Skip to main content
Log in

Exact Constructions in the (Non-linear) Planar Theory of Elasticity: From Elastic Crystals to Nematic Elastomers

  • Published:
Archive for Rational Mechanics and Analysis Aims and scope Submit manuscript

Abstract

In this article we deduce necessary and sufficient conditions for the presence of “Conti-type”, highly symmetric, exactly stress-free constructions in the geometrically non-linear, planar n-well problem, generalising results of Conti et al. (Proc R Soc A 473(2203):20170235, 2017). Passing to the limit \(n\rightarrow \infty \), this allows us to treat solid crystals and nematic elastomer differential inclusions simultaneously. In particular, we recover and generalise (non-linear) planar tripole star type deformations which were experimentally observed in Kitano and Kifune (Ultramicroscopy 39(1–4):279–286, 1991), Manolikas and Amelinckx (Physica Status Solidi (A) 60(2):607–617, 1980; Physica Status Solidi (A) 61(1):179–188, 1980). Furthermore, we discuss the corresponding geometrically linearised problem.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17

Similar content being viewed by others

References

  1. Agostiniani, V., Dal Maso, G., DeSimone, A.: Attainment results for nematic elastomers. Proc. R. Soc. Edinb. Sect. A Math. 145(4), 669–701, 2015

    Article  MathSciNet  Google Scholar 

  2. Ball, J.M.: Mathematical models of martensitic microstructure. Mater. Sci. Eng. A378(1–2), 61–69, 2004

    Article  Google Scholar 

  3. Bhattacharya, K.: Microstructure of Martensite: Why It Forms and How It Gives Rise to the Shape-Memory Effect. Oxford Series on Materials Modeling. Oxford University Press, Oxford 2003

    MATH  Google Scholar 

  4. Ball, J.M., James, R.D.: Fine phase mixtures as minimizers of energy. Arch. Ration. Mech. Anal. 100(1), 13–52, 1987

    Article  MathSciNet  Google Scholar 

  5. Bladon, P., Warner, M., Terentjev, E.M.: Orientational order in strained nematic networks. Macromolecules27, 7067–7075, 1994

    Article  ADS  Google Scholar 

  6. Cesana, P., DeSimone, A.: Quasiconvex envelopes of energies for nematic elastomers in the small strain regime and applications. J. Mech. Phys. Solids59(4), 787–803, 2011

    Article  ADS  MathSciNet  Google Scholar 

  7. Conti, S., Dolzmann, G., Kirchheim, B.: Existence of lipschitz minimizers for the three-well problem in solid–solid phase transitions. Annales de l’Institut Henri Poincaré (C) Non Linear Analysis24(6), 953–962, 2007

    Article  ADS  MathSciNet  Google Scholar 

  8. Cesana, P.: Relaxation of multiwell energies in linearized elasticity and applications to nematic elastomers. Arch. Ration. Mech. Anal. 197(3), 903–923, 2010

    Article  MathSciNet  Google Scholar 

  9. Curnoe, S.H., Jacobs, A.E.: Time evolution of tetragonal–orthorhombic ferroelastics. Phys. Rev. B64(6), 064101, 2001

    Article  ADS  Google Scholar 

  10. Cui, Y.-W., Koyama, T., Ohnuma, I., Oikawa, K., Kainuma, R., Ishida, K.: Simulation of hexagonal–orthorhombic phase transformation in polycrystals. Acta Mater. 55(1), 233–241, 2007

    Article  Google Scholar 

  11. Conti, S., Klar, M., Zwicknagl, B.: Piecewise affine stress-free martensitic inclusions in planar nonlinear elasticity. Proc. R. Soc. A473(2203), 20170235, 2017

    Article  ADS  MathSciNet  Google Scholar 

  12. Conti, S.: Quasiconvex functions incorporating volumetric constraints are rank-one convex. Journal de mathématiques pures et appliquées90(1), 15–30, 2008

    Article  MathSciNet  Google Scholar 

  13. Cesana, P., Plucinsky, P., Bhattacharya, K.: Effective behavior of nematic elastomer membranes. Arch. Ration. Mech. Anal. 218(2), 863–905, 2015

    Article  MathSciNet  Google Scholar 

  14. Cesana, P., Porta, M., Lookman, T.: Asymptotic analysis of hierarchical martensitic microstructure. J. Mech. Phys. Solids72, 174–192, 2014

    Article  ADS  MathSciNet  Google Scholar 

  15. Conti, S., Theil, F.: Single-slip elastoplastic microstructures. Arch. Ration. Mech. Anal. 178(1), 125–148, 2005

    Article  MathSciNet  Google Scholar 

  16. Dacorogna, B.: Direct Methods in the Calculus of Variations, vol. 78. Springer, Berlin 2007

    MATH  Google Scholar 

  17. DeSimone, A., Dolzmann, G.: Macroscopic response of nematic elastomers via relaxation of a class of SO(3)-invariant energies. Arch. Ration. Mech. Anal. 161(3), 181–204, 2002

    Article  MathSciNet  Google Scholar 

  18. Dacorogna, B., Marcellini, P.: Implicit Partial Differential Equations, vol. 37. Springer, Berlin 2012

    MATH  Google Scholar 

  19. Girault, V., Raviart, P.-A.: Finite Element Methods for Navier–Stokes Equations. Theory and Algorithms. Springer Series in Computational Mathematics, vol. 5. Springer, Berlin 1986

    Book  Google Scholar 

  20. Jacobs, A.E., Curnoe, S.H., Desai, R.C.: Landau theory of domain patterns in ferroelastics. Mater. Trans. 45(4), 1054–1059, 2004

    Article  Google Scholar 

  21. Kundler, I., Finkelmann, H.: Strain-induced director reorientation in nematic liquid single crystal elastomers. Macromol. Rapid Commun. 16(9), 679–686, 1995

    Article  Google Scholar 

  22. Kirchheim, B.: Rigidity and Geometry of Microstructures. MPI-MIS Lecture Notes, 2003

  23. Kitano, Y., Kifune, K.: HREM study of disclinations in MgCd ordered alloy. Ultramicroscopy39(1–4), 279–286, 1991

    Article  Google Scholar 

  24. Kitano, Y., Kifune, K., Komura, Y.: Star-lisclination in a ferro-elastic material B19 MgCd alloy. Le Journal de Physique Colloques49(C5), C5–201, 1988

    Google Scholar 

  25. Kirchheim, B., Müller, S., Šverák, V.: Studying nonlinear PDE by geometry in matrix space. Geometric Analysis and Nonlinear Partial Differential Equations. Springer, 347–395, 2003

  26. Manolikas, C., Amelinckx, S.: Phase transitions in ferroelastic lead orthovanadate as observed by means of electron microscopy and electron diffraction. I. Static observations. Physica Status Solidi (A)60(2), 607–617, 1980

    Article  ADS  Google Scholar 

  27. Manolikas, C., Amelinckx, S.: Phase transitions in ferroelastic lead orthovanadate as observed by means of electron microscopy and electron diffraction. II. Dynamic Observations. Physica Status Solidi (A)61(1), 179–188, 1980

    Article  ADS  Google Scholar 

  28. Müller, S., Šverák, V.: Unexpected solutions of first and second order partial differential equations. Proceedings of the International Congress of Mathematicians, volume 2 of Documents. Mathematica, Berlin, 691–702, 1998

  29. Müller, S., Šverák, V.: Convex integration with constraints and applications to phase transitions and partial differential equations. J. Eur. Math. Soc. 1, 393–422, 1999. https://doi.org/10.1007/s100970050012

    Article  MathSciNet  MATH  Google Scholar 

  30. Patching, S.: Microstructures in the hexagonal-to-rhombic phase transformation. OxPDE summer research project, 2014

  31. Plucinsky, P., Bhattacharya, K.: Interplay of microstructure and wrinkling in nematic elastomer membranes. XXIV ICTAM, 2016

  32. Porta, M., Lookman, T.: Heterogeneity and phase transformation in materials: energy minimization, iterative methods and geometric nonlinearity. Acta Mater. 61(14), 5311–5340, 2013

    Article  Google Scholar 

  33. Pompe, W.: Explicit construction of piecewise affine mappings with constraints. Bull. Pol. Acad. Sci. Math. 58(3), 209–220, 2010

    Article  MathSciNet  Google Scholar 

  34. Rüland, A., Taylor, J.M., Zillinger, C.: Convex integration arising in the modelling of shape-memory alloys: some remarks on rigidity, flexibility and some numerical implementations. J. Nonlinear Sci. 29, 2137–2184, 2018

    Article  ADS  MathSciNet  Google Scholar 

  35. Rüland, A.: The cubic-to-orthorhombic phase transition: rigidity and non-rigidity properties in the linear theory of elasticity. Arch. Ration. Mech. Anal. 221(1), 23–106, 2016

    Article  MathSciNet  Google Scholar 

  36. Rüland, A., Zillinger, C., Zwicknagl, B.: Higher Sobolev regularity of convex integration solutions in elasticity: the Dirichlet problem with affine data in int\((\text{ K }^{lc})\). SIAM J. Math. Anal. 50(4), 3791–3841, 2018

    Article  MathSciNet  Google Scholar 

  37. Rüland, A., Zillinger, C., Zwicknagl, B.: Higher Sobolev regularity of convex integration solutions in elasticity: the planar geometrically linearized hexagonal-to-rhombic phase transformation. J. Elast. 136, 1–76, 2019

    Article  MathSciNet  Google Scholar 

  38. Vicens, J., Delavignette, P.: A particular domain configuration observed in a new phase of the Ta–N system. Physica Status Solidi (A)33(2), 497–509, 1976

    Article  ADS  Google Scholar 

  39. Warner, M., Terentjev, E.M.: Liquid Crystal Elastomers. International Series of Monographs on Physics. Oxford University Press, Oxford 2003

    Google Scholar 

  40. Wen, Y.H., Wang, Y., Chen, L.-Q.: Effect of elastic interaction on the formation of a complex multi-domain microstructural pattern during a coherent hexagonal to orthorhombic transformation. Acta Mater. 47(17), 4375–4386, 1999

    Article  Google Scholar 

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

    Article  Google Scholar 

Download references

Acknowledgements

P.C. is supported by JSPS Grant-in-Aid for Young Scientists (B) 16K21213 and partially by JSPS Innovative Area Grant 19H05131. P.C. holds an honorary appointment at La Trobe University and is a member of GNAMPA. C.Z. acknowledges a travel grant from the Simon’s foundation. B.Z. would like to thank Sergio Conti for helpful discussions, and acknowledges support by the Berliner Chancengleichheitsprogramm and by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) through SFB 1060 “The Mathematics of Emergent Effects” (Project 211504053).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Pierluigi Cesana.

Additional information

Communicated by K. Bhattacharya.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Necessary Relation Between the Radius of the Outer Polygon and the Radius of the Inner Polygon: The Solutions to (31)

In this first part of the appendix, we provide the remainder of the argument from Proposition 1.

To this end, we solve

$$\begin{aligned}&\Big (1+x^2-2x\cos \Big ( \frac{2\pi }{n}\alpha \Big )\Big ) \Big (1+x^2-2x\cos \Big (\frac{2\pi }{n}(1-\alpha ) \Big )\Big ) \cos \Big (\frac{(n-2)\pi }{2n }\Bigr )\nonumber \\&\quad = \Big (1+ x^2\cos \Big (\frac{2\pi }{n}\Big ) -2x\cos \Big (\frac{\pi }{n}\Big )\cos \Big (\frac{\pi }{n}(1-2\alpha )\Big )\Big )^2, \end{aligned}$$
(99)

which is (31) squared. We get the following four solutions of the equation (31) for x:

  • \(x=\frac{1}{\cos \frac{\pi }{n}}\left( \cos \left( \frac{\rho _n}{2}\right) - \sqrt{\sin (\frac{2\pi }{n}\alpha )\sin \left( \frac{2\pi }{n}(1-\alpha )\right) } \right) \),

  • \(x=\frac{1}{\cos \frac{\pi }{n}}\left( \cos \left( \frac{\rho _n}{2}\right) + \sqrt{\sin (\frac{2\pi }{n}\alpha )\sin \left( \frac{2\pi }{n}(1-\alpha )\right) } \right) \),

  • \(x=\frac{1}{\cos \bigl (\frac{3\pi }{n}\bigr )} \left( \cos \left( \frac{2\pi }{n}\right) \cos \left( \frac{\rho _n}{2} \right) -\sqrt{\cos ^2\left( \frac{2\pi }{n}\right) \cos ^2 \left( \frac{\rho _n}{2}\right) -\cos \left( \frac{3\pi }{n}\right) \cos \left( \frac{\pi }{n}\right) } \right) \),

  • \(x=\frac{1}{\cos \bigl (\frac{3\pi }{n}\bigr )}\left( \cos \left( \frac{2\pi }{n}\right) \cos \left( \frac{\rho _n}{2}\right) +\sqrt{\cos ^2\left( \frac{2\pi }{n}\right) \cos ^2\left( \frac{\rho _n}{2}\right) -\cos \left( \frac{3\pi }{n}\right) \cos \left( \frac{\pi }{n}\right) } \right) \),

where as in (iv), \({\rho _n}:=\frac{2\pi }{n}(1-2\alpha )\). We now claim that just the first solution is admissible for us. Here and below we define a solution x of (99) admissible if \(x\in (0,1)\) and it satisfies (31). In order to prove our claim, we can assume without loss of generality that

$$\begin{aligned} \sqrt{4\cos ^2\Bigl (\frac{2\pi }{n}\Bigr )\cos ^2\Bigl (\frac{\rho _n}{2}\Bigr ) -4\cos \Bigl (\frac{3\pi }{n}\Bigr )\cos \Bigl (\frac{\pi }{n}\Bigr )} \end{aligned}$$

is real, otherwise the third and fourth solutions are not admissible. The proof of the claim is as follows:

  • Second solution: We estimate

    $$\begin{aligned} x \geqq \frac{\cos \left( \frac{{\rho _n}}{2} \right) }{\cos \left( \frac{\pi }{n} \right) }. \end{aligned}$$

    Since \(\alpha \in \left( 0,1\right) \) it is clear that the second solution is such that \(x\geqq 1\) for any \(\alpha \in (0,1)\), any \(n\geqq 3\).

  • Third solution: \(x\geqq 1\) if \(n=3,4\) and \(\alpha \in [0,1].\) We can hence restrict to the case \(n>4\). We now claim that

    $$\begin{aligned} 1+ x^2\cos \left( \frac{2\pi }{n}\right) -2x\cos \left( \frac{\pi }{n}\right) \cos \left( \frac{\rho _n}{2}\right) <0 \end{aligned}$$
    (100)

    for any \(\alpha \in (0,1)\), and any \(n\geqq 4\). Since the left-hand side of (31) is always non-negative, the claim would imply that the third solution of (99) does not satisfy (31), and is hence not admissible. We plot \(1+ x^2\cos \bigl (\frac{2\pi }{n}\bigr )-2x\cos \bigl (\frac{\pi }{n}\bigr )\cos \bigl (\frac{{\rho _n}}{2}\bigr )\) for \(n\in \{5,\dots ,50\}\) in Figure 18. For large n, we have that

    $$\begin{aligned} x = 1-\frac{2\pi }{n}\sqrt{(\alpha -\alpha ^2)} +\frac{2\pi ^2}{n^2}(-\alpha ^2+\alpha +1) + O(n^{-3}), \end{aligned}$$

    and, therefore,

    $$\begin{aligned}&1+ x^2\cos \left( \frac{2\pi }{n}\right) -2x\cos \left( \frac{\pi }{n}\right) \\&\quad \cos \left( \frac{\rho _n}{2}\right) = -\frac{4\pi ^3}{n^3}\sqrt{\alpha (1-\alpha )} +O(n^{-4}) < 0 \end{aligned}$$

    for any \(\alpha \in (0,1),\) and for any n large enough.

  • Fourth solution: It is easy to see that it is negative for any \(\alpha \in [0,1]\) when \(n=3,4,5\). Indeed, \(\cos \frac{3\pi }{n}<0\). If \(n=6\) we get \(x=\infty ,\) while for \(n>6\) we have \(x>1\). Indeed, in this case,

    $$\begin{aligned} x\geqq \frac{2\cos \left( \frac{2\pi }{n}\right) \cos \left( \frac{\rho _n}{2}\right) }{2\cos \left( \frac{3\pi }{n}\right) } \geqq \frac{\cos \left( \frac{2\pi }{n}\right) \cos \left( \frac{\pi }{n}\right) }{\cos \left( \frac{3\pi }{n}\right) }= \frac{1}{2}\left( 1+\frac{1}{2\cos \left( \frac{2\pi }{n}\right) -1} \right) >1. \end{aligned}$$

    Therefore, for any \(\alpha \in [0,1]\) and any \(n\geqq 3\) the fourth solution is not admissible.

Fig. 18
figure 18

Numerical verification of the fact that, for \(n\in \{1,\dots ,50\}\) and \(\alpha \in (0,1)\) we have (100). The bigger n is, the closer to zero the convex curves in the pictures are (colour figure online)

Appendix B: Proof of Corollary 2.4

In this part of the appendix we show that equation (43)

$$\begin{aligned} P_0 {H} P_0 = Q_{\alpha } {H} Q_{1-\alpha }^T \end{aligned}$$

is satisfied. In order to simplify calculations, we express all matrices with respect to the basis \((e_{11}, e_{11}^\perp )\) and thus have to show that

$$\begin{aligned} \begin{pmatrix} a &{} - \frac{a^{-1}-a}{\tan (\phi )} \\ 0 &{} a^{-1} \end{pmatrix} Q_{1-\alpha } = Q_{\alpha } \begin{pmatrix} a &{} \frac{a^{-1}-a}{\tan (\phi )} \\ 0 &{} a^{-1} \end{pmatrix}. \end{aligned}$$
(101)

We further recall that

$$\begin{aligned} a^2&= \frac{\sin \left( \frac{2\pi }{n}(1-\alpha )\right) }{\sin \left( \frac{2\pi }{n}\alpha \right) }, \ \frac{1}{\tan (\phi )}= \tan \left( \frac{\pi }{n}\right) . \end{aligned}$$

In particular, since \(\alpha \in (0,1)\), we may multiply the claimed equation with \(a \sin (\frac{2\pi }{n}\alpha ) \ne 0 \) and for simplicity of notation introduce \(t:=\frac{2\pi }{n}\alpha \) and \(s=\frac{2\pi }{n}(1-\alpha ) = \frac{2\pi }{n}-t\). With this notation, we have to show that

$$\begin{aligned}&\begin{pmatrix} \sin (s) &{}\quad -(\sin (t)-\sin (s))\tan \left( \frac{\pi }{n}\right) \\ 0 &{}\quad \sin (t) \end{pmatrix} \begin{pmatrix} \cos (s) &{}\quad -\sin (s) \\ \sin (s) &{}\quad \cos (s) \end{pmatrix} \\&\quad = \begin{pmatrix} \cos (t) &{}\quad -\sin (t) \\ \sin (t) &{}\quad \cos (t) \end{pmatrix} \begin{pmatrix} \sin (s) &{}\quad (\sin (t)-\sin (s))\tan \left( \frac{\pi }{n}\right) \\ 0 &{} \quad \sin (t) \end{pmatrix}. \end{aligned}$$

We consider each matrix entry separately. The claimed equality for the upper left entry is given by

$$\begin{aligned} \sin (s)\cos (s)- \sin (s)\tan \left( \frac{\pi }{n}\right) (\sin (t) -\sin (s))=\cos (t)\sin (s). \end{aligned}$$
(102)

In order to show this, we may factor out \(\sin (s)\) and use the angle addition formulas:

$$\begin{aligned} \cos (s)&=\cos (t)\cos \left( \frac{2\pi }{n}\right) + \sin (t) \sin \left( \frac{2\pi }{n}\right) , \\ \sin (s)&=\cos (t)\sin \left( \frac{2\pi }{n}\right) - \sin (t) \cos \left( \frac{2\pi }{n}\right) . \end{aligned}$$

We then collect terms involving \(\cos (t)\) and \(\sin (t)\) as

$$\begin{aligned}&\cos (t)\cos \left( \frac{2\pi }{n}\right) +\sin (t)\sin \left( \frac{2\pi }{n}\right) \\&\qquad -\tan \left( \frac{\pi }{n}\right) \left( \sin (t)-\cos (t)\sin \left( \frac{2\pi }{n}\right) +\sin (t)\cos \left( \frac{2\pi }{n}\right) \right) \\&\quad = \cos (t)\left( \cos \left( \frac{2\pi }{n}\right) +\tan \left( \frac{\pi }{n}\right) \sin \left( \frac{2\pi }{n}\right) \right) \\&\qquad + \sin (t)\left( \sin \left( \frac{2\pi }{n}\right) -\tan \left( \frac{\pi }{n}\right) (1+\cos \left( \frac{2\pi }{n}\right) \right) \\&\quad = \cos (t) 1+ \sin (t) 0 = \cos (t), \end{aligned}$$

where we used the half angle identities for \(\cos (2x)\) and \(\sin (2x)\) in the last equality.

The calculation for the bottom right-entry is analogous with the role of s and t and the sign of \((\sin (t)-\sin (s))\tan (\frac{\pi }{n})\)) interchanged. The bottom left equality \(\sin (t)\sin (s)=\sin (t)\sin (s)\) is always satisfied. It thus only remains to verify equality of the upper right entry, which can be simplified to read

$$\begin{aligned}&-\sin ^2(s)-\cos (s)(\sin (t)-\sin (s))\tan \left( \frac{\pi }{n}\right) \\&\quad = -\sin ^2(t)+ \cos (t)(\sin (t)-\sin (s))\tan \left( \frac{\pi }{n}\right) \\&\quad \Leftrightarrow \sin ^2(t)-\sin ^2(s) - (\cos (t)+\cos (s))(\sin (t) -\sin (s))\tan \left( \frac{\pi }{n}\right) =0. \end{aligned}$$

Factoring out the factor \((\sin (t)-\sin (s))\), it suffices to prove

$$\begin{aligned} \sin (t)+\sin (s) - (\cos (t)+\cos (s))\tan \left( \frac{\pi }{n}\right) =0. \end{aligned}$$

As above, the claimed equality then again follows by using angle addition formulas.

Appendix C: Reduction to Cauchy–Green Tensors Used in the Proof of Proposition 2.6

Last but not least, we provide the argument (used in the proof of Proposition 2.6) that it is possible to reduce the differential inclusion (23) to an inclusion for the associated Cauchy–Green tensors.

Lemma C.1

Suppose that \(\det (M)=\det ({H})>0\), then the inclusion

$$\begin{aligned} M \in \bigcup _{P \in {\mathcal {P}}_n}SO(2) P^T {H} P \end{aligned}$$
(103)

is satisfied, if and only if

$$\begin{aligned} M^TM \in \bigcup _{P \in {\mathcal {P}}_n} P^T {H}^T {H} P. \end{aligned}$$
(104)

This characterisation follows from basic properties of the singular value decomposition.

Proof

We observe that (103) implies (104). Thus, we only consider the converse and assume that

$$\begin{aligned} M^TM = (P^T{H}^TP) (P^T {H} P)=: M_1^T M_1. \end{aligned}$$

for some \(P \in {\mathcal {P}}_n\). Since \(M^TM\) is symmetric, there exists \(Q \in SO(2)\) and a diagonal matrix \({{\,\mathrm{diag}\,}}(\lambda _1,\lambda _2)\), with \(\lambda _1 \lambda _2=\det (M)^2\ne 0\), \(\lambda _1,\lambda _2>0\), such that

$$\begin{aligned} M^T M= Q^T {{\,\mathrm{diag}\,}}(\lambda _1,\lambda _2)Q. \end{aligned}$$

It follows that

$$\begin{aligned}&\tilde{M}:= M Q^T {{\,\mathrm{diag}\,}}\left( \frac{1}{\sqrt{\lambda _1}}, \frac{1}{\sqrt{\lambda _2}}\right) , \\&\tilde{M_1}:= M_1 Q^T {{\,\mathrm{diag}\,}}\left( \frac{1}{\sqrt{\lambda _1}}, \frac{1}{\sqrt{\lambda _2}}\right) , \end{aligned}$$

satisfy

$$\begin{aligned} \tilde{M}^T\tilde{M}= I= \tilde{M_1}^T\tilde{M_1} \end{aligned}$$

and thus \(\tilde{M}, \tilde{M_1} \in SO(2)\). Here we used that \(\det (M)=\det ({H})>0\). In particular,

$$\begin{aligned} M_1&= \tilde{M_1} Q^T{{\,\mathrm{diag}\,}}(\sqrt{\lambda _1}, \sqrt{\lambda _2}), \\ M&= \tilde{M} Q^T{{\,\mathrm{diag}\,}}(\sqrt{\lambda _1}, \sqrt{\lambda _2}) \\&= \tilde{M} \tilde{M}^T_1 M_1, \end{aligned}$$

where \(\tilde{M}\tilde{M}^T_1 \in SO(2)\), which implies the result. \(\quad \square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cesana, P., Della Porta, F., Rüland, A. et al. Exact Constructions in the (Non-linear) Planar Theory of Elasticity: From Elastic Crystals to Nematic Elastomers. Arch Rational Mech Anal 237, 383–445 (2020). https://doi.org/10.1007/s00205-020-01511-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00205-020-01511-9

Navigation