Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Next Article in Journal
The Acoustic and Cultural Heritage of the Banda Primitiva de Llíria Theater: Objective and Subjective Evaluation
Previous Article in Journal
Thermal Effects on Prestress Loss in Pretensioned Concrete Girders
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Frame-Type Seismic Surface Wave Barrier with Ultra-Low-Frequency Bandgaps for Rayleigh Waves

1
College of Hydraulic and Civil Engineering, Xinjiang Agricultural University, Urumqi 830052, China
2
College of Civil Engineering, Hefei University of Technology, Hefei 230009, China
*
Author to whom correspondence should be addressed.
Buildings 2024, 14(8), 2328; https://doi.org/10.3390/buildings14082328
Submission received: 18 June 2024 / Revised: 18 July 2024 / Accepted: 25 July 2024 / Published: 27 July 2024
(This article belongs to the Section Building Structures)

Abstract

:
Seismic surface waves carry significant energy that poses a major threat to structures and may trigger damage to buildings. To address this issue, the implementation of periodic barriers around structures has proven effective in attenuating seismic waves and minimizing structural dynamic response. This paper introduces a framework for seismic surface wave barriers designed to generate multiple ultra-low-frequency band gaps. The framework employs the finite-element method to compute the frequency band gap of the barrier, enabling a deeper understanding of the generation mechanism of the frequency band gap based on vibrational modes. Subsequently, the transmission rates of elastic waves through a ten-period barrier were evaluated through frequency–domain analysis. The attentional effects of the barriers were investigated by the time history analysis using site seismic waves. Moreover, the influence of the soil damping and material damping are separately discussed, further enhancing the assessment. The results demonstrate the present barrier can generate low-frequency band gaps and effectively attenuate seismic surface waves. These band gaps cover the primary frequencies of seismic surface waves, showing notable attenuation capabilities. In addition, the soil damping significantly contributes to the attenuation of seismic surface waves, resulting in an attenuation rate of 50%. There is promising potential for the application of this novel isolation technology in seismic engineering practice.

1. Introduction

Earthquakes release vast amounts of energy that can inflict significant damage to buildings [1,2,3,4], with the primary source of impact being the seismic waves they generate. Seismic waves can be categorized into surface waves and body waves, while seismic surface waves have the majority of the released energy [5,6,7]. However, the surface waves pose a significant threat to the safety of structures due to the propagation and attenuation characteristics of the seismic surface waves. Seismic barriers have proven to be an effective solution to mitigate the effect of seismic waves on existing buildings. However, conventional barriers with dimensions of tens of meters can only attenuate low-frequency seismic surface waves, which has a limitation in practical engineering [8,9]. Therefore, it is necessary to develop a small-sized barrier for reducing seismic waves.
Phononic crystals [10,11,12,13,14,15,16,17] represent periodic structures composed of two or more elastic solids. These crystals exhibit the ability to influence the propagation of elastic waves by generating distinctive dispersion relationships and frequency bandgaps [18,19,20,21,22,23,24,25]. The bandgap mechanism can be classified into two principal types: Bragg scattering [11,26,27,28,29,30,31,32,33] and local resonance [8,30,34,35,36,37]. Bragg scattering involves the scattering of elastic waves when they encounter the barrier structure. On the other hand, local resonance is the vibration within the barrier structure when an elastic wave occurs, attenuating the energy carried by the elastic wave. The phononic crystals can reduce the size of the barrier against the attenuating seismic surface waves and meet practical engineering applications. Consequently, numerous researchers [38,39,40,41,42] have incorporated phononic crystals into the design of seismic surface wave barriers.
Numerous researchers have focused on the development of seismic surface wave barriers. A 1D periodic composite wall structure composed of concrete and rubber was used to attenuate seismic waves due to the low-frequency band gaps. Enhanced attenuation capabilities were observed when these structures were arranged in positive and negative gradients [34,43,44]. However, the practical implementation of such barriers with a large size poses significant engineering challenges. Consequently, scholars [45,46,47,48,49] have turned their attention towards the exploration of 2D and 3D barrier configurations, seeking innovative solutions to overcome size constraints and maximize effectiveness in mitigating seismic wave impacts.
Huang et al. [28] and Cheng et al. [26] have presented results showing that square columns exhibit superior efficacy in attenuating elastic waves compared to cylindrical columns, whereas hexagonal columns yielded wider bandgaps. Du et al. [27] and Zeng et al. [33] have conducted comparisons on the bandgap properties of hollow, filled, and solid columns, revealing that hollow and filled columns possess the capacity to generate wider bandgaps. Hollow columns filled with various materials can generate different sizes of frequency band gaps [27,29,30,33,50]. The results revealed that a greater difference in Young’s modulus between the filler material and the outer frame of the composited structure leads to some wider frequency bandgaps at low frequencies.
To expand the low-frequency bandgap of elastic waves, a cross-shaped trenchless structure was developed to generate a low-frequency bandgap [30], which offered insights for designing shielding structures against seismic surface waves. The soil body at the real site consists of multiple soil layers rather than a single layer of soil. Some researchers [31,51] have constructed a multi-layer soil model to obtain a more realistic numerical simulation result. It was indicated that the elastic wave is concentrated in the first two layers of soil in the multi-layer soil model, while the lower layer of soil is almost unaffected by the propagation of the elastic waves. Refraction and reflection phenomena of elastic waves occur predominantly within these upper layers, forming the accumulation of energy at the soil surface. Analysis of the transmission spectrum revealed a notable mitigation in the efficacy of barriers for attenuating elastic waves, which explains the complexity of seismic wave propagation through stratified soil structures.
Furthermore, researchers [21,31,52,53,54,55] have designed two-dimensional columnar periodic structures such as “I” and “T” types as well as columnar periodic structures with bases. However, the findings show that these structures can produce some small ranges of bandgap within the ultra-low frequency. Li et al. [56] proposed a resonator structure featuring two internal vibrators based on local resonance theory. This configuration demonstrated the ability of the structure to generate an ultra-low frequency bandgap. Nonetheless, it was noteworthy that the bandwidth of the bandgap remained relatively narrow. Bai et al. [43] proposed a seismic metamaterial wall based on soil expended with concrete and rubber, which could generate several ultra-low frequency bandgaps in the range of 0 Hz–10 Hz. It was indicated that this metamaterial wall could effectively attenuate the frequency from 0 Hz to 10 Hz, but the size of it was too large to be used in engineering. A novel meter-scale seismic metamaterial [29] could generate one wide ultra-low bandgap with frequency from 1.132 Hz to 12.695 Hz. Although this metamaterial was designed for Lamb waves, and its ability to attenuate real seismic waves was not tested, this barrier provides a new reference for the study of barriers at ultra-low frequencies. Chen et al. [46] proposed a 3D metaconcrete that could generate a wide ultra-low frequency bandgap with a width of 4.6 Hz when the barriers were arranged according to decreasing or increasing dimensions. However, its size is also too large for engineering. Brûlé et al. [57] conducted a full-scale test on periodic cylindrical barriers, and the results showed that the barriers could attenuate elastic waves within the bandgap. It is proved that the cylindrical barriers can attenuate surface seismic waves and provided a practical basis for subsequent research. Many researchers [8,51,54] have conducted bench-scale testing to verify the attenuation effect of the numerically simulated barrier structure on elastic waves. The results demonstrated that the barrier effectively attenuated elastic waves within the bandgap, providing a foundation for practical applications.
Although previous studies have shown that 1D [34,43,58,59], 2D [29,60,61,62,63], or 3D [45,46,47] barriers can create narrow bandgaps, there is still a need for further exploration, particularly regarding ultra-low frequency bandgaps. Therefore, we studied a seismic surface wave barrier with ultra-low frequency bandgaps. Most of these studies have focused on cylindrical barriers, leaving other shapes underexplored. References [26,27] compared cylindrical barriers with square barriers and concluded that the square barrier had wider bandgaps than the cylindrical barrier. Therefore, we propose investigating the potential of a hollow frame-type square column combined with a solid square column barrier. Furthermore, it is worth noting that previous research has typically assumed soil to be linearly elastic, ignoring its damping properties in reality. To address this, we conducted time history analyses of the barrier structure on damped soil and barrier materials, aiming to provide more realistic and applicable insights.
This study is organized as follows. Section 2 describes the theoretical method for bandgaps. Section 3 shows the finite-element calculation method used in this study, calculates the bandgap, and analyzes the mechanism of the bandgap. Then, the transmission spectrum and time history analysis of the 10-cycle barrier are calculated. The effect of the soil and material damping are discussed. Finally, the main conclusions are given in Section 4.

2. Theory and Models of the Novel Barrier

This study proposes a novel frame-type barrier for Rayleigh waves. The unit cell of the novel frame-type barrier is shown schematically in Figure 1a–d. a is defined as the lattice constant, b and c are defined as the outer width and the thickness of the hollow column, d and e represent the width and length of the four columns, and h is the height of the barrier, respectively. Oudich et al. [64] pointed out that if the height of the unit cell is greater than 10 times the lattice constant, the height will not affect the bandgap results; therefore, the height used in this study is 16 times the lattice constant, H = 16a. The light yellow, blue, and orange areas in Figure 1b–d represent soil, rubber, and steel, respectively, and all of materials are linearly elastic and undamped. As shown in Figure 1d, each frame-type barrier is periodically arranged in the x–y directions to form a complete seismic surface wave barrier around the building. The geometric parameters and material properties [60] corresponding to this study are shown in Table 1 and Table 2, respectively.
The novel periodically arranged frame-type barrier can be simplified to a square unit cell in Figure 1e. The shaded area is the first Brillouin zone, and the eigenfrequencies are obtained by scanning the eigenvectors along Γ-Χ-Μ-Γ. The soil and material properties are assumed to be linearly elastic, homogeneous, and isotropic, so the governing equations for elastic wave propagation can be written as follows [61]:
ρ (r) u (r) = ∇ [(λ (r) + 2μ (r)) (∇∙u (r))] − ∇ × [μ (r) ∇ × u (r)]
where r (r = (x, y, z)) is the position vector, u (u = (u, v, w)) is the displacement vector, ρ is the mass density, and t is the time variable. According to the Floquet–Bloch theorem, the displacement field of an elastic wave in a periodic structure can be written as below:
u (r, t) = exp (i (krω t)) uk (r)
where i2 = −1, and k is the elastic wave vector. The periodic boundary condition can be set on the unit cell and combined with Equation (2) to solve Equation (1). The displacement amplitude condition can be written as follows:
(r + a) = exp (i (ka)) u (r)
where a is the lattice constant. The eigenvalue equation for the unit cell can be written as given below:
(Kω2 M) U = 0
where K and M are the stiffness and mass matrices of the unit cell, respectively. The eigenfrequencies of the unit cell can be obtained by solving Equations (1), (3), and (4).
To eliminate the effect of the dimensions, the relative bandgap width fw can be used to describe the bandgap characteristics of the unit cell and can be written as follows:
fw = 2 (fufl)/(fu + fl)
where fu and fl are the upper- and lower-boundary frequencies of the bandgap, respectively.
In this study, COMSOL Multiphysics 5.6 finite-element software was used to calculate the bandgaps, transmission spectrum, and time history analysis.

3. Results and Discussion

3.1. Frequency Bandgaps

Seismic waves are divided into body waves and surface waves, and most energy released by earthquakes is concentrated in Rayleigh waves [5,6,7]. Rayleigh waves are interference waves generated by the body waves at the free surface and propagated along the free surface, and their orbits are inverse elliptical. Rayleigh waves have almost no attenuation in the horizontal direction and decay exponentially with depth. Moreover, the bandgaps for Rayleigh waves can be obtained by the eigenfrequencies of the unit cell. The eigenfrequencies can be calculated by the eigenfrequency in the Solid Mechanics module using COMSOL.
To verify the accuracy of the finite-element software used in this study, the model and the bandgaps are plotted in Figure 2 [65]. The periodic constant of the unit, the radius of pile, and the parameter about depth are taken as a = 0.8 m, r = 0.3 m, and h = 20a, respectively. The model comprises foam columns that are completely buried in the soil. The Young’s modulus E, mass density ρ, and the Poisson ratio υ of foam and soil are taken as 37 Mpa and 20 Mpa, 60 kg/m3 and 1800 kg/m3, and 0.32 and 0.35, respectively. The two materials are also assumed to be linearly elastic and undamped. The COMSOL was used for the eigenfrequencies calculation, and the energy distribution parameter identification method was used for surface wave identification. It can be seen from Figure 2b that the bandgaps of the barrier are in agreement with the previous paper, which proves that the numerical method used in this study is correct and can be used to calculate the bandgap of the surface wave barrier.
To simulate the periodically aligned barrier structures, it is necessary to set the periodic boundary condition (Periodic BC) around the unit cell based on Bloch’s period theory, and a fixed boundary condition (Fixed BC) and a free boundary condition (Free BC) are set at the bottom and top of the unit cell to simulate an infinitely deep soil and surface soil, respectively, in Figure 3a. The eigenfrequencies calculated by COMSOL include both surface and body waves and pseudo-surface waves. The surface waves can be identified using the energy distribution parameter identification method [65].
The size of the grid cell has a significant impact on the accuracy of the results. It is essential that the cell size be considerably smaller than the minimum geometric size and the shortest wavelength of the model. In order to verify the convergence of the mesh size, the maximum cell size was chosen as 1/2, 1/4, and 1/8 of the minimum wavelength. The cell sizes at the minimum geometry, i.e., the thickness of the frame, were c, c/2, and c/4, respectively. These cell sizes were defined as mesh 1, mesh 2, and mesh 3, as shown in Figure 3b. Mesh 1 contains 12,058 domain cells, 4644 boundary cells, and 568 edge cells. Mesh 2 contains 14,568 domain cells, 5586 boundary cells, and 700 edge cells. Mesh 3 contains 40,060 domain cells, 8492 boundary cells, and 884 edge cells. The dispersion curves were calculated using the aforementioned three gridding methods, and the results are presented in Figure 3c. Figure 3c illustrates that the discrepancy between the eigenvalues of mesh 1 and mesh 2 and 3 at high frequencies are considerable, whereas the eigenvalues of mesh 2 and mesh 3 are identical. This suggests that the eigenvalues of mesh 2 are highly satisfactory with regard to convergence, and mesh 2 was selected for this paper.
Figure 3d shows the frequency bandgap of the barrier using the Rayleigh wave theory of the unit cell. As shown in Figure 3d, the surface waves are expressed by the dot–dash line. It can be seen that the unit cell has a complete bandgap with a relative width of 0.47 Hz according to Equations (2) and (5), and the upper- and lower-boundary frequencies are 2.7922 Hz and 4.4875 Hz, respectively. Simultaneously, the surface wave barriers produce seven directional bandgaps in the Γ–Χ direction. The second bandgap is a directional bandgap with a relative width of 0.24 Hz, and its upper- and lower-boundary frequencies are 6.7351 Hz and 8.5685 Hz, respectively.

3.2. Vibration Modes

In this section, the bandgap generation mechanism of the unit cell is investigated by analyzing the vibration modes of the first two bandgaps’ boundaries in the Γ–Χ direction. Figure 3 and Figure 4 show the points selected for the vibration mode diagram, and the vibration mode diagrams are shown in Figure 4.
It can be seen from Figure 4a that the displacements of points A1, A2, A3, and A4 are mainly concentrated in the central square column. The internal structure rotates clockwise in the xz plane, and the magnitude of the displacement decreases from top to bottom. In Figure 4b, the displacement is mainly concentrated in the rubber, and the internal structure rotates counter-clockwise in the xy plane. In Figure 4c, the displacement is mainly concentrated in the rubber, while the internal structure moves in a positive direction along the z-axis, and the magnitude of the displacement gradually decreases from top to bottom. In Figure 4d, the displacement is mainly concentrated in the two rubbers along the y-axis direction, while these two rubbers move in a positive direction along the y-axis. The reason is that the energy is absorbed by the internal structure through horizontal displacement at points A1 and A4 and rotational and vertical displacement at points A2 and A3, respectively. Additionally, the rotational displacement at A2 is uniformly distributed in the i-axis direction, indicating its independence from the height of the barrier structure. The displacements at points A1, A3, and A4 are related to the height of the structure, which are mainly concentrated in the upper part of the structure and gradually decrease with increasing depth. Furthermore, the vibration mode diagrams indicate that the displacements are concentrated in the upper part of the unit cell structure, particularly in the internal structure. These four vibration modes are the main causes of the Rayleigh wave bandgaps.

3.3. Transmission Spectrum Analysis

In this section, the attenuation effects of the surface wave barrier are studied by using frequency domain analysis. In theory, the periodic barrier is infinitely arranged in a half-infinite space of soil, while the finite-unit cell of the barrier is applied in practical engineering. The transmission spectrum of finite-period barriers on elastic waves must be calculated. The array transmission model of the barrier, which consists of ten-unit cells periodically embedded in the soil along the horizontal x direction, is shown in Figure 5. The load excitation location is 16a away from the barrier, a distance that should be sufficient for dissipating bulk waves in the soil, thereby ensuring that the barrier is exclusively affected by Raleigh waves; the output response position is 2a away from the last unit cell. Periodic boundary conditions are adopted along the z directions to meet the requirement of infinite periodicity (Oudich and Badreddine Assouar 2012), and a perfectly matched layer (PML) of 3a thickness around the soil is applied to mitigate the impact of waves on the soil surface. The transmission spectrum is defined as follows:
FRF = 20 × log10 (u2/u1)
where u1 and u2 are the displacements at the pickup points with and without the barriers, respectively. The transmission spectrum is shown in Figure 6.
In Figure 6, it can be seen that the barrier structure can effectively attenuate the elastic wave when the main frequency falls in the bandgap of the barrier. In particular, in the second bandgap range (from 6.7351 Hz and 8.5685 Hz), the attenuation value reaches −33.17 dB at 7.55 Hz.
The attenuation mechanism at special points can be better analyzed by the overall and internal vibration modals. The special points are 2 Hz, which is outside of the bandgaps, and 7.5 Hz, which is within the two bandgaps, and these are shown in Figure 7. At 2 Hz, the barrier structure has little effect on the vibration displacement of the soil surface at the pick-up point, which indicates that the barrier cannot attenuate the elastic wave at 2 Hz. Figure 7b shows that the soil around the barrier experiences a significant vibrational displacement compared to that not around the barrier. Additionally, Figure 7e indicates that only the first six unit-cell barriers are disturbed by the elastic wave. The combined effect of the barrier structure and the surrounding soil does not change the overall propagation of the elastic wave. This indicates that the barrier neither attenuates the elastic wave nor creates an elastic wave bandgap, which is consistent with the bandgap results obtained in Section 2. As depicted in Figure 7c, the elastic wave gradually transforms into body and surface waves when the frequency is 7.5 Hz, with the majority of the energy concentrated in the surface wave. Figure 7d demonstrates that the vibration displacement on the soil surface behind the barrier structure is considerably reduced. The barrier structure not only disrupts the propagation path of the elastic wave but also transforms a portion of the surface wave into the body wave. The vibrational displacement generated by the body wave in the soil does not increase significantly, and there is no significant vibrational displacement in the soil at the barrier. Figure 7f shows that the energy of the surface wave is concentrated in the internal structure of the barrier, particularly in the first eight unit-cell structures. The vibration displacement generated by the internal structure absorbs the energy of the surface wave, indicating that the barrier structure effectively blocks the propagation of the elastic wave and attenuates its energy. Therefore, the barrier can be used to mitigate the surface waves and protect the buildings.
Figure 6 shows the effects of the barrier in attenuating Rayleigh waves for all the frequencies, except for the 2 Hz. Specifically, it was found that the FRF is small in 0 Hz–7.4 Hz and large in 7.4 Hz–20 Hz. The reasons for the phenomenon are that the elastic waves propagate in the form of alternating maximum and minimum, and the barrier structure alters the propagation path of the elastic wave, thereby changing the vibration displacement value at the pickup point. Therefore, the FRF of the surface waves may also appear negative when the main frequency falls outside the bandgap. In addition, the lower the frequency, the larger the wavelength and the wider the range of vibration amplitudes. As illustrated in Figure 7e, the first six unit-cell structures in the barrier structure at 2 Hz can be considered one cycle of the Rayleigh wave, and the two unit-cell structures in the barrier structure at 7.5 Hz similarly constitute one cycle of the Rayleigh wave, as shown in Figure 7f. Therefore, the attenuation effect of the barrier on low-frequency elastic waves is better than the effect on ultra-low frequency waves.

3.4. Time History Analysis

To further verify the attenuation of the 10-cycle barrier on the surface waves, the time history analysis for the artificial waves and El Centro seismic wave are calculated in this section. The mode for the time history analysis is the same as in Section 3.3, but PML cannot be used in time history analysis; thus, the low-reflection boundary condition was set as the boundary condition, as shown in Figure 8. The Hanning modulation window function was selected for the artificial wave, allowing for the application of a center frequency excitation at specific frequencies, which can be written as follows:
A i n = { 1 A max [ 1 cos ( 2 π f c t C ) ] × sin ( 2 π f c t ) , 0 t C f c 0 , t C f c }
where Amax is the maximum amplitude of the artificial wave; fc and C are the central frequency and the periodic number, respectively. In this section, the main frequencies of the artificial waves are 2 Hz and 7.5 Hz, as shown in Figure 9. The El Centro seismic waves are shown in Figure 10.
The artificial and seismic waves are normalized to eliminate the differences in magnitude. The artificial and El Centro seismic waves are applied at the excitation point, and the resulting acceleration in the x-direction is measured at the pickup point. The corresponding results are depicted in Figure 11 and Figure 12, respectively.
As shown in Figure 11a, at the central frequency of 2 Hz, the acceleration response at the output position is larger than that of the structure without barriers, which means the surface wave barrier may amplify the acceleration when the main frequency falls outside the bandgap, with a maximum amplification of 67.64%, as confirmed by the transmission spectrum results. At the same time, the amplification continuously affects the soil surface at a rate of 35.5%, resulting in a vibratory displacement of the soil. As shown in Figure 11a, the surface wave can be attenuated effectively by the barrier when the central frequency is 7.5 Hz. In particular, the maximum attenuation of the acceleration reaches 84.48%, and the total attenuation reaches 80.49% in the same range of time (0.9 s–1.82 s). However, the barrier structure amplifies the elastic wave with a maximum amplification of 23.19% and continually acts on the soil surface with an amplification rate of 8% when the time is more than 1.82 s. Although the barrier may amplify the surface wave, the maximum acceleration decreases by 76.81% with the barrier. In addition, the barrier can effectively attenuate the elastic wave with a central frequency of 7.5 Hz, and the small-amplitude vibration cannot damage the building.
The time history analysis and acceleration response spectrum of El Centro are shown in Figure 12. In Figure 12a, it can be seen that the barrier structure significantly attenuates the maximum acceleration with a reduction ratio of 90%. Although the barrier structure amplifies the El Centro seismic waves when the frequency falls in the frequency range of 0–3.01 Hz, the maximum acceleration is only 0.45 and is evenly distributed. It can be concluded that the barrier structure attenuates 55% of the El Centro seismic waves. As seen from Figure 12b, the El Centro seismic waves can be significantly attenuated with a reduction of 14.91% when the main frequency falls into the bandgap at 8.5–20 Hz. Additionally, the attenuation ratio of the barrier reaches 65.53% in 2.8–3.56 Hz and 75.84% in 6.2–20 Hz. The barrier structure amplifies the acceleration response, which is consistent with the transmission spectrum results only by 27.15%, and the maximum acceleration value is only 0.2484. Overall, the barrier’s mitigation ratio of the acceleration of the El Centro seismic wave is 14.91%. It is concluded that the barrier structure can effectively attenuate El Centro seismic waves when the main frequency falls into the bandgap.

3.5. Effect of Damping

As shown above, the soil parameters are assumed to be linearly elastic and undamped, whereas the actual soil is damping. The factors influencing soil damping include the physical properties of the soil (density, viscosity, and angle), the geometry of the soil body (depth and shape), and the characteristics of the dynamic loads (frequency, magnitude, and duration). In the finite-element software COMSOL, Rayleigh damping is used, and the equation for Rayleigh damping is written as follows:
C = αdMM + βdKK
where αdM is the mass damping coefficient, and βdK is the stiffness damping coefficient. Rayleigh damping is affected by the frequency range and damping ratio of the structure. In this study, the frequency range of the structure is 0.1–20 Hz, and they can be written as follows:
{ f 1 = 0.1   Hz f 2 = 20   Hz
where f1 and f2 are the lower and upper boundaries of the structural frequency, respectively. The angular frequencies can be calculated:
{ ω 1 = 2 π f 1 = 0.628 ω 2 = 2 π f 2 = 125.664
where ω1 and ω2 are the lower and upper boundaries of the angular frequency of the barrier, respectively. The Rayleigh damping coefficients are written as given below:
{ α d M 2 ω 1 + β d K ω 1 2 = ξ α d M 2 ω 2 + β d K ω 2 2 = ξ
where ζ is the damping ratio.
The soil damping ratio is from 0.01 to 0.08, with a general value of 0.05 taken in this study. The mass and stiffness damping coefficient of Rayleigh damping was calculated to be 0.0625 and 0.0008, respectively. The time history analysis of the El Centro seismic wave was calculated with soil damping, and the results are shown in Figure 13. Figure 13a illustrates that the soil damping significantly attenuates the El Centro seismic wave, with an attenuation value reaching 50%. Figure 13b illustrates that soil damping attenuates frequencies within 3.66–20 Hz, with the greatest attenuation observed within 4.9–2.13 Hz, reaching a value of 62.45%. However, soil damping does not attenuate frequencies within 0–3.66 Hz. It can be concluded that the presence of soil damping can accelerate the attenuation of elastic waves.
At the same time, the rubber and steel are still damping. The damping coefficient of a composite material is closely related to its internal structure and composition, primarily due to the internal friction of the material itself and the friction between the material and the external environment. It is challenging to determine the composite material damping for a barrier with rubber and steel, as proposed in this study. Consequently, the general damping ratio of the material was used in this study, which is consistent with the soil damping. The time history analysis of the barrier to El Centro was conducted with both soil and barrier damping, and the results are shown in Figure 14. The attenuation by the barrier with both soil and material damping is almost the same as it is with only soil damping. It is indicated that material damping has a negligible effect on the attenuation of seismic waves.

4. Conclusions

The present study introduces a novel framework for a seismic surface wave barrier composed of steel and rubber. The band structure of the barrier was computed using finite-element analysis. Additionally, frequency–domain and time–domain analyses were conducted for a 10-cycle potential barrier to verify its attenuation characteristics on seismic waves. The influence of soil damping and material damping on the attenuation effects was also considered. The results indicate the follows:
(1)
The barrier exhibits all directional ultra-low-frequency band gaps ranging from 2.7922 to 4.4875 Hz, along with multiple directional band gaps. These gaps effectively cover the 3–8 Hz dominant frequency of seismic waves such as those from the El Centro earthquake. And the soil damping plays a beneficial role in attenuating seismic waves;
(2)
It can significantly reduce structural damage during earthquakes and holds promising potential as a novel seismic isolation technology for engineering applications. It is of great significance for the seismic protection of existing important buildings, especially ultra-high-rise buildings and places where precision instruments are manufactured;
(3)
The buried frame-type seismic surface wave barrier proposed in this paper has good attenuation only for seismic wave-dominant frequency in the range of 2.7922 Hz–4.4875 Hz, but there is almost no attenuation for lower-frequency seismic waves. Therefore, some further research is needed on barriers with ultra-low-frequency bandgaps;
(4)
The buried frame-type barrier proposed in this paper is of the Bragg scattering type, and there is also a possibility for a local resonance-type barrier according to the mechanism of bandgap generation. Therefore, the local resonance-type barrier will be researched in future work.

Author Contributions

Conceptualization, C.Z. and H.J.; methodology, C.Z. and H.J.; software, H.J.; validation, C.Z., Y.C. and J.L.; formal analysis, C.Z.; investigation, H.J.; resources, C.Z.; data curation, H.J.; writing—original draft preparation, H.J.; writing—review and editing, C.Z.; visualization, Y.C.; supervision, J.L.; project administration, Y.C.; funding acquisition, C.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by Xinjiang Natural Science Key Fund under Grant No. 2022D01D33 and the National Natural Science Foundation of China Upper-level Project under Grant No. 52278302.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ariyaratana, C.; Fahnestock, L.A. Evaluation of buckling-restrained braced frame seismic performance considering reserve strength. Eng. Struct. 2011, 33, 77–89. [Google Scholar] [CrossRef]
  2. Jara, J.M.; Hernandez, E.J.; Olmos, B.A. Effect of epicentral distance on the applicability of base isolation and energy dissipation systems to improve seismic behavior of RC buildings. Eng. Struct. 2021, 230, 111727. [Google Scholar] [CrossRef]
  3. Zhang, Y.; Ren, X.; Zhang, X.Y.; Huang, T.T.; Sun, L.; Xie, Y.M. A novel buckling-restrained brace with auxetic perforated core: Experimental and numerical studies. Eng. Struct. 2021, 249, 113223. [Google Scholar] [CrossRef]
  4. Zhou, X.Y.; Yan, W.M.; Yang, R.L. Seismic base isolation, energy dissipation and vibration control of building structures. J. Build. Struct. 2002, 23, 26. [Google Scholar]
  5. Ben-Menahem, A.; Singh, S.J. Seismic Waves and Sources; Springer Science & Business Media: Cham, Switzerland, 2012. [Google Scholar]
  6. Dobrin, M.B.; Savit, C.H. Introduction to Geophysical Prospecting; McGraw-Hill: New York, NY, USA, 1960; Volume 4. [Google Scholar]
  7. White, J.E. Seismic Waves: Radiation, Transmission, and Attenuation; McGraw-Hill: New York, NY, USA, 1965. [Google Scholar]
  8. Palermo, A.; Krodel, S.; Marzani, A.; Daraio, C. Engineered metabarrier as shield from seismic surface waves. Sci. Rep. 2016, 6, 39356. [Google Scholar] [CrossRef] [PubMed]
  9. Wen, X.; Wen, J.; Yu, D. Phononic Crystals; National Defence Industry Press: Beijing, China, 2009. [Google Scholar]
  10. Fu, Z.; Li, L.; Gu, Y. A localized meshless collocation method based on semi-analytical basis functions for bandgap calculation of elastic waves in phononic crystals. Appl. Math. Mech. 2020, 20, e202000021. [Google Scholar] [CrossRef]
  11. Kushwaha, M.; Halevi, P.; Dobrzynski, L.; Djafari-Rouhani, B. Acoustic band structure of periodic elastic composites. Phys. Rev. Lett. 1993, 71, 2022–2025. [Google Scholar] [CrossRef] [PubMed]
  12. Kushwaha, M.S.; Halevi, P.; Martinez, G.; Dobrzynski, L.; Djafari-Rouhani, B. Theory of acoustic band structure of periodic elastic composites. Phys. Rev. B 1994, 49, 2313. [Google Scholar] [CrossRef] [PubMed]
  13. Laude, V.; Wilm, M.; Benchabane, S.; Khelif, A. Full band gap for surface acoustic waves in a piezoelectric phononic crystal. Phys. Rev. E 2005, 71, 036607. [Google Scholar] [CrossRef]
  14. Trzaskowska, A.; Hakonen, P.; Wiesner, M.; Mielcarek, S. Generation of a mode in phononic crystal based on 1D/2D structures. Ultrasonics 2020, 106, 106146. [Google Scholar] [CrossRef]
  15. Wang, G.; Wen, X.; Wen, J.; Shao, L.; Liu, Y. Two-dimensional locally resonant phononic crystals with binary structures. Phys. Rev. Lett. 2004, 93, 154302. [Google Scholar] [CrossRef] [PubMed]
  16. Wu, T.T.; Wu, L.C.; Huang, Z.G. Frequency band-gap measurement of two-dimensional air/silicon phononic crystals using layered slanted finger interdigital transducers. J. Appl. Phys. 2005, 97, 094916. [Google Scholar] [CrossRef]
  17. Zhou, X.; Wang, L. Opening complete band gaps in two dimensional locally resonant phononic crystals. J. Phys. Chem. Solids 2018, 116, 174–179. [Google Scholar] [CrossRef]
  18. Li, X.; Ning, S.; Liu, Z. Designing phononic crystal with anticipated band gap through a deep learning based data-driven method. Comput. Methods Appl. Mech. Eng. 2019, 361, 112737. [Google Scholar] [CrossRef]
  19. Wang, K.; Liu, Y.; Wang, B. Ultrawide band gap design of phononic crystals based on topological optimization. Phys. B Phys. Condens. Matter 2019, 571, 263–272. [Google Scholar] [CrossRef]
  20. Wang, Y.F.; Guo, J.G.; Zhang, Z. The deformation induced tunable topology in controlling of band gap characteristics for stepped phononic crystals. Solid State Commun. 2022, 351, 114809. [Google Scholar] [CrossRef]
  21. Zeng, C.; Zhao, C.F.; Chen, C. Research on attenuation of seismic surface waves by buried local resonance metabarriers. Eng. Mech. 2023, 40, 1–12. [Google Scholar]
  22. Zeng, C.; Zhao, C.F.; Zeighami, F. Seismic surface wave attenuation by resonant metasurfaces on stratified soil. Earthq. Eng. Struct. Dyn. 2022, 51, 1201–1223. [Google Scholar] [CrossRef]
  23. Zhao, C.F.; Zeng, C.; Witarto, W. Attenuation characteristics and isolation performance of one-dimensional periodic foundation. J. Build. Struct. 2020, 41, 77–85. [Google Scholar]
  24. Zhao, C.F.; Wang, Y.; Chu, F. Research on the isolation performance of cruciform gradient seismic metamaterial. Eng. Mech. 2024, 1–14. [Google Scholar] [CrossRef]
  25. Zhao, C.F.; Zeng, C.; Mo, Y. Attenuation performance of periodic foundation and its experimental verification for seismic waves. J. Vib. Eng. 2022, 35, 1471–1480. [Google Scholar]
  26. Cheng, Z.; Shi, Z. Vibration attenuation properties of periodic rubber concrete panels. Constr. Build. Mater. 2014, 50, 257–265. [Google Scholar] [CrossRef]
  27. Du, Q.J.; Zeng, Y.; Huang, G.L.; Yang, H.W. Elastic metamaterial-based seismic shield for both Lamb and surface waves. AIP Adv. 2017, 7, 075015. [Google Scholar] [CrossRef]
  28. Huang, J.; Shi, Z. Application of Periodic Theory to Rows of Piles for Horizontal Vibration Attenuation. Int. J. Geomech. 2013, 13, 132–142. [Google Scholar] [CrossRef]
  29. Luo, Y.M.; Huang, T.T.; Zhang, Y.; Xu, H.H.; Xie, Y.M.; Ren, X. Novel meter-scale seismic metamaterial with low-frequency wide bandgap for Lamb waves. Eng. Struct. 2023, 275, 115321. [Google Scholar] [CrossRef]
  30. Miniaci, M.; Krushynska, A.; Bosia, F.; Pugno, N.M. Large scale mechanical metamaterials as seismic shields. New J. Phys. 2016, 18, 083041. [Google Scholar] [CrossRef]
  31. Pu, X.B.; Shi, Z.F. Surface-wave attenuation by periodic pile barriers in layered soils. Constr. Build. Mater. 2018, 180, 177–187. [Google Scholar] [CrossRef]
  32. Witarto, W.; Wang, S.J.; Yang, C.Y.; Mo, Y.L.; Chang, K.C.; Tang, Y. Three-dimensional periodic materials as seismic base isolator for nuclear infrastructure. AIP Adv. 2019, 9, 045014. [Google Scholar] [CrossRef]
  33. Zeng, Y.; Xu, Y.; Yang, H.; Muzamil, M.; Xu, R.; Deng, K.K.; Peng, P.; Du, Q.J. A Matryoshka-like seismic metamaterial with wide band-gap characteristics. Int. J. Solids Struct. 2020, 185–186, 334–341. [Google Scholar] [CrossRef]
  34. Geng, Q.; Zhu, S.Y.; Chong, K.P. Issues in design of one-dimensional metamaterials for seismic protection. Soil Dyn. Earthq. Eng. 2018, 107, 264–278. [Google Scholar] [CrossRef]
  35. Guo, J.J.; Cao, J.Z.; Xiao, Y.; Shen, H.J.; Wen, J.H. Interplay of local resonances and Bragg band gaps in acoustic waveguides with periodic detuned resonators. Phys. Lett. A 2020, 384, 126253. [Google Scholar] [CrossRef]
  36. Liu, Z.Y.; Zhang, X.X.; Mao, Y.; Zhu, Y.Y.; Yang, Z.Y.; Chan, C.T.; Sheng, P. Locally Resonant Sonic Materials. Science 2000, 289, 1734–1736. [Google Scholar] [CrossRef] [PubMed]
  37. Zhou, X.; Liu, X.; Hu, G. Elastic metamaterials with local resonances: An overview. Theor. Appl. Mech. Lett. 2012, 2, 041001. [Google Scholar] [CrossRef]
  38. Cheng, Z.; Shi, Z.; Palermo, A.; Xiang, H.; Guo, W.; Marzani, A. Seismic vibrations attenuation via damped layered periodic foundations. Eng. Struct. 2020, 211, 110427. [Google Scholar] [CrossRef]
  39. Liu, Z.; Rumpler, R.; Feng, L. Locally resonant metamaterial curved double wall to improve sound insulation at the ring frequency and mass-spring-mass resonance. Mech. Syst. Signal Pr. 2021, 149, 107179. [Google Scholar] [CrossRef]
  40. Luo, C.; Han, C.Z.; Zhang, X.Y.; Zhang, X.G.; Ren, X.; Xie, Y.M. Design, manufacturing and applications of auxetic tubular structures: A review. Thin-Walled Struct. 2021, 163, 107682. [Google Scholar] [CrossRef]
  41. Xu, R.; Muzamil, M.; Fan, L.; Yuan, K.K.; Yang, H.W.; Du, Q.J. Broadband seismic metamaterial with an improved cylinder by introducing plus-shaped structure. Europhys. Lett. 2021, 133, 37001. [Google Scholar] [CrossRef]
  42. Zhang, X.Y.; Wang, X.Y.; Ren, X.; Xie, Y.M.; Wu, Y.; Zhou, Y.Y. A novel type of tubular structure with auxeticity both in radial direction and wall thickness. Thin-Walled Struct. 2021, 163, 107758. [Google Scholar] [CrossRef]
  43. Bai, Y.T.; Li, X.L.; Zhou, X.H.; Li, P.; Beer, M. Soil-expended seismic metamaterial with ultralow and wide bandgap. Mech. Mater. 2023, 180, 104601. [Google Scholar] [CrossRef]
  44. Liu, Z.; Qin, K.Q.; Yu, G.L. Partially Embedded Gradient Metabarrier: Broadband Shielding from Seismic Rayleigh Waves at Ultralow Frequencies. J. Eng. Mech. 2020, 146, 04020032. [Google Scholar] [CrossRef]
  45. Chen, C.; Lei, J.C.; Liu, Z.S. A Ternary Seismic Metamaterial for Low Frequency Vibration Attenuation. Materials 2022, 15, 1246. [Google Scholar] [CrossRef] [PubMed]
  46. Chen, Z.Y.; Wang, G.F.; Lim, C.W. Artificially Engineered Metaconcrete with Wide Bandgap for Seismic Surface Wave Manipulation. Eng. Struct. 2023, 276, 115375. [Google Scholar] [CrossRef]
  47. Liu, Z.X.; Zhang, M.K.; Huang, L.; Zhang, H. Broadband Seismic Isolation Effect of Three-Dimensional Seismic Metamaterials in a Semi-Infinite Foundation: Modelled by Fast Multipole Indirect Boundary Element Method. Eng. Anal. Bound. Elem. 2023, 154, 7–20. [Google Scholar] [CrossRef]
  48. Masoud, B.; Kiarasi, B.; Marashi, S.M.H.; Ebadati, M.; Masoumi, F.; Asemi, K. Stress wave propagation and natural frequency analysis of functionally graded graphene platelet-reinforced porous joined conical–cylindrical–conical shell. Waves Random Complex Media 2021, 1–33. [Google Scholar] [CrossRef]
  49. Mojtaba, K.; Jafari, J.; Asemi, K. Low-velocity impact analysis of functionally graded porous circular plate reinforced with graphene platelets. Waves Random Complex Media 2022, 1–27. [Google Scholar] [CrossRef]
  50. Hajjaj, M.M.; Tu, J.W. A seismic metamaterial concept with very short resonators using depleted uranium. Arch. Appl. Mech. 2021, 91, 2279–2300. [Google Scholar] [CrossRef]
  51. Chen, Y.Y.; Qian, F.; Scarpa, F.; Zuo, L.; Zhuang, X.Y. Harnessing multi-layered soil to design seismic metamaterials with ultralow frequency band gaps. Mater. Des. 2019, 175, 107813. [Google Scholar] [CrossRef]
  52. Achaoui, Y.; Ungureanu, B.; Enoch, S.; Brule, S.; Guenneau, S. Seismic waves damping with arrays of inertial resonators. Extrem. Mech. Lett. 2016, 8, 30–37. [Google Scholar] [CrossRef]
  53. Ji, D.X.; Yu, G.L. Shielding performance of T-shaped periodic barrier for surface waves in transversely isotropic soil. Proc. Inst. Mech. Eng. Part L J. Mater. Des. Appl. 2022, 236, 2242–2254. [Google Scholar] [CrossRef]
  54. Maheshwari, H.K.M.; Rajagopal, P. Novel locally resonant and widely scalable seismic metamaterials for broadband mitigation of disturbances in the very low frequency range of 0–33 Hz. Soil Dyn. Earthq. Eng. 2022, 161, 107409. [Google Scholar] [CrossRef]
  55. Xu, L.; Wang, J.; Dai, G.; Yang, S.; Yang, F.; Wang, G. Geometric phase, effective conductivity enhancement, and invisibility cloak in thermal convection conduction. Int. J. Heat Mass Transf. 2021, 165, 120659. [Google Scholar] [CrossRef]
  56. Li, L.X.; Wang, Q.; Liu, H.X.; Li, L.; Yang, Q.; Zhu, C. Seismic metamaterials based on coupling mechanism of inertial amplification and local resonance. Phys. Scr. 2023, 98, 045024. [Google Scholar] [CrossRef]
  57. Brule, S.; Javelaud, E.H.; Enoch, S.; Guenneau, S. Experiments on Seismic Metamaterials: Molding Surface Waves. Phys. Rev. Lett. 2014, 112, 133901. [Google Scholar] [CrossRef]
  58. Lim, C.W.; Muhammad. Elastic waves propagation in thin plate metamaterials and evidence of low frequency pseudo and local resonance bandgaps. Phys. Lett. A 2019, 383, 2789–2796. [Google Scholar]
  59. Van, H.C.; Schevenels, M.; Lombaert, G. Double Wall Barriers for the Reduction of Ground Vibration Transmission. Soil Dyn. Earthq. Eng. 2017, 97, 1–13. [Google Scholar] [CrossRef]
  60. Huang, T.T.; Ren, X.; Zeng, Y.; Zhang, Y.; Luo, C.; Zhang, X.Y.; Xie, Y.M. Based on auxetic foam: A novel type of seismic metamaterial for Lamb waves. Eng. Struct. 2021, 246, 112976. [Google Scholar] [CrossRef]
  61. Mei, J.; Liu, Z.Y.; Shi, J.; Tian, D.C. Theory for elastic wave scattering by a two-dimensional periodical array of cylinders: An ideal approach for band-structure calculations. Phys. Rev. B 2003, 67, 245107. [Google Scholar] [CrossRef]
  62. Muhammad; Lim, C.W.; Reddy, J.N. Built-up structural steel sections as seismic metamaterials for surface wave attenuation with low frequency wide bandgap in layered soil medium. Eng. Struct. 2019, 188, 440–451. [Google Scholar] [CrossRef]
  63. Zhao, C.F.; Zeng, C.; Wang, Y.Z.; Bai, W.; Dai, J.W. Theoretical and Numerical Study on the Pile Barrier in Attenuating Seismic Surface Waves. Buildings 2022, 12, 1488. [Google Scholar] [CrossRef]
  64. Oudich, M.; Badreddine, A.M. Surface acoustic wave band gaps in a diamond-based two-dimensional locally resonant phononic crystal for high frequency applications. J. Appl. Phys. 2012, 111, 014504. [Google Scholar] [CrossRef]
  65. Pu, X.B.; Shi, Z.F. A novel method for identifying surface waves in periodic structures. Soil Dyn. Earthq. Eng. 2017, 98, 67–71. [Google Scholar] [CrossRef]
Figure 1. (a) Overall layout of barrier structure; (b) lateral view of the unit cell; (c) internal view of the unit cell; (d) layout of the barrier structure; (e) the first Brillouin zone.
Figure 1. (a) Overall layout of barrier structure; (b) lateral view of the unit cell; (c) internal view of the unit cell; (d) layout of the barrier structure; (e) the first Brillouin zone.
Buildings 14 02328 g001
Figure 2. Verification for bandgaps in the previous paper: (a) the model; (b) the bandgaps [65].
Figure 2. Verification for bandgaps in the previous paper: (a) the model; (b) the bandgaps [65].
Buildings 14 02328 g002
Figure 3. (a) The boundary conditions of the unit cell; (b) finite-element mesh size of unit cell; (c) dispersion relations of Rayleigh wave calculated by three kinds of mesh generations; (d) bandgaps under the Rayleigh wave theory of the unit cell.
Figure 3. (a) The boundary conditions of the unit cell; (b) finite-element mesh size of unit cell; (c) dispersion relations of Rayleigh wave calculated by three kinds of mesh generations; (d) bandgaps under the Rayleigh wave theory of the unit cell.
Buildings 14 02328 g003
Figure 4. The vibration mode of: (a) A1; (b) A2; (c) A3; (d) A4.
Figure 4. The vibration mode of: (a) A1; (b) A2; (c) A3; (d) A4.
Buildings 14 02328 g004
Figure 5. The transmission spectrum analysis model.
Figure 5. The transmission spectrum analysis model.
Buildings 14 02328 g005
Figure 6. The transmission spectrum.
Figure 6. The transmission spectrum.
Buildings 14 02328 g006
Figure 7. Stress modes: (a) without barrier at 2 Hz; (b) with a 10-cycle barrier at 2 Hz; (c) without barrier at 7.5 Hz; (d) with a 10-cycle barrier at 7.5 Hz; (e) of the internal structures at 2 Hz; (f) of the internal structures at 7.5 Hz. The color is used to represent the amount of stress, with red indicating the maximum level of stress and blue indicating the minimum level of stress.
Figure 7. Stress modes: (a) without barrier at 2 Hz; (b) with a 10-cycle barrier at 2 Hz; (c) without barrier at 7.5 Hz; (d) with a 10-cycle barrier at 7.5 Hz; (e) of the internal structures at 2 Hz; (f) of the internal structures at 7.5 Hz. The color is used to represent the amount of stress, with red indicating the maximum level of stress and blue indicating the minimum level of stress.
Buildings 14 02328 g007
Figure 8. The time history analysis model.
Figure 8. The time history analysis model.
Buildings 14 02328 g008
Figure 9. The waveform diagram of artificial wave at (a) 2 Hz and (b) 7.5 Hz.
Figure 9. The waveform diagram of artificial wave at (a) 2 Hz and (b) 7.5 Hz.
Buildings 14 02328 g009
Figure 10. El Centro wave.
Figure 10. El Centro wave.
Buildings 14 02328 g010
Figure 11. Time history diagram of artificial wave at (a) 2 Hz and (b) 7.5 Hz.
Figure 11. Time history diagram of artificial wave at (a) 2 Hz and (b) 7.5 Hz.
Buildings 14 02328 g011
Figure 12. (a) Time history diagram of El Centro and (b) response spectrum of El Centro.
Figure 12. (a) Time history diagram of El Centro and (b) response spectrum of El Centro.
Buildings 14 02328 g012
Figure 13. (a) Time history diagram of El Centro with soil damping and (b) response spectrum of El Centro with soil damping.
Figure 13. (a) Time history diagram of El Centro with soil damping and (b) response spectrum of El Centro with soil damping.
Buildings 14 02328 g013
Figure 14. (a) Time history diagram of El Centro with soil and materials damping and (b) response spectrum of El Centro with soil and materials damping.
Figure 14. (a) Time history diagram of El Centro with soil and materials damping and (b) response spectrum of El Centro with soil and materials damping.
Buildings 14 02328 g014
Table 1. Geometric parameters of the unit cell.
Table 1. Geometric parameters of the unit cell.
a (m)b (m)c (m)d (m)e (m)h (m)f (m)H (m)
2.520.10.21.230.640
Table 2. Material parameters of the unit cell.
Table 2. Material parameters of the unit cell.
MaterialDensity (kg/m3)Young’s Modulus (Pa)Poisson’s Ratio
Clayey silt19394.4 × 1070.4
Rubber13007.7 × 1050.49
Steel77842.07 × 10110.3
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

Jiang, H.; Zhao, C.; Chen, Y.; Liu, J. Novel Frame-Type Seismic Surface Wave Barrier with Ultra-Low-Frequency Bandgaps for Rayleigh Waves. Buildings 2024, 14, 2328. https://doi.org/10.3390/buildings14082328

AMA Style

Jiang H, Zhao C, Chen Y, Liu J. Novel Frame-Type Seismic Surface Wave Barrier with Ultra-Low-Frequency Bandgaps for Rayleigh Waves. Buildings. 2024; 14(8):2328. https://doi.org/10.3390/buildings14082328

Chicago/Turabian Style

Jiang, Hui, Chunfeng Zhao, Yingjie Chen, and Jian Liu. 2024. "Novel Frame-Type Seismic Surface Wave Barrier with Ultra-Low-Frequency Bandgaps for Rayleigh Waves" Buildings 14, no. 8: 2328. https://doi.org/10.3390/buildings14082328

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