ]\! ]\dot{\boldsymbol{F}}_{-})^{T}\boldsymbol{a}} \end{array}\displaystyle \right ], $$
(A.3)

where \(\mathbb{ C}_{\pm}\) is given by (3.1). Observe that

$$\left [ \textstyle\begin{array}{c} {({}[\!{}[\mathsf{L} ]\!]\dot{\boldsymbol {F}})\boldsymbol{n}}\\ {({}[\!{}[\mathsf{L} ]\! ]\dot{\boldsymbol{F}}_{-})^{T}\boldsymbol{a}} \end{array}\displaystyle \right ]\cdot \left [ \textstyle\begin{array}{c} {\boldsymbol{a}}\\ {-\boldsymbol{n}} \end{array}\displaystyle \right ]=0. $$

By the non-degeneracy assumption the right-hand side of (A.3) is orthogonal to the null-space of \(\mathbb{ C}_{+}\). Thus, (A.3) has a solution. This solution is determined uniquely by the property \(\dot{\boldsymbol{n}}\cdot\boldsymbol{n}=0\), since the vector \([\mathbf{0},\boldsymbol{n}]\) does not belong to \((\mathbb{ R}[\boldsymbol{a},-\boldsymbol{n}])^{\perp}\). Thus, there is a well defined operator denoted \(\mathbb{ C}_{+}^{-1}:(\mathbb{ R}[\boldsymbol{a},-\boldsymbol {n}])^{\perp}\to(\mathbb{ R}[\mathbf{0},\boldsymbol{n}])^{\perp}\) with the property \(\mathbb{ C}_{+}(\boldsymbol{a},\boldsymbol{n})(\mathbb { C}_{+}^{-1}\boldsymbol{z})=\boldsymbol{z}\) for any \(\boldsymbol{z}\in(\mathbb{ R}[\boldsymbol{a},-\boldsymbol {n}])^{\perp}\). We can write

$$\left [ \textstyle\begin{array}{c} {\dot{\boldsymbol{a}}}\\ {\dot{\boldsymbol{n}}} \end{array}\displaystyle \right ]=-\mathbb{ C}_{+}^{-1}\left [ \textstyle\begin{array}{c} {({}[\!{}[\mathsf{L} ]\!]\dot{\boldsymbol {F}}_{-})\boldsymbol{n}}\\ {({}[\!{}[\mathsf{L} ]\! ]\dot{\boldsymbol{F}}_{-})^{T}\boldsymbol{a}} \end{array}\displaystyle \right ]. $$

Substituting this into the third equation in (A.2) we obtain

$$ \frac{1}{2}\left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{+}(\boldsymbol{n})\boldsymbol{a}}\\ {\boldsymbol {A}_{+}^{*}(\boldsymbol{a})\boldsymbol{n}} \end{array}\displaystyle \right ]\cdot\mathbb{ C}_{+}^{-1} \left [ \textstyle\begin{array}{c} {({}[\!{}[\mathsf{L} ]\!]\dot{\boldsymbol {F}}_{-})\boldsymbol{n}}\\ {({}[\!{}[\mathsf{L} ]\! ]\dot{\boldsymbol{F}}_{-})^{T}\boldsymbol{a}} \end{array}\displaystyle \right ]+ \bigl({}\bigl[\!{}[\boldsymbol{P} ]\!\bigr]-\bigl\lbrace \! \! \lbrace\mathsf{L}\rbrace\!\!\bigr\rbrace {}\bigl[\!{}[\boldsymbol {F} ]\!\bigr], \dot{\boldsymbol{F}}_{-}\bigr)=0. $$
(A.4)

Observe that \(\mathbb{ C}_{+}\) is a symmetric matrix and that

$$\left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{+}\boldsymbol{a}}\\ {\boldsymbol {A}_{+}^{*}\boldsymbol{n}} \end{array}\displaystyle \right ]\cdot \left [ \textstyle\begin{array}{c} {\boldsymbol{a}}\\ {-\boldsymbol{n}} \end{array}\displaystyle \right ]=0. $$

Therefore,

$$\left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{+}\boldsymbol{a}}\\ {\boldsymbol {A}_{+}^{*}\boldsymbol{n}} \end{array}\displaystyle \right ]\cdot\mathbb{ C}_{+}^{-1} \left [ \textstyle\begin{array}{c} {({}[\!{}[\mathsf{L} ]\!]\dot{\boldsymbol {F}}_{-})\boldsymbol{n}}\\ {({}[\!{}[\mathsf{L} ]\! ]\dot{\boldsymbol{F}}_{-})^{T}\boldsymbol{a}} \end{array}\displaystyle \right ]= \mathbb{ C}_{+}^{-1}\left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{+}\boldsymbol{a}}\\ {\boldsymbol {A}_{+}^{*}\boldsymbol{n}} \end{array}\displaystyle \right ]\cdot \left [ \textstyle\begin{array}{c} {({}[\!{}[\mathsf{L} ]\!]\dot{\boldsymbol {F}}_{-})\boldsymbol{n}}\\ {({}[\!{}[\mathsf{L} ]\! ]\dot{\boldsymbol{F}}_{-})^{T}\boldsymbol{a}} \end{array}\displaystyle \right ]. $$

We easily compute

$$\mathbb{ C}_{+}^{-1}\left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{+}\boldsymbol{a}}\\ {\boldsymbol {A}_{+}^{*}\boldsymbol{n}} \end{array}\displaystyle \right ]=\left [ \textstyle\begin{array}{c} {\boldsymbol{a}}\\ {\mathbf{0}} \end{array}\displaystyle \right ]. $$

Hence, (A.4) becomes

$$\biggl(\frac{1}{2}{}[\!{}[\mathsf{L} ]\!] {}[\!{}[ \boldsymbol{F} ]\!]+{}[\!{}[ \boldsymbol{P} ]\!]-\lbrace\!\! \lbrace\mathsf {L}\rbrace\!\!\rbrace{}[\!{}[\boldsymbol{F} ]\! ], \dot{\boldsymbol{F}}_{-} \biggr)=0. $$

In other words, \((\boldsymbol{N}_{-},\dot{\boldsymbol{F}}_{-})=0\), where

$$ \boldsymbol{N}_{-}=\mathsf{L}_{-}{}[ \!{}[\boldsymbol{F} ]\!]-{}[\!{}[\boldsymbol{P} ]\!]. $$
(A.5)

1.2 A.2 Foliation of the Simple Laminate Region

Now, suppose that \(\boldsymbol{F}_{-}\) varies over an open subset \(G\) of \(\mathfrak{J}_{-}\) that is sufficiently small, so that due to the non-degeneracy assumption, the functions \(\boldsymbol{a}(\boldsymbol{F}_{-})\) and \(\boldsymbol {n}(\boldsymbol{F}_{-})\) are well-defined and smooth on \(G\). Let us show that the line segments joining \(\boldsymbol{F}_{-}\) and \(\boldsymbol{F}_{+}=\boldsymbol{F}_{-}+\boldsymbol{a}(\boldsymbol {F}_{-})\otimes\boldsymbol{n}(\boldsymbol{F}_{-})\) foliate an open subset of \(\mathbb{ M}\), as \(\boldsymbol{F}_{-}\) ranges over \(G\). We have a map \(\boldsymbol{F}:G\times[0,1]\to \mathbb{ M}\) given by

$$ \boldsymbol{F}=\boldsymbol{F}_{-}+t\boldsymbol{a}( \boldsymbol {F}_{-})\otimes\boldsymbol{n}(\boldsymbol{F}_{-}). $$
(A.6)

If \(G\) is sufficiently small then the only possibility that the map \(\boldsymbol{F}\) is not injective is for \(\boldsymbol{F}\) to be locally non-injective, i.e., for \(d\boldsymbol{F}(\boldsymbol{F}_{-},t)\) to be singular for some \(t\in[0,1]\). If this is the case, then there exists \(\dot{\boldsymbol{F}}_{-}\in T_{\boldsymbol{F}_{-}}\mathfrak{J}_{-}\) and \(t\in[0,1]\) such that

$$\dot{\boldsymbol{F}}_{-}+t(\dot{\boldsymbol{a}}\otimes\boldsymbol {n}+\boldsymbol{a}\otimes\dot{\boldsymbol{n}})=\boldsymbol {a}\otimes \boldsymbol{n}, $$

where \([\dot{\boldsymbol{a}},\dot{\boldsymbol{n}}]\) solves (A.3). If (A.6) is satisfied then there exist \(\boldsymbol{u}\in\mathbb{ R}^{m}\) and \(\boldsymbol{\eta}\in\mathbb{ R}^{d}\) such that

$$ \dot{\boldsymbol{F}}_{-}=\boldsymbol{u}\otimes \boldsymbol {n}+\boldsymbol{a}\otimes\boldsymbol{\eta},\qquad \boldsymbol{u}+t \dot{\boldsymbol{a}}=\lambda\boldsymbol{a},\qquad \boldsymbol{\eta}+t\dot{ \boldsymbol{n}}=(1-\lambda)\boldsymbol{n} $$
(A.7)

for some \(\lambda\in\mathbb{ R}\). Substituting (A.7) into (A.3) we obtain

$$\widetilde{\mathbb{ C}}_{t}\left [ \textstyle\begin{array}{c} {\boldsymbol{u}}\\ {\boldsymbol{\eta}} \end{array}\displaystyle \right ]=\left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{+}(\boldsymbol{n})\boldsymbol{a}}\\ {\boldsymbol {A}_{+}^{*}(\boldsymbol{a})\boldsymbol{n}} \end{array}\displaystyle \right ],\quad \widetilde{ \mathbb{ C}}_{t}=(1-t)\mathbb{ C}_{+}+t\mathbb{ C}_{-}. $$

The solution \([\boldsymbol{u},\boldsymbol{\eta}]\) is determined up to a multiple of \([\boldsymbol{a},-\boldsymbol{n}]\). However, any choice of solution \([\boldsymbol{u},\boldsymbol{\eta}]\) gives one and the same value \(\dot{\boldsymbol{F}}_{-}\) in (A.7). We therefore, write

$$\left [ \textstyle\begin{array}{c} {\boldsymbol{u}}\\ {\boldsymbol{\eta}} \end{array}\displaystyle \right ]=\widetilde{\mathbb{ C}}_{t}^{-1}\left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{+}(\boldsymbol{n})\boldsymbol{a}}\\ {\boldsymbol {A}_{+}^{*}(\boldsymbol{a})\boldsymbol{n}} \end{array}\displaystyle \right ], $$

where the inverse of the positive semidefinite matrix \(\widetilde {\mathbb{ C}}_{t}\) is computed on \((\mathbb{ R}[\boldsymbol{a},-\boldsymbol{n}])^{\perp}\), where it is positive definite by assumption of local stability. The resulting \(\dot{\boldsymbol{F}}_{-}\) has to belong to \(T_{\boldsymbol{F}_{-}}\mathfrak{J}_{-}\), therefore \((\dot {\boldsymbol{F}}_{-},\boldsymbol{N}_{-})=0\), i.e.,

$$ q(t)=\left (\widetilde{\mathbb{ C}}_{t}^{-1} \left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{+}(\boldsymbol{n})\boldsymbol{a}}\\ {\boldsymbol {A}_{+}^{*}(\boldsymbol{a})\boldsymbol{n}} \end{array}\displaystyle \right ], \left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{-}(\boldsymbol{n})\boldsymbol{a}}\\ {\boldsymbol {A}_{-}^{*}(\boldsymbol{a})\boldsymbol{n}} \end{array}\displaystyle \right ] \right )=0. $$
(A.8)

Let us show that (A.8) is impossible. We observe that

$$q(0)=\bigl(\boldsymbol{A}_{-}(\boldsymbol{n})\boldsymbol{a}, \boldsymbol {a}\bigr)>0,\qquad q(1)=\bigl(\boldsymbol{A}_{+}( \boldsymbol{n})\boldsymbol {a},\boldsymbol{a}\bigr)>0. $$

For any \(t\in(0,1)\) we have

$$\widetilde{\mathbb{ C}}_{t}>t\mathbb{ C_{-}}>0, $$

where the inequality is strict on \((\mathbb{ R}[\boldsymbol {a},-\boldsymbol{n}])^{\perp}\). Therefore,

$$ t\widetilde{\mathbb{ C}}_{t}^{-1}< \mathbb{ C}_{-}^{-1}. $$
(A.9)

We have

$$\widetilde{\mathbb{ C}}_{t}\left [ \textstyle\begin{array}{c} {\boldsymbol{a}}\\ {\boldsymbol{n}} \end{array}\displaystyle \right ]=2(1-t)\left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{+}(\boldsymbol{n})\boldsymbol{a}}\\ {\boldsymbol {A}_{+}^{*}(\boldsymbol{a})\boldsymbol{n}} \end{array}\displaystyle \right ] +2t\left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{-}(\boldsymbol{n})\boldsymbol{a}}\\ {\boldsymbol {A}_{-}^{*}(\boldsymbol{a})\boldsymbol{n}} \end{array}\displaystyle \right ]. $$

Therefore,

$$\left [ \textstyle\begin{array}{c} {\boldsymbol{a}}\\ {\boldsymbol{n}} \end{array}\displaystyle \right ]-\lambda \left [ \textstyle\begin{array}{c} {\boldsymbol{a}}\\ {-\boldsymbol{n}} \end{array}\displaystyle \right ]= 2(1-t)\widetilde{\mathbb{ C}}_{t}^{-1}\left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{+}(\boldsymbol{n})\boldsymbol{a}}\\ {\boldsymbol {A}_{+}^{*}(\boldsymbol{a})\boldsymbol{n}} \end{array}\displaystyle \right ] +2t\widetilde{\mathbb{ C}}_{t}^{-1}\left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{-}(\boldsymbol{n})\boldsymbol{a}}\\ {\boldsymbol {A}_{-}^{*}(\boldsymbol{a})\boldsymbol{n}} \end{array}\displaystyle \right ], $$

where \(\lambda\in\mathbb{ R}\) is chosen such that the left-hand side is in \((\mathbb{ R}[\boldsymbol{a},-\boldsymbol{n}])^{\perp}\). Taking inner product with \([\boldsymbol{A}_{-}(\boldsymbol{n})\boldsymbol{a},\boldsymbol {A}_{-}^{*}(\boldsymbol{a})\boldsymbol{n}]\) we obtain

$$\bigl(\boldsymbol{A}_{-}(\boldsymbol{n})\boldsymbol{a},\boldsymbol{a} \bigr)=(1-t)q(t)+ t\left (\widetilde{\mathbb{ C}}_{t}^{-1} \left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{-}(\boldsymbol{n})\boldsymbol{a}}\\ {\boldsymbol {A}_{-}^{*}(\boldsymbol{a})\boldsymbol{n}} \end{array}\displaystyle \right ], \left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{-}(\boldsymbol{n})\boldsymbol{a}}\\ {\boldsymbol {A}_{-}^{*}(\boldsymbol{a})\boldsymbol{n}} \end{array}\displaystyle \right ] \right ). $$

By (A.9) we have

$$\bigl(\boldsymbol{A}_{-}(\boldsymbol{n})\boldsymbol{a},\boldsymbol{a} \bigr)< (1-t)q(t)+ \left (\mathbb{ C}_{-}^{-1}\left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{-}(\boldsymbol{n})\boldsymbol{a}}\\ {\boldsymbol {A}_{-}^{*}(\boldsymbol{a})\boldsymbol{n}} \end{array}\displaystyle \right ], \left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{-}(\boldsymbol{n})\boldsymbol{a}}\\ {\boldsymbol {A}_{-}^{*}(\boldsymbol{a})\boldsymbol{n}} \end{array}\displaystyle \right ] \right )=(1-t)q(t)+\bigl(\boldsymbol{A}_{-}( \boldsymbol {n})\boldsymbol{a},\boldsymbol{a}\bigr). $$

Thus, \(q(t)>0\) for all \(t\in[0,1]\). It follows that \({\mathcal {G}}=\boldsymbol{F}(G\times(0,1))\) is an open subset of \(\mathbb{ M}\).

Appendix B: Formula for the Second Derivative of \(\overline{W}(\boldsymbol{F})\)

Let \(\boldsymbol{F}_{0}\in\mathfrak{O}\) and let \(\boldsymbol{F}(s)\) be any curve in \(\mathfrak{O}\) passing through \(\boldsymbol{F}_{0}\) when \(s=0\). Then, for sufficiently small \(s\) there exist unique \(\boldsymbol{F}_{-}(s)\in \mathfrak{J}\), \(\boldsymbol{a}(s)\otimes\boldsymbol{n}(s)\) and \(t(s)\in(0,1)\), such that \(\boldsymbol{F}(s)=t(s)\boldsymbol{F}_{+}(s)+(1-t(s))\boldsymbol {F}_{-}(s)\) and (2.5) holds, where \(\boldsymbol{F}_{+}(s)=\boldsymbol{F}_{-}(s)+\boldsymbol {a}(s)\otimes\boldsymbol{n}(s)\in\mathfrak{J}\). Let \(\boldsymbol{\xi}=\dot{\boldsymbol{F}}(0)\). In the calculations below dot over a symbol denotes derivative in \(s\) at \(s=0\), while a symbol without an argument \(s\) refers to \(s=0\). For example, \(\boldsymbol{a}\) denotes \(\boldsymbol{a}(0)\).

Differentiating (4.1) at \(s=0\) we obtain

$$ \bigl\langle \overline{W}_{\boldsymbol{F}}(\boldsymbol {F}_{0}),\boldsymbol{\xi} \bigr\rangle =t\langle\boldsymbol{P}_{+}, \dot {\boldsymbol{F}}_{+} \rangle+(1-t)\langle\boldsymbol{P}_{-}, \dot {\boldsymbol{F}}_{-} \rangle- \dot{t} {}[\!{}[ W ]\!]. $$
(B.1)

To compute \(\dot{\boldsymbol{F}}_{\pm}\) we need to express \(\boldsymbol{F}_{\pm}(s)\) in terms of \(\boldsymbol{F}(s)\) and \(\boldsymbol{a}(s)\otimes\boldsymbol{n}(s)\):

$$ \boldsymbol{F}_{-}(s)=\boldsymbol{F}(s)-t(s) \boldsymbol{a}(s)\otimes \boldsymbol{n}(s),\qquad \boldsymbol{F}_{+}(s)= \boldsymbol{F}(s)+\bigl(1-t(s)\bigr)\boldsymbol {a}(s)\otimes\boldsymbol{n}(s). $$
(B.2)

Differentiating in \(s\) at \(s=0\) we get

$$ \dot{\boldsymbol{F}}_{+}=\boldsymbol{\xi}-\dot{t} {} [\!{}[ \boldsymbol{F} ]\!]+(1-t){}[\!{}[\dot {\boldsymbol{F}} ]\! ],\qquad \dot{\boldsymbol{F}}_{-}=\boldsymbol{\xi}-\dot{t} {} [\!{}[ \boldsymbol{F} ]\!]-t{}[\!{}[\dot{\boldsymbol {F}} ]\! ], $$
(B.3)

where \({}[\!{}[\dot{\boldsymbol{F}} ]\!]\) is just a compact form of

$$ {}[\!{}[\dot{\boldsymbol{F}} ]\!]=\boldsymbol {a} \otimes\dot{\boldsymbol{n}}+\dot{\boldsymbol{a}}\otimes \boldsymbol{n}. $$
(B.4)

Substituting (B.3) into (B.1) we obtain

$$\begin{aligned} \bigl\langle \overline{W}_{\boldsymbol{F}}(\boldsymbol {F}_{0}), \boldsymbol{\xi} \bigr\rangle =&\bigl\langle t\boldsymbol {P}_{+}+(1-t) \boldsymbol{P}_{-},\boldsymbol{\xi} \bigr\rangle +\dot{t}\bigl\{ {} [\!{}[ W ]\!]-\bigl\langle t\boldsymbol {P}_{+}+(1-t) \boldsymbol{P}_{-},{}[\!{}[\boldsymbol{F} ]\!] \bigr\rangle \bigr\} \\ &{}+t(1-t)\bigl\langle {}[\!{}[ \boldsymbol{P} ]\!],{}[\!{}[ \dot{\boldsymbol {F}} ]\!] \bigr\rangle . \end{aligned}$$

Using (2.5) we conclude that

$$ \overline{W}_{\boldsymbol{F}}\bigl(\boldsymbol{F}(s) \bigr)=t(s)W_{\boldsymbol {F}}\bigl(\boldsymbol{F}_{+}(s)\bigr)+ \bigl(1-t(s)\bigr)W_{\boldsymbol{F}}\bigl(\boldsymbol {F}_{-}(s)\bigr). $$
(B.5)

Differentiating (B.5) at \(s=0\) and using (B.3) we see that

$$ \overline{W}_{\boldsymbol{F}\boldsymbol{F}}(\boldsymbol {F}_{0}) \boldsymbol{\xi}=\mathsf{L}_{t}\boldsymbol{\xi}-\dot {t} \boldsymbol{N}_{t}+t(1-t){}[\!{}[\mathsf{L} ]\! ]{}[ \!{}[\dot{\boldsymbol{F}} ]\!], $$
(B.6)

where we use the shorthand \(X_{t}=tX_{+}+(1-t)X_{-}\).

In order to express \(\dot{\boldsymbol{a}}\), \(\dot{\boldsymbol{n}}\) and \(\dot{t}\) in terms of \(\boldsymbol{\xi}\) we differentiate the last two equations in (2.5) at \(s=0\) and obtain

$$ \widetilde{\mathbb{ C}}_{t}\left [ \textstyle\begin{array}{c} {\dot{\boldsymbol{a}}}\\ {\dot{\boldsymbol{n}}} \end{array}\displaystyle \right ]-\dot{t}\left [ \textstyle\begin{array}{c} {{}[\!{}[\boldsymbol{A} ]\!]\boldsymbol{a}}\\ {{}[\!{}[\boldsymbol{A}^{*} ]\!]\boldsymbol{n}} \end{array}\displaystyle \right ]= -\left [ \textstyle\begin{array}{c} {({}[\!{}[\mathsf{L} ]\!]\boldsymbol{\xi })\boldsymbol{n}}\\ {({}[\!{}[\mathsf{L} ]\!] \boldsymbol{\xi})^{T}\boldsymbol{a}} \end{array}\displaystyle \right ]. $$
(B.7)

Differentiating the Maxwell relation we obtain \(\langle\boldsymbol{N}_{\pm},\dot{\boldsymbol{F}}_{\pm} \rangle =0\) (which is expressing the fact that \(\boldsymbol{N}_{\pm}\) are orthogonal to the jump set at \(\boldsymbol {F}_{\pm}\)). Using (B.3) to eliminate \(\dot{\boldsymbol{F}}_{\pm}\) we obtain

$$\dot{t}\boldsymbol{A}_{-}(\boldsymbol{n})\boldsymbol{a}\cdot \boldsymbol{a}=\langle\boldsymbol{N}_{-},\boldsymbol{\xi} \rangle -t \bigl\langle \boldsymbol{N}_{-},{}[\!{}[\dot{\boldsymbol{F}} ]\! ] \bigr\rangle ,\qquad \dot{t}\boldsymbol{A}_{+}(\boldsymbol{n}) \boldsymbol{a}\cdot \boldsymbol{a}=\langle\boldsymbol{N}_{+}, \boldsymbol{\xi} \rangle +(1-t)\bigl\langle \boldsymbol{N}_{+},{}[ \!{}[\dot{\boldsymbol {F}} ]\!] \bigr\rangle . $$

We combine the two relations into a more symmetric form by multiplying the first equation by \(1-t\), the second one by \(t\) and adding:

$$\dot{t}= \frac{\langle\boldsymbol{N}_{t},\boldsymbol{\xi} \rangle +t(1-t)\langle{}[\!{}[\boldsymbol{N} ]\!] ,{}[\!{}[\dot{\boldsymbol{F}} ]\!] \rangle }{\boldsymbol{A}_{t}\boldsymbol{a}\cdot\boldsymbol{a}}. $$

Substituting this formula into (B.6) and (B.7) we obtain

$$ \langle\overline{W}_{\boldsymbol{F}\boldsymbol{F}}\boldsymbol{\xi },\boldsymbol{ \xi} \rangle=\langle\mathsf{L}_{t}\boldsymbol{\xi },\boldsymbol{\xi} \rangle-\frac{\langle\boldsymbol {N}_{t},\boldsymbol{\xi} \rangle^{2}}{\boldsymbol{A}_{t}\boldsymbol {a}\cdot\boldsymbol{a}}-t(1-t) \biggl(\bigl\langle {}[\!{}[ \mathsf{L} ]\! ]\boldsymbol{\xi},{}[\!{}[\dot {\boldsymbol{F}} ]\!] \bigr\rangle - \frac{\langle \boldsymbol{N}_{t},\boldsymbol{\xi} \rangle\langle{}[\! {}[\boldsymbol{N} ]\!],{}[\!{}[\dot {\boldsymbol{F}} ]\!] \rangle}{\boldsymbol {A}_{t}\boldsymbol{a}\cdot\boldsymbol{a}} \biggr), $$
(B.8)

and

$$ \mathbb{ D}(t)\left [ \textstyle\begin{array}{c} {\dot{\boldsymbol{a}}}\\ {\dot{\boldsymbol{n}}} \end{array}\displaystyle \right ]= -\left [ \textstyle\begin{array}{c} {({}[\!{}[\mathsf{L} ]\!]\boldsymbol{\xi })\boldsymbol{n}}\\ {({}[\!{}[\mathsf{L} ]\!] \boldsymbol{\xi})^{T}\boldsymbol{a}} \end{array}\displaystyle \right ]+\frac{\langle\boldsymbol{N}_{t},\boldsymbol{\xi} \rangle }{\boldsymbol{A}_{t}\boldsymbol{a}\cdot\boldsymbol{a}} \left [ \textstyle\begin{array}{c} {{}[\!{}[\boldsymbol{A} ]\!]\boldsymbol{a}}\\ {{}[\!{}[\boldsymbol{A}^{*} ]\!]\boldsymbol{n}} \end{array}\displaystyle \right ], $$
(B.9)

respectively, where

$$ \mathbb{ D}(t)=\widetilde{\mathbb{ C}}_{t}- \frac {t(1-t)}{\boldsymbol{A}_{t}\boldsymbol{a}\cdot\boldsymbol{a}}\left [ \textstyle\begin{array}{c} {{}[\!{}[\boldsymbol{A} ]\!]\boldsymbol{a}}\\ {{}[\!{}[\boldsymbol{A}^{*} ]\!]\boldsymbol{n}} \end{array}\displaystyle \right ]\otimes \left [ \textstyle\begin{array}{c} {{}[\!{}[\boldsymbol{A} ]\!]\boldsymbol{a}}\\ {{}[\!{}[\boldsymbol{A}^{*} ]\!]\boldsymbol{n}} \end{array}\displaystyle \right ]. $$
(B.10)

Hence,

$$ \bigl\langle \overline{W}_{\boldsymbol{F}\boldsymbol{F}}(\boldsymbol {F}_{0})\boldsymbol{\xi},\boldsymbol{\xi} \bigr\rangle =\langle\mathsf {L}_{t}\boldsymbol{\xi},\boldsymbol{\xi} \rangle-\frac{\langle \boldsymbol{N}_{t},\boldsymbol{\xi} \rangle^{2}}{\boldsymbol {A}_{t}\boldsymbol{a}\cdot\boldsymbol{a}}+t(1-t) \mathbb{ D}(t)^{-1}\boldsymbol{Z}(t)\cdot\boldsymbol{Z}(t), $$
(B.11)

where

$$\boldsymbol{Z}(t)=-\left [ \textstyle\begin{array}{c} {({}[\!{}[\mathsf{L} ]\!]\boldsymbol{\xi })\boldsymbol{n}}\\ {({}[\!{}[\mathsf{L} ]\!] \boldsymbol{\xi})^{T}\boldsymbol{a}} \end{array}\displaystyle \right ]+ \frac{\langle\boldsymbol{N}_{t},\boldsymbol{\xi} \rangle }{\boldsymbol{A}_{t}\boldsymbol{a}\cdot\boldsymbol{a}}\left [ \textstyle\begin{array}{c} {{}[\!{}[\boldsymbol{A} ]\!]\boldsymbol{a}}\\ {{}[\!{}[\boldsymbol{A}^{*} ]\!]\boldsymbol{n}} \end{array}\displaystyle \right ]. $$

In the last term of (B.11) \(\mathbb{ D}(t)^{-1}\boldsymbol{Z}(t)\) is understood in the orthogonal complement to \([\boldsymbol{a},-\boldsymbol{n}]\). This is justified by the following lemma.

Lemma B.1

Assume that \(\boldsymbol{F}_{\pm}\) is a non-degenerate pair in the sense of Definition  4.1. Assume that (3.2) holds. Then \(\mathbb{ D}(t)\) is a positive-definite self-adjoint operator on \(V=(\mathbb{ R}[\boldsymbol{a},-\boldsymbol{n}])^{\perp}\).

Proof

It is obvious that \(\mathbb{ D}(t)\) is self-adjoint and that \(\mathbb{ D}(t)[\boldsymbol{a},-\boldsymbol{n}]=0\). Thus, \(\mathbb{ D}(t)\) is a self-adjoint operator on \(V\). By assumption of non-degeneracy, \(\widetilde{\mathbb{ C}}_{t}\) is a positive-definite self-adjoint operator on \(V\). Thus, in order to prove the lemma we need to show that

$$ \widetilde{\mathbb{ C}}_{t}^{-1} \boldsymbol{X}\cdot\boldsymbol {X}< \frac{\boldsymbol{A}_{t}\boldsymbol{a}\cdot\boldsymbol {a}}{t(1-t)},\quad \boldsymbol{X}= \left [ \textstyle\begin{array}{c} {{}[\!{}[\boldsymbol{A} ]\!]\boldsymbol{a}}\\ {{}[\!{}[\boldsymbol{A}^{*} ]\!]\boldsymbol{n}} \end{array}\displaystyle \right ] $$
(B.12)

for every \(t\in[0,1]\).

We observe that

$$\widetilde{\mathbb{ C}}_{t}\left [ \textstyle\begin{array}{c} {\boldsymbol{a}}\\ {\boldsymbol{n}} \end{array}\displaystyle \right ]=2\left [ \textstyle\begin{array}{c} {\widetilde{\boldsymbol{A}}_{t}\boldsymbol{a}}\\ {\widetilde {\boldsymbol{A}}_{t}^{*}\boldsymbol{n}} \end{array}\displaystyle \right ]. $$

We have

$$\boldsymbol{X}=\frac{1}{t}\left (\left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{+}\boldsymbol{a}}\\ {\boldsymbol {A}_{+}^{*}\boldsymbol{n}} \end{array}\displaystyle \right ]-\left [ \textstyle\begin{array}{c} {\widetilde{\boldsymbol{A}}_{t}\boldsymbol{a}}\\ {\widetilde {\boldsymbol{A}}_{t}^{*}\boldsymbol{n}} \end{array}\displaystyle \right ] \right ). $$

Hence,

$$\widetilde{\mathbb{ C}}_{t}^{-1}\boldsymbol{X}\cdot \boldsymbol {X}=\frac{\widetilde{\boldsymbol{A}}_{t}\boldsymbol{a}\cdot \boldsymbol{a}-2\boldsymbol{A}_{+}\boldsymbol{a}\cdot\boldsymbol {a}+\widetilde{\mathbb{ C}}_{t}^{-1}\boldsymbol{Y}_{+}\cdot \boldsymbol{Y}_{+}}{t^{2}}, $$

where

$$\boldsymbol{Y}_{+}=\left [ \textstyle\begin{array}{c} {\boldsymbol{A}_{+}\boldsymbol{a}}\\ {\boldsymbol {A}_{+}^{*}\boldsymbol{n}} \end{array}\displaystyle \right ]. $$

When \(t=0\) or 1, the statement is obvious. For \(t\in(0,1)\) we have

$$\widetilde{\mathbb{ C}}_{t}>(1-t)\mathbb{ C}_{+}, $$

in the sense of quadratic forms on \(V\), so that

$$\widetilde{\mathbb{ C}}_{t}^{-1}\boldsymbol{Y}_{+} \cdot\boldsymbol {Y}_{+}< \frac{1}{1-t}\mathbb{ C}_{+}^{-1}\boldsymbol{Y}_{+}\cdot \boldsymbol{Y}_{+}=\frac{\boldsymbol{A}_{+}\boldsymbol{a}\cdot \boldsymbol{a}}{1-t}. $$

Thus, we obtain

$$\widetilde{\mathbb{ C}}_{t}^{-1}\boldsymbol{X}\cdot \boldsymbol {X}< \frac{\boldsymbol{A}_{t}\boldsymbol{a}\cdot\boldsymbol{a}}{t(1-t)}, $$

as required. □

Finally, let us show that the acoustic tensor of \(\overline {W}(\boldsymbol{F})\) at \(\boldsymbol{F}_{0}=t\boldsymbol{F}_{+}+(1-t)\boldsymbol{F}_{-}\) is nonnegative whenever it is nonnegative at both \(\boldsymbol{F}_{+}\) and \(\boldsymbol{F}_{-}\). Indeed, the last term in (B.11) is nonnegative, by Lemma B.1. It remains to observe that the function

$$\phi(t)=\langle\mathsf{L}_{t}\boldsymbol{\xi},\boldsymbol{\xi} \rangle-\frac{\langle\boldsymbol{N}_{t},\boldsymbol{\xi} \rangle ^{2}}{\boldsymbol{A}_{t}\boldsymbol{a}\cdot\boldsymbol{a}} $$

is concave in \(t\), which we can see by differentiating \(\phi(t)\) twice:

$$\phi''(t)=-\frac{2(\langle\boldsymbol{N}_{+},\boldsymbol{\xi} \rangle\boldsymbol{A}_{-}\boldsymbol{a}\cdot\boldsymbol{a}-\langle \boldsymbol{N}_{-},\boldsymbol{\xi} \rangle\boldsymbol {A}_{+}\boldsymbol{a}\cdot\boldsymbol{a})^{2}}{(\boldsymbol {A}_{t}\boldsymbol{a}\cdot\boldsymbol{a})^{3}}. $$

Thus, \(\phi(t)\) attains its minimum at \(t=0\) or \(t=1\). Setting \(t=0\) and 1 in (B.11) we obtain

$$ \overline{W}_{\boldsymbol{F}\boldsymbol{F}}(\boldsymbol{F}_{\pm })= \mathsf{L}_{\pm}-\frac{\boldsymbol{N}_{\pm}\otimes\boldsymbol {N}_{\pm}}{\boldsymbol{A}_{\pm}\boldsymbol{a}\cdot\boldsymbol{a}}. $$
(B.13)

Formula (4.4) now follows easily from (B.13).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Grabovsky, Y., Truskinovsky, L. Legendre-Hadamard Conditions for Two-Phase Configurations. J Elast 123, 225–243 (2016). https://doi.org/10.1007/s10659-015-9557-y

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10659-015-9557-y

Keywords

Mathematics Subject Classification

Navigation