Skip to main content
Log in

Global bifurcations and homoclinic chaos in nonlinear panel optomechanical resonators under combined thermal and radiation stresses

  • Original paper
  • Published:
Nonlinear Dynamics Aims and scope Submit manuscript

Abstract

In this paper, we investigate the nonlinear interaction and time-evolution of confined optical, thermal and mechanical modes in a three-dimensional optomechanical resonator. We model a high-finesse cavity with a thin thermo-elastic panel placed at the center of two highly reflective mirrors that is subjected to an incoming continuous-wave electromagnetic field. The latter is constructed by constraining the light-structure interaction to a first-order scattering phenomenon in the classical interpretation modeled as a spatio-temporal perturbation around a time-harmonic field. We employ a Galerkin-based separation of variables decomposition on the resultant fields and replace them with their nonlinear modal counterparts. The resulting dynamical system is thus governed by the combined effects of thermal and radiation stresses which yield a complex spatially dependent self-excited bifurcation structure where Hopf bifurcations give rise to periodic limit-cycle solutions. In regions where coexisting solutions are found, homoclinic connections ensue codimension-two Bogdanov–Takens and Double-Hopf bifurcations and that for a range of control parameters a global homoclinic Shilnikov bifurcation culminates with a distinct period-doubling route to chaos. We note that the current formulation demonstrates the essential contribution of coupled thermal and radiation stresses to the bifurcation structure of nonlinear light-structure interaction systems and may shed light to modal energy transfer mechanisms and scattering phenomena.

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
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22

Similar content being viewed by others

References

  1. Jacobs-Cook, A.J.: MEMS versus MOMS from a systems point of view. J. Micromech. Microeng. 6, 148 (1996). https://doi.org/10.1088/0960-1317/6/1/035

    Article  Google Scholar 

  2. Kippenberg, T.J., Vahala, K.J.: Cavity optomechanics: back-action at the mesoscale. Science 321, 1172–1176 (2008). https://doi.org/10.1126/science.1156032

    Article  Google Scholar 

  3. Poosanaas, P., Tonooka, K., Uchino, K.: Photostrictive actuators. Mechatronics 10, 467–487 (2000). https://doi.org/10.1016/S0957-4158(99)00073-2

    Article  Google Scholar 

  4. Stokes, N.A.D., Fatah, R.M.A., Venkatesh, S.: Self-excited vibrations of optical microresonators. Electron. Lett. 24, 777 (1988)

    Article  Google Scholar 

  5. Restrepo, J., Gabelli, J., Ciuti, C., Favero, I.: Classical and quantum theory of photothermal cavity cooling of a mechanical oscillator. C. R. Phys. 12, 860–870 (2011). https://doi.org/10.1016/j.crhy.2011.02.005

    Article  Google Scholar 

  6. Stahl, S.W., Puchner, E.M., Gaub, H.E.: Photothermal cantilever actuation for fast single-molecule force spectroscopy. Rev. Sci. Instrum. 80, 073702 (2009). https://doi.org/10.1063/1.3157466

    Article  Google Scholar 

  7. Ratcliff, G.C., Erie, D.A., Superfine, R.: Photothermal modulation for oscillating mode atomic force microscopy in solution. Appl. Phys. Lett. 72, 1911–1913 (1998). https://doi.org/10.1063/1.121224

    Article  Google Scholar 

  8. Jourdan, G., Comin, F., Chevrier, J.: Mechanical mode dependence of bolometric backaction in an atomic force microscopy microlever. Phys. Rev. Lett. 101, 133904 (2008). https://doi.org/10.1103/PhysRevLett.101.133904

    Article  Google Scholar 

  9. Hölscher, H., Milde, P., Zerweck, U., Eng, L.M., Hoffmann, R.: The effective quality factor at low temperatures in dynamic force microscopes with Fabry–Pérot interferometer detection. Appl. Phys. Lett. 94, 223514 (2009). https://doi.org/10.1063/1.3149700

    Article  Google Scholar 

  10. Labuda, A., Kobayashi, K., Miyahara, Y., Grütter, P.: Retrofitting an atomic force microscope with photothermal excitation for a clean cantilever response in low Q environments. Rev. Sci. Instrum. 83, 053703 (2012). https://doi.org/10.1063/1.4712286

    Article  Google Scholar 

  11. Tabib-Azar, M.: Optically controlled silicon microactuators. Nanotechnology 1, 81 (1990). https://doi.org/10.1088/0957-4484/1/1/013

    Article  Google Scholar 

  12. Kiracofe, D., Kobayashi, K., Labuda, A., Raman, A., Yamada, H.: High efficiency laser photothermal excitation of microcantilever vibrations in air and liquids. Rev. Sci. Instrum. 82, 013702 (2011). https://doi.org/10.1063/1.3518965

    Article  Google Scholar 

  13. Carmon, T., Rokhsari, H., Yang, L., Kippenberg, T.J., Vahala, K.J.: Temporal behavior of radiation-pressure-induced vibrations of an optical microcavity phonon mode. Phys. Rev. Lett. 94, 223902 (2005). https://doi.org/10.1103/PhysRevLett.94.223902

    Article  Google Scholar 

  14. Meyer, T.R., Pryor, W.R., McKay, C.P., McKenna, P.M.: Laser elevator: momentum transfer using an optical resonator. J. Spacecraft Rockets 39, 258–266 (2002). https://doi.org/10.2514/2.3807

    Article  Google Scholar 

  15. Meystre, P., Wright, E.M., McCullen, J.D., Vignes, E.: Theory of radiation-pressure-driven interferometers. J. Opt. Soc. Am. B. 2, 1830–1840 (1985). https://doi.org/10.1364/JOSAB.2.001830

    Article  Google Scholar 

  16. Anetsberger, G., Arcizet, O., Unterreithmeier, Q.P., Rivière, R., Schliesser, A., Weig, E.M., Kotthaus, J.P., Kippenberg, T.J.: Near-field cavity optomechanics with nanomechanical oscillators. Nat. Phys. 5, 909–914 (2009). https://doi.org/10.1038/nphys1425

    Article  Google Scholar 

  17. Ludwig, M., Safavi-Naeini, A.H., Painter, O., Marquardt, F.: Enhanced quantum nonlinearities in a two-mode optomechanical system. Phys. Rev. Lett. 109, 063601 (2012). https://doi.org/10.1103/PhysRevLett.109.063601

    Article  Google Scholar 

  18. Molloy, J.E., Padgett, M.J.: Lights, action: optical tweezers. Contemp. Phys. 43, 241–258 (2002). https://doi.org/10.1080/00107510110116051

    Article  Google Scholar 

  19. Aubin, K., Zalalutdinov, M., Alan, T., Reichenbach, R.B., Rand, R., Zehnder, A., Parpia, J., Craighead, H.: Limit cycle oscillations in CW laser-driven NEMS. J. Microelectromech. Syst. 13, 1018–1026 (2004). https://doi.org/10.1109/JMEMS.2004.838360

    Article  Google Scholar 

  20. Marino, F., Marin, F.: Chaotically spiking attractors in suspended-mirror optical cavities. Phys. Rev. E 83, 015202 (2011). https://doi.org/10.1103/PhysRevE.83.015202

    Article  Google Scholar 

  21. Lee, D., Underwood, M., Mason, D., Shkarin, A.B., Hoch, S.W., Harris, J.G.E.: Multimode optomechanical dynamics in a cavity with avoided crossings. Nat. Commun. 6, 6232 (2015). https://doi.org/10.1038/ncomms7232

    Article  Google Scholar 

  22. Chia, C.-Y.: Nonlinear Analysis of Plates. McGraw-Hill Inc., New York (1980)

    Google Scholar 

  23. Nayfeh, A.H., Pai, P.F.: Linear & Nonlinear Structural Mechanics. Wiley-VCH, Hoboken (2002)

    MATH  Google Scholar 

  24. Wu, C.-I., Vinson, J.R.: Influences of large amplitudes, transverse shear deformation, and rotatory inertia on lateral vibrations of transversely isotropic plates. J. Appl. Mech. 36, 254 (1969). https://doi.org/10.1115/1.3564617

    Article  MATH  Google Scholar 

  25. Yu, Y.Y., Lai, J.L.: Influence of transverse shear and edge condition on nonlinear vibration and dynamics buckling of homogeneous and sandwich plates. J. Appl. Mech. 33, 934–936 (1966). https://doi.org/10.1115/1.3625205

    Article  Google Scholar 

  26. Jones, R., Mazumdar, J., Cheung, Y.K.: Vibration and buckling of plates at elevated temperatures. Int. J. Solids Struct. 16, 61–70 (1980). https://doi.org/10.1016/0020-7683(80)90095-5

    Article  MathSciNet  MATH  Google Scholar 

  27. Shabana, A.: Vibration of discrete and continuous systems. Springer, Berlin (2012)

    MATH  Google Scholar 

  28. Jackson, J.D.: Classical Electrodynamics, 3rd edn. Wiley, Hoboken (1998)

    MATH  Google Scholar 

  29. Baruch, G., Fibich, G., Tsynkov, S.: A high-order numerical method for the nonlinear Helmholtz equation in multidimensional layered media. J. Comput. Phys. 228, 3789–3815 (2009). https://doi.org/10.1016/j.jcp.2009.02.014

    Article  MathSciNet  MATH  Google Scholar 

  30. Xuereb, A., Domokos, P., Asbóth, J., Horak, P., Freegarde, T.: Scattering theory of cooling and heating in optomechanical systems. Phys. Rev. A 79, 053810 (2009). https://doi.org/10.1103/PhysRevA.79.053810

    Article  Google Scholar 

  31. Gorodetsky, M.L., Ilchenko, V.S.: Optical microsphere resonators: optimal coupling to high-Q whispering-gallery modes. J. Opt. Soc. Am. B 16, 147 (1999). https://doi.org/10.1364/JOSAB.16.000147

    Article  Google Scholar 

  32. Law, C.K.: Interaction between a moving mirror and radiation pressure: a Hamiltonian formulation. Phys. Rev. A 51, 2537–2541 (1995). https://doi.org/10.1103/PhysRevA.51.2537

    Article  Google Scholar 

  33. Janowicz, M.: Evolution of wave fields and atom-field interactions in a cavity with one oscillating mirror. Phys. Rev. A 57, 4784–4790 (1998). https://doi.org/10.1103/PhysRevA.57.4784

    Article  Google Scholar 

  34. Crocce, M., Dalvit, D.A.R., Mazzitelli, F.D.: Resonant photon creation in a three-dimensional oscillating cavity. Phys. Rev. A 64, 013808 (2001). https://doi.org/10.1103/PhysRevA.64.013808

    Article  Google Scholar 

  35. Cheung, H.K., Law, C.K.: Nonadiabatic optomechanical Hamiltonian of a moving dielectric membrane in a cavity. Phys. Rev. A 84, 023812 (2011). https://doi.org/10.1103/PhysRevA.84.023812

    Article  Google Scholar 

  36. Gil-Santos, E., Ramos, D., Pini, V., Llorens, J., Fernández-Regúlez, M., Calleja, M., Tamayo, J., Paulo, A.S.: Optical back-action in silicon nanowire resonators: bolometric versus radiation pressure effects. New J. Phys. 15, 035001 (2013). https://doi.org/10.1088/1367-2630/15/3/035001

    Article  Google Scholar 

  37. Mansuripur, M.: Radiation pressure and the linear momentum of the electromagnetic field. Opt. Express 12, 5375 (2004). https://doi.org/10.1364/OPEX.12.005375

    Article  Google Scholar 

  38. Zakharian, A.R., Mansuripur, M., Moloney, J.V.: Radiation pressure and the distribution of electromagnetic force in dielectric media. Opt. Express 13, 2321 (2005). https://doi.org/10.1364/OPEX.13.002321

    Article  Google Scholar 

  39. Rakhmanov, M.: Doppler-induced dynamics of fields in Fabry-Perot cavities with suspended mirrors. Appl. Opt. 40, 1942 (2001). https://doi.org/10.1364/AO.40.001942

    Article  Google Scholar 

  40. Vial, B., Zolla, F., Nicolet, A., Commandré, M.: Quasimodal expansion of electromagnetic fields in open two-dimensional structures. Phys. Rev. A 89, 023829 (2014). https://doi.org/10.1103/PhysRevA.89.023829

    Article  Google Scholar 

  41. Aspelmeyer, M., Kippenberg, T.J., Marquardt, F.: Cavity optomechanics. Rev. Mod. Phys. 86, 1391–1452 (2014). https://doi.org/10.1103/RevModPhys.86.1391

    Article  Google Scholar 

  42. Bhat, R.B.: Natural frequencies of rectangular plates using characteristic orthogonal polynomials in rayleigh-ritz method. J. Sound Vib. 102, 493–499 (1985). https://doi.org/10.1016/S0022-460X(85)80109-7

    Article  Google Scholar 

  43. Huang, C.-H., Chen, Y.-Y.: Vibration analysis for piezoceramic rectangular plates using Ritz’s method with equivalent constants. IEEE Trans. Ultrason. Ferroelectr. Freq. Control 53, 265–273 (2006). https://doi.org/10.1109/TUFFC.2006.1593364

    Article  Google Scholar 

  44. Crawford, J.D., Knobloch, E.: Symmetry and symmetry-breaking bifurcations in fluid dynamics. Annu. Rev. Fluid Mech. 23, 341–387 (1991). https://doi.org/10.1146/annurev.fl.23.010191.002013

    Article  MathSciNet  MATH  Google Scholar 

  45. Strogatz, S.H.: Nonlinear Dynamics and Chaos: With Applications to Physics, Biology, Chemistry, and Engineering. Westview Press, Cambridge (2001)

    MATH  Google Scholar 

  46. Wurl, C.: Symmetry-breaking oscillations in membrane optomechanics. Phys. Rev. A. (2016). https://doi.org/10.1103/PhysRevA.94.063860

    Article  Google Scholar 

  47. PrasannaV enkatesh, B., Larson, J., O’Dell, D.H.J.: Band-structure loops and multistability in cavity QED. Phys. Rev. A. 83, 063606 (2011). https://doi.org/10.1103/PhysRevA.83.063606

    Article  Google Scholar 

  48. Guckenheimer, J., Holmes, P.: Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vector Fields. Springer, Berlin (1983)

    Book  Google Scholar 

  49. Wiggins, S.: Introduction to Applied Nonlinear Dynamical Systems and Chaos. Springer, New York (2003)

    MATH  Google Scholar 

  50. Nayfeh, A.H., Balachandran, B.: Applied Nonlinear Dynamics: Analytical, Computational and Experimental Methods. Wiley-VCH, New York (1995)

    Book  Google Scholar 

  51. Guckenheimer, J., Myers, M., Sturmfels, B.: Computing Hopf Bifurcations I. SIAM J. Numer. Anal. 34, 1–21 (1997). https://doi.org/10.1137/S0036142993253461

    Article  MathSciNet  MATH  Google Scholar 

  52. Gross, T., Feudel, U.: Analytical search for bifurcation surfaces in parameter space. Phys. D 195, 292–302 (2004). https://doi.org/10.1016/j.physd.2004.03.019

    Article  MathSciNet  MATH  Google Scholar 

  53. Shkarin, A.B., Flowers-Jacobs, N.E., Hoch, S.W., Kashkanova, A.D., Deutsch, C., Reichel, J., Harris, J.G.E.: Optically mediated hybridization between two mechanical modes. Phys. Rev. Lett. 112, 013602 (2014). https://doi.org/10.1103/PhysRevLett.112.013602

    Article  Google Scholar 

  54. Lü, X.-Y.: PT symmetry-breaking chaos in optomechanics. Phys. Rev. Lett. (2015). https://doi.org/10.1103/PhysRevLett.114.253601

    Article  Google Scholar 

  55. Kuznetsov, Y.A.: Elements of Applied Bifurcation Theory. Springer, Berlin (2013)

    Google Scholar 

  56. Hollander, E., Gottlieb, O.: Self-excited chaotic dynamics of a nonlinear thermo-visco-elastic system that is subject to laser irradiation. Appl. Phys. Lett. 101, 133507 (2012). https://doi.org/10.1063/1.4755844

    Article  Google Scholar 

  57. Hollander, E.: Self-excited oscillations bifurcations and chaos in nonlinear optomechanical thermo-visco-elastic panel resonators (2017)

  58. Bakemeier, L., Alvermann, A., Fehske, H.: Route to Chaos in Optomechanics. Phys. Rev. Lett. 114, 013601 (2015). https://doi.org/10.1103/PhysRevLett.114.013601

    Article  Google Scholar 

  59. Zaitsev, S., Gottlieb, O., Buks, E.: Nonlinear dynamics of a microelectromechanical mirror in an optical resonance cavity. Nonlinear Dyn. 69, 1589–1610 (2012). https://doi.org/10.1007/s11071-012-0371-9

    Article  MathSciNet  Google Scholar 

  60. Herrero, R., Pons, R., Farjas, J., Pi, F., Orriols, G.: Homoclinic dynamics in experimental Shil’nikov attractors. Phys. Rev. E 53, 5627–5636 (1996). https://doi.org/10.1103/PhysRevE.53.5627

    Article  Google Scholar 

  61. Shil’nikov, L.P.: Methods of Qualitative Theory in Nonlinear Dynamics. World Scientific, London (2001)

    Book  Google Scholar 

  62. Afraimovich, V.S., Gonchenko, S.V., Lerman, L.M., Shilnikov, A.L., Turaev, D.V.: Scientific heritage of L.P. Shilnikov. Regul. Chaot. Dyn. 19, 435–460 (2014). https://doi.org/10.1134/S1560354714040017

    Article  MathSciNet  MATH  Google Scholar 

  63. Arneodo, A., Coullet, P., Tresser, C.: Oscillators with chaotic behavior: an illustration of a theorem by Shil’nikov. J. Stat. Phys. 27, 171–182 (1982). https://doi.org/10.1007/BF01011745

    Article  MATH  Google Scholar 

  64. Gottlieb, O., Habib, G.: Non-linear model-based estimation of quadratic and cubic damping mechanisms governing the dynamics of a chaotic spherical pendulum. J. Vib. Control 18, 536–547 (2012). https://doi.org/10.1177/1077546310395969

    Article  MathSciNet  Google Scholar 

  65. Gomis-Bresco, J., Navarro-Urrios, D., Oudich, M., El-Jallal, S., Griol, A., Puerto, D., Chavez, E., Pennec, Y., Djafari-Rouhani, B., Alzina, F., Martínez, A., Torres, C.M.S.: A one-dimensional optomechanical crystal with a complete phononic band gap. Nat. Commun. 5, 4452 (2014). https://doi.org/10.1038/ncomms5452

    Article  Google Scholar 

  66. Ventsel, E., Krauthammer, T.: Thin Plates and Shells: Theory: Analysis, and Applications. CRC Press, Boca Raton (2001)

    Book  Google Scholar 

  67. Orfanidis, S.J.: Electromagnetic waves and antennas. Rutgers University, 2002. APA, New Brunswick, NJ (2002)

Download references

Acknowledgements

This research was supported in part by the Israel Science Foundation (136/16) founded by the Israel Academy of Science and the Technion Russell Berrie Nanotechnology Institute. EH is grateful to the 2014 Guthwirth Fellowship and the 2018 Shavit Award in Mechanical Engineering.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to O. Gottlieb.

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.

Appendices

Appendix A: non-dimensional constants

The non-dimensional constants of (12) are:

$$ \begin{aligned} & \tilde{\alpha }_{1} = \frac{{2a^{2} }}{{b^{2} }} ,\tilde{\alpha }_{2} = \frac{{a^{4} }}{{b^{4} }} , \tilde{\rho }_{1} = \frac{{\left| {\tilde{\omega }_{\text{in}} \epsilon_{0} \left( {\epsilon_{r} - 1} \right)} \right|\eta_{0} \hat{I}_{0} }}{{2\rho h\omega_{s}^{2} c_{0} }} \\ & \tilde{\beta }_{1} = \frac{{6\lambda_{\text{in}}^{2} }}{{h^{2} }} , \tilde{\beta }_{2} = \frac{{6\nu \lambda_{\text{in}}^{2} a^{2} }}{{b^{2} h^{2} }} ,\tilde{\beta }_{3} = \frac{{12\left( {1 - \nu } \right)\lambda_{\text{in}}^{2} a^{2} }}{{b^{2} h^{2} }} ,\tilde{\beta }_{4} = \frac{{6\lambda_{\text{in}}^{2} a^{4} }}{{b^{4} h^{2} }} , \\ & \tilde{\gamma }_{1} = \frac{{12\alpha_{T} a^{2} \left( {1 + \nu } \right)}}{{h^{2} }} ,\tilde{\gamma }_{2} = \frac{{a^{2} \tilde{\gamma }_{1} }}{{b^{2} }} , \tilde{\gamma }_{3} = \frac{{h\tilde{\gamma }_{1} }}{{\lambda_{\text{in}} }} ,\tilde{\gamma }_{4} = \frac{{a^{2} h\tilde{\gamma }_{1} }}{{b^{2} \lambda_{\text{in}} }} , \\ \end{aligned} $$
(66)

where \( \hat{I}_{0} = \frac{{\left| {\hat{e}_{0}^{2} } \right|}}{{\eta_{0} }} \) is the laser intensity in W/m2 where \( \eta_{0} = \frac{1}{{\epsilon_{0} c_{0} }} = 376.73\left[ \varOmega \right] \) is free space impedance.

The non-dimensional constants related to the thermal field in (13) are:

$$ \begin{aligned} & \tilde{\eta }_{1} = \frac{k}{{\bar{C}_{p} a^{2} \omega_{s} }} ,\tilde{\eta }_{2} = \frac{k}{{\bar{C}_{p} b^{2} \omega_{s} }} , \tilde{\eta }_{3} = \frac{k}{{\bar{C}_{p} h^{2} \omega_{s} }} , \tilde{\rho }_{2} = \frac{{\text{Re} \left( {\tilde{\sigma }} \right)\eta_{0} \hat{I}_{0} }}{{2\bar{C}_{p} T_{0} \omega_{s} }} \\ & \tilde{\sigma }_{1} = \frac{{E\alpha_{T} \lambda_{\text{in}} h}}{{\bar{C}_{p} \left( {1 - \nu } \right)a^{2} }} ,\tilde{\sigma }_{2} = \frac{{a^{2} \tilde{\sigma }_{1} }}{{b^{2} }} ,\tilde{\sigma }_{3} = \frac{{\lambda_{\text{in}} \tilde{\sigma }_{1} }}{h} , \tilde{\sigma }_{4} = \frac{{\lambda_{\text{in}} a^{2} \tilde{\sigma }_{1} }}{{hb^{2} }} \\ \end{aligned} $$
(67)

where \( \bar{C}_{p} = \rho c_{p} + \frac{{E\alpha_{T}^{2} T_{0} \left( {1 + \nu } \right)}}{{\left( {1 - \nu } \right)\left( {1 - 2\nu } \right)}} \) is an equivalent heat capacity per unit volume at constant stress.

The non-dimensional constants in (22) are:

$$ \chi_{1} = \frac{{c_{0}^{2} }}{{a^{2} \omega_{s}^{2} }} , \chi_{2} = \frac{{c_{0}^{2} }}{{b^{2} \omega_{s}^{2} }} , \chi_{3} = \frac{{c_{0}^{2} }}{{\lambda_{in}^{2} \omega_{s}^{2} }} $$
(68)

Appendix B: mode functions derivation

B.1 Panel mechanical modes

A Ritz method solution to (46)a as expressed in (48) is based on building the solution as a linear combination of one-dimensional modes in (x) and (y) directions with Boundary conditions of S–S and F–F beams, respectively:

$$ \varphi_{l}^{\left( x \right)} = \sin \left( {k_{l}^{\left( x \right)} x} \right) $$
(69)
$$ \begin{aligned} & \varphi_{1}^{\left( y \right)} = 1 , \varphi_{2}^{\left( y \right)} = \sqrt 3 \left( {1 - \frac{2y}{{\alpha_{r} }}} \right) \\ & \varphi_{m \ge 3}^{\left( y \right)} = \sin \left( {k_{m}^{\left( y \right)} y} \right) + \sinh \left( {k_{m}^{\left( y \right)} y} \right) + B_{m} \left( {\cos \left( {k_{m}^{\left( y \right)} y} \right) + \cosh \left( {k_{m}^{\left( y \right)} y} \right)} \right) \\ \end{aligned} $$
(70)

where \( k_{l}^{\left( x \right)} = l\pi \), \( l = 1,2, \ldots ,M \) are the solutions to \( \sin \left( k \right)\sinh \left( k \right) = 0 \); \( k_{1}^{\left( y \right)} = k_{2}^{\left( y \right)} = 0 \) for \( m = 1 \) and \( m = 2 \), \( k_{m}^{\left( y \right)} = 4.73,7.85,10.99, \ldots \) for \( m = 3,4, \ldots ,M \) are the solutions to \( \cos \left( k \right)\cosh \left( k \right) = 1 \) and \( B_{m} = \left( {\cos \left( {k_{m}^{\left( y \right)} y} \right) - \cosh \left( {k_{m}^{\left( y \right)} y} \right)} \right)/\left( {\sin \left( {k_{m}^{\left( y \right)} y} \right) + \sinh \left( {k_{m}^{\left( y \right)} y} \right)} \right) \).

The kinetic (\( T \)) and potential (\( V \)) energies of the panel [66] for a specific Eigen-mode are defined as:

$$ \begin{aligned} & T = \varOmega_{n}^{2} \mathop \int \limits_{ - 1/2}^{1/2} \mathop \int \limits_{ - 1/2}^{1/2} \varphi_{n}^{2} {\text{d}}x{\text{d}}y, \\ & V = \mathop \int \limits_{ - 1/2}^{1/2} \mathop \int \limits_{ - 1/2}^{1/2} \left( {\varphi_{n,xx}^{2} + \alpha_{r}^{4} \varphi_{n,yy}^{2} + 2\nu \alpha_{r}^{2} \varphi_{n,xx} \varphi_{n,yy} + 2\left( {1 - \nu } \right)\alpha_{r}^{2} \varphi_{n,xy}^{2} } \right){\text{d}}x{\text{d}}y . \\ \end{aligned} $$
(71)

From conservation of energy, V and T are equated to yield:

$$ \varOmega_{n}^{2} = \frac{V}{{\mathop \int \nolimits_{ - 1/2}^{1/2} \mathop \int \nolimits_{ - 1/2}^{1/2} \varphi_{n}^{2} {\text{d}}x{\text{d}}y}} $$
(72)

Substituting (48) into (72) and minimizing with respect to \( a_{lm}^{n} \) yield a generalized eigenvalue problem with \( \varOmega_{n} \) and \( a_{lm}^{n} \) as eigenvalues and eigenvectors, respectively:

$$ \mathop \sum \limits_{m = 1}^{M} \mathop \sum \limits_{l = 1}^{M} \left( {C_{lmij} - \varOmega_{n}^{2} B_{lmij}^{{\left( {0,0,0,0} \right)}} } \right)a_{lm}^{n} = 0 $$
(73)

where

$$ B_{lmij}^{{\left( {p,q,s,r} \right)}} = \mathop \int \limits_{ - 1/2}^{1/2} \frac{{{\text{d}}^{p} }}{{{\text{d}}x^{p} }}\left( {\varphi_{l}^{\left( x \right)} } \right)\frac{{{\text{d}}^{q} }}{{{\text{d}}x^{q} }}\left( {\varphi_{i}^{\left( x \right)} } \right){\text{d}}x\mathop \int \limits_{ - 1/2}^{1/2} \frac{{{\text{d}}^{s} }}{{{\text{d}}y^{s} }}\left( {\varphi_{m}^{\left( y \right)} } \right)\frac{{{\text{d}}^{r} }}{{{\text{d}}y^{r} }}\left( {\varphi_{j}^{\left( y \right)} } \right){\text{d}}y, $$
(74)
$$ C_{lmij} = B_{lmij}^{{\left( {2,2,0,0} \right)}} + \alpha_{r}^{2} B_{lmij}^{{\left( {0,0,2,2} \right)}} + \nu \alpha_{r}^{2} \left( {B_{lmij}^{{\left( {2,0,0,2} \right)}} + B_{lmij}^{{\left( {0,2,2,0} \right)}} } \right) + 2\left( {1 - \nu } \right)\alpha_{r}^{2} B_{lmij}^{{\left( {1,1,1,1} \right)}} , $$
(75)

and \( l,m,i,j = 1,2, \ldots M \). Some of the mode shapes of an S–F–S–F panel are showed in Fig. 23.

Fig. 23
figure 23

1st three mechanical modes of an S–F–S–F panel (n = 1 (1, 1), n = 2 (1, 2) and n = 3 (1, 3))

B.2 Panel thermal modes

The constants for the thermal field modes (49) are:

$$ \begin{aligned} & \hat{\lambda }_{lmn} = \sqrt {\left( {\pi^{2} \left( {l^{2} \tilde{\eta }_{1} + m^{2} \tilde{\eta }_{2} } \right) + i\varOmega_{n} } \right)/\tilde{\eta }_{3} } \\ & A_{lmn} = \mathop \int \limits_{ - 1/2}^{1/2} \mathop \int \limits_{ - 1/2}^{1/2} \left( {\tilde{\sigma }_{1} \varphi_{nxx} + \tilde{\sigma }_{2} \varphi_{nyy} } \right)\sin \left( {l\pi \left( {x + 0.5} \right)} \right)\sin \left( {m\pi \left( {y + 0.5} \right)} \right){\text{d}}x{\text{d}}y \\ \end{aligned} $$
(76)

where some of the thermal modes shapes of an S–F–S–F panel are shown in Fig. 24.

Fig. 24
figure 24

Top: 1st three thermal modes (arbitrary \( \zeta \)) of an S–F–S–F panel. Bottom: \( \psi \left( {x,y = 0,\zeta } \right) \) for \( a = b = 40\;\upmu{\text{m}} \) and \( h = 30\;\upmu{\text{m}} \)

B.3 EM field modes

Omitting the Gaussian profile \( \varGamma \left( {x,y} \right) \) in (36), the 1D electric (\( \mathcal{E}_{lm} \)) and magnetic (\( {\mathcal{H}}_{lm} \)) fields modes of the multilayered cavity in Fig. 25 are [67]:

$$ \begin{aligned} \left[ {\begin{array}{*{20}c} {\mathcal{E}_{lm} \left( {z,w} \right)} \\ {{\mathcal{H}}_{lm} \left( {z,w} \right)} \\ \end{array} } \right] & = \left[ {\begin{array}{*{20}c} 1 \\ {\frac{{n_{l} }}{{\eta_{0} }}} \\ \end{array} } \right]E_{lm}^{ + } \left( {w\left( {x,y,t} \right)} \right)e^{{ - ik_{m} n_{l} \left( {z - z_{l} } \right)}} \\ & \quad + \left[ {\begin{array}{*{20}c} 1 \\ { - \frac{{n_{l} }}{{\eta_{0} }}} \\ \end{array} } \right]E_{lm}^{ - } \left( {w\left( {x,y,t} \right)} \right)e^{{ + ik_{m} n_{l} \left( {z - z_{l} } \right)}} , \\ \end{aligned} $$
(77)

where \( m = 0, \pm 1, \pm 2, \ldots \) is mode number, \( l = 0,..,4 \) is layer number and \( z_{l} = a_{l} + b_{l} w \) is the z location at the interface, as depicted in Fig. 25 (\( z_{0} = 0 \), \( z_{1} = \tilde{V}_{1} - \tilde{h}/2 + w \), \( z_{2} = \tilde{V}_{1} + \tilde{h}/2 + w \) and \( z_{3} = \tilde{V}_{2} \)). Note that the EM field modes have only outgoing fields, thus \( E_{0}^{ + } = E_{4}^{ - } = 0 \). The remaining \( E_{lm}^{ \pm } \) terms for \( l = 0,..,3 \) (\( E_{4m}^{ + } = 1 \)) are:

Fig. 25
figure 25

2D sketch of a lossy cavity with panel in the center

$$ \begin{aligned} & \left[ {\begin{array}{*{20}c} 0 \\ {E_{0m}^{ - } } \\ \end{array} } \right] = M_{1} M_{2} M_{3} M_{4} \left[ {\begin{array}{*{20}c} 1 \\ 0 \\ \end{array} } \right] \\ & \left[ {\begin{array}{*{20}c} {E_{1m}^{ + } } \\ {E_{1m}^{ - } } \\ \end{array} } \right] = M_{2} M_{3} M_{4} \left[ {\begin{array}{*{20}c} 1 \\ 0 \\ \end{array} } \right] \\ & \left[ {\begin{array}{*{20}c} {E_{2m}^{ + } } \\ {E_{2m}^{ - } } \\ \end{array} } \right] = M_{3} M_{4} \left[ {\begin{array}{*{20}c} 1 \\ 0 \\ \end{array} } \right] \\ & \left[ {\begin{array}{*{20}c} {E_{3m}^{ + } } \\ {E_{3m}^{ - } } \\ \end{array} } \right] = M_{4} \left[ {\begin{array}{*{20}c} 1 \\ 0 \\ \end{array} } \right] \\ \end{aligned} $$
(78)

where

$$ \begin{aligned} & M_{1} = \left[ {\begin{array}{*{20}c} {\left( {1 - \frac{{ik_{m} \nu_{1} }}{2}} \right)e^{{ik_{m} \left( {\tilde{V}_{1} - \frac{{\tilde{h}}}{2} - w} \right)}} } & {\frac{{ik_{m} \nu_{1} }}{2}e^{{ - ik_{m} \left( {\tilde{V}_{1} - \frac{{\tilde{h}}}{2} - w} \right)}} } \\ { - \frac{{ik_{m} \nu_{1} }}{2}e^{{ik_{m} \left( {\tilde{V}_{1} - \frac{{\tilde{h}}}{2} - w} \right)}} } & {\left( {1 + \frac{{ik_{m} \nu_{1} }}{2}} \right)e^{{ - ik_{m} \left( {\tilde{V}_{1} - \frac{{\tilde{h}}}{2} - w} \right)}} } \\ \end{array} } \right], \\ & M_{2} = \left[ {\begin{array}{*{20}c} {e^{{ik_{m} n_{2} \tilde{h}}} } & {\rho_{2} e^{{ - ik_{m} n_{2} \tilde{h}}} } \\ {\rho_{2} e^{{ik_{m} n_{2} \tilde{h}}} } & {e^{{ - ik_{m} n_{2} \tilde{h}}} } \\ \end{array} } \right], \\ \end{aligned} $$
$$ \begin{aligned} & M_{3} = \left[ {\begin{array}{*{20}c} {e^{{ik_{m} \left( {\tilde{V}_{2} - \tilde{V}_{1} - \frac{{\tilde{h}}}{2} - w} \right)}} } & {\rho_{3} e^{{ - ik_{m} \left( {\tilde{V}_{2} - \tilde{V}_{1} - \frac{h}{2} - w} \right)}} } \\ {\rho_{3} e^{{ik_{m} \left( {\tilde{V}_{2} - \tilde{V}_{1} - \frac{h}{2} - w} \right)}} } & {e^{{ - ik_{m} \left( {\tilde{V}_{2} - \tilde{V}_{1} - \frac{h}{2} - w} \right)}} } \\ \end{array} } \right], \\ & M_{4} = \left[ {\begin{array}{*{20}c} {1 - \frac{{ik_{m} \nu_{4} }}{2}} & {\frac{{ik_{m} \nu_{4} }}{2}} \\ {\frac{{ik_{m} \nu_{4} }}{2}} & {1 + \frac{{ik_{m} \nu_{4} }}{2}} \\ \end{array} } \right]. \\ \end{aligned} $$

The cavity modes frequencies are found by equating the first line in (78)a to zero (\( \left\{ {M_{1} M_{2} M_{3} M_{4} } \right\}_{11} = 0 \)), yielding the transcendental equation of (38).

Substituting \( {\mathbb{E}}_{m} \) into (32) and assuming \( \omega_{s} \ll \tilde{\omega }_{\text{in}} \) yields the expression for the spatio-temporal current density:

$$ {\mathbb{J}}_{m} = {\mathcal{J}}_{m} e_{m} \left( t \right). $$
(79)

where \( {\mathcal{J}}_{m} = \frac{{\left( {\epsilon_{r} - 1} \right)}}{{i\left| {\left( {\epsilon_{r} - 1} \right)} \right|}}\mathcal{E}_{m} \).

To find the spatio-temporal magnetic field \( {\mathbb{H}}_{m} ,{\mathbb{E}}_{m} \) is substituted into (33), the result is multiplied by \( {\hat{\mathcal{E}}}_{m}^{*} \), integrated over the entire volume and rearranged to yield (assuming \( \omega_{s} \ll \tilde{\omega }_{\text{in}} \)) the reduced parameter \( h_{m} \left( t \right) \):

$$ h_{m} \left( t \right) = \frac{2\pi }{i}\frac{{\iiint {\epsilon_{r} \mathcal{E}_{m} {\hat{\mathcal{E}}}_{m}^{*} {\text{d}}x{\text{d}}y{\text{d}}z}}}{{\iiint {\left( {\partial_{z} {\mathcal{H}}_{m} } \right){\hat{\mathcal{E}}}_{m}^{*} {\text{d}}x{\text{d}}y{\text{d}}z}}}e_{m} \left( t \right). $$
(80)

Thus:

$$ {\mathbb{H}}_{m} \left( t \right) = {\mathcal{H}}_{m} h_{m} \left( t \right) = {\hat{\mathcal{H}}}_{m} e_{m} \left( t \right) , $$
(81)

where \( {\hat{\mathcal{H}}}_{m} = \frac{2\pi }{i}{\mathcal{H}}_{m} \frac{{\iiint {\epsilon_{r} \mathcal{E}_{m} {\hat{\mathcal{E}}}_{m}^{*} {\text{d}}x{\text{d}}y{\text{d}}z}}}{{\iiint {\left( {\partial_{z} {\mathcal{H}}_{m} } \right){\hat{\mathcal{E}}}_{m}^{*} {\text{d}}x{\text{d}}y{\text{d}}z}}} \).

Appendix C: reduced model constants and functions

C.1 Parameters related to the EM field

The optical functions in (44) are:

$$ {\mathcal{K}}_{m}^{\left( p \right)} \left( {q_{1} , \ldots ,q_{N} } \right) = \frac{{{\hat{\mathcal{K}}}_{m}^{\left( p \right)} \left( {q_{1} , \ldots ,q_{N} } \right)}}{{{\mathcal{K}}_{m}^{\left( 5 \right)} \left( {q_{1} , \ldots ,q_{N} } \right)}} $$
(82)

where \( p = 1, \ldots ,5 \) and \( {\hat{\mathcal{K}}}_{m}^{\left( i \right)} \left( {q_{1} , \ldots ,q_{N} } \right) \) are expressed through the following integrals:

$$ \begin{aligned} & {\hat{\mathcal{K}}}_{mi}^{\left( 1 \right)} \left( {q_{1, \ldots ,N} } \right) = \iiint {\epsilon_{r} \left( z \right)\left( {\chi_{1} \left( {\varphi_{ix} \mathcal{E}_{m,wx} + \varphi_{ixx} \mathcal{E}_{m,w} } \right) + \chi_{2} \left( {\varphi_{iy} \mathcal{E}_{m,wy} + \varphi_{iyy} \mathcal{E}_{m,w} } \right)} \right){\hat{\mathcal{E}}}_{m}^{*} } \\ & {\hat{\mathcal{K}}}_{mij}^{\left( 2 \right)} \left( {q_{1, \ldots ,N} } \right) = \iiint {\epsilon_{r} \left( z \right)\left( {\chi_{1} \varphi_{ix} \varphi_{jx} + \chi_{2} \varphi_{iy} \varphi_{jy} } \right)\mathcal{E}_{m,ww} {\hat{\mathcal{E}}}_{m}^{*} {\text{d}}^{3} r} \\ & {\hat{\mathcal{K}}}_{mi}^{\left( 3 \right)} \left( {q_{1, \ldots ,N} } \right) = \iiint {\epsilon_{r} \left( z \right)\varphi_{i} \mathcal{E}_{m,w} {\hat{\mathcal{E}}}_{m}^{*} {\text{d}}^{3} r} \\ & {\hat{\mathcal{K}}}_{m}^{\left( 4 \right)} \left( {q_{1, \ldots ,N} } \right) = \iiint {\epsilon_{r} \left( z \right)\left( {\frac{1}{{n_{z}^{2} \omega_{in}^{2} }}\left( {\chi_{1} \mathcal{E}_{in,xx} + \chi_{2} \mathcal{E}_{in,yy} + \chi_{3} \mathcal{E}_{in,zz} } \right) + \mathcal{E}_{in} } \right){\hat{\mathcal{E}}}_{m}^{*} {\text{d}}^{3} r} \\ & {\hat{\mathcal{K}}}_{m}^{\left( 5 \right)} \left( {q_{1, \ldots ,N} } \right) = \iiint {\epsilon_{r} \left( z \right)\delta \omega_{m} \left( {1 + \frac{{\delta \omega_{m} }}{{2\omega_{in} }}} \right)\mathcal{E}_{m} {\hat{\mathcal{E}}}_{m}^{*} {\text{d}}^{3} r} \\ \end{aligned} $$
(83)

Since \( \mathcal{E}_{m} \) are functions of the transverse coordinates \( \left( {x,y} \right) \) and of the spatio-temporal displacement \( w\left( {x,y,t} \right) = \mathop \sum \limits_{i} \varphi_{i} \left( {x,y} \right)q_{i} \left( t \right) \), an analytic integral on \( {\hat{\mathcal{K}}}_{m}^{\left( p \right)} \) is not applicable. However, we can approximate it by using a two-dimensional trapezoidal rule in the \( \left( {x, y} \right) \) axes, then solve the \( z \) axis separately:

$$ \begin{aligned} {\hat{\mathcal{K}}}_{m}^{\left( i \right)} \left( {q_{1, \ldots ,N} } \right) & = \mathop \int \limits_{{0^{ - } }}^{{V_{2}^{ + } }} \left( {\mathop \int \limits_{ - 0.5}^{0.5} \mathop \int \limits_{ - 0.5}^{0.5} \epsilon_{r} \left( z \right){\text{d}}K_{m}^{\left( p \right)} \left( {x,y,z,\mathop \sum \limits_{k = 1}^{N} \varphi_{n} \left( {x,y} \right)q_{n} \left( t \right)} \right){\text{d}}x{\text{d}}y} \right){\text{d}}z \\ & = \frac{1}{4LK}\mathop \sum \limits_{l = - L}^{L} \mathop \sum \limits_{k = - K}^{K} \alpha_{lk} \mathop \int \limits_{{0^{ - } }}^{{V_{2}^{ + } }} \int_{r} \left( z \right){\text{d}}K_{m}^{\left( p \right)} \left( {x_{l} ,y_{l} ,z,\mathop \sum \limits_{k = 1}^{N} \varphi_{n} \left( {x_{l} ,y_{k} } \right)q_{n} \left( t \right)} \right){\text{d}}z \\ \end{aligned} $$
(84)

where \( \epsilon_{r} {\text{d}}K_{m}^{\left( p \right)} \) is any one of the integrands in (83), \( \left( {x_{l} ,y_{k} } \right) \) are the transverse integration points, \( \alpha_{lk} \) are weighting parameters (\( \alpha_{ \pm L, \pm K} = 1 \) or \( \alpha_{l, \pm K} = \alpha_{ \pm L,k} = 2 \) at the edges and \( \alpha_{lk} = 4 \) at the inner points) and (2L + 1,2 K + 1) are the number of integration points in each axis. Note that the integration in z must be analytical or at least solved by a fast-converging numerical scheme in order for it to be practical to use in each iteration of a FDTD solver.

A simplification of the above integrals is done by assuming the integrand \( {\text{d}}K_{m}^{\left( p \right)} \) can be separated to a multiplication of functions with different variables, i.e. \( {\text{d}}K_{m}^{\left( p \right)} = {\text{d}}K_{xy} \left( {x,y} \right){\text{d}}K_{z} \left( {z,w\left( {x,y,t} \right)} \right) \). The integrand \( {\text{d}}K_{xy} \) is assumed to be integrable and is solved outside the FDTD scheme and \( {\text{d}}K_{z} \) is assumed to have weak dependence on \( w\left( {x,y,t} \right) \) and is replaced with \( {\text{d}}K_{z} \left( {z,w\left( {x_{0} ,y_{0} ,t} \right)} \right) \). Thus, \( {\hat{\mathcal{K}}}_{m}^{\left( p \right)} \) are simplified to:

$$ {\hat{\mathcal{K}}}_{m}^{\left( p \right)} \left( {q_{1, \ldots ,N} } \right) = \mathop \int \limits_{ - 0.5}^{0.5} \mathop \int \limits_{ - 0.5}^{0.5} {\text{d}}K_{xy} \left( {x,y} \right){\text{d}}x{\text{d}}y\mathop \int \limits_{{0^{ - } }}^{{V_{2}^{ + } }} \left( {\epsilon_{r} \left( z \right){\text{d}}K_{z} \left( {z,\mathop \sum \limits_{k = 1}^{N} \varphi_{n} \left( {x_{0} ,y_{0} } \right)q_{n} \left( t \right)} \right)} \right){\text{d}}z. $$
(85)

C.2 Parameters related to the elastic field

The elastic field (50) constants are:

$$ \varOmega_{n}^{2} = {\iint }\left( {\varphi_{n,xxxx} + \tilde{\alpha }_{1} \varphi_{n,xxyy} + \tilde{\alpha }_{2} \varphi_{n,yyyy} } \right)\varphi_{n} {\text{d}}x{\text{d}}y , $$
(86)
$$ \begin{aligned} \beta_{nijk} & = - {\iint }\left( {\left( {\tilde{\beta }_{1} \varphi_{i,x} \varphi_{j,x} + \tilde{\beta }_{2} \varphi_{i,y} \varphi_{j,y} } \right) + \tilde{\beta }_{3} \varphi_{i,x} \varphi_{j,y} \varphi_{k,xy} } \right. \\ & \quad \left. { + \;\left( {\tilde{\beta }_{2} \varphi_{i,x} \varphi_{j,x} + \tilde{\beta }_{4} \varphi_{i,y} \varphi_{j,y} } \right)} \right)\varphi_{n} {\text{d}}x{\text{d}}y, \\ \end{aligned} $$
(87)
$$ \begin{aligned} & \gamma_{n}^{\left( 1 \right)} = \iiint {\zeta \left( {\tilde{\gamma }_{3} \psi_{n,xx} + \tilde{\gamma }_{4} \psi_{n,yy} } \right)\varphi_{n} {\text{d}}x{\text{d}}y{\text{d}}\zeta ,} \\ & \gamma_{nij}^{\left( 2 \right)} = \iiint {\left( {\tilde{\gamma }_{1} \varphi_{i,xx} + \tilde{\gamma }_{2} \varphi_{i,yy} } \right)\psi_{j} \varphi_{n} {\text{d}}x{\text{d}}y{\text{d}}\zeta .} \\ \end{aligned} $$
(88)

The spatially averaged variable radiation pressure on the panel is:

$$ {\mathcal{F}}_{nkl} \left( {q_{1, \ldots ,N} } \right) = \mathop \int \limits_{ - 1/2}^{1/2} \mathop \int \limits_{ - 1/2}^{1/2} \varphi_{n} \mathop \int \limits_{{\tilde{V}_{1} - \frac{{\tilde{h}}}{2} + w}}^{{\tilde{V}_{1} + \frac{{\tilde{h}}}{2} + w}} {\mathcal{J}}_{k} {\hat{\mathcal{H}}}_{l}^{*} {\text{d}}x{\text{d}}y{\text{d}}z , $$
(89)

where \( {\mathcal{J}}_{m} \left( {q_{1, \ldots ,N} } \right) \) and \( {\hat{\mathcal{H}}}_{m} \left( {q_{1, \ldots ,N} } \right) \) are derived in “Appendix B”.

C.3 Parameters related to the thermal field

The thermal field (51) constants are:

$$ \eta_{n} = \iiint {\left( {\tilde{\eta }_{1} \psi_{n,xx} + \tilde{\eta }_{2} \psi_{n,yy} + \tilde{\eta }_{3} \psi_{n,\zeta \zeta } } \right)\psi_{n} {\text{d}}x{\text{d}}y{\text{d}}\zeta } $$
(90)
$$ \begin{aligned} & \sigma_{n}^{\left( 1 \right)} = \iiint {\zeta \left( {\tilde{\sigma }_{1} \varphi_{n,xx} + \tilde{\sigma }_{2} \varphi_{n,yy} } \right)\psi_{n} {\text{d}}x{\text{d}}y{\text{d}}\zeta } \\ & \sigma_{nij}^{\left( 2 \right)} = \iiint {\left( {\tilde{\sigma }_{3} \varphi_{i,x} \varphi_{j,x} + \tilde{\sigma }_{4} \varphi_{i,y} \varphi_{j,y} } \right)\psi_{n} {\text{d}}x{\text{d}}y{\text{d}}\zeta } \\ & \sigma_{nij}^{\left( 3 \right)} = \iiint {\zeta \psi_{i} \left( {\tilde{\sigma }_{1} \varphi_{j,xx} + \tilde{\sigma }_{2} \varphi_{j,yy} } \right)\psi_{n} {\text{d}}x{\text{d}}y{\text{d}}\zeta } \\ & \sigma_{nijk}^{\left( 2 \right)} = - \iiint {\psi_{i} \left( {\tilde{\sigma }_{3} \varphi_{j,x} \varphi_{k,x} + \tilde{\sigma }_{4} \varphi_{j,y} \varphi_{k,y} } \right)\psi_{n} {\text{d}}x{\text{d}}y{\text{d}}\zeta } \\ \end{aligned} $$
(91)

The spatially averaged variable heat function is:

$$ {\mathcal{Q}}_{nkl} \left( {q_{1, \ldots ,N} } \right) = \mathop \int \limits_{ - 1/2}^{1/2} \mathop \int \limits_{ - 1/2}^{1/2} \mathop \int \limits_{ - 1/2}^{1/2} \psi_{n}^{*} \mathcal{E}_{k} \mathcal{E}_{l}^{*} {\text{d}}x{\text{d}}y{\text{d}}\zeta . $$
(92)

C.4 Parameters related to the limiting adiabatic system

The linear and nonlinear damping coefficients in (58) are:

$$ \begin{aligned} & \delta_{01}^{\left( T \right)} = - \frac{{\gamma_{1} \sigma_{1} }}{\eta } ,\delta_{11}^{\left( T \right)} = - \frac{{\gamma_{2} \sigma_{1} + \gamma_{1} \sigma_{2} }}{\eta } , \delta_{21}^{\left( T \right)} = - \frac{{\gamma_{2} \sigma_{2} }}{\eta } \\ & \delta_{02}^{\left( T \right)} = - \frac{{\gamma_{1} \sigma_{1} \sigma_{3} }}{{\eta^{2} }} , \delta_{12}^{\left( T \right)} = - \frac{{\gamma_{1} \sigma_{1} \sigma_{4} }}{{\eta^{2} }} , \delta_{03}^{\left( T \right)} = - \frac{{\gamma_{1} \sigma_{1} \sigma_{3}^{2} }}{{\eta^{3} }} \\ \end{aligned} $$
(93)
$$ \delta_{ij}^{\left( W \right)} = - \left. {\frac{{\partial^{i + j} \Delta {\mathcal{W}}}}{{\partial x_{1}^{i} \partial x_{2}^{j} }}} \right|_{{\left( {0,0} \right)}} $$
(94)

where \( x_{1} = q \) and \( x_{2} = \dot{q} \).

Appendix D: stability analysis definitions

The coefficients of the characteristic polynomial related to (61) are:

$$ \begin{aligned} c_{0} & = \left( {\eta \rho_{1} {\mathcal{F}}_{e} + \tilde{\gamma }_{e} \rho_{2} {\mathcal{Q}}_{e} } \right)\left( {\left( {G_{q} r_{e} + V_{q} s_{e} } \right)\kappa_{e} + 2\left( {G_{q} s_{e} - V_{q} r_{e} } \right)\left( {\Delta \omega_{0} + \Delta_{e} } \right)} \right) \\ & \quad + \;\left( {\rho_{2} \tilde{\gamma }_{e} \left( {r_{e}^{2} + s_{e}^{2} } \right){\mathcal{Q}}_{e} - \omega_{q}^{2} \eta } \right)\omega_{r}^{2} \\ c_{1} & = \left( {\omega_{q}^{2} + \tilde{\sigma }_{e} \tilde{\gamma }_{e} - \delta \eta } \right)\omega_{r}^{2} + 2\left( {G_{q} r_{e} + V_{q} s_{e} } \right)\left( {\left( {\eta - \frac{{\kappa_{e} }}{2}} \right)\rho_{1} {\mathcal{F}}_{e} + \tilde{\gamma }_{e} \rho_{2} {\mathcal{Q}}_{e} } \right) \\ & \quad + \;\left( {\rho_{2} \tilde{\gamma }_{e} \left( {r_{e}^{2} + s_{e}^{2} } \right){\mathcal{Q}}_{e} - \omega_{q}^{2} \eta } \right)\kappa_{e} - 2\left( {\Delta \omega_{0} + \Delta_{e} } \right)\rho_{1} {\mathcal{F}}_{e} \left( {G_{q} s_{e} - V_{q} r_{e} } \right) \\ c_{2} & = \left( {\delta - \eta } \right)\omega_{r}^{2} + \left( {\omega_{q}^{2} + \tilde{\sigma }_{e} \tilde{\gamma }_{e} - \delta \eta } \right)\kappa_{e} + \rho_{2} \tilde{\gamma }_{e} \left( {r_{e}^{2} + s_{e}^{2} } \right){\mathcal{Q}}_{e} \\ & \quad - \;\eta \omega_{q}^{2} - 2\rho_{1} {\mathcal{F}}_{e} \left( {G_{q} r_{e} + V_{q} s_{e} } \right) \\ c_{3} & = \omega_{r}^{2} + \left( {\delta - \eta } \right)\kappa_{e} + \omega_{q}^{2} + \tilde{\sigma }_{e} \tilde{\gamma }_{e} - \delta \eta \\ c_{4} & = \kappa_{e} + \delta - \eta \\ \end{aligned} $$
(95)

For an fifth-order dynamical system (\( {\dot{\mathbf{x}}} = {\mathbf{F}}\left( {\mathbf{x}} \right) ;{\mathbf{x}} \in {\mathbb{R}}^{5} \)), the resultant (\( {\mathcal{R}} \)) of the characteristic polynomial \( \mathop \sum \nolimits_{n = 0}^{5} c_{n} \lambda^{n} \) is defined by (Sylvester form) [51]:

$$ {\mathcal{R}} = \left| {\begin{array}{*{20}c} {c_{0} } & {c_{2} } & {c_{4} } & 0 \\ 0 & {c_{0} } & {c_{2} } & {c_{4} } \\ {c_{1} } & {c_{3} } & 1 & 0 \\ 0 & {c_{1} } & {c_{3} } & 1 \\ \end{array} } \right| $$
(96)

In addition, the sub-resultants \( {\mathcal{S}}_{i} \) (i = 1, 2) for a fifth-order system are:

$$ {\mathcal{S}}_{1} = \left| {\begin{array}{*{20}c} {c_{2} } & {c_{4} } \\ {c_{3} } & 1 \\ \end{array} } \right| , {\mathcal{S}}_{2} = \left| {\begin{array}{*{20}c} {c_{0} } & {c_{4} } \\ {c_{1} } & 1 \\ \end{array} } \right| $$
(97)

Thus, the sub-resultant criterion is defined by:

$$ \fancyscript{h} = {\mathcal{S}}_{1} \cdot {\mathcal{S}}_{2} $$
(98)

According to theory, codimension-one bifurcation points are found for \( {\mathcal{R}} = 0 \). Higher-order bifurcations are determined by the sign of \( \fancyscript{h} \), as explained in the main text.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hollander, E., Gottlieb, O. Global bifurcations and homoclinic chaos in nonlinear panel optomechanical resonators under combined thermal and radiation stresses. Nonlinear Dyn 103, 3371–3405 (2021). https://doi.org/10.1007/s11071-020-05977-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-020-05977-w

Keywords

Navigation