Skip to main content
Log in

Wave propagation properties of one-dimensional acoustic metamaterials with nonlinear diatomic microstructure

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

Abstract

Acoustic metamaterials are artificial microstructured media, typically characterized by a periodic locally resonant cell. The cellular microstructure can be functionally customized to govern the propagation of elastic waves. A one-dimensional diatomic lattice with cubic inter-atomic coupling—described by a Lagrangian model—is assumed as minimal mechanical system simulating the essential undamped dynamics of nonlinear acoustic metamaterials. The linear dispersion properties are analytically determined by solving the linearized eigenproblem governing the free wave propagation in the small-amplitude oscillation range. The dispersion spectrum is composed by a low-frequency acoustic branch and a high-frequency optical branch. The two frequency branches are systematically separated by a stop band, whose amplitude is analytically derived. Superharmonic 3:1 internal resonances can occur within a wavenumber-dependent locus defined in the mechanical parameter space. A general asymptotic approach, based on the multiple scale method, is employed to determine the nonlinear dispersion properties. Accordingly, the nonlinear frequencies and waveforms are obtained for the two fundamental cases of non-resonant and superharmonically 3:1 resonant or nearly resonant lattices. Moreover, the invariant manifolds associated with the nonlinear waveforms are parametrically determined in the space of the two principal coordinates. Finally, some examples of non-resonant and resonant lattices are selected to discuss their nonlinear dispersion properties from a qualitative and quantitative viewpoint.

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

Similar content being viewed by others

References

  1. Bacigalupo, A., Lepidi, M.: High-frequency parametric approximation of the Floquet-Bloch spectrum for anti-tetrachiral materials. Int. J. Solids Struct. 97–98, 575–592 (2016)

    Google Scholar 

  2. Bacigalupo, A., Lepidi, M.: Acoustic wave polarization and energy flow in periodic beam lattice materials. Int. J. Solids Struct. 147, 183–203 (2018)

    Google Scholar 

  3. Bacigalupo, A., Lepidi, M., Gnecco, G., Gambarotta, L.: Optimal design of auxetic hexachiral metamaterials with local resonators. Smart Mater. Struct. 25(5), 054,009 (2016)

    Google Scholar 

  4. Balmes, E.: High modal density, curve veering, localization—a different perspective on the structural response. J. Sound Vib. 161, 358–363 (1993)

    Google Scholar 

  5. Bigoni, D., Guenneau, S., Movchan, A., Brun, M.: Elastic metamaterials with inertial locally resonant structures: application to lensing and localization. Phys. Rev. B 87(17), 174 303 (2013)

    Google Scholar 

  6. Brillouin, L.: Wave Propagation in Periodic Structures: Electric Filters and Crystal Lattices. Dover Publications, New York (1946)

    MATH  Google Scholar 

  7. Colombi, A., Colquitt, D., Roux, P., Guenneau, S., Craster, R.V.: A seismic metamaterial: the resonant metawedge. Sci. Rep. 6(27), 717 (2016)

    Google Scholar 

  8. Craster, R.V., Guenneau, S.: Acoustic Metamaterials: Negative Refraction, Imaging, Lensing and Cloaking, vol. 166. Springer, Heidelberg (2012)

    Google Scholar 

  9. Cummer, S.A., Christensen, J., Alù, A.: Controlling sound with acoustic metamaterials. Nat. Rev. Mater. 1(3), 16,001 (2016)

    Google Scholar 

  10. Deymier, P.A.: Acoustic Metamaterials and Phononic Crystals, vol. 173. Springer, Berlin (2013)

    Google Scholar 

  11. Diaz, A., Haddow, A., Ma, L.: Design of band-gap grid structures. Struct. Multidiscip. Optim. 29(6), 418–431 (2005)

    Google Scholar 

  12. Fang, X., Wen, J., Yin, J., Yu, D.: Wave propagation in nonlinear metamaterial multi-atomic chains based on homotopy method. AIP Adv. 6(12), 121,706 (2016)

    Google Scholar 

  13. Fang, X., Wen, J., Yin, J., Yu, D., Xiao, Y.: Broadband and tunable one-dimensional strongly nonlinear acoustic metamaterials: theoretical study. Phys. Rev. E 94(5), 052,206 (2016)

    Google Scholar 

  14. Gattulli, V., Lepidi, M.: Localization and veering in the dynamics of cable-stayed bridges. Comput. Struct. 85(21–22), 1661–1678 (2007)

    Google Scholar 

  15. Georgiou, I.T., Vakakis, A.F.: An invariant manifold approach for studying waves in a one-dimensional array of non-linear oscillators. Int. J. Non-linear Mech. 31(6), 871–886 (1996)

    MathSciNet  MATH  Google Scholar 

  16. Guenneau, S., Movchan, A., Pétursson, G., Ramakrishna, S.A.: Acoustic metamaterials for sound focusing and confinement. New J. Phys. 9(11), 399 (2007)

    Google Scholar 

  17. Huang, H., Sun, C., Huang, G.: On the negative effective mass density in acoustic metamaterials. Int. J. Eng. Sci. 47(4), 610–617 (2009)

    Google Scholar 

  18. King, M., Vakakis, A.: An energy-based approach to computing resonant nonlinear normal modes. J. Appl. Mech. 63(3), 810–819 (1996)

    MathSciNet  MATH  Google Scholar 

  19. Kovacic, I., Brennan, M.J.: The Duffing Equation: Nonlinear Oscillators and Their Behaviour. Wiley, Hoboken (2011)

    MATH  Google Scholar 

  20. Lacarbonara, W., Camillacci, R.: Nonlinear normal modes of structural systems via asymptotic approach. Int. J. Solids Struct. 41(20), 5565–5594 (2004)

    MATH  Google Scholar 

  21. Lacarbonara, W., Rega, G.: Resonant non-linear normal modes. Part ii: activation/orthogonality conditions for shallow structural systems. Int. J. Non-Linear Mech. 38(6), 873–887 (2003)

    MATH  Google Scholar 

  22. Lacarbonara, W., Rega, G., Nayfeh, A.: Resonant non-linear normal modes. Part i: analytical treatment for structural one-dimensional systems. Int. J. Non-Linear Mech. 38(6), 851–872 (2003)

    MATH  Google Scholar 

  23. Lacarbonara, W., Arafat, H.N., Nayfeh, A.H.: Non-linear interactions in imperfect beams at veering. Int. J. Non-Linear Mech. 40(7), 987–1003 (2005)

    MATH  Google Scholar 

  24. Lazarov, B.S., Jensen, J.S.: Low-frequency band gaps in chains with attached non-linear oscillators. Int. J. Non-Linear Mech. 42(10), 1186–1193 (2007)

    Google Scholar 

  25. Lepidi, M., Bacigalupo, A.: Multi-parametric sensitivity analysis of the band structure for tetrachiral acoustic metamaterials. Int. J. Solids Struct. 136–137, 186–202 (2018)

    Google Scholar 

  26. Lepidi, M., Bacigalupo, A.: Parametric design of the band structure for lattice materials. Meccanica 53(3), 613–628 (2018)

    MathSciNet  Google Scholar 

  27. Liu, Z., Zhang, X., Mao, Y., Zhu, Y., Yang, Z., Chan, C.T., Sheng, P.: Locally resonant sonic materials. Science 289(5485), 1734–1736 (2000)

    Google Scholar 

  28. Lu, M.H., Feng, L., Chen, Y.F.: Phononic crystals and acoustic metamaterials. Mater. Today 12(12), 34–42 (2009)

    Google Scholar 

  29. Luongo, A., Romeo, F.: A transfer-matrix-perturbation approach to the dynamics of chains of nonlinear sliding beams. J. Vib. Acoust. 128(2), 190–196 (2006)

    Google Scholar 

  30. Ma, G., Sheng, P.: Acoustic metamaterials: from local resonances to broad horizons. Sci. Adv. 2(2), e1501,595 (2016)

    MathSciNet  Google Scholar 

  31. Mace, B.R., Manconi, E.: Wave motion and dispersion phenomena: veering, locking and strong coupling effects. J. Acoust. Soc. Am. 131(2), 1015–1028 (2012)

    Google Scholar 

  32. Maldovan, M.: Sound and heat revolutions in phononics. Nature 503(7475), 209 (2013)

    Google Scholar 

  33. Manimala, J.M., Sun, C.: Microstructural design studies for locally dissipative acoustic metamaterials. J. Appl. Phys. 115(2), 023,518 (2014)

    Google Scholar 

  34. Manktelow, K., Leamy, M.J., Ruzzene, M.: Multiple scales analysis of wave-wave interactions in a cubically nonlinear monoatomic chain. Nonlinear Dyn. 63(1–2), 193–203 (2011)

    MathSciNet  MATH  Google Scholar 

  35. Matlack, K.H., Serra-Garcia, M., Palermo, A., Huber, S.D., Daraio, C.: Designing perturbative metamaterials from discrete models. Nat. Mater. 17(4), 323 (2018)

    Google Scholar 

  36. Narisetti, R.K., Leamy, M.J., Ruzzene, M.: A perturbation approach for predicting wave propagation in one-dimensional nonlinear periodic structures. J. Vib. Acoust. 132(3), 031,001 (2010)

    Google Scholar 

  37. Narisetti, R.K., Ruzzene, M., Leamy, M.J.: A perturbation approach for analyzing dispersion and group velocities in two-dimensional nonlinear periodic lattices. J. Vib. Acoust. 133(6), 061,020 (2011)

    Google Scholar 

  38. Natsiavas, S.: Mode localization and frequency veering in a non-conservative mechanical system with dissimilar components. J. Sound Vib. 165(1), 137–147 (1993)

    MathSciNet  MATH  Google Scholar 

  39. Nayfeh, A., Chin, C., Nayfeh, S.: On nonlinear normal modes of systems with internal resonance. J. Vib. Acoust. 118(3), 340–345 (1996)

    Google Scholar 

  40. Nayfeh, A.H., Lacarbonara, W., Chin, C.M.: Nonlinear normal modes of buckled beams: three-to-one and one-to-one internal resonances. Nonlinear Dyn. 18(3), 253–273 (1999)

    MathSciNet  MATH  Google Scholar 

  41. Perkins, N., Mote Jr., C.: Comments on curve veering in eigenvalue problems. J. Sound Vib. 106(3), 451–463 (1986)

    Google Scholar 

  42. Pierre, C.: Mode localization and eigenvalue loci veering phenomena in disordered structures. J. Sound Vib. 126(3), 485–502 (1988)

    MathSciNet  Google Scholar 

  43. Reda, H., Karathanasopoulos, N., Ganghoffer, J., Lakiss, H.: Wave propagation characteristics of periodic structures accounting for the effect of their higher order inner material kinematics. J. Sound Vib. 431, 265–275 (2018)

    Google Scholar 

  44. Romeo, F., Rega, G.: Wave propagation properties in oscillatory chains with cubic nonlinearities via nonlinear map approach. Chaos Solitons Fractals 27(3), 606–617 (2006)

    MathSciNet  MATH  Google Scholar 

  45. Romeo, F., Rega, G.: Propagation properties of bi-coupled nonlinear oscillatory chains: analytical prediction and numerical validation. Int. J. Bifurc. Chaos 18(07), 1983–1998 (2008)

    MathSciNet  Google Scholar 

  46. Romeo, F., Rega, G.: Periodic and localized solutions in chains of oscillators with softening or hardening cubic nonlinearity. Meccanica 50(3), 721–730 (2015)

    MathSciNet  Google Scholar 

  47. Rosenberg, R.: On nonlinear vibrations of systems with many degrees of freedom. In: Advances in Applied Mechanics, vol. 9, Elsevier, pp. 155–242 (1966)

  48. Rothos, V., Vakakis, A.: Dynamic interactions of traveling waves propagating in a linear chain with an local essentially nonlinear attachment. Wave Mot. 46(3), 174–188 (2009)

    MathSciNet  MATH  Google Scholar 

  49. Shaw, S., Pierre, C.: Non-linear normal modes and invariant manifolds. J. Sound Vib. 150(1), 170–173 (1991)

    Google Scholar 

  50. Shaw, S., Pierre, C.: Normal modes for non-linear vibratory systems. J. Sound Vib. 164(1), 85–124 (1993)

    MATH  Google Scholar 

  51. Triantafyllou, M., Triantafyllou, G.: Frequency coalescence and mode localization phenomena: a geometric theory. J. Sound Vib. 150(3), 485–500 (1991)

    MathSciNet  Google Scholar 

  52. Vakakis, A.F., King, M.E.: Nonlinear wave transmission in a monocoupled elastic periodic system. J. Acoust. Soc. Am. 98(3), 1534–1546 (1995)

    Google Scholar 

  53. Vakakis, A.F., King, M.E.: Resonant oscillations of a weakly coupled, nonlinear layered system. Acta Mech. 128(1–2), 59–80 (1998)

    MathSciNet  MATH  Google Scholar 

  54. Zhang, S., Yin, L., Fang, N.: Focusing ultrasound with an acoustic metamaterial network. Phys. Rev. Lett. 102(19), 194,301 (2009)

    Google Scholar 

Download references

Acknowledgements

The authors acknowledge financial support of the (MURST) Italian Department for University and Scientific and Technological Research in the framework of the research MIUR Prin15 Project 2015LYYXA8, “Multi-scale mechanical models for the design and optimization of micro-structured smart materials and metamaterials”, coordinated by prof. A. Corigliano. The authors also acknowledge financial support by National Group of Mathematical Physics (GNFM-INdAM).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Marco Lepidi.

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: Model formulation

1.1 Hamilton’s principle

The kinetic energy for the one-dimensional lattice characterized by the diatomic periodic cell reads

$$\begin{aligned} {\mathcal {K}}=\tfrac{1}{2}M({\dot{U}})^2+\tfrac{1}{2}M_r({\dot{V}})^2 \end{aligned}$$
(67)

whereas the elastic potential energy takes the form

$$\begin{aligned} {\mathcal {U}}=\tfrac{1}{2}K(U-U_2)^2+\tfrac{1}{2}K(U_3-U)^2+K_r\varepsilon _r^2 \end{aligned}$$
(68)

where—according to a second-order kinematics—the secondary spring extension reads

$$\begin{aligned} \varepsilon _r=\varepsilon ^\circ _r+\frac{(V-U)^2}{2R} \end{aligned}$$
(69)

and \(\varepsilon ^\circ _r=H/K_r\) is the spring preextension responsible for the spring pretension.

Introducing the work done by the external forces \({\mathcal {W}}=F_2U_2+F_3U_3\), the Hamiltonian action between two instants of time \(t_1\) and \(t_2\) takes the integral expression

$$\begin{aligned} {\mathcal {H}}=\int _{t_1}^{t_2}\left( {\mathcal {K}}-{\mathcal {V}}\right) \hbox {d}t \end{aligned}$$
(70)

where \({\mathcal {V}}={\mathcal {U}}-{\mathcal {W}}\) is the total potential energy. According to the Hamilton’s Principle, the Hamiltonian action can be imposed to be stationary to obtain the nonlinear equations of motion (3)-(4) and (5), where nondimensional variables and parameters have been introduced.

1.2 Matrices and coefficients of the Lagrangian model

The 2-by-2 mass and stiffness matrices governing the equation of motion (10) are

$$\begin{aligned} {\mathbf {M}}= \left[ \begin{array}{cc} 1+\varrho ^2 &{} \varrho ^2\\ \varrho ^2 &{} \varrho ^2 \end{array}\right] ,\quad {\mathbf {K}}= \left[ \begin{array}{cc} 1-\cos \beta &{} 0\\ 0 &{} 2\mu \end{array}\right] \end{aligned}$$
(71)

whereas the 2-by-1 operator accounting for the cubic nonlinearities reads

$$\begin{aligned} {\mathbf {n}}({\mathbf {u}},{\mathbf {u}},{\mathbf {u}})=\left( \begin{array}{c} 0 \\ \eta \, w^3 \end{array}\right) \end{aligned}$$
(72)

The transformed 2-by-1 operator accounting for the cubic nonlinearities reads

$$\begin{aligned} {\mathbf {c}}({\mathbf {q}},{\mathbf {q}},{\mathbf {q}})=-\eta \,{\varvec{\varphi }}_2\left( {\varvec{\varphi }}_2^\top {\mathbf {q}}\right) \left( {\varvec{\varphi }}_2^\top {\mathbf {q}}\right) \left( {\varvec{\varphi }}_2^\top {\mathbf {q}}\right) \end{aligned}$$
(73)

where \({\varvec{\varphi }}_2=\left( {\varphi }^-_2,{\varphi }^+_2\right) \) is the vector collecting the second components of the mass-normalized waveforms

$$\begin{aligned} {\varphi }^-_2=\frac{2}{D^-_2},\qquad {\varphi }^+_2=\frac{2}{D^+_2} \end{aligned}$$
(74)

where the denominators are

$$\begin{aligned} D^\mp _2\!=&\!\Big [1\!+\!\varrho ^2-2\big (1-\varrho ^2\big )I_2^\mp (\beta )+\nonumber \\&+\,\big (1\!+\!\varrho ^2\big )(I_2^\mp (\beta ))^2\Big ]^{\!\tfrac{1}{2}} \end{aligned}$$
(75)

and the \(\beta \)-dependent auxiliary parameters

$$\begin{aligned} I^\mp _2(\beta )=\frac{2\mu \big (1+\varrho ^2\big )\pm S(\beta )}{\varrho ^2(1-\cos \beta )} \end{aligned}$$
(76)

The components of the nonlinearity vector \({\mathbf {c}}({\mathbf {q}},{\mathbf {q}},{\mathbf {q}})=\left( {c}^-({\mathbf {q}},{\mathbf {q}},{\mathbf {q}}),{c}^+({\mathbf {q}},{\mathbf {q}},{\mathbf {q}})\right) \) are complete cubic polinomials

$$\begin{aligned}&{c}^-({\mathbf {q}},{\mathbf {q}},{\mathbf {q}})=c_{111}(q^-)^3+c_{112}(q^-)^2q^++\\&\qquad \qquad \qquad +\,c_{122}q^-(q^+)^2+c_{222}(q^+)^3\nonumber \\&{c}^+({\mathbf {q}},{\mathbf {q}},{\mathbf {q}})=d_{111}(q^-)^3+d_{112}(q^-)^2q^++\nonumber \\&\qquad \qquad \qquad +\,d_{122}q^-(q^+)^2+d_{222}(q^+)^3\nonumber \end{aligned}$$
(77)

where the coefficients are

$$\begin{aligned}&c_{111}=-\eta (\varphi ^-_2)^4,&c_{112}=-3\eta (\varphi ^-_2)^3(\varphi ^+_2)\\&c_{122}=-3\eta (\varphi ^-_2)^2(\varphi ^+_2)^2,&c_{222}=-\eta (\varphi ^-_2)(\varphi ^+_2)^3\nonumber \\&d_{111}=-\eta (\varphi ^-_2)^3(\varphi ^+_2),&d_{112}=-3\eta (\varphi ^-_2)^2(\varphi ^+_2)^2\nonumber \\&d_{122}=-3\eta (\varphi ^-_2)(\varphi ^+_2)^3,&d_{222}=-\eta (\varphi ^+_2)^4\nonumber \end{aligned}$$
(78)

1.3 Maxima of the resonant locus

The analytical expressions for the \(\beta \)-dependent maxima of the parametric locus \(\mathcal {S}\) portrayed in Fig. 4 are

\(\bullet \):

for the \(\varrho ^2\)-maxima

$$\begin{aligned} \varrho ^2&=\frac{16}{15129}\!\left( 400\sqrt{2(1-\cos \beta )}\csc \frac{\beta }{2}+881\!\right) \\ \mu&=\frac{8}{25}(1-\cos \beta )\nonumber \end{aligned}$$
(79)
\(\bullet \):

for the \(\mu \)-maxima

$$\begin{aligned} \varrho ^2&=\frac{16}{25}\\ \mu&=\frac{8}{9}(1-\cos \beta )\nonumber \end{aligned}$$
(80)

and for \(\beta =\pi \) correspond to the mechanical parameter combinations \((\varrho ^2_\mathrm {max},\mu )=(16/9,16/25)\) and \((\varrho ^2,\mu _\mathrm {max})=(16/25,16/9)\) respectively.

Appendix: Alternative approach

The asymptotic strategy proposed in Sect. 2 requires the enforcement of the quasi-periodicity conditions on the nonlinear equations of motion (3), (4) governing the free vibration dynamics of the diatomic cell. Differently, the common procedure requires to expand the nonlinear equations according to the multiple scale method and, subsequently, to enforce the quasi-periodicity conditions in the linear equations governing each order of the asymptotic expansion [15, 37]. In order to demonstrate the substantial equivalence of the two approaches, the latter procedure can be demonstrated to originate the same formal hierarchy of differential linear problems. To this purpose, the nonlinear equations (3)–(5) can be partitioned in the form

$$\begin{aligned} \left[ {\begin{array}{cc} {\mathbf{M}}&{}{\mathbf{O}}\\ {\mathbf{O}}&{}{\mathbf{O}} \end{array}} \right] \!\! \left( \!{\begin{array}{c} {\ddot{\mathbf{u}}_a}\\ {\ddot{\mathbf{u}}_p} \end{array}}\!\right) \! +\!\left[ {\begin{array}{cc} {{{\mathbf{K}}_{aa}^{n}}}&{}{{{\mathbf{K}}_{ap}}}\\ {{{\mathbf{K}}_{pa}}}&{}{{{\mathbf{K}}_{pp}}} \end{array}} \right] \!\! \left( \!{\begin{array}{c} {{{\mathbf{u}}_a}}\\ {{{\mathbf{u}}_p}} \end{array}}\!\right) \!=\! \left( \!{\begin{array}{c} {{{\mathbf{0}}}}\\ {{{\mathbf{f}}_p}} \end{array}}\!\right) \end{aligned}$$
(81)

where the active displacement subvector \({\mathbf{u}}_a=(u,w)\) has been distinguished from the passive displacement subvector \({\mathbf{u}}_p=(u_2,u_3)\) and the passive force subvector \({\mathbf{f}}_p=(f_2,f_3)\). The submatrix \({\mathbf{K}}_{aa}^n={\mathbf{K}}_{aa}+{\mathbf{N}}({\mathbf{u}}_a)\), where \({\mathbf{N}}({\mathbf{u}}_a)\) accounts for the nonlinearities. Expanding all the variables in series of the small \(\epsilon \)-parameter

$$\begin{aligned} {{\mathbf{u}}_a}=&\,\epsilon {{\mathbf{u}}_{a1}}({T_0},{T_1},{T_2})+{\epsilon ^2}{{\mathbf{u}}_{a2}}({T_0},{T_1},{T_2})+\nonumber \\&+{\epsilon ^3}{{\mathbf{u}}_{a3}}({T_0},{T_1},{T_2}) + {\mathcal {O}}(\epsilon ^4)\nonumber \\ {{\mathbf{u}}_p}=&\,\epsilon {{\mathbf{u}}_{p1}}({T_0},{T_1},{T_2})+{\epsilon ^2}{{\mathbf{u}}_{p2}}({T_0},{T_1},{T_2})+\nonumber \\ {}&+{\epsilon ^3}{{\mathbf{u}}_{p3}}({T_0},{T_1},{T_2}) +\, {\mathcal {O}}(\epsilon ^4)\nonumber \\ {{\mathbf{f}}_p}=&\,\epsilon {{\mathbf{f}}_{p1}}({T_0},{T_1},{T_2})+{\epsilon ^2}{{\mathbf{f}}_{p2}}({T_0},{T_1},{T_2})+\nonumber \\ {}&+\,{\epsilon ^3}{{\mathbf{f}}_{p3}}({T_0},{T_1},{T_2}) + {\mathcal {O}}(\epsilon ^4) \end{aligned}$$
(82)

Introducing the expansion into the nonlinear differential equation (81) and equating all the terms of like \(\epsilon \)-powers yields a ordered hierarchy of linear differential equation

\(\bullet \):

Order \(\epsilon \)

$$\begin{aligned} \left[ {\begin{array}{cc} {\mathbf{M}}&{}{\mathbf{O}}\\ {\mathbf{O}}&{}{\mathbf{O}} \end{array}} \right] \!\! \left( \!{\begin{array}{c} {D_0^2 {\mathbf{u}}_{a1}}\\ {D_0^2 {\mathbf{u}}_{p1}} \end{array}}\!\right) \! +\!\left[ {\begin{array}{cc} {{{\mathbf{K}}_{aa}}}&{}{{{\mathbf{K}}_{ap}}}\\ {{{\mathbf{K}}_{pa}}}&{}{{{\mathbf{K}}_{pp}}} \end{array}} \right] \!\! \left( \!{\begin{array}{c} {{{\mathbf{u}}_{a1}}}\\ {{{\mathbf{u}}_{p1}}} \end{array}}\!\right) \!=\! \left( \!{\begin{array}{c} {{{\mathbf{0}}}}\\ {{{\mathbf{f}}_{p1}}} \end{array}}\!\right) \end{aligned}$$
(83)
\(\bullet \):

Order \(\epsilon ^2\)

$$\begin{aligned} \left[ {\begin{array}{cc} {\mathbf{M}}&{}{\mathbf{O}}\\ {\mathbf{O}}&{}{\mathbf{O}} \end{array}} \right] \!\! \left( \!{\begin{array}{c} {D_0^2 {\mathbf{u}}_{a2}}\\ {D_0^2 {\mathbf{u}}_{p2}} \end{array}}\!\right) \! +\!\left[ {\begin{array}{cc} {{{\mathbf{K}}_{aa}}}&{}{{{\mathbf{K}}_{ap}}}\\ {{{\mathbf{K}}_{pa}}}&{}{{{\mathbf{K}}_{pp}}} \end{array}} \right] \!\! \left( \!{\begin{array}{c} {{{\mathbf{u}}_{a2}}}\\ {{{\mathbf{u}}_{p2}}} \end{array}}\!\right) \!=\! \left( \!{\begin{array}{c} {{{\mathbf{f}}_{a2}}}\\ {{{\mathbf{f}}_{p2}}} \end{array}}\!\right) \end{aligned}$$
(84)
\(\bullet \):

Order \(\epsilon ^3\)

$$\begin{aligned} \left[ {\begin{array}{cc} {\mathbf{M}}&{}{\mathbf{O}}\\ {\mathbf{O}}&{}{\mathbf{O}} \end{array}} \right] \!\! \left( \!{\begin{array}{c} {D_0^2 {\mathbf{u}}_{a3}}\\ {D_0^2 {\mathbf{u}}_{p3}} \end{array}}\!\right) \! +\!\left[ {\begin{array}{cc} {{{\mathbf{K}}_{aa}}}&{}{{{\mathbf{K}}_{ap}}}\\ {{{\mathbf{K}}_{pa}}}&{}{{{\mathbf{K}}_{pp}}} \end{array}} \right] \!\! \left( \!{\begin{array}{c} {{{\mathbf{u}}_{a3}}}\\ {{{\mathbf{u}}_{p3}}} \end{array}}\!\right) \!=\! \left( \!{\begin{array}{c} {{{\mathbf{f}}_{a3}}}\\ {{{\mathbf{f}}_{p3}}} \end{array}}\!\right) \end{aligned}$$
(85)

where the force vectors \({\mathbf{f}}_{a2}=-2{\mathbf{M}}D_0D_1{\mathbf{u}}_{a1}\) and \({\mathbf{f}}_{a3}=-2{\mathbf{M}}{D_0}{D_1}{{\mathbf{q}}_{a2}}-2{\mathbf{M}}{D_0}{D_2}{{\mathbf{q}}_{a1}}-{\mathbf{M}}D_1^2{{\mathbf{q}}_{a1}}+{{\mathbf{N}}_2}{{\mathbf{q}}_{a1}}\).

The linear differential problem (83) at the \(\epsilon \)-order can be reformulated by separating the dynamic (upper) part from the quasi-static (lower) part, yielding

$$\begin{aligned}&{\mathbf{M}}{D_0^2 {\mathbf{u}}_{a1}}+{\mathbf{K}}_{aa}{\mathbf{u}}_{a1}+{\mathbf{K}}_{ap}{\mathbf{u}}_{p1}={\mathbf{0}}\\&{\mathbf{K}}_{pa}{\mathbf{u}}_{a1}+{\mathbf{K}}_{pp}{\mathbf{u}}_{p1}={\mathbf{f}}_{p1}\nonumber \end{aligned}$$
(86)

where the quasi-periodicity conditions (6) can be applied on the passive displacement and force vectors. Consequently, employing the quasi-static condensation rules \({\mathbf{u}}_{p1}={\mathbf{L}}_{pa}{\mathbf{u}}_{a1}\) and \({\mathbf{f}}_{p1}=\left( {\mathbf{K}}_{pa}+{\mathbf{K}}_{pp}{\mathbf{L}}_{pa}\right) {\mathbf{u}}_{a1}\), the condensed homogeneous equation in the active displacement vector reads

$$\begin{aligned}&{\mathbf{M}}{D_0^2 {\mathbf{u}}_{a1}}+\left( {\mathbf{K}}_{aa}+{\mathbf{K}}_{ap}{\mathbf{L}}_{pa}\right) {\mathbf{u}}_{a1}={\mathbf{0}} \end{aligned}$$
(87)

where the auxiliary matrix

$$\begin{aligned}&{{\mathbf{L}}_{pa}}=B^{-1}\left( {\begin{array}{*{20}{c}} {{\mathbf{K}}_{pa}^{(2)}+{\mathbf{K}}_{pa}^{(1)}{\text {e}}^{-\imath \beta }}\\ {{\mathbf{K}}_{pa}^{(2)}{\text {e}}^{-\imath \beta }+{\mathbf{K}}_{pa}^{(1)}{\text {e}}^{-2\imath \beta }} \end{array}}\right) \end{aligned}$$
(88)

where \({\mathbf{K}}_{pa}^{(i)}\) is the generic row of the 2-by-2 matrix \({\mathbf{K}}_{pa}\) (\(i=1,2\)) and the scalar quantity

$$\begin{aligned}&B\!=\!-\!\left[ {K_{pp}^{(21)}+\left( K_{pp}^{(22)}+K_{pp}^{(11)}\right) {\text {e}}^{-\imath \beta }+K_{pp}^{(12)}{\text {e}}^{-2\imath \beta }}\right] \end{aligned}$$
(89)

where \(K_{pp}^{(ij)}\) is the generic component of the 2-by-2 matrix \({\mathbf{K}}_{pp}\) (\(i,j=1,2\)). Since \(({\mathbf{K}}_{aa}+{\mathbf{K}}_{ap}{\mathbf{L}}_{pa})\) can be verified to coincide with \({\mathbf{K}}(\beta )\), the linear equation (87) states a differential problem that formally coincides with the linear part of Eq. (10). Consequently, the same eigenpairs \((\lambda ,{\varvec{\phi }})\) satisfy the eigenproblems associated to both the equations. Thus, recalling the matrix \({\varvec{\Phi }}\) of the mass-normalized eigenvectors (\({\varvec{\Phi }}^\top {\mathbf{M}}{\varvec{\Phi }}={\mathbf{I}}\)) and introducing the change of variable \({\mathbf{u}}_{a1}={\varvec{\Phi }}{\mathbf{q}}_{1}\) from the active displacement vector \({\mathbf{u}}_{a1}\) to the principal coordinate vector \({\mathbf{q}}_{1}\), Eq. (87) becomes

$$\begin{aligned} D_0^2\,{\mathbf {q}}_1+{\varvec{\Lambda }}{\mathbf {q}}_1={\mathbf {0}} \end{aligned}$$
(90)

which exactly coincides with the \(\epsilon \)-order equation (17). Moreover, the passive displacements are related to the principal coordinates through the relation \({\mathbf{u}}_{p1}={\mathbf{L}}_{pa}{\varvec{\Phi }}{\mathbf{q}}_{1}\).

The homogeneous problem associated with the linear differential equation (84) at the \(\epsilon ^2\)-order is formally coincident with that associated with the linear differential equation (83). Therefore, the same eigenvector matrix \({\varvec{\Phi }}\) can conveniently be employed to perform the change of variables \({\mathbf{u}}_{a2}={\varvec{\Phi }}{\mathbf{q}}_{2}\), yielding

$$\begin{aligned} D_0^2\,{\mathbf {q}}_2+{\varvec{\Lambda }}{\mathbf {q}}_2=-2{\varvec{\Phi }}^\top {\mathbf{M}}{\varvec{\Phi }}D_0D_1{\mathbf {q}}_1 \end{aligned}$$
(91)

which, recalling the mass-normalization rule \({\varvec{\Phi }}^\top {\mathbf{M}}{\varvec{\Phi }}={\mathbf{I}}\), actually coincides with the \(\epsilon ^2\)-order equation (18).

The homogeneous problem associated with the linear differential equation (85) at the \(\epsilon ^3\)-order is again formally coincident with that associated with the linear differential equation (83). Therefore, the eigenvector matrix \({\varvec{\Phi }}\) can conveniently be employed to perform the change of variables \({\mathbf{u}}_{a3}={\varvec{\Phi }}{\mathbf{q}}_{3}\), yielding

$$\begin{aligned}&D_0^2\,{\mathbf {q}}_3+{\varvec{\Lambda }}{\mathbf {q}}_3=-2{\varvec{\Phi }}^\top \!{\mathbf{M}}{\varvec{\Phi }}D_0D_1{\mathbf {q}}_2+\\&\quad -\,2{\varvec{\Phi }}^\top {\mathbf{M}}{\varvec{\Phi }}D_0D_2{\mathbf {q}}_1-{\varvec{\Phi }}^\top \!{\mathbf{M}}{\varvec{\Phi }}D_1^2{\mathbf {q}}_1 -{\varvec{\Phi }}^\top {\mathbf{N}}_2{\varvec{\Phi }}{\mathbf {q}}_1\nonumber \end{aligned}$$
(92)

which, recalling the mass-normalization rule \({\varvec{\Phi }}^\top {\mathbf{M}}{\varvec{\Phi }}={\mathbf{I}}\) and verifying that \({\varvec{\Phi }}^\top {\mathbf{N}}_2{\varvec{\Phi }}{\mathbf {q}}_1=-{\mathbf{c}}({\mathbf {q}}_1,{\mathbf {q}}_1,{\mathbf {q}}_1)\), actually coincides with the \(\epsilon ^3\)-order equation (19).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lepidi, M., Bacigalupo, A. Wave propagation properties of one-dimensional acoustic metamaterials with nonlinear diatomic microstructure. Nonlinear Dyn 98, 2711–2735 (2019). https://doi.org/10.1007/s11071-019-05032-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-019-05032-3

Keywords

Navigation