Next Article in Journal
The Denseness of the Closure of Some Nyman–Beurling Linear Manifolds Implies the Absence of Zeroes of Certain Combinations of Riemann Zeta-Functions in the Critical Strip
Previous Article in Journal
Asymptotic Convergence of Solutions for Singularly Perturbed Linear Impulsive Systems with Full Singularity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Design of Stretch-Dominated Metamaterials Avoiding Bandgap Resonance

Space Structures Research Centre, College of Civil Engineering and Architecture, Zhejiang University, Hangzhou 310058, China
Symmetry 2025, 17(9), 1390; https://doi.org/10.3390/sym17091390
Submission received: 20 July 2025 / Revised: 16 August 2025 / Accepted: 19 August 2025 / Published: 26 August 2025
(This article belongs to the Section Engineering and Materials)

Abstract

Mechanical metamaterials subjected to dynamic loads within their bandgaps can still experience significant undesired structural responses, which are referred to as bandgap resonances. This paper proposes a design method for stretch-dominated metamaterials to avoid such a phenomenon. The metamaterials are modeled as pin-jointed bar structures, and the sufficient condition for preventing their bandgap resonances is derived: they must exhibit spatial inversion symmetry and satisfy certain boundary conditions. A matrix-form perturbation expression of the bandgap is then provided to generate expected bandgaps by adjusting the node coordinates and element cross-sectional areas of the unit cells under symmetry constraint. As an example, a two-dimensional metamaterial is designed to achieve an expected bandgap of 600–1000 Hz. The frequency-response analyses show that the sufficient condition ensures the suppression of bandgap resonances.

1. Introduction

Bandgap is a fundamental property of materials in solid-state physics governing their electrical conductivity characteristics. In the context of mechanical metamaterials, the presence of a bandgap in the dispersion function signifies that mechanical waves within a specific frequency range cannot propagate, constituting the basis for vibration isolation materials and topological insulators. Mechanical wave bandgaps are commonly categorized into Bragg bandgaps [1] and local resonant bandgaps [2], originating from the Bragg scattering and local resonator–wave interaction, respectively. Owing to the ability to isolate mechanical waves, bandgap mechanical metamaterials are broadly applicable in engineering vibration and sound insulation. Vasconcelos et al. [3] proposed a metamaterial-based interface to attenuate pressure waves induced by the impact hammer acting on the top of an offshore monopile. Zuo et al. [4] applied a star-shaped metamaterial to the shell of an underwater vehicle to reduce engine noise emission. A lattice-structured metamaterial capable of both sound insulation and ventilation was developed by Li et al. [5] and can be used as a roadside noise barrier. Seismic metamaterials composed of periodically arranged building foundations have been validated for shielding seismic waves and train-induced vibrations [6]. Replacing conventional aggregates with local resonant masses [7] enables concrete structural members to function as metamaterials capable of dissipating elastic waves induced by dynamic loads.
Dynamic loads within the bandgaps may still induce significant structural responses in the metamaterials, which are known as bandgap resonances [8]. For example, Zhang et al. [9] designed a negative stiffness metamaterial that exhibits a peak in its frequency-response curve within the bandgap range. Jiang et al. [10] designed negative stiffness metamaterials whose frequency-response curves do not exactly match their bandgaps. The chiral and hexagonal lattices proposed by Li et al. [11,12] also experience large frequency responses within the bandgap ranges. Bandgap resonances can similarly be observed in the frequency-response curves of the metamaterials studied in Refs. [13,14,15,16]. The presence of bandgap resonances suggests that certain eigenfrequencies of the metamaterial lie within the bandgap range, thereby compromising its vibration isolation performance. Therefore, it is necessary to prevent this phenomenon in the engineering design of bandgap metamaterials.
Bandgap resonance attracted early attention in the field of solid-state physics. In calculating the specific heat capacity of crystals, periodically arranged atoms are modeled as vibrating one-dimensional atomic chains. Thus, some early studies focused on the correspondence between the eigenfrequency spectra and the dispersion functions of finite atomic chains. For example, Wallis [17] identified an eigenfrequency within the bandgap while analyzing the spectrum of a finite diatomic chain. Recent studies continue to concentrate on the bandgap resonances in simple metamaterials. Based on topology energy band theory, Rosa et al. [18] proposed a method to analyze and manipulate the bandgap resonance of an elastic beam. Ba’ba’a et al. derived closed-form expressions for the characteristic equations governing the natural frequencies of a finite one-dimensional rod [19] and a mass-spring chain [20] using the transfer matrix method and analyzed the factors contributing to bandgap resonances. Park et al. [21] numerically and experimentally investigated how unit cell symmetry, boundary conditions, material/geometric properties, and the number of unit cells affect the existence of bandgap resonances in bilayer beams. The modified perturbed tridiagonal n-Toeplitz method is used by Ramakrishnan et al. [22] to estimate the bandgap resonance characteristics in 1D and 2D monoatomic phononic crystal lattices.
Calculating the dispersion function via Bloch’s theorem assumes that the metamaterials exhibit translational symmetry [23], necessitating that they be either infinitely extended or subjected to periodic boundary conditions [24]. However, finite metamaterials in engineering possess non-periodic boundaries, so their spectra deviate from the dispersion functions. This suggests that the existence of bandgap resonances is determined by the boundary conditions. Bastawrous et al. [8] introduced boundary conditions as perturbations to the dynamic stiffness matrix and analytically derived a closed-form condition for the existence of bandgap resonance in finite diatomic chains. Guo et al. [25] and Sugino et al. [26] solved dynamic differential equations with boundary conditions substituted to obtain the spectra of finite two-phase plates. Jin et al. [27] incorporated boundary conditions in the form of a Fourier series into the dynamic stiffness matrix and solved the spectrum of a two-phase metamaterial plate. By comparing the obtained spectra with the dispersion functions of the metamaterials, the existence of bandgap resonances can be identified.
Analytical calculations of the spectra of metamaterials require explicit expressions for the eigenfrequencies, which pose a substantial mathematical challenge. As a result, existing studies have primarily focused on simple systems [8,17,18,19,20,21,22,23,24,25,26,27], such as two-phase plates or diatomic chains. However, metamaterials designed for specific bandgaps are usually complex, making it difficult to analytically solve their spectra. Therefore, to ensure the vibration isolation performance of a designed metamaterial, bandgap resonance should preferably be avoided without solving the spectrum, which is one of the topics of this paper.
The application of mechanical metamaterials requires the design of material microstructures tailored to specific bandgap properties. Recent studies have employed topology optimization or deep learning to design metamaterials with expected bandgaps. Cool et al. [28] presented a topology optimization framework to obtain bandgaps simultaneously insolating acoustic and structural waves. Zhang et al. [29] employed a deep learning network to predict curved rod lattices with expected bandgap characteristics. Wan et al. [30] predicted two-dimensional phononic crystal lattices that satisfy expected bandgaps with a deep learning network. Metamaterials designed via these methods can exhibit expected bandgaps, but they do not necessarily avoid bandgap resonances. It is essential to conduct a theoretical analysis on the conditions for avoiding bandgap resonance, which can then be incorporated into the design of metamaterials.
This paper proposes a design method for stretch-dominated metamaterials that achieve expected bandgaps while avoiding bandgap resonance. The paper is organized as follows: Section 2 derives the sufficient condition for avoiding the bandgap resonance in stretch-dominated metamaterials. Section 3 presents the equations for designing the unit cell parameters under symmetry constraint. A design example is illustrated in Section 4, which is used to validate both the sufficient condition and the corresponding design method. Finally, Section 5 draws some conclusions.

2. Sufficient Condition for Avoiding Bandgap Resonance

A two-dimensional stretch-dominated metamaterial is used as an example to demonstrate the derivation of the sufficient condition for avoiding bandgap resonance. As shown in Figure 1, a 3 × 3 metamaterial, denoted by m, is modeled as a pin-jointed bar structure and subjected to a boundary condition that can be practically realized in engineering. Figure 2a and Figure 3a show the distributions of modal forces and displacements on the boundaries corresponding to any eigenfrequency ωj of m. Except for the corner nodes, the horizontal components of the node forces are zero on the top and bottom edges, while the vertical components are zero on the left and right edges. Similarly, the vertical components of the node displacements are zero on the top and bottom edges, and the horizontal components are zero on the left and right edges.
According to Figure 2a and Figure 3a, the boundary node displacement and force vectors of m can be written as d t / b = [0, 0, d t / b 2 x , 0, …, d t / b ( n t / b 1 ) x , 0, 0, 0]T, d l / r = [0, 0, 0, d l / r 2 y , …, 0, d l / r ( n l / r 1 ) y , 0, 0]T, f t / b = [ f t / b 1 x , f t / b 1 y 0, f t / b 2 y , …, 0, f t / b ( n t / b 1 ) y , f t / b ( n t / b ) x , f t / b ( n t / b ) y ]T, f l / r = [ f l / r 1 x , f l / r 1 y , f l / r 2 x , 0, …, f l / r ( n l / r 1 ) x , 0, f l / r ( n l / r ) x , f l / r ( n l / r ) y ]T, where d t / b and f t / b are the displacement and force vectors of the nodes on the top or bottom edge of m; d l / r and f l / r are the displacement and force vectors of the nodes on the left or right edge of m; d t / b u x and d t / b u y (u = 1, …, nt/b) are the x and y components of the displacement of the u-th node on the top or bottom edge, respectively, and f t / b u x and f t / b u y (u = 1, …, nt/b) are the corresponding force components; d l / r u x and d l / r u y (u = 1, …, nl/r) are the x and y components of the displacement of the u-th node on the left or right edge, and f l / r u x and f l / r u y (u = 1, …, nl/r) are the corresponding force components; nt/b is the number of nodes on the top or bottom edge, and nl/r is that on the left or right edge; and the superscript T denotes the transpose of a vector or matrix.
Spatial inversion of m in the x-direction (i.e., mirroring across the y-axis) yields the metamaterial mx. The distributions of modal forces and displacements on the boundaries are shown in Figure 2b and Figure 3b, and the corresponding frequency remains constant at ωj. According to Figure 2b and Figure 3b, the boundary node displacement and force vectors of mx can be written as d t / b x = [0, 0, d t / b ( n t / b 1 ) x , 0, …, d t / b 2 x , 0, 0, 0]T, d r / l x = [0, 0, 0, d l / r 2 y , …, 0, d l / r ( n l / r 1 ) y , 0, 0]T, f t / b x = [ f t / b ( n t / b ) x , f t / b ( n t / b ) y , 0, f t / b ( n t / b 1 ) y , …, 0, f t / b 2 y , f t / b 1 x , f t / b 1 y ]T, f r / l x = [ f l / r 1 x , f l / r 1 y , f l / r 2 x , 0, …, f l / r ( n l / r 1 ) x , 0, f l / r ( n l / r ) x , f l / r ( n l / r ) y ]T, where d t / b x and f t / b x are the displacement and force vectors of the nodes on the top or bottom edge of mx; and d r / l x and f r / l x are the displacement and force vectors of the nodes on the right or left edge of mx.
Spatial inversion of m in the y-direction (i.e., mirroring across the x-axis) yields the metamaterial my. The distributions of modal forces and displacements on the boundaries are shown in Figure 2c and Figure 3c, and the corresponding frequency remains constant at ωj. According to Figure 2c and Figure 3c, the boundary node displacement and force vectors of my can be written as d b / t y = [0, 0, d t / b 2 x , 0, …, d t / b ( n t / b 1 ) x , 0, 0, 0]T, d l / r y = [0, 0, 0, d l / r ( n l / r 1 ) y , …, 0, d l / r 2 y , 0, 0]T, f b / t y = [ f t / b 1 x , f t / b 1 y , 0, f t / b 2 y , …, 0, f t / b ( n t / b 1 ) y , f t / b ( n t / b ) x , f t / b ( n t / b ) y ]T, f l / r y = [ f l / r ( n l / r ) x , f l / r ( n l / r ) y ,  f l / r ( n l / r 1 ) x , 0, …, f l / r 2 x , 0, f l / r 1 x , f l / r 1 y ]T, where d b / t y and f b / t y are the displacement and force vectors of the nodes on the bottom or top edge of my; and d l / r y and f l / r y are the displacement and force vectors of the nodes on the left or right edge of my.
Spatial inversion of mx in the y-direction yields the metamaterial mxy. The distributions of modal forces and displacements on the boundaries are shown in Figure 2d and Figure 3d, and the corresponding frequency remains constant at ωj. According to Figure 2d and Figure 3d, the boundary node displacement and force vectors of mxy can be written as d b / t xy = [0, 0, d t / b ( n t / b 1 ) x , 0, …, d t / b 2 x , 0, 0, 0]T, d r / l xy = [0, 0, 0, d l / r ( n l / r 1 ) y , …, 0, d l / r 2 y , 0, 0]T, f b / t xy = [ f t / b ( n t / b ) x , f t / b ( n t / b ) y , 0, f t / b ( n t / b 1 ) y , …, 0, f t / b 2 y , f t / b 1 x , f t / b 1 y ]T, f r / l xy = [ f l / r ( n l / r ) x , f l / r ( n l / r ) y , f l / r ( n l / r 1 ) x , 0, …, f l / r 2 x , 0, f l / r 1 x , f l / r 1 y ]T, where d b / t xy and f b / t xy are the displacement and force vectors of the nodes on the bottom or top edge of mxy; and d r / l xy and f r / l xy are the displacement and force vectors of the nodes on the right or left edge of mxy.
As can be seen from Figure 2, the boundary node displacement vectors satisfy
d r = d l x ,   d r y = d l xy ,   d b = d t y ,   d b x = d t xy
Moreover, Figure 3 indicates that the boundary node force vectors satisfy the following relationships except for corner nodes:
f r + f l x = 0 ,   f r y + f l xy = 0 ,   f b + f t y = 0 ,   f b x + f t xy = 0
Let nodes 1, 2, 3, and 4 denote the lower right corner node of m, the lower left corner node of mx, the upper right corner node of my, and the upper left corner node of mxy, respectively. Then, their node forces satisfy
f 1 + f 2 + f 3 + f 4 = f b x n b f b y n b + f b x n b f b y n b + f b x n b f b y n b + f b x n b f b y n b = 0
Equations (1)–(3) indicate that the corresponding nodes on the bottom edge of m and the top edge of my have identical displacements, and their node forces are in equilibrium. The same relationship holds between the right edge of m and the left edge of mx, the top edge of mxy and the bottom edge of mx, and the right edge of my and the left edge of mxy. This suggests that m, mx, my, and mxy can be merged into a new structure M. Figure 4a,b show the distributions of its modal forces and displacements on the boundaries, and the corresponding frequency is still ωj.
The boundary node displacement and force vectors of M can be written as
d t M = [ 0 , 0 , d t 2 x , 0 , , d t ( n t 1 ) x , 0 , 0 , 0 , d t ( n t 1 ) x , 0 , , d t 2 x , 0 , 0 , 0 ] T
d b M = [ 0 , 0 , d t 2 x , 0 , , d t ( n t 1 ) x , 0 , 0 , 0 , d t ( n t 1 ) x , 0 , , d t 2 x , 0 , 0 , 0 ] T
d l M = [ 0 , 0 , 0 , d l 2 y , , 0 , d l ( n l 1 ) y , 0 , 0 , 0 , d l ( n l 1 ) y , , 0 , d l 2 y , 0 , 0 ] T
d r M = [ 0 , 0 , 0 , d l 2 y , , 0 , d l ( n l 1 ) y , 0 , 0 , 0 , d l ( n l 1 ) y , , 0 , d l 2 y , 0 , 0 ] T
f t M = [ f t 1 x , f t 1 y , 0 , f t 2 y , , , 0 , f t ( n t ) y , , 0 , f t 2 y , f t 1 x , f t 1 y ] T
f b M = [ f t 1 x , f t 1 y , 0 , f t 2 y , , , 0 , f t ( n t ) y , , 0 , f t 2 y , f t 1 x , f t 1 y ] T
f l M = [ f l 1 x , f l 1 y , f l 2 x , 0 , , f l ( n l / r ) x , 0 , , f l 2 x , 0 , f l 1 x , f l 1 y ] T
f l M = [ f l 1 x , f l 1 y , f l 2 x , 0 , , f l ( n l / r ) x , 0 , , f l 2 x , 0 , f l 1 x , f l 1 y ] T
where d t / b / l / r M and f t / b / l / r M are the displacement and force vectors of the nodes on the top, bottom, left, or right edge of M.
According to Equations (4)–(11), the node displacements and forces on the boundaries of M satisfy the periodic boundary condition [28]. However, M may not be a translationally symmetric metamaterial because m is not necessarily identical to mx, my, and mxy. Conversely, if m remains unchanged after the spatial inversions in both the x- and y-directions, M is a metamaterial satisfying the periodic boundary condition. This requires that the unit cell of m exhibits spatial inversion symmetry along the x- and y-axes. In this case, the spectrum of M must coincide with the dispersion function obtained via Bloch’s theorem, indicating that none of the eigenfrequencies of M, including ωj, lie within the bandgap. Since any eigenfrequency ωj of m lies outside the bandgap, the metamaterial m can avoid bandgap resonance.
The above derivation can be generalized to three-dimensional metamaterials and some other boundary conditions. For two- and three-dimensional cases, the applicable boundary conditions are listed in Table 1 and Table 2, respectively.
In Table 1 and Table 2, x, y, and z represent nodes on the boundaries normal to the x-, y- and z-directions, respectively; c1, c2 and c3 represent constraints applied in the x-, y- and z- directions, respectively.
In summary, the sufficient condition for avoiding bandgap resonance can be expressed as
Theorem 1.
A d-dimensional metamaterial can avoid bandgap resonance if both of the following conditions are satisfied: 1. the unit cell exhibits spatial inversion symmetry along all d-coordinate axes; 2. one of the boundary conditions listed in Table 1 or Table 2 is imposed.
The mechanism by which the above sufficient condition suppresses bandgap resonance is explained as follows. Bandgap resonances are in fact defect-induced modes localized at symmetry-breaking defects [22]. The spatial inversion symmetry of the metamaterial ensures the absence of bulk defects and their associated defect modes. The boundaries of a finite metamaterial also act as defects, giving rise to vibrational modes localized at the edges. Imposing the boundary conditions listed in Table 1 or Table 2 constrains the displacements of the boundary nodes, which eliminates these modes and suppresses bandgap resonances.

3. Perturbation of Bandgap

The following generalized characteristic equation [31,32,33,34] can be used to solve the dispersion function of the metamaterial consisting of unit cells with spatial inversion symmetry:
D ^ ( k ) d ^ ( k ) = 0
where D ^ ( k ) = T d ( k ) H [ K ω ( k ) 2 M ] T d ( k ) ; K and M are the d·n × d·n stiffness and mass matrices of the unit cell, respectively, and their explicit forms can be found in Ref. [34]; n is the number of nodes in the unit cell; d is the dimension of the unit cell; the superscript H denotes the Hermite transpose; T d ( k ) is the d·n × d·r reduction matrix introducing Bloch’s theorem [34], and r is the number of inequivalent nodes [34]; ω(k) is the circular frequency, and k is the d × 1 wave vector [35]; d ^ ( k ) = T d ( k ) d and d is the d·n × 1 node displacement vector of the unit cell. Once k is given, the generalized eigenvalues ωj(k)2 (j = 1, 2, …, d·r) and their corresponding eigenvectors d ^ j ( k ) (d·r × 1) can be solved via Equation (12). Thus, the dispersion function of the metamaterial, i.e., the kω relation, is established with a total of d·r branches of kωj.
According to matrix perturbation theory [33,34], the perturbation of branch j caused by the increments of unit cell parameters, including node coordinates and element cross-sectional areas, can be calculated by
Δ ω j ( k ) = O j ( k ) Δ ε
where Δ ε = { Δ A 1 , Δ A 2 , , Δ A b , Δ x 1 1 , , Δ x 1 d , , Δ x n 1 , , Δ x n d } T is the (b + d·n) × 1 incremental unit cell parameter vector; b is the number of elements in the unit cell; Ai (i = 1,…, b) is the cross-sectional area of the i-th element; and x p k (p = 1, …, n, k = 1, …, d) is the k-th coordinate of the p-th node. Oj(k) is a 1 × (b + d·n) row vector whose v-th component is
( O j ( k ) ) v = 1 2 ω j ( k ) d ^ j ¯ ( k ) H D ^ ( k ) A v d ^ j ¯ ( k ) ,   v b 1 2 ω j ( k ) d ^ j ¯ ( k ) H D ^ ( k ) x p k d ^ j ¯ ( k ) ,   v > b
where v = b + d·(p−1) + k for the second term; d ^ j ¯ ( k ) is the eigenvector normalized by d ^ j ¯ ( k ) H T d ( k ) H M T d ( k ) d ^ j ¯ ( k ) = 1 ; D ^ ( k ) A v = T d ( k ) H [ K A v ω j ( k ) 2 M A v ] T d ( k ) , D ^ ( k ) x p k = T d ( k ) H [ K x p k ω j ( k ) 2 M x p k ] T d ( k ) . The explicit forms of K A v , K x p k , M A v , and M x p k are given in Ref. [34].
An illustrative bandgap between the lower branch ωl(k) and the upper branch ωu(k) is shown in Figure 5. The range of the bandgap is [ωl(kl), ωu(ku)], where kl and ku are the wave vectors corresponding to the maximum value of ωl(k) and the minimum value of ωu(k), respectively. Substituting kl and ku into Equation (13), the perturbation of ωl(kl) and ωu(ku) can be expressed as
Δ ω = Δ ω u ( k u ) Δ ω l ( k l ) = O u ( k u ) O l ( k l ) Δ ε = O Δ ε
The difference between the expected bandgap [ωl, ωu] shown in Figure 5 and the current bandgap is Δωt = {ωuωu(ku), ωlωl(kl)}T. By substituting Δωt into Equation (15), Δε for adjusting unit cell parameters and tuning bandgap can be solved. However, Δε must satisfy spatial inversion symmetry, or it will break the symmetry of the unit cell. Figure 6 shows a unit cell with eight nodes and 12 elements. An increment vector adjusting the coordinates of nodes 5, 6, 7, and 8 with Δ x 5 1 = 1 , Δ x 6 1 = 1 , Δ x 7 1 = 1 , and Δ x 8 1 = 1 satisfies spatial inversion symmetry and is denoted as Δεg. Similarly, ΔA5 = 1, ΔA6 = 1, ΔA7 = 1, and ΔA8 = 1 can be assembled into another Δεg satisfying spatial inversion symmetry. If there are at most τ linearly independent Δεg, then any incremental unit cell parameter vector satisfying spatial inversion symmetry can be expressed as
Δ ε sym = { Δ ε g 1 , Δ ε g 1 , Δ ε g τ } { κ 1 , κ 2 , , κ τ } T = Ε g Δ κ
where Δκ is an arbitrary τ × 1 coefficient vector.
Substituting Equation (16) and Δωt into Equation (15), it can be obtained that
Δ ω t = O Ε g Δ κ
The least squares solution [36] of Equation (17) is
Δ κ = ( O Ε g ) T [ ( O Ε g ) ( O Ε g ) T ] 1 Δ ω t
By substituting Equation (18) into Equation (16), the resulting Δεsym can be used to adjust the unit cell parameters and tune the bandgap under the constraint of spatial inversion symmetry.
The unit cell parameters should be modified in small steps due to the strong nonlinearity in their relationship with the dispersion function. Thus, Δεsym is scaled by
Δ ε s = s Δ ε sym | | Δ ε sym | |
where s is a specified scaling factor; ||•|| denotes the two-norm of the vector. The expected bandgap can be obtained by iteratively adjusting the unit cell parameters with Δεs until ||Δωt|| falls below a specified tolerance c.

4. Example

Figure 7 shows a square unit cell with a side length of 10 mm that exhibits spatial inversion symmetry along the x- and y-axes. The origin of the coordinate system is at the center of the unit cell with the axes shown in Figure 7. The node coordinates are listed in Table 3. All the elements of the unit cell have a circular cross-section with a radius of 0.3 mm (area A = 0.283 mm2) and are made of plastic with a Young’s modulus of E = 1.2 GPa and a density of ρ = 0.9 g/cm3.
For this unit cell, the dispersion function obtained using Equation (12) consists of 52 branches, whose significant overlap makes it difficult to depict their surfaces on a plane. To illustrate the band structure, a frequency distribution function for branch j can be constructed as follows:
D j ( ω ) = Π d A ω j ( k ) = ω ( j = 1 , 2 , , d r )
where Π denotes the first Brillouin zone [35]; dA = dkxdky is the area of the infinitesimal element of Π; and kx and ky are the x and y components of k, respectively. According to Equation (20), Dj(ω) represents the total area of the infinitesimal elements for which ωj = ω. If Dj(ω) ≠ 0, there must exist a wave vector k such that ωj(k) = ω. Therefore, ωj(k) lies within the frequency domain with nonzero Dj(ω). The frequency distribution of the dispersion function can be visualized by plotting Dj(ω) for all branches, with ω as the horizontal axis and area as the vertical axis. Figure 8 shows 27 lower-frequency branches distributed between 0 and 3100 Hz.
This example aims to achieve a bandgap of 600–1000 Hz by designing the coordinates of nodes 5–29 and the cross-sectional areas of all elements. According to Figure 8, the expected bandgap is most likely generated between branches 3 and 4. The unit cell parameters are iteratively adjusted using Equations (15)–(19) with a tolerance of c = 2 and a scaling factor of s = 0.01. The resulting unit cell is shown in Figure 9, whose element cross-sectional areas and node coordinates are listed in Table 4 and Table 5, respectively. The frequency distribution of the updated dispersion function is shown in Figure 10, where a bandgap of 600.0–998.6 Hz can be found.
A metamaterial consisting of 4 × 4 designed unit cells is shown in Figure 11a. The metamaterial is denoted as PBC, FBC, or SBC when the periodic, free, or specific boundary condition is imposed. The specific boundary condition is illustrated in Figure 11b and corresponds to case 3 in Table 1. In addition, the symmetry of the designed unit cell can be slightly broken by adjusting the coordinate of node 15 to (−4.03, 1.00), resulting in a bandgap of 601.0–904.7 Hz. The 4 × 4 metamaterial, consisting of this asymmetric unit cell and subjected to the specific boundary condition, is denoted as SBCA.
A frequency-response function is defined as follows for PBC, FBC, SBC, or SBCA:
y ( ω ) = 20 log 10 ( | | d out ( ω ) | | | | d in ( ω ) | | )
where ||din(ω)|| is the displacement amplitude at input node 1, and ||dout(ω)|| is that at output node 2 or 3. These nodes are labeled in Figure 11a. With a modal damping of 0.02, ||din(ω)|| and ||dout(ω)|| are calculated using the modal superposition method (MS) [35] and the finite element software ANSYS 2021 R2, respectively.
The MS is implemented in MATLAB 2024a, and the procedure includes the following steps:
  • Set up the dynamic equation for the metamaterial, with its specific form given in Ref. [34].
  • Calculate the eigenfrequencies and eigenmodes of the metamaterial.
  • Calculate the displacements at the input and output nodes, i.e., din and dout, under a simple harmonic load using the displacement equation [37] of the modal superposition method.
The metamaterials are also modeled in ANSYS with all the elements simulated by Link180. A harmonic response analysis is conducted using the command HROPT to obtain the displacements din and dout, with the solution method [38] set to FULL.
The frequency-response curves and eigenfrequencies of PBC, FBC, SBC, and SBCA are shown in Figure 12, Figure 13, Figure 14 and Figure 15, where the dashed lines indicate that the dout obtained by ANSYS is zero. As illustrated in Figure 12, the eigenfrequencies and frequency-response curves of PBC are consistent with the bandgap, confirming that the metamaterial with the expected bandgap has been successfully designed. However, when the metamaterial is subjected to the free boundary condition, eigenfrequencies emerge within the bandgap, as illustrated in Figure 13. Figure 14 shows that the bandgap resonances in SBC are effectively suppressed by imposing the specific boundary condition, thereby validating the second condition of Theorem 1. For the asymmetric metamaterial, even when the specific boundary condition is imposed, bandgap resonances still emerge in the frequency-response curves of Figure 15. This validates the first condition of Theorem 1.

5. Conclusions and Outlook

This paper proposes a design method for stretch-dominated metamaterials that exhibit expected bandgaps while avoiding bandgap resonance. The conclusions are summarized as follows:
First, the sufficient condition for avoiding bandgap resonance is derived, specifically including spatial inversion symmetry and the boundary conditions listed in either Table 1 or Table 2. The symmetry constraint should be incorporated into the design process of bandgap metamaterials to ensure their vibration isolation performance.
Second, equations for adjusting unit cell parameters are formulated to obtain expected bandgaps. The matrix-form perturbation expression of the bandgap is established, which represents the sensitivity of the bandgap with respect to node coordinates and element cross-sectional areas. It is then used to design the unit cell under the symmetry constraint to generate the expected bandgap.
Third, a two-dimensional metamaterial example is presented to validate the proposed sufficient condition and design method. The designed metamaterial, satisfying spatial inversion symmetry and the boundary condition in Table 1, achieves both the expected bandgap and the suppression of bandgap resonance, thus validating the proposed design method. However, bandgap resonances emerge once either of these two conditions is not satisfied. This validates the sufficient condition for avoiding bandgap resonance.
Although the derived condition is sufficient for avoiding bandgap resonance, the symmetry constraint is overly restrictive, thus significantly reducing the number of design parameters. In addition, although the boundary conditions listed in Table 1 and Table 2 are feasible in engineering practice, they substantially increase the complexity of the metamaterials. Future work should further investigate more readily implementable conditions for avoiding bandgap resonance, including weaker symmetry constraints and simpler boundary conditions.
Stretch-dominated metamaterials represent one class of mechanical metamaterials with significant application potential, alongside bend-dominated, origami, kirigami, honeycomb, and sandwich structures. The sufficient condition for avoiding bandgap resonance and the corresponding design method proposed in this study cannot be directly extended to these metamaterials. For instance, the perturbation expression for designing the bandgap of bend-dominated metamaterials must be re-derived based on beam elements. Future work will further investigate the bandgap resonances and design methods for metamaterials of other structural forms.

Funding

This research received no external funding.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Acknowledgments

Sincere thanks to Deng Hua for the guidance on my research during my graduate studies.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Sigalas, M.M.; Economou, E.N. Elastic and acoustic wave band structure. J. Sound Vib. 1992, 158, 377–382. [Google Scholar] [CrossRef]
  2. 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]
  3. Azevedo Vasconcelos, A.C.; Valiya Valappil, S.; Schott, D.; Jovanova, J.; Aragón, A.M. A Metamaterial-Based Interface for the Structural Resonance Shielding of Impact-Driven Offshore Monopiles. Eng. Struct. 2024, 300, 117261. [Google Scholar] [CrossRef]
  4. Zuo, Y.; Yang, D.; Zhuang, Y. Broadband Transient Vibro-Acoustic Prediction and Control for the Underwater Vehicle Power Cabin with Metamaterial Components. Ocean Eng. 2024, 298, 117121. [Google Scholar] [CrossRef]
  5. Li, X.; Zhao, M.; Yu, X.; Wei Chua, J.; Yang, Y.; Lim, K.M.; Zhai, W. Multifunctional and Customizable Lattice Structures for Simultaneous Sound Insulation and Structural Applications. Mater. Des. 2023, 234, 112354. [Google Scholar] [CrossRef]
  6. Zheng, H.; Miao, L.; Xiao, P.; Lei, K.; Wang, Q. Novel Metamaterial Foundation with Multi Low-Frequency Bandgaps for Isolating Earthquakes and Train Vibrations. Structures 2024, 61, 106070. [Google Scholar] [CrossRef]
  7. Mitchell, S.J.; Pandolfi, A.; Ortiz, M. Metaconcrete: Designed Aggregates to Enhance Dynamic Performance. J. Mech. Phys. Solids 2014, 65, 69–81. [Google Scholar] [CrossRef]
  8. Bastawrous, M.V.; Hussein, M.I. Closed-Form Existence Conditions for Bandgap Resonances in a Finite Periodic Chain under General Boundary Conditions. J. Acoust. Soc. Am. 2022, 151, 286–298. [Google Scholar] [CrossRef]
  9. Zhang, K.; Qi, L.; Zhao, P.; Zhao, C.; Deng, Z. Buckling Induced Negative Stiffness Mechanical Metamaterial for Bandgap Tuning. Compos. Struct. 2023, 304, 116421. [Google Scholar] [CrossRef]
  10. Jiang, T.; Han, Q.; Li, C. Design and Compression-Induced Bandgap Evolution of Novel Polygonal Negative Stiffness Metamaterials. Int. J. Mech. Sci. 2024, 261, 108658. [Google Scholar] [CrossRef]
  11. Li, X.; Cheng, S.; Wang, R.; Yan, Q.; Wang, B.; Sun, Y.; Yan, H.; Zhao, Q.; Xin, Y. Design of Novel Two-Dimensional Single-Phase Chiral Phononic Crystal Assembly Structures and Study of Bandgap Mechanism. Results Phys. 2023, 48, 106431. [Google Scholar] [CrossRef]
  12. Li, X.; Cheng, S.; Yang, H.; Yan, Q.; Wang, B.; Xin, Y.; Sun, Y.; Ding, Q.; Yan, H.; Li, Y.; et al. Integrated Analysis of Bandgap Optimization Regulation and Wave Propagation Mechanism of Hexagonal Multi-Ligament Derived Structures. Eur. J. Mech. A-Solid 2023, 99, 104952. [Google Scholar] [CrossRef]
  13. Yang, H.; Cheng, S.; Li, X.; Yan, Q.; Wang, B.; Xin, Y.; Sun, Y.; Ding, Q.; Yan, H.; Zhao, Q. Propagation Mechanism of Low-Frequency Elastic Waves and Vibrations in a New Tetragonal Hybrid Metamaterial. Int. J. Solids Struct. 2023, 285, 112536. [Google Scholar] [CrossRef]
  14. Cheng, S.; Yang, H.; Yan, Q.; Wang, B.; Sun, Y.; Xin, Y.; Ding, Q.; Yan, H.; Wang, L. Study on the Band Gap and Directional Wave Propagation Mechanism of Novel Single-Phase Metamaterials. Phys. B 2023, 650, 414545. [Google Scholar] [CrossRef]
  15. Yao, D.; Xiong, M.; Luo, J.; Yao, L. Flexural Wave Mitigation in Metamaterial Cylindrical Curved Shells with Periodic Graded Arrays of Multi-Resonator. Mech. Syst. Signal. Process. 2022, 168, 108721. [Google Scholar] [CrossRef]
  16. Ruan, H.; Hou, J.; Li, D. Wave Propagation Characterization of 2D Composite Chiral Lattice Structures with Circular Plate Inclusions. Eng. Struct. 2022, 264, 114466. [Google Scholar] [CrossRef]
  17. Wallis, R.F. Effect of Free Ends on the Vibration Frequencies of One-Dimensional Lattices. Phys. Rev. 1957, 105, 540. [Google Scholar] [CrossRef]
  18. Rosa, M.I.N. Material vs. Structure: Topological Origins of Band-Gap Truncation Resonances in Periodic Structures. Phys. Rev. Mater. 2023, 7, 124201. [Google Scholar] [CrossRef]
  19. Ba’ba’a, H.B.A.; Willey, C.L.; Chen, V.W.; Juhl, A.T.; Nouh, M. Theory of Truncation Resonances in Continuum Rod-Based Phononic Crystals with Generally Asymmetric Unit Cells. Adv. Theor. Simul. 2023, 6, 2200700. [Google Scholar] [CrossRef]
  20. Ba’ba’a, H.B.; Yousef, H.; Nouh, M. A Blueprint for Truncation Resonance Placement in Elastic Diatomic Lattices with Unit Cell Asymmetry. JASA Express Lett. 2024, 4, 077501. [Google Scholar] [CrossRef]
  21. Park, S.; Yan, R.F.; Matlack, K.H. Characteristics of Truncation Resonances in Periodic Bilayer Rods and Beams with Symmetric and Asymmetric Unit Cells. J. Acoust. Soc. Am. 2024, 155, 791–802. [Google Scholar] [CrossRef]
  22. Ramakrishnan, V.; Matlack, K.H. A Quantitative Study of Energy Localization Characteristics in Defect-Embedded Monoatomic Phononic Crystals. J. Sound Vib. 2025, 614, 119164. [Google Scholar] [CrossRef]
  23. Huang, K. Solid State Physics; Higher Education Press: Beijing, China, 1988. [Google Scholar]
  24. Hussein, M.I.; Leamy, M.J.; Ruzzene, M. Dynamics of Phononic Materials and Structures: Historical Origins, Recent Progress, and Future Outlook. Appl. Mech. Rev. 2014, 66, 040802. [Google Scholar] [CrossRef]
  25. Guo, Z.; Sheng, M.; Pan, J. Effect of Boundary Conditions on the Band-Gap Properties of Flexural Waves in a Periodic Compound Plate. J. Sound Vib. 2017, 395, 102–126. [Google Scholar] [CrossRef]
  26. Sugino, C.; Ruzzene, M.; Erturk, A. An Analytical Framework for Locally Resonant Piezoelectric Metamaterial Plates. Int. J. Solids Struct. 2020, 182–183, 281–294. [Google Scholar] [CrossRef]
  27. Jin, G.; Zhang, C.; Ye, T.; Zhou, J. Band Gap Property Analysis of Periodic Plate Structures under General Boundary Conditions Using Spectral-Dynamic Stiffness Method. Appl. Acoust. 2017, 121, 1–13. [Google Scholar] [CrossRef]
  28. Cool, V.; Sigmund, O.; Aage, N. Metamaterial Design with Vibroacoustic Bandgaps through Topology Optimization. Comput. Methods Appl. Mech. Eng. 2025, 436, 117744. [Google Scholar] [CrossRef]
  29. Zhang, K.; Guo, Y.; Liu, X.; Hong, F.; Hou, X.; Deng, Z. Deep Learning-Based Inverse Design of Lattice Metamaterials for Tuning Bandgap. Extrem. Mech. Lett. 2024, 69, 102165. [Google Scholar] [CrossRef]
  30. Wan, X.-H.; Zhang, Y.; Guo, Q.-H.; Zheng, L.-Y. Deep Learning-Based Inverse Design of Irregular Phononic Crystals. Int. J. Mech. Sci. 2025, 297–298, 110335. [Google Scholar] [CrossRef]
  31. Phani, A.S.; Woodhouse, J.; Fleck, N.A. Wave Propagation in Two-Dimensional Periodic Lattices. J. Acoust. Soc. Am. 2006, 119, 1995–2005. [Google Scholar] [CrossRef] [PubMed]
  32. Messner, M.C.; Barham, M.I.; Kumar, M.; Barton, N.R. Wave Propagation in Equivalent Continuums Representing Truss Lattice Materials. Int. J. Solids Struct. 2015, 73–74, 55–66. [Google Scholar] [CrossRef]
  33. Chen, S. Matrix Perturbation Theory in Structural Dynamic Design; Science Press: Beijing, China, 1999. [Google Scholar]
  34. Wang, Z.; Deng, H.; Liu, H. A Matrix Method for Designing Bandgap of Stretch-Dominated Mechanical Metamaterials. J. Sound Vib. 2025, 617, 119257. [Google Scholar] [CrossRef]
  35. Grosso, G.; Parravicini, G.P. Solid State Physics; Elsevier: Amsterdam, The Netherlands, 2000. [Google Scholar]
  36. Ben-Israel, A.; Greville, T.N.E. Generalized Inverses: Theory and Applications; Springer Science & Business Media: New York, NY, USA, 2003; ISBN 978-0-387-00293-4. [Google Scholar]
  37. Chopra, A.K. Dynamics of Structures; Pearson Education: London, UK, 2007. [Google Scholar]
  38. Coupled Field Harmonic, Harmonic Acoustics, and Harmonic Response Options. Available online: https://ansyshelp.ansys.com/public/account/secured?returnurl=/Views/Secured/corp/v242/en/wb_sim/ds_options_harmonic.html?q=hropt (accessed on 8 March 2025).
Figure 1. A 3 × 3 metamaterial with specific boundary condition.
Figure 1. A 3 × 3 metamaterial with specific boundary condition.
Symmetry 17 01390 g001
Figure 2. Distributions of modal forces on the boundaries of 3 × 3 metamaterials: (a) m; (b) mx; (c) my; (d) mxy.
Figure 2. Distributions of modal forces on the boundaries of 3 × 3 metamaterials: (a) m; (b) mx; (c) my; (d) mxy.
Symmetry 17 01390 g002
Figure 3. Distributions of modal displacements on the boundaries of 3 × 3 metamaterials: (a) m; (b) mx; (c) my; (d) mxy.
Figure 3. Distributions of modal displacements on the boundaries of 3 × 3 metamaterials: (a) m; (b) mx; (c) my; (d) mxy.
Symmetry 17 01390 g003
Figure 4. Distributions of modal forces and displacements on the boundaries of M: (a) forces; (b) displacements.
Figure 4. Distributions of modal forces and displacements on the boundaries of M: (a) forces; (b) displacements.
Symmetry 17 01390 g004
Figure 5. An illustrative bandgap.
Figure 5. An illustrative bandgap.
Symmetry 17 01390 g005
Figure 6. A unit cell.
Figure 6. A unit cell.
Symmetry 17 01390 g006
Figure 7. Initial unit cell.
Figure 7. Initial unit cell.
Symmetry 17 01390 g007
Figure 8. Frequency distribution of the dispersion function for the initial unit cell.
Figure 8. Frequency distribution of the dispersion function for the initial unit cell.
Symmetry 17 01390 g008
Figure 9. The unit cell satisfying expected bandgap.
Figure 9. The unit cell satisfying expected bandgap.
Symmetry 17 01390 g009
Figure 10. Frequency distribution of the dispersion function for the unit cell satisfying expected bandgap.
Figure 10. Frequency distribution of the dispersion function for the unit cell satisfying expected bandgap.
Symmetry 17 01390 g010
Figure 11. Metamaterial consisting of designed unit cells and a boundary condition satisfying Table 1: (a) 4 × 4 metamaterial; (b) boundary condition satisfying Table 1.
Figure 11. Metamaterial consisting of designed unit cells and a boundary condition satisfying Table 1: (a) 4 × 4 metamaterial; (b) boundary condition satisfying Table 1.
Symmetry 17 01390 g011
Figure 12. Frequency-response curves and eigenfrequencies of PBC.
Figure 12. Frequency-response curves and eigenfrequencies of PBC.
Symmetry 17 01390 g012
Figure 13. Frequency-response curves and eigenfrequencies of FBC.
Figure 13. Frequency-response curves and eigenfrequencies of FBC.
Symmetry 17 01390 g013
Figure 14. Frequency-response curves and eigenfrequencies of SBC.
Figure 14. Frequency-response curves and eigenfrequencies of SBC.
Symmetry 17 01390 g014
Figure 15. Frequency-response curves and eigenfrequencies of SBCA.
Figure 15. Frequency-response curves and eigenfrequencies of SBCA.
Symmetry 17 01390 g015
Table 1. Boundary conditions of two-dimensional metamaterials for avoiding bandgap resonance.
Table 1. Boundary conditions of two-dimensional metamaterials for avoiding bandgap resonance.
CaseBoundary Condition
xy
1c1c1
2c2c1
3c1c2
4c2c2
Table 2. Boundary conditions of three-dimensional metamaterials for avoiding bandgap resonance.
Table 2. Boundary conditions of three-dimensional metamaterials for avoiding bandgap resonance.
CaseBoundary Condition
xyz
1c2, c3c1, c3c1, c2
2c1c1, c3c1, c2
3c2, c3c2c1, c2
4c1c2c1, c2
5c2, c3c1, c3c3
6c1c1, c3c3
7c2, c3c2c3
8c1c2c3
Table 3. Node coordinates of initial unit cell (mm).
Table 3. Node coordinates of initial unit cell (mm).
NodeCoordinateNodeCoordinateNodeCoordinate
x-y-x-y-x-y-
1−5.005.0011−2.001.0021−2.00−1.00
25.005.00120.001.00220.00−1.00
35.00−5.00132.001.00232.00−1.00
4−5.00−5.00144.001.00244.00−1.00
5−4.003.0015−4.000.0025−4.00−3.00
6−2.003.0016−2.000.0026−2.00−3.00
70.003.00170.000.00270.00−3.00
82.003.00182.000.00282.00−3.00
94.003.00194.000.00294.00−3.00
10−4.001.0020−4.00−1.00
Table 4. Element cross-sectional areas of the unit cell satisfying expected bandgap (mm2).
Table 4. Element cross-sectional areas of the unit cell satisfying expected bandgap (mm2).
ElementAreaElementAreaElementAreaElementArea
1-70.1071-20.00113-140.22023-180.276
1-60.2293-40.00110-150.23324-180.229
1-50.1592-30.00010-160.22924-190.233
1-100.1051-40.00011-160.27620-210.220
1-150.0385-60.23111-170.62221-220.458
2-70.1076-70.30512-170.32122-230.458
2-80.2297-80.30513-170.62223-240.220
2-90.1598-90.23113-180.27625-200.157
2-140.1055-100.15714-180.22925-210.387
2-190.0385-110.38714-190.23326-210.250
3-270.1076-110.25015-160.20226-220.509
3-280.2296-120.50916-170.26127-220.246
3-290.1597-120.24617-180.26128-220.509
3-240.1058-120.50918-190.20228-230.250
3-190.0388-130.25020-150.23329-230.387
4-270.1079-130.38720-160.22929-240.157
4-260.2299-140.15721-160.27625-260.231
4-250.15910-110.22021-170.62226-270.305
4-200.10511-120.45822-170.32127-280.305
4-150.03812-130.45823-170.62228-290.231
Table 5. Node coordinates of the unit cell satisfying expected bandgap (mm).
Table 5. Node coordinates of the unit cell satisfying expected bandgap (mm).
NodeCoordinateNodeCoordinateNodeCoordinate
x-y-x-y-x-y-
1−5.005.0011−1.871.0121−1.87−1.01
25.005.00120.001.21220.00−1.21
35.00−5.00131.871.01231.87−1.01
4−5.00−5.00143.950.98243.95−0.98
5−3.922.8715−4.030.0025−3.92−2.87
6−1.853.0116−2.090.0026−1.85−3.01
70.003.14170.000.00270.00−3.14
81.853.01182.090.00281.85−2.99
93.922.87194.030.00293.92−2.87
10−3.950.9820−3.95−0.98
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

Wang, Z. Design of Stretch-Dominated Metamaterials Avoiding Bandgap Resonance. Symmetry 2025, 17, 1390. https://doi.org/10.3390/sym17091390

AMA Style

Wang Z. Design of Stretch-Dominated Metamaterials Avoiding Bandgap Resonance. Symmetry. 2025; 17(9):1390. https://doi.org/10.3390/sym17091390

Chicago/Turabian Style

Wang, Zijian. 2025. "Design of Stretch-Dominated Metamaterials Avoiding Bandgap Resonance" Symmetry 17, no. 9: 1390. https://doi.org/10.3390/sym17091390

APA Style

Wang, Z. (2025). Design of Stretch-Dominated Metamaterials Avoiding Bandgap Resonance. Symmetry, 17(9), 1390. https://doi.org/10.3390/sym17091390

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