Next Article in Journal
Preparation and Load-Bearing Capacity of Lattice Cell Warren Truss Slot Resin-Stiffener-Reinforced Foam Sandwich Material
Previous Article in Journal
Study on Preparation and Performance of CO2 Foamed Concrete for Heat Insulation and Carbon Storage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Vibration and Bandgap Behavior of Sandwich Pyramid Lattice Core Plate with Resonant Rings

School of Mechatronics Engineering, Harbin Institute of Technology, Harbin 150001, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(7), 2730; https://doi.org/10.3390/ma16072730
Submission received: 8 March 2023 / Revised: 25 March 2023 / Accepted: 27 March 2023 / Published: 29 March 2023
(This article belongs to the Section Mechanics of Materials)

Abstract

:
The vibration suppression performance of the pyramid lattice core sandwich plates is receiving increasing attention and needs further investigation for technical upgrading of potential engineering applications. Inspired by the localized resonant mechanism of the acoustic metamaterials and considering the integrity of the lattice sandwich plate, we reshaped a sandwich pyramid lattice core with resonant rings (SPLCRR). Finite element (FE) models are built up for the calculations of the dispersion curves and vibration transmission. The validity of the bandgap of the SPLCRR and remarkable vibration suppression are verified by experimental observations and the numerical methods. Furthermore, the effects of geometric parameters, material parameters and period parameters on the bandgaps of the SPLCRR are systematically investigated, which offers a deeper understanding of the underlying mechanism of bandgap and helps the SPLCRR structure meet the technological update requirements of practical engineering design.

1. Introduction

The continuous development and technological update requirements of aerospace platforms and high-speed rails have led to great efforts for designing new structures and materials that integrate both lightweight and special multifunctional properties [1,2,3,4,5,6]. However, sandwich plates generate harmful vibration and radiated noise that can negatively impact both equipment functionality and operator safety. Therefore, it is crucial to explore innovative methods and theories for controlling vibration and noise in plate structures. The lattice sandwich structure is a multi-layer structure consisting of two panels and a truss core. In recent years, the lattice sandwich structures attract huge attention due to their high stiffness, low density and light weight [7,8,9,10,11]. Interestingly, various grid-truss core structures such as tetrahedron, three-dimensional kagome, and vertebral body structures have been developed [7,11]. At present, there are many studies on mechanical properties, manufacturing process, vibration characteristics, and dynamic responses of lattice sandwich structures [12,13,14]. Nevertheless, the elastic wave propagates at lower frequencies due to the open cell sandwich structure and high stiffness to mass ratio, leading to poor vibration attenuation and acoustic performance [8,10]. The low-frequency mode contributes the most to sound power, and thus, effective reduction in sound power can be achieved by suppressing vibrations in the low-frequency range [15]. Therefore, it is still a challenge for lattice sandwich structures to simultaneously achieve the balance of vibration suppression and light weight.
Fortunately, in 2000, Liu and co-workers [16] proposed a localized resonant mechanism, also known as elastic metamaterials (EM), to break the conventional mass-density law of sound transmission, significantly improving sound transmission loss (STL) below 400 Hz. Since then, numerous studies have been carried out relating to the properties of unusual structures, such as negative modulus [17,18,19], energy harvesting [20,21,22], negative refraction [23], acoustic invisibility [24,25], vibration attenuation [26,27,28,29,30], topological acoustics [31] and so forth, which could enhance wave modulation of classical lattice sandwich structures.
Nowadays, there is plenty of research on the propagation of flexural waves and band structures in thin plates. It is noteworthy that the structures showed mechanisms of low frequency forbidden bands and Bragg bandgaps in parallel with local resonance and Bragg scattering. The bandgaps, in turn, are directly influenced by the geometry and lattice symmetry of the resonator array [32]. Yong et al. [33] extended the well-known plane wave expansion method (PWEM) to deal with plate systems with periodic arrays of spring-mass resonators, and they determined that the bandwidth of the bandgaps is greatly affected by the frequencies of the causes of the local resonators. Furthermore, considering the Kirchhoff–Love thin-plate theory, Miranda et al. investigated the band structure of flexural waves propagating in thin metamaterial plates and the effect of periodic arrays of multi-degree-of-freedom local resonators in square and triangular lattices [34].
Using finite element model calculations, Massimo et al. [35] investigated the dynamic response of a three-layer sandwich panel with a honeycomb core. They also investigated the effect of individually periodic placement of cores of different geometries in the two-dimensional space of the structure. Song et al. [36] studied the sound transmission of sandwich panels and its reduction using the bandgap concept. The results showed that sound transmission was significantly reduced over the bandgaps of periodic sandwich panels. Li and An [27,37] proposed improved three-dimensional truss lattice structures for low and broadband elastic wave absorption, respectively.
The abovementioned works are concerned with the bandgap, vibration reduction, and acoustic characteristics of various plated structures. Although some achievements have been chosen in resolving the vibration attenuation and acoustic characteristics of thin or sandwich plates by attaching mass block, the uniformity of lattice sandwich plates was challenged [27]. In this work, inspired by the localized resonant mechanism of the acoustic metamaterials and considering the integrity of the lattice sandwich plate, we reshaped a sandwich pyramid lattice core with resonant rings (SPLCRR). Compared to the typical lattice sandwich plate, the SPLCRR addition of a resonant ring onto the rod enables the sandwich plate to fulfill the requirements for a smooth outer surface in engineering applications while also providing increased stiffness, structural simplicity, and ease to design. The dispersion curves of the SPLCRR and evolution of the bandgaps for the SPLCRR are investigated to achieve a wide bandgap in the low-frequency range. The effect of bandgaps on vibration attenuation is further investigated by theoretical and experimental verification. Furthermore, the effects of geometric parameters, material parameters and period parameters on the bandgap of the SPLCRR are systematically analyzed, which offers a deeper understanding of the underlying mechanism of bandgap and helps the SPLCRR structure meet the technological update requirements of practical engineering design.

2. Model and Theory

Figure 1a illustrates a diagram depicting a typical lattice sandwich structure plate with periodic lattice-truss-core. As shown in Figure 1b,c, the typical lattice unit cell and the SPLCRR unit cell are considered in this work, respectively. Different from the system of the traditional elastic metamaterial (EM) unit cell which is usually made by attaching different materials to the surface of the plate, the unit cell of SPLCRR has two radius jump discontinuities in the core rods. The equation that governs the propagation of elastic waves in solids is expressed as follows:
j = 1 3 x λ u j x j + x j μ u i x j + u j x i = ρ 2 u i t 2 ; i , j = x , y , z ,
where r represents the mass density, u and v denote the displacement vector, t denotes time, and λ and μ represent the Lame constants. Additionally, x, y and z represent the Cartesian coordinate variables.
In order to investigate the bandgaps of the proposed SPLCRR plate, a series of calculations of dispersion relations are conducted with the FEM based on Solid Mechanics Module of the COMSOL. Based on the periodicity of the proposed plate, we assumed that in both the x and y directions, single unit cell was considered, as depicted in Figure 1c. An adaptive mesh was selected to achieve sufficiently good convergence while minimizing computational costs within a reasonable time constraint.
Due to the Bloch–Floquet theorem, interface between adjacent cells using Bloch periodic boundary conditions can be expressed as follows:
u r = u k r e i k r ,
where u is the displacement and r refers to the coordinate vectors.
After scanning all Bloch wave vectors along the edge of ΓXM in various lattice unit cells, the dispersion curves for the periodic sandwich plates can be obtained separately. By modulating the wave vector k within the first irreducible Brillouin zone (as depicted in the left inset of Figure 2) of the unit cell [30], one can derive the dispersion curves of the unit cell with respect to its wave number and frequency. These curves facilitate the determination of the band structure of an infinite periodic plate, which in turn enables the identification of its bandgaps.
To further verify the bandgap properties of the SPLCRR, the vibration attenuation characteristics of this plate were calculated by FEM. As shown in the inset on the right panel of Figure 2, the sandwich plate comprises 7 × 7 cells, a harmonic excitation is loaded at the edge of the panel, and the response occurs on the other side of the panel. It is assumed that the deflection of the sandwich panel under the transverse load is far less than its thickness. As a result, this study only takes into account the out-of-plane bandgaps and disregards the displacement parameters in the x-y plane.
At point P0 of the plate, an input acceleration excitation (denoted as ain ) is introduced, and the resulting acceleration response (denoted as aout) is measured at the response acquisition point P1, as depicted in the inset on the right of Figure 2. Ultimately, the vibration attenuation T can be obtained by altering the frequency of the input acceleration excitation:
T = 20 l o g a o u t a i n .

3. Results and Discussion

In this section, a series of numerical simulations are conducted to analyze the dispersion relations properties of the proposed SPLCRR. The dispersion relations and vibration attenuation of the SPLCRR design matched up with the typical lattice sandwich plate, respectively. The influence of geometric parameters, material parameters, and period parameters on bandgap characteristics is systematically discussed.

3.1. Simulation Parameters

As shown in Figure 1c, the SPLCRR is composed of two face-sheets and a core layer which is four diagonals with additional resonant rings, wherein the sandwich plate with truss core is characterized by the total thickness represented by h, where the upper and lower face sheets possess equal thickness hf‘, and the core height is denoted by hc. The length, radius and inclination angle of the truss core and the resonant ring are represented by l, hm, rc, rm and θ, respectively. The geometrical parameters and material properties of the model are listed in Table 1 and Table 2.
Figure 2 presents the calculated dispersion curves of the proposed SPLCRR and the typical lattice sandwich plate. According to plate wave theory, a finite thickness plate can support the propagation of various of anti-symmetric (A-mode) and symmetric (S-mode) Lamb waves and shear-horizontal (SH-mode) waves with cutoff frequencies, including three fundamental plate modes. Compared with the typical lattice sandwich plate, the band structure of the SPLCRR presents a complete bandgap from 780 Hz to 990 Hz (the dark blue region in Figure 2) and a flexural wave bandgap from 600 Hz to 660 Hz (the light blue region in Figure 2). In a complete bandgap, the propagation of elastic waves will be significantly suppressed. Regarding various localized modes of guided elastic waves in solids, the dispersion curves of longitudinal and shear modes remain straight since their wave velocity is constant across frequencies. Conversely, the dispersion curves of flexural waves are non-linear because the wave velocity is dependent on frequency. Generally, flexural waves carry more vibrational energy than other types of waves. As a result, the complete flexural-wave bandgap can effectively reduce vibrations [34].
To further illustrate the physical mechanism of the bandgap, the eigenmodes of the two models at the bandgap boundary frequencies were analyzed and are shown in Figure 3. The frequencies A, B and C are located on the first and second order dispersion curves. As depicted in Figure 2, the SPLCRR reduces the frequencies of the eigenmodes and opens and widens the bandgap at point B. More complex coupling effects are obtained for the eigenmodes at the upper and lower edges of the bending wave bandgap.
As depicted in Figure 2, the eigenmodes at frequencies D, F represent the upper edge of the complete bandgap and the flexural wave bandgap, which are at fourth and ninth dispersion curves, respectively. The intermediate state E1 and E2 represent the eigenmodes of the fourth to eighth dispersion curve. The face-sheets of the SPLCRR primarily exhibit lateral motion, with negligible movement in the vertical direction. The motion characteristics of the complete bandgap and the flexural wave bandgap will significantly affect vibration attenuation, as we will show later.
To obtain vibration attenuation, one can calculate T by sweeping a range of excitation frequencies. Figure 4 shows a comparison of the vibration attenuation calculated for a typical lattice sandwich plate and SPLCRR. The dispersion spectra yield two distinct bandgaps, represented by dark blue and light blue regions, corresponding to the complete bandgap and the flexural-wave bandgap, respectively. The vibration attenuation in the SPLCRR is significantly higher than in the typical lattice sandwich plate, both within the complete bandgap and the flexural-wave bandgap (represented by the light blue regions). Moreover, the vibration attenuation bands agree with the bandgaps shown in the dispersion curve. In contrast, the flexural-wave bandgap in the other region (680 Hz–760 Hz) does not produce significant vibration attenuation. Corresponding to the eigenmodes (see Figure 3) of the upper and lower boundaries of this bandgap, it can be seen that the suppression of out-of-plane vibrations does not play a significant role.

3.2. Experimental Verification

In Section 3.1, we calculated the vibration transmission of a typical lattice plate and a SPLCRR plate using FEM and compared it with their dispersion curve to verify the bandgaps and the performance of the vibration attenuation. To further verify the validity of the bandgaps and vibration attenuation performance, in this section, a typical lattice plate and a SPLCRR plate are used for experimental verification.
The specimens contain 7 × 7 units (360 mm × 360 mm) with the same unit geometry and material parameters as in the FEM in Section 3.1. The experimental setup and experimental specimens are shown in Figure 5. Two end sides of the specimen are simply supported by foam materials to mimic the simply supported boundary condition. The vibration transmission spectra of the sandwich plates specimen were obtained by collecting the impact hammer (Brüel and Kjær, 8206-002) of the exciting point and the vibration acceleration (Brüel and Kjær, 4507) responses of the acquisition point with date acquisition (Brüel and Kjær, 3040). The numerical bandgaps detected in Figure 6 are represented by the blue regions. It was observed that the measured vibration suppression ranges exhibited a strong agreement with both the predicted bandgaps and the previous FEM simulation. The vibration suppression performance is experimentally verified in the proposed SPLCRR plate.

3.3. Effect of Geometric Parameters on Bandgap

Structural geometric parameters are one of the main factors affecting the bandgap of metamaterials. To further study the effect of geometric parameters on the bandgap, the dispersion curves of the SPLCRR are calculated under different rod diameters, angles between rods and panels, panel thicknesses, oscillator diameters, and oscillator lengths. Throughout the study, unless explicitly mentioned, the material and geometry parameters employed are the same as those specified in Section 3.1.
In this section, we present an analysis of bandgap behaviors of SPLCRR with varying lattice core and face-sheets dimensions. The investigation encompasses nine discrete lattice core rod radius values (rc = 1 mm–3 mm) and angle values (θ = 20–32°), and face-sheet thickness values (hf = 2 mm–3 mm) are chosen. The bandgap behaviors of SPLCRR are significantly influenced by changes in the lattice core and face-sheet dimensions, as evident from Figure 7, wherein the bandgaps move to higher frequencies with the increase in the rod radius and face-sheets thickness, but the increase in the lattice core rod angle causes the bandgaps to move to low frequencies. This is expected, because the change in lattice core parameters and the face-sheets parameter results in a modification stiffness of lattice sandwich layer (eq. X). It is worth noting that the width of the second bandgap of our interest reaches its maximum when rc = 2.5 mm, θ = 22° and hf = 2.5 mm, respectively. In addition, the change in rod radius is more sensitive to width of the second bandgap as compared to other two parameters.
As the dimensions of resonant ring are regarded as highly significant parameters in the whole SPLCRR, it is essential to evaluate its effect on the bandgap behaviors of SPLCRR. Thus, the influences of two geometric parameters rm and hm (defined in Section 3.1) of resonant ring on bandgaps are discussed in detail here, where nine different geometric parameters (rm = 4–6 mm) and (hm = 5–15 mm) chosen in this example. From Figure 8, it can be observed that the bandgap behaviors are strongly influenced by resonant ring radius and height, and increasing the radius and height leads to the bandgaps moving to lower frequency. This is expected, because the increase in resonant ring radius and height results in a modification of equivalent mass of lattice sandwich layer. In addition, the change in the resonant ring radius and height has no significant effect on the width of the second bandgap. In other words, the bandgap frequency range can be adjusted by changing geometric parameters of the resonant ring without affecting the bandgaps width.

3.4. Effect of Material Parameters on Bandgap

Material parameters are one of the main factors affecting the bandgap of metamaterials. To further investigate the effect of material parameters on the bandgap, the dispersion curves of the SPLCRR were calculated for different rod, panel, resonant ring elastic modulus and resonant ring damping ratio.
As can be seen in Figure 9, nine different elastic modulus of lattice rod (Ec = 2.5–250 GPa) and face-sheets (Ef = 0.25–25 GPa) are chosen in this example; the other properties are the same as those in Section 3.1. From Figure 9, it can be observed that the change in the elastic modulus of lattice rod and face-sheets exerts a significant impact on the bandgap behaviors of SPLCRR, wherein the bandgaps move to higher frequencies with the increase in the elastic modulus of lattice rod and face-sheets. This is reasonable that the increase in the elastic modulus of lattice rod and face-sheets results in the increase in equivalent stiffness of lattice sandwich layer. It is worth noting that the width of the second bandgap of our interest reaches its maximum when Ec = 7.5 GPa and Ef = 2.5 GPa, respectively. In addition, the width of the second bandgap first increases and then gradually disappears with the increase in the elastic modulus Ec, while the increase in the elastic modulus Ef shows the opposite trend.
As the material parameters of resonant rings are regarded as highly significant parameters in the whole SPLCRR, it is essential to evaluate its effect on the bandgap behaviors of SPLCRR. Thus, the influence of two material parameters Em and ηm (defined in Section 3.1) of resonant ring on bandgaps are discussed in detail here, where nine different material parameters (Em = 2.5–250 GPa) and (ηm = 0.01–0.3) are chosen in this example. From Figure 10, it can be obtained that with the increase of Young's modulus and damping ratio of the resonant ring, there is no significant change in the upper and lower edge frequencies of the band gap and no significant change in the bandgap width. Therefore, the Young's modulus and damping ratio of the resonant ring has little effect on the band gap of the metamaterial plate.

3.5. Effect of Period Parameters on Bandgap

The period parameter is one of the main factors affecting the bandgap of metamaterials. To further investigate the effect of the period parameter on the bandgap, the dispersion curves of the SPLCRR are calculated for different period lengths and different distribution forms.
From Figure 11a, it can be obtained that the frequencies of the bandgap are gradually decreasing as the period length increases, which is because the core does not change as the period length increases, resulting in a decrease in the stiffness of the sandwich plate, which causes the bandgap to move to lower frequencies. The width of the first bandgap increases gradually with the period length and then tends to be unchanged. The widths of the second and third bandgaps do not change significantly with increasing period length and then disappear rapidly.
The dispersion curves of the SPLCRR with different distribution forms are calculated while keeping other parameters constant. The distribution forms are shown in Figure 12a,b. From Figure 11b, it can be obtained that the frequency of the upper and lower edges of the bandgap under the square distribution is higher than that of the hexagonal distribution, where the width of the first and second bandgap of the hexagonal distribution is wider than that of the square distribution, and the opposite is true for the third bandgap. From the effect of the period length on the bandgap in the previous section, it can be seen that the hexagonal distribution increases the distance between the nodes, producing a result that is commensurate with the variation of the period length.

4. Conclusions

In this work, inspired by the localized resonant mechanism of classical acoustic metamaterials and considering the integrity of the lattice sandwich plate, we reshaped a sandwich pyramid lattice core with resonant rings (SPLCRR). Based on the experimental verifications and numerical simulations, the effects of geometric parameters, material parameters and period parameters on the bandgaps of the SPLCRR were systematically investigated. Several key observations obtained from the detailed parametric investigation can be summarized as:
(a)
Remarkable vibration suppression and bandgaps were verified by comparisons between numerical and experimental results.
(b)
Different geometric parameters were discussed. The thickness of face-sheets and the rod radius of the core had significant effects on the frequency range and width of the bandgap, which moved to higher frequency with the increase in the two values. The rod radius was more sensitive to width of the second bandgap compared to other parameters.
(c)
Different material parameters were discussed. The elastic modulus of lattice core and face-sheets had significant effects on the frequency range and width of the bandgap, wherein the bandgaps moved to higher frequencies with the increase in the elastic modulus of lattice rod and face-sheets.
(d)
Different period parameters were discussed. The frequencies of the bandgap gradually decrease as the period length increases. The frequency of the upper and lower edges of the bandgap under the square distribution was higher than that of the hexagonal distribution.
The research results are expected to be of theoretic significance and offer engineering application prospects.

Author Contributions

C.L.: investigation and writing—original draft. Z.C.: supervision and funding acquisition. Y.J.: editing and supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This reserch was funded by by the National Key R&D Program of China (No.2019YFE0116200) and National Natural Science Foundation of China (No. 11772103).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data used to support the findings of this study are available from the corresponding author upon request.

Acknowledgments

The authors would like to thank the National Natural Science Foundation of China and National Key R&D Program of China.

Conflicts of Interest

The authors have no financial or proprietary interests in any material discussed in this article.

References

  1. Casalino, D.; Diozzi, F.; Sannino, R.; Paonessa, A. Aircraft noise reduction technologies: A bibliographic review. Aerosp. Sci. Technol. 2008, 12, 1–17. [Google Scholar] [CrossRef]
  2. Ivanov, N.; Boiko, I.; Shashurin, A. The Problem of High-Speed Railway Noise Prediction and Reduction. Procedia Eng. 2017, 189, 539–546. [Google Scholar] [CrossRef]
  3. Ma, Q.; Rejab, M.R.M.; Siregar, J.P.; Guan, Z. A review of the recent trends on core structures and impact response of sandwich panels. J. Compos. Mater. 2021, 55, 2513–2555. [Google Scholar] [CrossRef]
  4. Li, Y.; Shen, C.; Xie, Y.; Li, J.; Wang, W.; Cummer, S.A.; Jing, Y. Tunable Asymmetric Transmission via Lossy Acoustic Metasurfaces. Phys. Rev. Lett. 2017, 119, 035501. [Google Scholar] [CrossRef] [Green Version]
  5. Sobhani, E.; Masoodi, A.R.; Ahmadi-Pari, A.R. Circumferential vibration analysis of nano-porous-sandwich assembled spheri-cal-cylindrical-conical shells under elastic boundary conditions. Eng. Struct. 2022, 273, 115094. [Google Scholar] [CrossRef]
  6. Li, Q.; Xie, B.; Sahmani, S.; Safaei, B. Surface stress effect on the nonlinear free vibrations of functionally graded composite nanoshells in the presence of modal interaction. J. Braz. Soc. Mech. Sci. Eng. 2020, 42, 237. [Google Scholar] [CrossRef]
  7. Fan, H.; Meng, F.; Yang, W. Sandwich panels with Kagome lattice cores reinforced by carbon fibers. Compos. Struct. 2007, 81, 533–539. [Google Scholar] [CrossRef]
  8. Fu, T.; Chen, Z.; Yu, H.; Zhu, X.; Zhao, Y. Sound transmission loss behavior of sandwich panel with different truss cores under external mean airflow. Aerosp. Sci. Technol. 2019, 86, 714–723. [Google Scholar] [CrossRef]
  9. Wang, D.-W.; Ma, L.; Wang, X.-T.; Wen, Z.-H.; Glorieux, C. Sound transmission loss of laminated composite sandwich structures with pyramidal truss cores. Compos. Struct. 2019, 220, 19–30. [Google Scholar] [CrossRef]
  10. Wen, Z.-H.; Wang, D.-W.; Ma, L. Sound transmission of composite sandwich panel with face-centered cubic core. Mech. Adv. Mater. Struct. 2021, 28, 1663–1676. [Google Scholar] [CrossRef]
  11. Yang, G.; Hou, C.; Zhao, M.; Mao, W. Comparison of convective heat transfer for Kagome and tetrahedral truss-cored lattice sand-wich panels. Sci. Rep. 2019, 9, 3731. [Google Scholar] [CrossRef] [Green Version]
  12. Fan, H.-L.; Zeng, T.; Fang, D.-N.; Yang, W. Mechanics of advanced fiber reinforced lattice composites. Acta Mech. Sin. 2010, 26, 825–835. [Google Scholar] [CrossRef]
  13. Wu, Q.; Ma, L.; Wu, L.; Xiong, J. A novel strengthening method for carbon fiber composite lattice truss structures. Compos. Struct. 2016, 153, 585–592. [Google Scholar] [CrossRef]
  14. Lu, L.; Song, H.; Huang, C. Effects of random damages on dynamic behavior of metallic sandwich panel with truss core. Compos. Part B Eng. 2016, 116, 278–290. [Google Scholar] [CrossRef] [Green Version]
  15. Maury, C.; Gardonio, P.; Elliott, S.J. Model for Active Control of Flow-Induced Noise Transmitted Through Double Partitions. AIAA J. 2002, 40, 1113–1121. [Google Scholar] [CrossRef]
  16. Liu, Z.; Zhang, X.; Mao, Y.; Zhu, Y.Y.; Yang, Z.; Chan, C.T.; Sheng, P. Locally Resonant Sonic Materials. Science 2000, 289, 1734–1736. [Google Scholar] [CrossRef] [PubMed]
  17. Cummer, S.A.; Christensen, J.; Alù, A. Controlling sound with acoustic metamaterials. Nat. Rev. Mater. 2016, 1, 16001. [Google Scholar] [CrossRef] [Green Version]
  18. Ding, C.; Hao, L.; Zhao, X. Two-dimensional acoustic metamaterial with negative modulus. J. Appl. Phys. 2010, 108, 074911. [Google Scholar]
  19. Lee, S.H.; Park, C.M.; Seo, Y.M.; Wang, Z.G.; Kim, C.K. Acoustic metamaterial with negative modulus. J. Physics Condens. Matter 2009, 21, 175704. [Google Scholar] [CrossRef] [Green Version]
  20. Wen, Z.; Zeng, S.; Wang, D.; Jin, Y.; Djafari-Rouhani, B. Robust edge states of subwavelength chiral phononic plates. Extreme Mech. Lett. 2021, 44, 101209. [Google Scholar] [CrossRef]
  21. Wen, Z.; Jin, Y.; Gao, P.; Zhuang, X.; Rabczuk, T.; Djafari-Rouhani, B. Topological cavities in phononic plates for robust energy harvesting. Mech. Syst. Signal Process. 2021, 162, 108047. [Google Scholar] [CrossRef]
  22. Wen, Z.; Wang, W.; Khelif, A.; Djafari-Rouhani, B.; Jin, Y. A perspective on elastic metastructures for energy harvesting. Appl. Phys. Lett. 2022, 120, 020501. [Google Scholar] [CrossRef]
  23. Park, C.M.; Lee, S.H. Zero-reflection acoustic metamaterial with a negative refractive index. Sci. Rep. 2019, 9, 3372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Choudhury, B.; Jha, R.M. A Review of Metamaterial Invisibility Cloaks. Comput. Mater. Contin. 2013, 33, 275–308. [Google Scholar]
  25. Islam, S.S.; Hasan, M.M.; Faruque, M.R.I. A new metamaterial-based wideband rectangular invisibility cloak. Appl. Phys. A 2018, 124, 160. [Google Scholar] [CrossRef]
  26. Wang, Z.; Zhang, Q.; Zhang, K.; Hu, G. Tunable Digital Metamaterial for Broadband Vibration Isolation at Low Frequency. Adv. Mater. 2016, 28, 9857–9861. [Google Scholar] [CrossRef]
  27. Li, H.; Hu, Y.; Huang, H.; Chen, J.; Zhao, M.; Li, B. Broadband low-frequency vibration attenuation in 3D printed composite meta-lattice sandwich structures. Compos. Part B Eng. 2021, 215, 108772. [Google Scholar] [CrossRef]
  28. Li, Z.-Y.; Ma, T.-X.; Wang, Y.-Z.; Li, F.-M.; Zhang, C. Vibration isolation by novel meta-design of pyramid-core lattice sandwich structures. J. Sound Vib. 2020, 480, 115377. [Google Scholar] [CrossRef]
  29. An, X.; Fan, H.; Zhang, C. Elastic wave and vibration bandgaps in two-dimensional acoustic metamaterials with resonators and disorders. Wave Motion 2018, 80, 69–81. [Google Scholar] [CrossRef]
  30. Tian, H.; Wang, X.; Zhou, Y.-H. Theoretical model and analytical approach for a circular membrane–ring structure of locally resonant acoustic metamaterial. Appl. Phys. A 2013, 114, 985–990. [Google Scholar] [CrossRef]
  31. Zhang, Z.; Wei, Q.; Cheng, Y.; Zhang, T.; Wu, D.; Liu, X. Topological Creation of Acoustic Pseudospin Multipoles in a Flow-Free Sym-metry-Broken Metamaterial Lattice. Phys. Rev. Lett. 2017, 118, 084303. [Google Scholar] [CrossRef] [Green Version]
  32. Hsu, J.-C. Local resonances-induced low-frequency band gaps in two-dimensional phononic crystal slabs with periodic stepped resonators. J. Phys. Appl. Phys. 2011, 44, 055401. [Google Scholar] [CrossRef]
  33. Xiao, Y.; Wen, J.; Wen, X. Flexural wave band gaps in locally resonant thin plates with periodically attached spring–mass resonators. J. Phys. Appl. Phys. 2012, 45, 195401. [Google Scholar] [CrossRef]
  34. Miranda, E.; Nobrega, E.; Ferreira, A.; Dos Santos, J. Flexural wave band gaps in a multi-resonator elastic metamaterial plate using Kirchhoff-Love theory. Mech. Syst. Signal Process. 2018, 116, 480–504. [Google Scholar] [CrossRef]
  35. Ruzzene, M.; Mazzarella, L.; Tsopelas, P.; Scarpa, F. Wave Propagation in Sandwich Plates with Periodic Auxetic Core. J. Intell. Mater. Syst. Struct. 2002, 13, 587–597. [Google Scholar] [CrossRef]
  36. Song, Y.; Feng, L.; Wen, J.; Yu, D.; Wen, X. Reduction of the sound transmission of a periodic sandwich plate using the stop band concept. Compos. Struct. 2015, 128, 428–436. [Google Scholar] [CrossRef]
  37. An, X.; Lai, C.; Fan, H.; Zhang, C. 3D acoustic metamaterial-based mechanical metalattice structures for low-frequency and broadband vibration attenuation. Int. J. Solids Struct. 2020, 191–192, 293–306. [Google Scholar] [CrossRef]
Figure 1. (a) Typical lattice sandwich plate, (b) typical lattice sandwich unit cell, (c) SPLCRR unit-cell.
Figure 1. (a) Typical lattice sandwich plate, (b) typical lattice sandwich unit cell, (c) SPLCRR unit-cell.
Materials 16 02730 g001
Figure 2. Dispersion curves of the sandwich pyramid lattice core plates without and with resonant rings on the left and right panels, respectively. Inset: The left panel is the first irreducible Brillouin zone, and the right panel is the input/output setting of numerical simulation.
Figure 2. Dispersion curves of the sandwich pyramid lattice core plates without and with resonant rings on the left and right panels, respectively. Inset: The left panel is the first irreducible Brillouin zone, and the right panel is the input/output setting of numerical simulation.
Materials 16 02730 g002
Figure 3. Eigenmodes of the SPLCRR unit cell at frequency (a) fA = 502 Hz (b) fB = 596 Hz (c) fC = 625 Hz (d) fD = 874 Hz (e) fF = 964 Hz and (f) fE = 1003 Hz.
Figure 3. Eigenmodes of the SPLCRR unit cell at frequency (a) fA = 502 Hz (b) fB = 596 Hz (c) fC = 625 Hz (d) fD = 874 Hz (e) fF = 964 Hz and (f) fE = 1003 Hz.
Materials 16 02730 g003
Figure 4. Vibration transmission spectra of non-dissipative typical and SPLCRR structures.
Figure 4. Vibration transmission spectra of non-dissipative typical and SPLCRR structures.
Materials 16 02730 g004
Figure 5. (a) The experiment set-up (b) typical sandwich plate (c) SPLCRR plate.
Figure 5. (a) The experiment set-up (b) typical sandwich plate (c) SPLCRR plate.
Materials 16 02730 g005
Figure 6. Comparison between the typical sandwich plate and SPLCRR plate for vibration transmission spectra.
Figure 6. Comparison between the typical sandwich plate and SPLCRR plate for vibration transmission spectra.
Materials 16 02730 g006
Figure 7. Effect of geometric parameters of lattice core and face-sheets on bandgaps of the SPLCRR. (a) lattice core rod radius values rc (b) angle values θ (c) face-sheet thickness values hf.
Figure 7. Effect of geometric parameters of lattice core and face-sheets on bandgaps of the SPLCRR. (a) lattice core rod radius values rc (b) angle values θ (c) face-sheet thickness values hf.
Materials 16 02730 g007
Figure 8. Effect of geometric parameters of resonant ring on bandgap of the SPLCRR. (a) resonant ring radius values rm (b) resonant ring height values hm.
Figure 8. Effect of geometric parameters of resonant ring on bandgap of the SPLCRR. (a) resonant ring radius values rm (b) resonant ring height values hm.
Materials 16 02730 g008
Figure 9. Effect of material parameters of lattice core and face-sheets on bandgaps of the SPLCRR. (a) elastic modulus of lattice rod Ec (b) elastic modulus of face-sheets Ef.
Figure 9. Effect of material parameters of lattice core and face-sheets on bandgaps of the SPLCRR. (a) elastic modulus of lattice rod Ec (b) elastic modulus of face-sheets Ef.
Materials 16 02730 g009
Figure 10. Effect of material parameters of resonant ring on bandgaps of the SPLCRR. (a) elastic modulus of resonant ring Em (b) damping ratio of the resonant ring ηm.
Figure 10. Effect of material parameters of resonant ring on bandgaps of the SPLCRR. (a) elastic modulus of resonant ring Em (b) damping ratio of the resonant ring ηm.
Materials 16 02730 g010
Figure 11. Effect of period parameters on bandgap of the SPLCRR. (a) period length (b) different distribution forms.
Figure 11. Effect of period parameters on bandgap of the SPLCRR. (a) period length (b) different distribution forms.
Materials 16 02730 g011
Figure 12. Distribution forms of SPLCRR plate for (a) square and (b) triangular lattices. First irreducible Brillouin zone in shaded region for (a) square and (b) triangular lattices.
Figure 12. Distribution forms of SPLCRR plate for (a) square and (b) triangular lattices. First irreducible Brillouin zone in shaded region for (a) square and (b) triangular lattices.
Materials 16 02730 g012
Table 1. Geometric parameters of the SPLCRR.
Table 1. Geometric parameters of the SPLCRR.
ahchfhmrcrmθ
40 mm38 mm2 mm10 mm2 mm5 mm30°
Table 2. Material parameters of the SPLCRR [27].
Table 2. Material parameters of the SPLCRR [27].
Density ρ (kg/m3)Modulus E (GPa)Poisson’s Ratio μ
Nylon12003.50.37
Steel78002100.3
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, C.; Chen, Z.; Jiao, Y. Vibration and Bandgap Behavior of Sandwich Pyramid Lattice Core Plate with Resonant Rings. Materials 2023, 16, 2730. https://doi.org/10.3390/ma16072730

AMA Style

Li C, Chen Z, Jiao Y. Vibration and Bandgap Behavior of Sandwich Pyramid Lattice Core Plate with Resonant Rings. Materials. 2023; 16(7):2730. https://doi.org/10.3390/ma16072730

Chicago/Turabian Style

Li, Chengfei, Zhaobo Chen, and Yinghou Jiao. 2023. "Vibration and Bandgap Behavior of Sandwich Pyramid Lattice Core Plate with Resonant Rings" Materials 16, no. 7: 2730. https://doi.org/10.3390/ma16072730

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop