Skip to main content
Log in

Stable Spatially Localized Configurations in a Simple Structure—A Global Symmetry-Breaking Approach

  • Published:
Journal of Elasticity Aims and scope Submit manuscript

Abstract

We revisit the classic stability problem of the buckling of an inextensible, axially compressed beam on a nonlinear elastic foundation with a semi-analytical approach to understand how spatially localized deformation solutions emerge in many applications in mechanics. Instead of a numerical search for such solutions using arbitrary imperfections, we propose a systematic search using branch-following and bifurcation techniques along with group-theoretic methods to find all the bifurcated solution orbits (primary, secondary, etc.) of the system and to examine their stability and hence their observability. Unlike previously proposed methods that use multi-scale perturbation techniques near the critical load, we show that to obtain a spatially localized deformation equilibrium path for the perfect structure, one has to consider the secondary bifurcating path with the longest wavelength and follow it far away from the critical load. The novel use of group-theoretic methods here illustrates a general methodology for the systematic analysis of structures with a high degree of symmetry.

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
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

Notes

  1. The reader should note that we employ a linear beam model for its simplicity, both analytically and numerically, in this work. Thus, undue weight should not be given to the physical interpretation of fine-scale details associated with the highly-deformed solutions obtained here. Instead, the primary focus here is an analysis of the simplest model leading to the existence of such spatially localized solutions. As we show, all of the important phenomena (primary bifurcation, cascades of secondary bifurcations, etc.) occur well within the small-displacement regime. Thus, the linear beam model poses no limitation on our ability to determine the correct mechanics of the nonlinear beam–foundation system of interest.

  2. Such as those illustrated in Fig. 8(a).

  3. Assuming adequate continuity, the fourth-order Euler–Lagrange equation in \([-L/2, L/2]\) requires four boundary conditions to be satisfied.

  4. In fact, for systems with Lie symmetry groups (continuous infinite groups), such as the one considered here, the continuous orbits of solutions are known as relative equilibria. In general, relative equilibria of the system’s equations of motion can correspond to an orbit of equilibria, traveling or rotating waves, dynamic trajectories that appear time-periodic in a suitable moving frame, or other more complicated motions. For a discussion of the theory of relative equilibria and their stability see [13]. In this work, we explore only orbits of equilibria, and so the distinction between a relative equilibrium of a system with continuous symmetry and an orbit of equilibria of a system with discrete symmetry is immaterial.

  5. In Sect. 4.1.1 it is shown that \(L_{c} = 2\pi \) for the perfect beam–foundation system.

  6. Indeed, with (as below) \(q=20\), the primary bifurcation branch will have period \(L_{c}\). Then each secondary bifurcating branch may be associated with one of the periods \(L_{m} = m L_{c}\) with \(m \in \{1,2,\dots ,q\}\).

  7. Actually, its symmetry is a subgroup of \(D_{\infty h}\) that is conjugate to \(D_{4d}\). Here, we will not be concerned with this distinction.

  8. As mentioned above, in the problem at hand, we find continuous orbits of equilibria, all having the same energy. This implies the existence of a zero eigenvalue of the stability operator \(\mathcal{E}_{,ww}\) in Eq. (2.9). Thus, all equilibria are, at best, neutrally stable. Accordingly, in this work we ignore the zero eigenvalue associated with the solution orbit and require that all other eigenvalues be positive for stability.

  9. This is adequate for correct integration of the higher-order gradient term in the stiffness matrix.

  10. As described in Sect. 2.3, the primary bifurcation orbit consists of an infinite set of configurations generated by the symmetries of \(D_{\infty h}\); here we select one specific representative of the orbit.

  11. The higher-order derivatives that must vanish by symmetry are thus only approximately satisfied.

  12. Note, as written it is necessary to take \(\delta w_{s0} := 0\).

  13. In fact, even the employed parameterization is problematic, since it is restricted to \(\xi \geq 0\), and is therefore unable to distinguish between the two “halves” of the bifurcated path.

  14. There are also simple bifurcations of symmetry \(D_{1}\) which occur at loads \(\lambda _{cia}\), but \(\lambda _{cia} > \lambda _{cis}\) (see Appendix A.4). Here we are interested in the first (i.e. lowest load) bifurcation point, so we do not present the imperfect bifurcations with \(D_{1}\) symmetry.

  15. From Appendix A.4, it is clear that \(2(1+\zeta )^{1/2} < \lambda _{cis} < 2\), for \(\zeta < 0\).

  16. See Footnote 8.

  17. Compare Eq. (A.6) to Eq. (4.6). In Eq. (4.6), no assumptions about symmetry are made, and one must ensure that the entire operator is non-singular. In Eq. (A.6), equivariance is assumed and this suffices to ensure that \(\mathcal{E}^{b}_{,w\lambda } {\stackrel{i}{w}} = 0\;,\ i=1,2,\dots ,n_{ \mu }\) and that the operator \(\mathcal{E}^{0}_{,ww}\) is a scalar multiple of the identity. Thus, the two criteria are equivalent when the assumptions of equivariant bifurcation theory are satisfied.

References

  1. Allgower, E.L., Georg, K.: Introduction to Numerical Continuation Methods. Classics in Applied Mathematics, vol. 45. SIAM, Philadelphia (2003)

    MATH  Google Scholar 

  2. Amazigo, J.C., Budiansky, B., Carrier, G.F.: Asymptotic analyses of the buckling of imperfect columns on nonlinear elastic foundations. Int. J. Solids Struct. 6(10), 1341–1356 (1970)

    MATH  Google Scholar 

  3. Audoly, B.: Localized buckling of a floating elastica. Phys. Rev. E 84, 011605 (2011)

    ADS  Google Scholar 

  4. Auguste, A., Jin, L., Hayward, R.C., Suo, Z.: Post-wrinkle bifurcations in elastic bilayers with modest contrast in modulus. Extreme Mech. Lett. 11, 30–36 (2017)

    Google Scholar 

  5. Beardmore, R., Peletier, M., Budd, C., Wadee, M.: Bifurcations of periodic solutions satisfying the zero-Hamiltonian constraint in reversible differential equations. SIAM J. Math. Anal. 36(5), 1461–1488 (2005)

    MathSciNet  MATH  Google Scholar 

  6. Biot, M.A.: Surface instability of rubber in compression. Appl. Sci. Res., Sect. A 12, 168–182 (1963)

    MATH  Google Scholar 

  7. Budd, C., Hunt, G., Kuske, R.: Asymptotics of cellular buckling close to Maxwell load. Proc. R. Soc. Lond., Ser. A, Math. Phys. Eng. Sci. 457(2016), 2935–2964 (2001)

    ADS  MathSciNet  MATH  Google Scholar 

  8. Cao, Y., Hutchinson, J.W.: From wrinkles to creases in elastomers: the instability and imperfection-sensitivity of wrinkling. Proc. R. Soc. Lond., Ser. A, Math. Phys. Eng. Sci. 468(2137), 94–115 (2012)

    ADS  MathSciNet  MATH  Google Scholar 

  9. Champneys, A.R., Toland, J.: Bifurcation of a plethora of multi-modal homoclinic orbits for autonomous Hamiltonian systems. Nonlinearity 6(5), 665–721 (1993)

    ADS  MathSciNet  MATH  Google Scholar 

  10. Chen, D., Cai, S., Suo, Z., Hayward, R.C.: Surface energy as a barrier to creasing of elastomer films: an elastic analogy to classical nucleation. Phys. Rev. Lett. 109, 038001 (2012)

    ADS  Google Scholar 

  11. Chen, D., Jin, L., Suo, Z., Hayward, R.C.: Controlled formation and disappearance of creases. Mater. Horiz. 1(2), 207–213 (2014)

    Google Scholar 

  12. Chen, Y.C., Yang, S., Wheeler, L.: Surface instability of elastic half-spaces by using the energy method. Proc. R. Soc. Lond., Ser. A, Math. Phys. Eng. Sci. 474(2213), 20170854 (2018).

    ADS  MathSciNet  MATH  Google Scholar 

  13. Chossat, P., Lauterbach, R.: Methods in Equivariant Bifurcations and Dynamical Systems. Advanced Series in Nonlinear Dynamics, vol. 15. World Scientific, Singapore (2000)

    MATH  Google Scholar 

  14. Ciarletta, P.: Matched asymptotic solution for crease nucleation in soft solids. Nat. Commun. 9(496), 1–7 (2018)

    Google Scholar 

  15. Coman, C.D.: Inhomogeneities and localised buckling patterns. IMA J. Appl. Math. 71(1), 133–152 (2006)

    MathSciNet  MATH  Google Scholar 

  16. Coman, C.D.: Localized elastic buckling: non-linearities versus inhomogeneities. IMA J. Appl. Math. 75(3), 461–474 (2010)

    MathSciNet  MATH  Google Scholar 

  17. Diab, M., Kim, K.S.: Ruga-formation instabilities of a graded stiffness boundary layer in a neo-Hookean solid. Proc. R. Soc. Lond., Ser. A, Math. Phys. Eng. Sci. 470(2168), 20140218 (2014)

    ADS  Google Scholar 

  18. Diamant, H., Witten, T.A.: Compression induced folding of a sheet: an integrable system. Phys. Rev. Lett. 107, 164302 (2011)

    ADS  Google Scholar 

  19. Everall, P.R., Hunt, G.: Mode jumping in the buckling of struts and plates: a comparative study. Int. J. Non-Linear Mech. 35(6), 1067–1079 (2000)

    MATH  Google Scholar 

  20. Gatermann, K., Hohmann, A.: Symbolic exploitation of symmetry in numerical pathfollowing. Impact Comput. Sci. Eng. 3(4), 330–365 (1991). https://doi.org/10.1016/0899-8248(91)90003-D

    Article  MathSciNet  MATH  Google Scholar 

  21. Gent, A.N., Cho, I.S.: Surface instabilities in compressed or bent rubber blocks. Rubber Chem. Technol. 72(2), 253–262 (1999)

    Google Scholar 

  22. Golublitsky, M., Stewart, I., Schaeffer, D.G.: Singularities and Groups in Bifurcation Theory. Applied Mathematical Sciences, vol. 2. Springer, Berlin (1988)

    Google Scholar 

  23. Healey, T.J.: A group-theoretic approach to computational bifurcation problems with symmetry. Comput. Methods Appl. Mech. Eng. 67(3), 257–295 (1988)

    ADS  MathSciNet  MATH  Google Scholar 

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

    ADS  Google Scholar 

  25. Hohlfeld, E., Mahadevan, L.: Scale and nature of sulcification patterns. Phys. Rev. Lett. 109, 025701 (2012)

    ADS  Google Scholar 

  26. Hong, W., Zhao, X., Suo, Z.: Formation of creases on the surfaces of elastomers and gels. Appl. Phys. Lett. 95, 111901 (2009)

    ADS  Google Scholar 

  27. Hunt, G., Everall, P.R.: Arnold tongues and mode-jumping in the supercritical post-buckling of an archetypal elastic structure. Proc. R. Soc. Lond., Ser. A, Math. Phys. Eng. Sci. 455(1981), 125–140 (1999)

    ADS  MathSciNet  MATH  Google Scholar 

  28. Hunt, G.W., Wadee, M.K.: Comparative Lagrangian formulations for localized buckling. Proc. R. Soc. Lond., Ser. A, Math. Phys. Eng. Sci. 434(1892), 485–502 (1991)

    ADS  MathSciNet  MATH  Google Scholar 

  29. Hunt, G.W., Bolt, H.M., Thompson, J.M.T.: Structural localization phenomena and dynamical phase-space analogy. Proc. R. Soc. Lond., Ser. A, Math. Phys. Eng. Sci. 425(1869), 245–267 (1989)

    ADS  MathSciNet  MATH  Google Scholar 

  30. Hunt, G.W., Wadee, M.K., Shiacolas, N.: Localized elasticae for the strut on the linear foundation. J. Appl. Mech. 60(4), 1033–1038 (1993)

    ADS  MATH  Google Scholar 

  31. Hunt, G., Peletier, M.A., Champneys, A.R., Woods, P.D., Wadee, M.A., Budd, C.J., Lord, G.J.: Cellular buckling in long structures. Nonlinear Dyn. 21(1), 3–29 (2000)

    MathSciNet  MATH  Google Scholar 

  32. Ikeda, K., Murota, K.: Imperfect Bifurcation in Structures and Materials: Engineering Use of Group-Theoretic Bifurcation Theory, 2nd edn. Applied Mathematical Sciences, vol. 149. Springer, Berlin (2010)

    MATH  Google Scholar 

  33. Iooss, G., Joseph, D.D.: Elementary Stability and Bifurcation Theory, 2nd edn. Springer, Berlin (1990)

    MATH  Google Scholar 

  34. Jin, L., Chen, D., Hayward, R.C., Suo, Z.: Creases on the interface between two soft materials. Soft Matter 10(2), 303–311 (2014)

    ADS  Google Scholar 

  35. Jin, L., Auguste, A., Hayward, R.C., Suo, Z.: Bifurcation diagrams for the formation of wrinkles or creases in soft bilayers. J. Appl. Mech. 82(6), 061008 (2015)

    ADS  Google Scholar 

  36. Keller, H.B.: Numerical Methods in Bifurcation Problems. Tata Institute of Fundamental Research/Springer, Bombay/New York (1987)

    Google Scholar 

  37. Li, B., Cao, Y.P., Feng, X.Q., Gao, H.: Mechanics of morphological instabilities and surface wrinkling in soft materials: a review. Soft Matter 8(21), 5728–5745 (2012)

    ADS  Google Scholar 

  38. Luongo, A.: On the amplitude modulation and localization phenomena in interactive buckling problems. Int. J. Solids Struct. 27(15), 1943–1954 (1991)

    MATH  Google Scholar 

  39. Makrides, E., Sandstede, B.: Existence and stability of spatially localized patterns. J. Differ. Equ. 266(2), 1073–1120 (2019). https://doi.org/10.1016/j.jde.2018.07.064

    Article  ADS  MathSciNet  MATH  Google Scholar 

  40. McWeeny, R.: Symmetry: An Introduction to Group Theory and Its Applications. Dover, New York (2002)

    MATH  Google Scholar 

  41. Peletier, M.: Sequential buckling: a variational analysis. SIAM J. Math. Anal. 32(5), 1142–1168 (2001)

    MathSciNet  MATH  Google Scholar 

  42. Pocivavsek, L., Dellsy, R., Kern, A., Johnson, S., Lin, B., Lee, K.Y.C., Cerda, E.: Stress and fold localization in thin elastic membranes. Science 320(5878), 912–916 (2008)

    ADS  Google Scholar 

  43. Potier-Ferry, M.: Amplitude Modulation, Phase Modulation and Localization of Buckling Patterns. Cambridge University Press, Cambridge (1983)

    Google Scholar 

  44. Potier-Ferry, M.: Foundations of Elastic Postbuckling Theory, vol. 288. Springer, Berlin (1987)

    MATH  Google Scholar 

  45. Rivetti, M.: Non-symmetric localized fold of a floating sheet. C. R., Méc. 341(3), 333–338 (2013)

    ADS  Google Scholar 

  46. Silling, S.A.: Creasing singularities in compressible elastic materials. J. Appl. Mech. 58(1), 70–74 (1991)

    ADS  Google Scholar 

  47. Triantafyllidis, N., Peek, R.: On stability and the worst imperfection shape in solids with nearly simultaneous eigenmodes. Int. J. Solids Struct. 29(18), 2281–2299 (1992). https://doi.org/10.1016/0020-7683(92)90216-G

    Article  MathSciNet  Google Scholar 

  48. Wadee, M.A.: Effects of periodic and localized imperfections on struts on nonlinear foundations and compression sandwich panels. Int. J. Solids Struct. 37(8), 1191–1209 (2000)

    MATH  Google Scholar 

  49. Wadee, M.K., Bassom, A.P.: Effects of exponentially small terms in the perturbation approach to localized buckling. Proc. R. Soc. Lond., Ser. A, Math. Phys. Eng. Sci. 455(1986), 2351–2370 (1999)

    ADS  MathSciNet  MATH  Google Scholar 

  50. Wadee, M.K., Bassom, A.P.: Restabilization in structures susceptible to localized buckling: an approximate method for the extended post-buckling regime. J. Eng. Math. 38(1), 77–90 (2000)

    MATH  Google Scholar 

  51. Wadee, M.K., Hunt, G.W., Whiting, A.I.M.: Asymptotic and Rayleigh-Ritz routes to localized buckling solutions in an elastic instability problem. Proc. R. Soc. Lond., Ser. A, Math. Phys. Eng. Sci. 453(1965), 2085–2107 (1997)

    ADS  MathSciNet  MATH  Google Scholar 

  52. Wadee, M.K., Coman, C.D., Bassom, A.P.: Solitary wave interaction phenomena in a strut buckling model incorporating restabilisation. Physica D 163(1–2), 26–48 (2002)

    ADS  MathSciNet  MATH  Google Scholar 

  53. Wang, Q., Zhao, X.: A three-dimensional phase diagram of growth-induced surface instabilities. Sci. Rep. 5(8887), 1–10 (2015)

    Google Scholar 

  54. Woods, P., Champneys, A.: Heteroclinic tangles and homoclinic snaking in the unfolding of a degenerate reversible Hamiltonian–Hopf bifurcation. Phys. D: Nonlinear Phenom. 129(3), 147–170 (1999)

    ADS  MathSciNet  MATH  Google Scholar 

  55. Zhao, R., Zhang, T., Diab, M., Gao, H., Kim, K.-S.: The primary bilayer ruga-phase diagram I: localizations in ruga evolution. Extreme Mech. Lett. 4, 76–82 (2015)

    Google Scholar 

Download references

Acknowledgements

All authors would acknowledge support from the École Polytechnique and its Laboratore de Mécanique des Solides (LMS). The work was initiated in part by grants from École Polytechnique and the C.N.R.S. (Centre National de Recherche Scientifique) during the AY 2017-2018, while TJH was a Distinguished Visiting Professor and SSP a visiting doctoral student. RSE acknowledges several visits to LMS during this period supported by the LMS. The work of SSP and TJH was also supported in part by the National Science Foundation (NSF) through grant DMS-1613753. The work of RSE was also partially supported by the NSF grant CMMI-1462826.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ryan S. Elliott.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher’s Note

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

Appendix: Group-Theoretic Considerations

Appendix: Group-Theoretic Considerations

The fundamental concept used to study the bifurcated equilibrium paths and their stability in any conservative elastic system is the existence of a group \(G\) of transformations that leave its energy \(\mathcal{E}(w;\lambda )\)—defined in Eq. (2.2) for the problem at hand—unchanged, i.e., invariant under the action of all transformations \(g \in G\). More specifically, to each element \(g \in G\) we associate an orthogonal transformation \(T_{g}\) (termed “representation” of \(g\)) acting on \(w(x) \in H\) with image \(T_{g}[w] \in H\) that satisfies

$$ \mathcal{E}(T_{g}[w];\lambda ) = \mathcal{E}(w ;\lambda ) \, ;\quad \forall \lambda \ge 0 \, , \ \forall w \in H \, , \ \forall g \in G\; , $$
(A.1)

where \(\lambda \) is the scalar load parameter (assumed positive) and \(H\) the space of admissible displacements.

It follows from Eq. (A.1) that the variation of ℰ with respect to its argument \(w\) (first order functional derivative \(\mathcal{E}_{,w}\)) possess the property of “equivariance

$$ T_{g}[\mathcal{E}_{,w}(w;\lambda )] \delta w = \mathcal{E}_{,w}(T_{g} [w] ;\lambda ) \delta w \, ; \quad \forall \lambda \ge 0 \, , \ \forall w, \delta w \in H \, ,\ \forall g \in G\; . $$
(A.2)

According to Eq. (2.3), the system’s equilibrium solutions \(w(x ; \lambda )\) are found by extremizing its energy; consequently all solutions of the system \(\mathcal{E}_{,w}(w;\lambda ) \delta w = 0\) must satisfy Eq. (A.2). It is more appropriate to talk about orbits of equilibrium paths since, in view of the equivariance described in (A.2), applying to an equilibrium solution \(w\) the transformation \(T_{g}\) automatically generates another equilibrium solution \(T_{g}[w]\).

A subset of these equilibrium solutions, termed “principal solutions” and denoted by \({\stackrel{0}{w}}(x; \lambda )\), are invariant under all transformations \(T_{g}\). These solutions belong to an invariant subspace of \(H\), denoted \({\mathcal{S}}_{G}\) and called the “fixed-point space

$$ \mathcal{E}_{,w}({\stackrel{0}{w}}(x; \lambda );\lambda ) \delta w = 0 \; , \ \forall \lambda \ge 0 \; ; \quad {\stackrel{0}{w}} \in { \mathcal{S}}_{G} := \{ w \in H \;|\; \ T_{g} [w] = {w} \; , \ \forall g \in G \} \; . $$
(A.3)

To determine the stability of a principal solution, one has to check the positive definiteness of the self-adjoint bilinear operator \({\mathcal{E}}^{0}_{,ww}\), evaluated on the principal path \({\stackrel{0}{w}}(x; \lambda )\), by finding its eigenvalues \(\beta (\lambda )\)

$$ ({\mathcal{E}}^{0}_{,ww}\Delta w)\delta w = \beta (\lambda ) < \Delta w, \delta w> \; ; \quad \forall \delta w \in H\; ; \ {\mathcal{E}}^{0}_{,ww} := {\mathcal{E}}_{,ww}({\stackrel{0}{w}}(x;\lambda );\lambda ) \; , $$
(A.4)

where \(\Delta w\) is the corresponding eigenvector and \(< \cdot \; ,\; \cdot >\) denotes an inner product in \(H\). A stable solution corresponds to a positive minimum eigenvalueFootnote 16\(\beta _{min} > 0\) (the number of eigenvalues depends on the dimension of \(H\)). When \({\stackrel{0}{w}}(x;\lambda )\) is periodic, the Bloch-wave representation may be used, as described in Sect. 2.4. For a well-posed problem, its stress-free (unloaded) configuration at \(\lambda =0\) is stable; as the load increases stability will be lost at the first bifurcation point encountered along the loading path at some \(\lambda _{b}\).

It can be shown, e.g., [22, 40], that the existence of the group \(G\) implies the existence of a symmetry basis with respect to which (i) the operator \({\mathcal{E}}^{0}_{,ww}\) defined in (A.4) can be block-diagonalized and (ii) the space of admissible functions \(H=\oplus _{\mu =1}^{h} V^{\mu }\) can be uniquely decomposed into a direct sum of mutually orthogonal invariant subspaces \(V^{\mu }\) (with \(h\) being the number of equivalence, or conjugacy, classes for \(G\)). Each subspace \(V^{\mu }\) is associated with an \(n_{\mu }\)-dimensional irreducible representation \(\tau ^{\mu }\) of \(G\), also termed “irrep” from which an appropriate projection operator can be constructed giving the \(V^{\mu }\) component of any function in \(H\).

Bifurcated equilibrium paths, termed primary, can emerge from the principal path at loads \(\lambda _{b}\) corresponding to zero eigenvalues of \({\mathcal{E}}^{b}_{,ww}\). That is, \(\beta (\lambda _{b})=0\) so that

$$ \begin{gathered} ({\mathcal{E}}^{b}_{,ww} {\stackrel{i}{w}})\delta w = 0 \; , \ { \stackrel{i}{w}} \in {\mathcal{N}}_{\mu }\; ,\ < {\stackrel{i}{w}},{ \stackrel{j}{w}}>=\delta _{ij}\; , \ i , j = 1,\dots ,n_{\mu }\; ; \\ \forall \delta w \in H\; ;\ {\mathcal{E}}^{b}_{,ww} := {\mathcal{E}}_{,ww}({\stackrel{0}{w}}(x; \lambda _{b});\lambda _{b}) \; , \end{gathered} $$
(A.5)

where the eigenmodes \(\Delta w =\stackrel{i}{w}\) span \({\mathcal{N}}_{\mu }\), the \(n_{\mu }\)-dimensional null space of the operator \({\mathcal{E}}^{b}_{,ww}\). Some additional conditions, termed “transversality” conditions must also hold to ensure that \(\lambda _{b}\) is a bifurcation and not a limit (i.e., turning) point:Footnote 17

$$ ([d{\mathcal{E}}^{0}_{,ww}/d\lambda ]_{b}{\stackrel{1}{w}}){ \stackrel{1}{w}} \neq 0\; . $$
(A.6)

It can be shown that the null space \({\mathcal{N}}_{\mu }\) of the stability operator \({\mathcal{E}}^{b}_{,ww}\) is an \(n_{\mu }\)-dimensional subspace of the invariant subspace \(V^{\mu }\) for some \(\mu \), i.e. \({ \mathcal{N}}_{\mu } \subseteq V^{\mu }\), associated to the irreducible representation \(\tau ^{\mu }\). The (primary) bifurcated paths \({\stackrel{b}{w}}(x; \lambda )\) emerging from \({\stackrel{0}{w}}(x; \lambda )\) at \(\lambda _{b}\), can be represented with the following asymptotic expressions, resulting from the projection of the equilibrium equations onto the null space \({\mathcal {N}}_{\mu }\) and termed “Lyapunov–Schmidt–Koiter” asymptotics:

$$ \lambda (\xi )=\lambda _{b}+ \lambda _{1}\xi + {\frac{1}{2}} \lambda _{2} \xi ^{2} + O(\xi ^{3})\; , \qquad {\stackrel{b}{w}}(x; \xi ) = { \stackrel{0}{w}}(x; \lambda (\xi )) + \xi \sum _{i=1}^{n_{\mu }} \alpha _{i} {\stackrel{i}{w}}(x) + O(\xi ^{2})\; , $$
(A.7)

where \(\xi \) is the path’s “bifurcation amplitude” parameter and \(\alpha _{i}\) the components of the bifurcated path’s initial tangent vector. From knowledge of the bifurcation irrep \(\tau ^{\mu }\), group theory results (such as the “lattice of isotropy subgroups” of \(G\) [13]) allow the direct computation of the \(\alpha _{i}\) tangents (one set for each bifurcating orbit) emerging at \(\lambda _{b}\). Once the asymptotic expression of the primary bifurcating equilibrium path \({\stackrel{b}{w}}(x; \xi )\) is established, the global solution path may be computed efficiently by using the path’s isotropy group, i.e. the elements of the subgroup of \(G\) satisfying \(T_{g} [{\stackrel{b}{w}}]={\stackrel{b}{w}}\), to find its corresponding fixed-point space. Along this path there may occur (secondary) bifurcation points. In such cases, the above procedure begins once again with \({\stackrel{b}{w}}(x; \lambda )\) as the new principal path from which—secondary with respect to \({\stackrel{0}{w}}(x; \lambda )\)—bifurcated orbits will emerge.

In order to determine the local (asymptotic) type—transverse or pitchfork—of the bifurcating paths (which may be different than the global type), we need to determine if \(\lambda _{1} = 0\) or \(\lambda _{1} \neq 0\) in Eq. (A.7)1. From an asymptotic analysis (e.g., [47]) \(\lambda _{1} = 0\) is guaranteed if

$$ \mathcal{E}_{ijk} := (([\mathcal{E}_{,www}({\stackrel{0}{w}}(x; \lambda _{b}); \lambda _{b})] {\stackrel{i}{w}}) {\stackrel{j}{w}}) { \stackrel{k}{w}} = 0, \qquad i,j,k = 1,\dots ,n_{\mu }. $$
(A.8)

To reiterate, the strategy followed in this work is to sequentially apply the above-described procedure to follow the bifurcating equilibrium orbits of the system by identifying, each time, their symmetry group and their corresponding fixed-point space. As we proceed from the principal solution to the primary bifurcations emerging from it, then to the secondary bifurcations emerging from the primary ones, the corresponding symmetry groups and fixed-point spaces change accordingly. Knowledge of the symmetries of a path allows for an efficient calculation of a unique solution in its own fixed-point space. The method adopted here follows the procedures introduced by [20, 23]. Moreover, following [13, 20], knowledge of the lattice of isotropy subgroups of the initial symmetry group guides the search for the bifurcated equilibrium paths in a systematic way and explains our findings.

1.1 A.1 Principal Solution, Irreps, and Bifurcations—Group \(G = D_{\infty h}\)

As described in Sect. 2.3, the fixed-point space of \(D_{\infty h}\) for the beam–foundation system consists of only the trivial principal solution \(\overset{0}{w}(x; \lambda ) = 0\). According to group theory (e.g., [32]) \(D_{\infty h}\) has four 1-dimensional irreps (one being the trivial identity irrep). These provide the possibility of simple bifurcations to paths with symmetry groups of \(C_{\infty v}\), \(C_{\infty h}\) or \(D_{\infty }\). There are also an infinity of 2-dimensional irreps, providing the possibility of double bifurcations. These correspond to bifurcating paths with symmetry groups \(D_{nh}\) or \(D_{nd}\), where \(n\in \mathbb{N}\), as shown in Table 1. In Sect. 4.1.2 we find from Eq. (4.5) only double bifurcations at \(\lambda _{n} \geq 2\) with corresponding eigenmodes \({\stackrel{ni}{w}}(x), i=1,2\)—as expected from the 2-dimensional irreps of \(D_{\infty h}\) of Table 1. Notice that no simple bifurcations exist, in spite of the existence of 1-dimensional irreps of this group. From the infinity of primary bifurcation paths, we follow next the bifurcation orbit emerging from the lowest load \(\lambda _{c} = 2\) which corresponds to \(n_{c}=q\) and \(L_{c} = 2\pi \) (recall that the periodicity of the model is \(L_{d}= L_{c} q\)).

Table 1 Table of irreps and bifurcated orbit symmetries for \(G = D_{\infty h}\). The first column gives the dimension \(n_{\mu }\) of the corresponding irrep; the second column gives a standard label/name for the irrep; the third through fifth columns provide the corresponding irrep matrix for the generators \(\tau ^{\mu }_{c(\phi )}\), \(\tau ^{\mu }_{\sigma _{v}}\), and \(\tau ^{\mu }_{\sigma _{h}}\), respectively; the sixth column gives the kernel of \(\tau ^{\mu }\) (i.e., \(G^{\mu } = \{ g \in G \;|\; \tau ^{\mu }_{g} = I\}\), where \(I\) is the \(n_{\mu }\)-dimensional identity matrix); the seventh column gives the symmetry group of the corresponding bifurcating equilibrium path(s)

1.2 A.2 Primary Bifurcation Orbit at \(\lambda _{c} = 2\), Irreps, and Bifurcations—Group \(G = D_{qd}\)

The bifurcated equilibrium paths emerging from the lowest critical load \(\lambda _{c}\) correspond according to Eq. (4.4) to a double bifurcation with eigenmodes \({\stackrel{c1}{w}}(x) = \sin (x)\) and \({\stackrel{c2}{w}}(x) = \cos (x)\). Since every linear combination of these eigenmodes is left invariant by the elements of the group \(S_{2q}\), the critical point corresponds to the \(\mu = E_{2q}\) irrep, and according to the general theory (see Table 1) the symmetry group of the bifurcating orbit is \(D_{qd}\). This symmetry group is finite and has the following two generators: \(\sigma _{h} c(\pi /q)\) and \(\sigma _{h}\sigma _{v}\).

As indicated in Table 2, this group has four 1-dimensional irreps (one being the trivial identity irrep). These provide the possibility of simple bifurcations to paths with symmetry groups of \(S_{2q}\), \(D_{q}\), or \(C_{qv}\). There are also \((q-1)\) 2-dimensional irreps, providing the possibility of double bifurcations. These correspond to bifurcating paths with symmetry groups \(D_{rd}\), \(C_{rv}\) and \(D_{r}\), where \(r:=\mathrm{gcd}(j,q)\), with \(j=1,\dots ,q-1\). The fixed-point space \(\mathcal{S}_{ D_{qd}} := \{w(x) \in H \;|\; T_{g}[w(x)] = w(x), \; \forall g \in D_{qd}\}\) of the primary bifurcation solutions \({\stackrel{1}{w}}\) with \(D_{qd}\) symmetry, as defined above, contains configurations \(w(x)\) of the \(L_{d}\)-periodic beam that remain unaltered under the transformations of this group. The primary bifurcation path \({\stackrel{1}{w}}(x; \lambda )\) is contained within \(\mathcal{S}_{D_{qd}}\) and a representative configuration along this path parameterized with respect to the bifurcation amplitude \(\xi \) (see Sect. 4.1.2) is plotted in Fig. 3.

Table 2 Table of irreps and bifurcated orbit symmetries for \(G = D_{qd}\). The first column gives the dimension \(n_{\mu }\) of the corresponding irrep; the second column gives a standard label/name for the irrep; the third and fourth columns provide the corresponding irrep matrix for the generators \(\tau ^{\mu }_{c(\pi /q)}\) and \(\tau ^{\mu }_{\sigma _{h} \sigma _{v}}\), respectively; the fifth column gives the kernel of \(\tau ^{\mu }\) (i.e., \(G^{\mu } = \{ g \in G \;|\; \tau ^{\mu }_{g} = I\}\), where \(I\) is the \(n_{\mu }\)-dimensional identity matrix); the sixth column gives the symmetry group of the corresponding bifurcating equilibrium path(s). The function \(\gcd (a,b)\) is the greatest common divisor of \(a\) and \(b\)

1.3 A.3 Secondary Bifurcation Orbits and Their Symmetry

Recall that for the numerical calculations reported here, we have selected \(q=20\) which gives \(L_{d} = L_{c} q = 40 \pi \). Due to the (energy-induced) symmetry of the dispersion curve discussed in Sect. 4.1.2, all bifurcations are double. Indeed, double bifurcation points corresponding to all of the 2-dimensional irreps \(E^{j}\) of \(D_{20d}\) are found along the primary bifurcation orbit \({\stackrel{1}{w}}(x; \lambda )\). The correspondence between the 2-dimensional irrep index \(j\) of Table 2 and the Bloch-wave wavenumber \(k\) is presented in Table 3, along with the symmetry group of each orbit bifurcating from a bifurcation point of this type. The \(j\)\(k\) correspondence is obtained by using standard group representation theory results to decompose the irreps of \(D_{20d}\) into direct sums of the irreps of its subgroup \(C_{20}\). In this way, we can find which wavenumbers \(k\) (used in the Bloch-wave calculations of Sect. 3) correspond to each 2-dimensional irrep of \(D_{20d}\). The global bifurcation type, either transverse or pitchfork, is obtained from a theorem in equivariant bifurcation theory [20], and indicates the nature of the global bifurcating equilibrium path. In particular, a globally transcritical path \(w(x; \xi )\) will have \(w(x; -\xi ) \neq w(x; \xi )\) for some value(s) of \(\xi \), whereas a globally pitchfork path will have \(w(x; -\xi ) = w(x; \xi )\) for all \(\xi \). See Sect. 4 for some discussion of the different types of paths found in the present work.

Table 3 Correspondence between irreps of \(D_{20d}\) and Bloch-wave wavenumber \(k\). Also shown are the symmetries and global bifurcation type (transverse or pitchfork) of the corresponding bifurcating orbits

In addition to the group-theory results in Table 3, we can also show by group-theoretic calculation that all the secondary bifurcating paths are locally pitchfork, i.e. \(\lambda _{1}=0\) in Eq. (A.7)1 since all the coefficients \(\mathcal{E}_{ijk}\) of Eq. (A.8) vanish. This result is obtained by constructing an appropriate tensor-product rep [40] for the vector space of third-order tensors to which the “vector” \(\mathcal{E}_{ijk}\) belongs. Due to equivariance, \(\mathcal{E}_{ijk}\) must be in the fixed-point space \(\mathcal{S}_{D_{20d}}\) of this tensor-product space. Further, it is easily found that \(\mathcal{S}_{D_{20d}} = \{0\}\). Thus, all paths bifurcating from the \(D_{20d}\) primary orbit are locally pitchfork. However, as indicated in Table 3, not all such bifurcating paths are globally pitchfork. Again, see Sect. 4 for some discussion of this issue.

1.4 A.4 Critical Load and Eigenmodes for the Imperfect Beam (Amplitude: \(\zeta < 0\))

We present here a calculation of the lowest load \(\lambda _{cis}\) corresponding to a bifurcation with a symmetric (with respect to \(x=0\)) eigenmode \({\stackrel{is}{w}}(x)\) for an imperfect, periodic beam of length \(L_{d}\). The calculation of the lowest load corresponding to a bifurcation with an antisymmetric (with respect to \(x=0\)) eigenmode is similar and gives a slightly higher critical load \(\lambda _{cia}\); the corresponding calculation is omitted here since we are interested in the bifurcated equilibrium paths emerging from the lowest load \(\lambda _{cis}\).

By taking the functional derivative of the equilibrium Eq. (2.7) and evaluating at \(w(x)=0\) we obtain the equation for the symmetric with respect to \(x=0\) eigenmode of the imperfect beam (\({\stackrel{is}{w}}(x)={ \stackrel{is}{w}}(-x)\)). The corresponding boundary conditions follow from Eq. (2.4) for \(p=1, 3\) and hence

$$ \begin{gathered} \frac{d^{4} {\stackrel{is}{w}}}{d x^{4}} + \lambda \frac{d^{2} {\stackrel{is}{w}}}{d x^{2}} + [1+z(x)]{ \stackrel{is}{w}}= 0\; ,\ x \in (0,\frac{L_{d}}{2}) ;\\ \frac{d {\stackrel{is}{w}}}{d x} (0) = \frac{d {\stackrel{is}{w}}}{d x} ({\frac{L_{d}}{2}}) = 0 \; ,\ \frac{d^{3} {\stackrel{is}{w}}}{d x^{3}} (0) = \frac{d^{3} {\stackrel{is}{w}}}{d x^{3}} ({\frac{L_{d}}{2}}) = 0\; . \end{gathered} $$
(A.9)

Solving the problem in Eq. (A.9) with the constant coefficient ordinary differential equation in the intervals \((0, x_{0})\) and \((x_{0}, L_{d}/2)\), one has

$$ {\stackrel{is}{w}}(x) = \left \{ \textstyle\begin{array}{l@{\quad }l} \left . \textstyle\begin{array}{l} \eta \cos (a_{\zeta }x) + \theta \cos (b_{\zeta }x) \\ a_{\zeta }:= \displaystyle \left [{ \frac{\lambda +[\lambda ^{2}-4(1+\zeta )]^{1/2}}{2}}\right ]^{1/2}, \quad b_{\zeta }:= \left [{ \frac{\lambda -[\lambda ^{2}-4(1+\zeta )]^{1/2}}{2}}\right ]^{1/2} \end{array}\displaystyle \right \} \\ \quad x \in [0, x_{0}] \\ \left . \textstyle\begin{array}{l} \displaystyle \delta \cosh [a(x-\frac{L_{d}}{2})] \cos [b(x- \frac{L_{d}}{2})] + \epsilon \sinh [a(x-\frac{L_{d}}{2})] \sin [b(x- \frac{L_{d}}{2})] \\ a := \displaystyle {\frac{[2 -\lambda ]^{1/2}}{2}}, \quad b := { \frac{[2 +\lambda ]^{1/2}}{2}} \end{array}\displaystyle \quad \quad \ \ \right \} \\ \quad \displaystyle x \in [x_{0} ,{ \frac{L_{d}}{2}}] \end{array}\displaystyle \right . $$
(A.10)

where \(a_{\zeta },\; b_{\zeta },\; a,\; b\) are all positive constants. Notice from Eq. (A.10) that the periodicity-imposed symmetry conditions at \(L_{d}/2\) according to Eq. (A.9) are automatically satisfied.

Using the continuity condition at \(x_{0}\) for \({\stackrel{is}{w}}(x)\) and its derivatives up to order three, one obtains a homogeneous linear system of four equations for the four constants \(\eta , \theta , \delta , \epsilon \) appearing in Eq. (A.10). The vanishing of the determinant of the corresponding \(4 \times 4\) matrix (not recorded here in view of the cumbersome expressions) provides the equation for the critical load \(2(1+\zeta )^{1/2} < \lambda _{cis} < 2\) (guaranteed by \(\zeta <0\)), which is obtained numerically.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pandurangi, S.S., Elliott, R.S., Healey, T.J. et al. Stable Spatially Localized Configurations in a Simple Structure—A Global Symmetry-Breaking Approach. J Elast 142, 163–199 (2020). https://doi.org/10.1007/s10659-020-09794-5

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10659-020-09794-5

Keywords

Mathematics Subject Classification

Navigation