Previous Article in Journal
Properties and Microstructure of TiSiC- and TiSiCN-Based Coatings Produced by RPS
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rayleigh Bound States in the Continuum in Shallow Surface Relief Phononic Crystals

Catalan Institute of Nanoscience and Nanotechnology (ICN2), CSIC and BIST, Campus UAB, Bellaterra, 08193 Barcelona, Spain
Crystals 2025, 15(12), 1030; https://doi.org/10.3390/cryst15121030
Submission received: 3 November 2025 / Revised: 26 November 2025 / Accepted: 28 November 2025 / Published: 1 December 2025
(This article belongs to the Section Crystal Engineering)

Abstract

We present an investigation into the band structure of acoustic waves in surface phononic crystals (SPnC), which comprise a square lattice of shallow cylinders on a mechanically isotropic semi-infinite substrate, utilizing the finite element method (FEM). The introduction of crystal periodicity to the surface modifies Rayleigh modes from non-dispersive to dispersive, thereby enabling the transformation of these modes into radiative or leaky forms. This spatial dispersion may facilitate the emergence of bound states in the continuum (BIC) by providing conditions appropriate for closing the radiative channels. A symmetry-protected BIC appears at the Γ point only when the periodicity of the crystal extends in the two dimensions of the surface plane. The decoupling from the radiative channels is due to symmetry incompatibility. An accidental BIC emerges in both one- and two-dimensional SPnCs at finite wave vectors. The partial-wave model applied to the empty lattice approximation shows that the underlying mechanism giving rise to the emergence of the accidental BIC is related to the simultaneous fulfillment of the nullification condition of the transverse radiative channel amplitude and the dispersion equation. Furthermore, the presence of the accidental BIC is not compromised by structural alterations that preserve the crystal symmetry, with only its frequency being influenced.

1. Introduction

The presence of surface acoustic waves (SAW), also known as Rayleigh waves, in mechanically isotropic half-space media is derived from the inhomogeneous or evanescent solutions of the Christoffel equation together with the free surface boundary condition [1]. Their displacement vector lies in the sagittal plane, and they decay exponentially with increasing distance from the surface into the medium. The nonradiative nature of the Rayleigh wave sets a limit on its velocity, which is always lower than the velocity of any bulk wave in the material. The isotropic half-space system also supports a radiative or leaky wave with a velocity between the transverse and longitudinal bulk wave velocities for a certain range of the Poisson’s ratio ( 0.263 < ν ) [2]; while radiating away from the surface into the bulk, the leaky wave is attenuated as it propagates along the surface, and its velocity is thus a complex-valued quantity. Rayleigh and leaky waves are nondispersive in the long wavelength limit. In addition, polarized surface acoustic waves perpendicular to the sagittal plane do not exist on these surfaces.
SAWs exist in any direction on the surface of anisotropic materials, with some differences compared to the isotropic case, such as in the variation of the velocity with the propagation direction and in the form that the amplitude of the wave decays with depth beneath the surface [3,4]. The dependence of the modes on the propagation direction in anisotropic media gives rise to the emergence of leaky surface waves that have no parallel in isotropic media. For example, anisotropy of the (001) surface of cubic crystals leads to the removal of the two-fold bulk transverse-mode degeneracy and the mixing of longitudinal and transverse polarizations. The coupling between the Rayleigh wave and the slowest quasi-transverse bulk wave when the propagation direction changes from the [100] to the [110] direction generates an anticrossing of the two branches [5,6]. The leaky upper branch exhibits the presence of modes for which the leakage to the bulk is suppressed, and they have been identified as the class of states known as bound states in the continuum (BIC). This occurs in a specific direction between the crystallographic [100] and [110] directions and in the exact [110] direction [7,8]. These are examples of parametric or accidental BIC, which is generated in this case by the coupling of modes, and symmetry-protected BIC, respectively, [9].
Leaky waves in anisotropic solids have been extensively studied and have become technologically relevant, as in some cases, their high phase velocity together with an optimal electromechanical coupling compensates for the propagating losses. Although not always identified as BICs, the low-propagation-loss modes found in certain orientations of LiNb03, quartz, and LiTa03 have been used for the development of radio-frequency devices [10,11,12]. More recently, high-frequency low-loss devices were realized using unattenuated SAWs present in leaky mode branches with a strong longitudinal character in layered structures combining thin films and anisotropic substrates [13,14,15].
The presence of an intrinsic characteristic length in half-infinite media, such as the thickness of an added layer [16] or the periodicity of a grating [17], leads to the appearance of spatial dispersion when the inhomogeneity of the properties in the media is introduced. In such a case, the nontrivial dispersion relations favor the transition of surface modes to leaky modes, enabling the presence of waves with a velocity larger than the limiting transverse bulk counterpart. Accidental BICs can exist in the branches of these leaky modes if the appropriate conditions are met. This applies to all isotropic layers in half-infinite media studied in systems comprising flat surfaces [18], surface gratings [19,20], and surface 2D-phononic crystal structures (SPnC) [21]. These studies on BICs in both 1D and 2D SPnCs were conducted in layered structures, in which guided modes such as Sezawa modes may be present in addition to the Rayleigh mode. BICs were also found in the intricate band structure of 2D SPnCs formed by a square lattice of cylindrical voids in a (001) GaAs half-space [22].
From a different perspective, BICs have been presented as polarization singularities of the radiation field in momentum space [23,24]. Research on the topological characteristics of BICs has largely concentrated on photonic dielectric structures, with relatively few studies focusing on acoustic waves [21].
In this work, we present a study of SPnCs formed in mechanically isotropic half-space media using the finite element method (FEM) with a special focus on the radiative part of the frequency spectrum. The surface roughness was considered a shallow relief to prevent the presence of local resonances [25]. Therefore, this represents the simplest case study, where isolated branches of surface and leaky modes allow the analysis of accidental BICs that do not originate from mode coupling [26].

2. Surface Phononic Crystal Model and Simulation Methods

We consider a semi-infinite isotropic elastic medium, unbounded in the x y plane, with a surface at z = 0 and extending to negative infinity. The flat surface is perturbed by a corrugated structure in the form of protruding cylinders extending periodically in the xy plane, forming a 2D surface-PnC in a square lattice ( C 4 v point group). The PnC structure establishes a periodic surface profile z = ζ ( x , y ) along the x y plane, replacing the flat surface z = 0 condition.
The FEM model using COMSOL Multiphysics 5.5 software has been shown to produce satisfactory results when simulating the band structure of surface and leaky acoustic waves in periodically patterned surfaces [27,28]. We seek for propagating elastic waves satisfying the governing motion equation of a linear elastic medium, which in the frequency domain reads
ρ ω 2 u = σ ,
where u = u ( r ) , with r = ( s , z ) = ( x , y , z ) , is the position-dependent component of the general solution of the displacement u ( r , t ) = u ( r ) e i ω t , i.e., a time-harmonic plane wave with angular frequency ω , ρ is the mass density, and σ is the stress tensor. Equation (1) is complemented by the boundary conditions derived from the presence of a free surface and the periodicity of the PnC lattice. The first requires the zero stress condition on the surface, thus
σ · n ^ = 0 ,
where n ^ is a unit vector normal to the surface at each point. The second introduces the Bloch-Floquet periodic boundary conditions on a single cell of the square SPnC [cf. Figure 1a] so that each u has associated a wave vector k = ( k x , k y ) such that
u ( r + R ) = u ( r ) e i k · R ,
for every R in the lattice.
The general reciprocal lattice vector G n , m of the 2D-square lattice characterized by a period a is given by
G m , n = m b 1 + n b 2 ,
where n,m are integers and b 1 = 2 π a x ^ , b 2 = 2 π a y ^ are primitive reciprocal vectors. The reciprocal lattice vectors and first Brillouin zone are shown in Figure 1b.
As a consequence of Equation (3), u is the product of e i k · r and a periodic function v ( r ) . By expanding the periodic function v in a Fourier series, a wave propagating with an in-plane 2D wave vector k can be represented as an infinite number of traveling wave components. The wave vectors of these wave components or space harmonic are expressed as
k , ( m , n ) = k + G m , n .
Even if periodic boundary conditions place a constraint on the lateral dimensions of the problem, it is necessary to address the truncation of the unit cell in the z direction when working with FEM. When searching for SAWs sustained by the SPnC, it is necessary to ensure that the length of the bounded domain in the z direction is greater than the penetration depth of the SAWs. Therefore, the depth of the structure must encompass several wavelengths to obtain a reliable solution that accurately describes the vanishing acoustic field in the direction perpendicular to the surface. An unintended consequence of domain truncation is the emergence of multiple solutions from the resulting plate-like structure. Different sorting methods based on the mechanical energy distribution inside the structure have been used to distinguish the surface modes from the modes that satisfy the top and bottom surface boundary conditions simultaneously.
Nevertheless, the search for solutions describing radiative waves relies on avoiding reflections from the bottom of the computational domain owing to leakage. Perfectly matched layers (PML) and viscoelastic absorbing regions (VAR), which were previously used to avoid undesired reflections from bounded domains in the material’s response to a mechanical excitation [29,30], have also been applied in the simulation of frequency spectra within the radiative region accessing radiative waves with accurate displacement profiles [22,31]. Moreover, adding an absorbing region results in complex-valued eigenfrequencies, ω , for the modes that propagate in this region, with a large imaginary part in the case of plate-like modes compared to radiative modes. Therefore, the magnitude of I m ( ω ) can be used to sort the FEM solutions by distinguishing between the true nonradiative and radiative waves and discarding unwanted plate-like modes. In addition, the attenuation ratio, γ written as I m ( ω ) / R e ( ω ) , quantitatively accounts for the radiation of power into the semi-infinite solid (see Appendix A).
COMSOL FEM eigenfrequency analysis was conducted for the single cell shown in Figure 1a with period a and forming a cylindrical pillar of diameter d and height h on the surface. It is built by a 95a deep, three-dimensional domain formed by a linear elastic and isotropic material. In this case, Equation (1) is
ω 2 v t 2 u = v l 2 v t 2 1 ( · u ) + 2 u ,
where we have introduced the relation between the speeds of sound for bulk longitudinal, v l , and transverse, v t sound waves and the components of the elastic constant tensor c, which relates stress and strain, i.e., σ = c · u . We have chosen an archetypical material with a Poisson ratio of ν = 0.33 ( v l / v t = 1.9852 ) (in the range of Al alloys and Sn) and, for convenience, we have taken v t = 1 m/s. To model the half-space condition of the simulated structure, an additional absorbing layer with a height of 15a was constructed by introducing imaginary parts to the material properties, as done in the improved version of the VAR model by Wei Ke [32].

3. Results

3.1. Surface-PnC Band Structure Simulation

Figure 2a shows the simulated band structure calculated for the primitive unit cell of a SPnC, which is formed by low-profile cylinders (diameter d = 0.75 a and height h = 0.05 a ), along the Γ X direction of the first Brillouin zone [see Figure 1a,b]. The diagram is expressed with the dimensionless frequency Ω = R e ( ω ) a / v t and wavenumber k a / π , and the color scale represents the value of γ . The dashed lines correspond to the empty lattice (EL) approximation, in which an infinitesimally small periodic perturbation is introduced into the system. This represents the dispersion relation for SAWs whose velocity is determined by solving the Rayleigh equation for a stress-free, flat surface, yielding a value of v SAW = 0.931 v t .
The true SAWs are indicated by black lines in Figure 2a, corresponding to γ = 0 within machine precision. These refer to modes below the first sound line, where the frequencies are less than ω = v t k . The band that nearly overlaps the first sound line and exhibits a cut-off frequency at k = π a corresponds to surface waves of shear horizontal polarization. These waves, characterized by displacements perpendicular to the sagittal plane and having no counterpart on the flat surface of an isotropic medium, can only exist when the surface is altered by inhomogeneities in the material properties [33]. Two more branches follow the EL approximation curves related to Rayleigh waves corresponding to ( n = 0 , m = 0 ) and ( n = 1 , m = 0 ) space harmonics. A frequency gap opens at the X point as a result of the phase-matching condition between these space harmonics, that is, | k π a , ( 0 , 0 ) | = | k π a , ( 1 , 0 ) | .
The convergence of the branch originating at the upper frequency boundary of the frequency gap with the first sound line signifies the threshold for the conversion of surface Rayleigh modes into radiative modes. In the latter, the dominant ( n = 1 , m = 0 ) space harmonic determines the frequency of the modes, whereas the ( n = 0 , m = 0 ) space harmonic introduces a transverse radiative channel. At this point, the mode attenuation departs from the typical values for Rayleigh waves, γ 1 × 10−16, to values above γ 1 × 10−5. The color scale in Figure 2a was configured to highlight the presence of an accidental bound state in the continuum (BIC) at wave vector k a / π = 0.877 and frequency Ω = 3.2758. Consequently, the radiative modes are discernible solely in proximity to this position within the branch, as shown in the section of the band structure near the Brillouin zone boundary [cf. Figure 2b]. The corresponding displacement profile of the BIC is shown in Figure 2d. As expected, there is no radiation within the bulk, in contrast to Figure 2e, where the displacement profile of the radiative mode shows wavefronts traveling at a specific angle from the surface. When compared to a SAW with the same k in the reduced zone [cf. Figure 2f], the BIC shows a greater confinement of the displacement field owing to its higher frequency. Four equivalent BIC states along the equivalent symmetry lines are present owing to the four-fold symmetry of the square lattice. In Section 4, a detailed account of the connection between the transverse radiative channel and the ( n = 0 , m = 0 ) harmonic is provided, together with the emergence of the accidental BIC when the amplitude of the radiative channel vanishes.
A second BIC is found at the edge of the second frequency gap, localized at the Γ point with frequency Ω = 5.848 [cf. Figure 2c]. This constitutes a symmetry-protected BIC that emerges as a result of the two-dimensional symmetry inherent to the crystal and thus has no analog in the one-dimensional grating. The basis functions of the B 1 representation of C 4 v , to which the BIC belongs, are not compatible with the coordinates ( x , y , z ) , which form the basis functions of the polarization components of the acoustic waves in the bulk. Therefore, the BIC is decoupled from the radiative modes and, consequently, localized on the surface [cf. Figure 2g]. A more detailed study is provided in Appendix B.

3.2. Topological Description of Bound States in the Continuum

The open transverse radiation channel of the radiative modes extends inside the bulk to the far zone (far field) over the evanescent components. The far field is defined on a specific band, which means that the frequency varies with k . The polarization state of the defined far field of each radiative mode in the band can be mapped to momentum space, i.e., as a function of k = ( k x , k y ) , constructing a far-field polarization field (for the details of the methodology used to obtain the polarization state, see Appendix C). Therefore, the polarization vector projected onto the xy-plane P ( k ) = p x ( k ) x ^ + p y ( k ) y ^ is introduced, where, in the rotated coordinate system, x ^ is collinear to k , and y ^ = z ^ x ^ . The state of polarization can be described using the Stokes parameters [34] defined as
S 0 = | p x | 2 + | p y | 2 S 1 = | p x | 2 | p y | 2 S 2 = 2 Re ( p x * p y ) S 3 = 2 Im ( p x * p y )
The orientation of a polarization ellipse, which is defined by the azimuth angle γ 0 formed between the major axis of the polarization ellipse and the x axis, and the ellipticity χ 0 are related to the Stokes parameters as follows:
γ 0 = 1 2 tan 1 S 2 S 1 χ 0 = 1 2 sin 1 S 3 S 0 .
In Figure 3, the distributions of the intensity, azimuth, and ellipticity are plotted in the k space in a small area around the position of the accidental BIC at k a / π = ( 0.877 , 0 ) . As shown in Figure 3a, the intensity in the far field is observed to cancel out, whereas both azimuth [cf. Figure 3b] and ellipticity [cf. Figure 3c] become undefined at the location of the accidental BIC (see Appendix C for details on the simulation of the far-field polarization distribution). Therefore, this is a type of singularity in which the state of polarization is indeterminate and falls under the type of Stokes point singularities [35].
Defining the complex Stokes field S 12 = S 1 + i S 2 , the azimuth is then related to the phase of S 12 [36]
ϕ 12 = tan 1 S 2 S 1 ,
so that ϕ 12 = 2 γ 0 . Consequently, the polarization singularities and BICs in the present case are the phase vortices of S 12 . These vortices are characterized by a quantized topological index given by
σ 12 = 1 2 π C d k · k ϕ 12 ,
where k ϕ 12 represents the phase gradient, and C is a closed loop in momentum space around the polarization singularity. Considering Figure 3b, the total winding angle along any anticlockwise loop around the singularity is + 2 π . Therefore, σ 12 = + 2 .

4. Discussion

In a good approximation, band structure in the k x axis of the 1st Brillouin zone of the square-lattice 2D SPnC is very similar to that of modes propagating in the direction normal to the stripes of a 1D grating or 1D SPnC (simulated band structure is discussed in Appendix D). This is appropriate for Rayleigh-like modes below and above the 1st frequency gap and for modes in the radiative region. In this section, we demonstrate that the similarity in the band structures of the two crystal configurations, including the presence of an accidental BIC mode in both instances [see Figure 2b and Figure A8a], is attributed to the contribution of equivalent space harmonics in the two crystal structures to the modes propagating along k x . Here, the partial-wave method applied to the 1D case [17] is used to achieve a conceptual understanding of how elastic waves propagate in a 1D grating, which can be extrapolated to the 2D case. In contrast, the region close to the 2nd frequency gap at k x = 0 requires a separate discussion, which is presented in Appendix B.
Seeking for solutions for the solid in the form of plane waves,
u ( r , t ) = U e i ( k · r ω t )
the motion Equation (Equation (6)) leads to
v t 2 k 2 ρ ω 2 U + v l 2 v t 2 k · U U .
The solutions to the characteristic equation of Equation (12), assuming an isotropic material, and propagation in the x z plane ( k y = 0 ), are
k x 2 ω 2 + k z 2 ω 2 = 1 v t 2 k x 2 ω 2 + k z 2 ω 2 = 1 v l 2
By substituting Equation (13) into Equation (12), the polarization vectors U for the six waves are determined:
U ( 1 , 4 ) = k x , 0 , k z ( 1 ) U ( 2 , 5 ) = 0 , k z ( 2 ) , 0 U ( 3 , 6 ) = k z ( 3 ) , 0 , k x .
These solutions correspond to two longitudinal waves of velocity v l and two transverse (shear vertical) waves of velocity v t polarized in the x z plane, and two transverse (shear horizontal) waves of velocity v t polarized in the y direction. Each pair is composed of upward and downward propagating waves with the same k z but opposite signs:
k z ( 1 ) ω = 1 v l 2 k x 2 ω 2 1 / 2 k z ( 2 ) ω = k z ( 3 ) ω = 1 v t 2 k x 2 ω 2 1 / 2
In addition to solutions with real-valued k z , solutions with imaginary-valued k z , corresponding to inhomogeneous evanescent waves, are also permitted.
The displacement vector of the waves propagating in a grating formed by a periodic surface configuration z = ζ ( x ) oriented along the x direction and characterized by a period of a is described by a weighted superposition of the three solutions traveling downward from the surface as
u i ( x , z , t ) = n = + p = 1 3 C n ( p ) U i , n ( p ) e i k z , n ( p ) z e i ( k x , n x ω t ) i = x , y , z
that satisfies the stress-free boundary condition. The partial waves in Equation (16) are expressed as an infinite series of traveling wave components. The wave vectors of these wave Fourier components or space harmonics are
k x , n = k x + n k g
where k g = 2 π / a x ^ denotes the primitive reciprocal lattice vector. This corresponds to the one-dimensional (1D) counterpart of Equation (5) [37]. Equation (13) are modified accordingly as
k z , n ( p ) ω = 1 v p 2 k x , n 2 ω 2 1 / 2 k x , n 2 ω 2 < 1 v p 2 = i k x , n 2 ω 2 1 v p 2 1 / 2 k x , n 2 ω 2 > 1 v p 2
with v p = v l for p = 1 and v p = v t for p = 2 , 3 . Although the general displacement of the waves propagating in the grating is represented as a linear combination of the three waves propagating into the bulk, when studying the propagation direction normal to the grooves, waves polarized in the sagittal plane
u i ( x , z , t ) = n = + C n ( 1 ) U i , n ( 1 ) e i k z , n ( 1 ) z + C n ( 3 ) U i , n ( 3 ) e i k z , n ( 3 ) z e i ( k x , n x ω t ) i = x , z
are decoupled from those polarized in the direction normal to it,
u i ( x , z , t ) = n = + C n ( 2 ) U i , n ( 2 ) e i k z , n ( 2 ) z e i ( k x , n x ω t ) i = y
Hereafter, only waves polarized in the sagittal plane are considered. Within the EL approximation, the zero-stress condition on the surface, as described by Equation (2) with n ^ = ( 001 ) , reduces to the well-known two boundary conditions for the free surface of an elastic half-space
σ z z | z = 0 ρ v l 2 u z z | z = 0 + ρ ( v l 2 2 v t 2 ) u x x | z = 0 = 0 σ x z | z = 0 ρ v t 2 u x z | z = 0 + u z x | z = 0 = 0 .
Substituting Equation (19) into Equation (21) gives a pair of homogeneous linear equations for the amplitudes C n ( 1 ) and C n ( 3 ) for each space harmonic as [38]:
Q 11 C n ( 1 ) + Q 12 C n ( 3 ) = 0 Q 21 C n ( 1 ) + Q 22 C n ( 3 ) = 0
with
Q 11 = Q 22 = v SAW 2 v t 2 2 , Q 12 = 2 v SAW 2 v t 2 1 1 / 2 , Q 21 = 2 v SAW 2 v l 2 2 1 / 2 .
For a nontrivial solution to exist, the determinant of the coefficients of C n ( 1 ) and C n ( 3 ) must vanish, and the solution to the resulting equation provides the phase velocity v SAW of the wave.
In Figure 4a, the dispersion curve for a Rayleigh wave on a flat surface is illustrated by the solid line, ω = v SAW k x . Upon introducing the EL approximation, the dispersion diagram features an infinite number of analogous linear branches for each space harmonic n separated by n k g . This configuration forms the repeated zone scheme shown in Figure 4a, where the space harmonics n = 1 and n = 1 are partially represented. For the case where n = 0 and k x < π / a , the dispersion relation derived from the EL approximation serves as an accurate representation of a weakly perturbed surface, such as the one examined in the current study. Although all space harmonics contribute to Equation (19) and in the form of evanescent wave components (where k z , n is imaginary), the n = 0 space harmonic predominates. For k x = π / a [point (a) in Figure 4a, the phase-matching condition between n = 0 and n = 1 harmonics,
| k x k g | = | k x | ,
determines the dispersion. The slowness curves of Figure 4b illustrates the phase-matching condition that occurs for the two pairs of evanescent partial waves (imaginary-valued k z / ω ) with transverse and longitudinal polarizations, respectively, from the n = 0 and n = 1 space harmonics. Therefore, these four terms must be considered in Equation (19), and the dispersion relation is strongly modified from the EL approximation, giving rise to the opening of the frequency gap discussed in the main text. The partial waves of n = 1 with the propagation wave vector k x + k g and imaginary-valued solutions k z / ω that are outside the plot range in Figure 4b do not contribute significantly to the total solution. This observation is equally applicable to all other harmonics except for n = 0 and n = 1 .
The modes of n = 0 with a frequency above the first sound line ω = v t ( k x k g ) [but below the second sound line ω = v l ( k x k g ) ] of n = 1 [point (b) in Figure 4a] are formed of partial waves of n = 0 with imaginary k z / ω values for both the longitudinal and transverse components, and partial waves of n = 1 harmonic with real-valued k z / ω transverse and imaginary-valued longitudinal components, as shown in Figure 4c. Therefore, while the components related to n = 0 harmonic are bound to the surface, i.e., evanescent waves, the transverse component of the n = 1 harmonic radiates into the bulk. The radiative component is marked with a dashed arrow in Figure 4c forming an angle θ
θ = arccos v t ω r k x k g
from the surface.
These properties established for the n = 0 space harmonic can be extended to other space harmonics owing to the periodicity. In the repeated-zone scheme illustrated in Figure 4a, the equivalent forward-traveling modes related to space harmonics other than n = 0 characterized by a positive group velocity and all sharing the same frequency ω r , have wave vectors that differ by ± n k g from the wave vector of the mode of n = 0 at the same frequency. When ω r exceeds the critical frequency, which is defined at the intersection of the space harmonic dispersion curve n with the first sound line of the subsequent space harmonic ( n 1 ), these modes consist of a radiative wave linked to the particular space harmonic ( n 1 ). Similarly, equivalent backward traveling modes with frequency ω r above the critical frequency, where the dispersion curve of the space harmonic n crosses the first sound line of the preceding space harmonic ( n + 1 ), are composed of a radiative wave associated with this specific space harmonic ( n + 1 ). This is illustrated in Figure 4d for the mode at point (c) of the dispersion curve for space harmonic n = 1 . The mode frequency is above the first sound line ω = v t k x of the space harmonic n = 0 . Consequently, the mode includes a radiative transverse wave component of the space harmonic n = 0 [the equivalent in 2D is ( n = 0 , m = 0 ) ] and both longitudinal and transverse evanescent components of the space harmonic n = 1 [the equivalent in 2D is ( n = 1 , m = 0 ) ]. The radiative component, indicated by a dashed arrow in Figure 4d, forms an angle from the surface that is supplementary to the angle θ , as given by Equation (25) with the sign of the argument inverted or
θ = arccos v t ω r k x .
Although this particular component does not significantly contribute to the amplitude and changes in the dispersion relation, it changes the nature of the modes by transforming them from modes bound to the surface to leaky modes, which decay as they propagate along the surface while radiating into the bulk. Consequently, the frequencies of these modes are complex valued, and the imaginary part accounts for the inverse of their lifetimes. When considering a complex frequency, the values of the solutions to Equation (18) are also complex. In this case, the radiative partial waves possess values of k z , n that are mainly real and have a small but non-negligible imaginary contribution. Likewise, the evanescent partial waves have k z , n values that are essentially imaginary, but contain a small real contribution. Therefore, the point in the dispersion diagram where the modes convert from truly surface waves to leaky waves will not be given exactly by the crossing point of the dispersion relations of the surface mode and first sound line.
When an accidental BIC arises, it results from the nullification of the radiative wave component of the leaky mode. This condition implies that the amplitude of the radiative partial wave must be zero. In the 1D and 2D PnC dispersion relations, a BIC emerges in the folded branch situated between the first and second sound lines. We previously assigned the radiative channel to the transverse component of the n = 0 harmonic based on the partial waves description and EL approximation. By setting C 0 ( 3 ) = 0 in Equation (22) implies Q 11 = 0 , which provides the condition v SAW = 2 v t . The position of the BIC is at the intersecting point of the condition ω = 2 v t k x and the dispersion relation of the folded branch ω = v SAW ( k x k g ) . Then the frequency ω a u x and wavevector k x ( a u x ) of the BIC are found to be
k x ( a u x ) = v SAW k g v SAW + 2 v t , ω a u x = 2 v t v SAW k g v SAW + 2 v t .
Figure 5 shows the evolution of the BIC frequency derived from the simulated band structure of a 1D SPnC by changing the groove width, w, with a constant height of h = 005a. As the PnC structure approaches the flat-surface condition, the BIC frequency converges to Ω a u x = ω a u x a / v t . The lower limit of the BIC frequency is the critical frequency Ω c = ω c a / v t , which marks the transition from leaky to true surface mode.
In conclusion, the occurrence of the accidental BIC in SPCs with shallow surface relief does not rely on the coupling between modes or the interference from multiple radiative paths, but on the fulfillment of the condition for null amplitude of the transverse radiative channel and the dispersion relation simultaneously. An analytical value for the location of the accidental BIC on the dispersion curve was obtained using the partial-wave model and the EL approximation. Consequently, this type of BIC originates from a singular state and is associated with a single radiative channel; thus, their existence does not necessitate coupling between modes or interference from multiple radiative paths.
Furthermore, Figure 5 illustrates the robustness of the BIC against variations in the structural parameters, specifically the groove width. The frequency of the BIC varies continuously from the existence condition, as determined by the EL approximation, to the critical value. The appearance of the accidental BIC was also proven in a full range of Poisson’s ratios covering 0 ν 0.5 . The robustness of BICs to changes in structural parameters and material properties, provided that the symmetry and periodicity of the structure are preserved, has been linked to the presence of the quantized topological index [23].
As real systems are never perfectly periodic owing to fabrication imperfections and disorder, the practical implementation of BICs in devices such as SAW resonators may be hindered by induced radiative leakage. In addition to fabrication tolerance requirements, integrating a phononic structure with transducers on the same surface leads to complex fabrication processes involving a trade-off with enhanced device performance.

Funding

This research received no external funding.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to author.

Acknowledgments

The ICN2 is funded by the CERCA programme/Generalitat de Catalunya.

Conflicts of Interest

The author declares no conflicts of interest.

Appendix A. Finite Element Method/COMSOL Model

Appendix A.1. Validation of the Attenuation Ratio

In addition to discriminating between surface and leaky waves, the addition of an absorbing region below the elastic material region provides a good quantitative account of the attenuation of the leaky waves present in the structure. This is shown in Figure A1, where we compare the attenuation of leaky acoustic modes for the (001) plane of Si as calculated in [39] by numerically solving the secular equation with a complex wave vector with the values obtained by FEM. The attenuation ratio can be equivalently defined as either Im( k )/Re( k ) or Im( ω )/Re( ω ), depending on whether the complex variable involved in the calculation is the wave vector, as in [39], or the frequency, as in this study. The two calculated attenuation ratios are practically superimposed in the range 35° to 45°. The narrow range around 37.3° where the value of the attenuation drops drastically, marks the position of a BIC. The leaky mode stops radiating into the bulk and becomes a true surface wave. The slight change in the position of the BIC when comparing the two models is possibly due to a small difference in the elastic parameters used in the calculation. An optimization procedure yields a minimal attenuation ratio value of 2 × 10−12, which is four orders of magnitude higher than the values attained by the Rayleigh modes (encircled values in Figure A1). The latter corresponds to zero attenuation up to numerical accuracy. By contrast, the attenuation of the BIC mode is limited by the numerical precision given by the seven significant figures of Im( ω ).
Figure A1. Comparison of the angular dependence of the attenuation ratio of leaky waves in the (001) plane of silicon obtained from the FEM model to that calculated in [39]. The encircled data correspond to attenuation values calculated for the true SAW.
Figure A1. Comparison of the angular dependence of the attenuation ratio of leaky waves in the (001) plane of silicon obtained from the FEM model to that calculated in [39]. The encircled data correspond to attenuation values calculated for the true SAW.
Crystals 15 01030 g0a1
The total displacement field profiles inside the material provide further details regarding the specific nature of the three modes featured in Figure A1. While the total displacement field of the Rayleigh mode decays away from the surface to values down to the floating-point precision at a depth of approximately 18 λ SAW [black dots in Figure A2a], an oscillatory profile characterizes the leaky mode owing to its radiative nature [blue dots in Figure A2a]. Although the BIC mode should show a decaying profile similar to the Rayleigh mode, the limiting numerical precision imposes a radiative behavior that is translated into the oscillatory profile [red dots in Figure A2a]. The three modes were studied for a given wave vector. Increasing the wave vector (decreasing the wavelength) without remeshing the domains results in discretization errors. Decreasing the wavelength increases the attenuation of the Rayleigh mode and, as shown in Figure A2b, the total displacement decays to a higher value than that at a larger wavelength. In contrast, the effects in the BIC and leaky modes are less evident in Figure A2b, and the change in attenuation for the latter is only 1.5%.

Appendix A.2. Periodic Boundary Conditions

In this subsection, we introduce the EL approximation to examine the effect of discretization on the calculation of surface modes. For this purpose, we first considered the flat surface of a structure with the same elastic properties used in the main text and introduced periodic boundary conditions. Figure A3 shows the depth profiles of the total displacement field of the Rayleigh modes in the ( 0 , 0 ) and folded-back ( 1 , 0 ) branches, both with the same wave vector k in the reduced zone scheme. The ordinary Rayleigh mode [ ( 0 , 0 ) branch] shows a characteristic exponential decay, reaching values as low as 1 × 10−14 (black line in Figure A3), with an attenuation of the order of 1 × 10−18. In contrast, the folded Rayleigh mode [ ( 1 , 0 ) branch] displays oscillatory behavior with a baseline of approximately 1 × 10−5 (green line in Figure A3) and an attenuation on the order of 1 × 10−12. The larger attenuation of the folded Rayleigh mode is an artifact of spatial discretization and is due to the smaller wavelength of this mode compared with that of the ordinary Rayleigh mode. Because the decay of the displacement field is inaccurate for a true Rayleigh mode and the mode lies in the radiative region as a result of the periodic boundary conditions, the total displacement shows an oscillatory behavior similar to that expected for a leaky mode. However, the attenuation value represents a lower limit for a given mesh size and mode wavelength. Therefore, leaky modes with attenuation higher than this lower limit value are described well by the model.
Figure A2. Total displacement as a function of depth of the BIC, SAW and leaky modes propagating along the (001) surface of Si with a wavenumber (a) k = 32,834 m−1 and (b) k = 4,590,095.5 m−1. The dotted vertical line marks the position of the interface between the elastic material and the VAR.
Figure A2. Total displacement as a function of depth of the BIC, SAW and leaky modes propagating along the (001) surface of Si with a wavenumber (a) k = 32,834 m−1 and (b) k = 4,590,095.5 m−1. The dotted vertical line marks the position of the interface between the elastic material and the VAR.
Crystals 15 01030 g0a2
Discretization errors can be reduced by re-meshing the structure and increasing the number of mesh elements. In this case, the reduction in the amplitude of the oscillations is noticeable in all components of the displacement field of the folded Rayleigh mode, as shown in Figure A4. However, there is a limit to the level of decay that can be reached by reducing the element size without compromising the results for the ordinary Rayleigh mode, owing to the occurrence of round-off errors.
Next, we focus on the results obtained for the SPnC. In this case, we found a trend similar to that discussed for the EL approximation, but with the distinction that the presence of folded and radiative modes is not an artifact of the boundary conditions but an outcome of the real periodicity. The Rayleigh mode in the SPnC shows a total displacement field in depth that is almost superposed to the corresponding mode in the EL approximation (red line in Figure A3). When the meshing error is reduced to a suitable value, the calculated attenuation is a genuine effect of the transformation of the nonradiative Rayleigh mode into a leaky mode. The calculated attenuation of the BIC mode is limited not only by the discretization error but also by the numerical precision, which has been discussed in the previous subsection. This is shown in Figure A3, where the total displacement field of the BIC mode decays to values higher than those reached by the folded Rayleigh mode in the EL approximation at the same wavelength. The attenuation of the BIC is determined by the numerical precision, which is limited by the significant figure of the imaginary part of the complex frequency and not by discretization.
Figure A3. Total displacement as a function of depth of the Rayleigh modes in the first (black line) and second (green line) branches in the flat-surface empty-crystal model. The total displacements of the Rayleigh mode (red line) and the accidental BIC (blue line) in the SPnC are also plotted. All modes propagate in the Γ -X direction with a wavenumber k = 4,590,095.5 m−1.
Figure A3. Total displacement as a function of depth of the Rayleigh modes in the first (black line) and second (green line) branches in the flat-surface empty-crystal model. The total displacements of the Rayleigh mode (red line) and the accidental BIC (blue line) in the SPnC are also plotted. All modes propagate in the Γ -X direction with a wavenumber k = 4,590,095.5 m−1.
Crystals 15 01030 g0a3
Figure A4. Displacement components along the x- (u), y- (v) and z- (w) direction of the Rayleigh (blue dashed-dotted line) and folded Rayleigh wave propagating in the Γ -X direction in the EL approximation with a wave vector k = 4,590,095.5 m−1. The effect of refining the mesh is showed in the displacement components of the folded Rayleigh wave by increasing the number of elements of the starting mesh (black line) by a 33.33% (red dotted line).
Figure A4. Displacement components along the x- (u), y- (v) and z- (w) direction of the Rayleigh (blue dashed-dotted line) and folded Rayleigh wave propagating in the Γ -X direction in the EL approximation with a wave vector k = 4,590,095.5 m−1. The effect of refining the mesh is showed in the displacement components of the folded Rayleigh wave by increasing the number of elements of the starting mesh (black line) by a 33.33% (red dotted line).
Crystals 15 01030 g0a4

Appendix B. 2D Surface Phononic Crystal: Band Structure Around the Center of the Brillouin Zone

We examine the band structure around the point Γ , which has the symmetry of the C 4 v point group. Specifically, we study the modes at the edges of the first frequency gap at this point and second in the Γ -X direction. Taking k = 0 , there are four G m , n reciprocal vectors yielding m 2 + n 2 = 1 . As a consequence, two wave vectors in the k x direction, k 0 , ( 1 , 0 ) and k 0 , ( 1 , 0 ) [see Figure A5a], and two wave vectors in the k y direction, k 0 , ( 0 , 1 ) and k 0 , ( 0 , 1 ) [see Figure A5b], satisfy the phase-matching condition. Therefore, the bands corresponding to the ( m , n ) space harmonics related to these wave vectors, that is ( 1 , 0 ) , ( 0 , 1 ) , ( 1 , 0 ) , and ( 0 , 1 ) , converge to a single point at k = 0 ( ω = 2 π a v S A W or Ω = 2 π v t v S A W ) in the EL approximation, as shown in Figure A6. When comparing 1D and 2D SPnCs, the addition of harmonics ( 0 , 1 ) and ( 0 , 1 ) first increases the degeneracy from two to four and then introduces polarization in the y z -plane.
Figure A5. Wave vectors of the space harmonics of a wave propagating with a k along the k x direction and reciprocal vectors G n , m yielding m 2 + n 2 = 1 . For k = 0, the two wave vectors in (a) contribute to the two-fold frequency degeneracy found in the 1D SPhC, and the four wave vectors in (a,b) contribute to the four-fold frequency degeneracy found in the 2D SPhC.
Figure A5. Wave vectors of the space harmonics of a wave propagating with a k along the k x direction and reciprocal vectors G n , m yielding m 2 + n 2 = 1 . For k = 0, the two wave vectors in (a) contribute to the two-fold frequency degeneracy found in the 1D SPhC, and the four wave vectors in (a,b) contribute to the four-fold frequency degeneracy found in the 2D SPhC.
Crystals 15 01030 g0a5
The characters of the four-dimensional Γ r e d reducible representation based on the four space harmonics are listed in Table A1. Γ r e d is decomposed into irreducible representations of C 4 v so that
Γ r e d = A 1 + B 1 + E .
Therefore, the four-fold degeneracy of these modes will be lifted according to this decomposition when the surface is perturbed by introducing the 2D grating (Figure 1a). Figure A6 shows that the splitting of this degeneracy results in the opening of the frequency gap at k = 0 with A 1 and B 1 modes in the upper edge and E modes in the lower edge. The modes were assigned based on the transformation properties of the displacement profiles (bottom of Figure A6) under the symmetry operations of the C 4 v group. The displacement profiles of the A 1 and E modes show energy leakage to the bulk through longitudinal (z-polarized) and transverse [ (x, y)-polarized] displacement radiative channels, respectively. This result is consistent because the coordinate components z and ( x , y ) are suitable basis of the representations A 1 and E, respectively. In contrast, none of the coordinate components transform as the B 1 representation. Consequently, the mode with the surface displacement transforming as B 1 does not radiate to the bulk and is an example of symmetry-protected BIC.
Figure A6. Dispersion relation around the center of zone of modes forming the second frequency gap in the 2D surface phononic crystal. The color coding of the solid lines indicates the attenuation of the waves. The dashed lines show the dispersion curves derived from the EL approximation. The total displacement profile together with the x- and z-components of the modes at the edges of the frequency gap and at the zone center are shown at bottom part of the figure.
Figure A6. Dispersion relation around the center of zone of modes forming the second frequency gap in the 2D surface phononic crystal. The color coding of the solid lines indicates the attenuation of the waves. The dashed lines show the dispersion curves derived from the EL approximation. The total displacement profile together with the x- and z-components of the modes at the edges of the frequency gap and at the zone center are shown at bottom part of the figure.
Crystals 15 01030 g0a6
Table A1. Character table and basis functions for the irrreducible representations of group C 4 v . Characters for the reducible representation, Γ r e d , of the degenerate modes in the EL approximation for ω = 2 π a v S A W ( Ω = 2 π v t v S A W ).
Table A1. Character table and basis functions for the irrreducible representations of group C 4 v . Characters for the reducible representation, Γ r e d , of the degenerate modes in the EL approximation for ω = 2 π a v S A W ( Ω = 2 π v t v S A W ).
C 4 v E C 2 2 C 4 2 σ v 2 σ d
Γ r e d 40020
A111111z
A2111−1−1
B111−11−1
B211−1−11
E2−2000(x,y)

Appendix C. Far Field Polarization

The radiative wave propagates in the far field inside the bulk of the material with a wave vector k = ( k , k z ) and transverse polarization P ( k ) = p x ( k ) x ^ + p y ( k ) y ^ + p z ( k ) z ^ . The components of P ( k ) are given by p x ( k ) = x ^ · u , p y ( k ) = y ^ · u , and p z ( k ) = z ^ · u , where brakets denote spatial averaging of the displacement vector u over a unit cell on an xy plane inside the material (cf. Figure A7). To discuss the distribution of the far-field polarization, we have to define a two-dimensional projection in the xy plane that preserves the state of polarization [40]. The projected polarization vector is expressed as P ( k ) = p x ( k ) x ^ + p y ( k ) y ^ , where x ^ = k ^ , y ^ = z ^ × x ^ , and k ^ = k / k .
Figure A7. Schematic illustration of the projection of the far-field polarization vector onto a x y plane of the 2D SPnC unit cell.
Figure A7. Schematic illustration of the projection of the far-field polarization vector onto a x y plane of the 2D SPnC unit cell.
Crystals 15 01030 g0a7

Appendix D. 1D Surface Phononic Crystal: Simulated Band Structure

The 1D SPnC consists of a shallow periodic modulation in the x direction of the surface of a semi-infinite isotropic material with period a and in the form of w wide and h thick strips ( w = 0.375 a and h = 0.05 a ) parallel to the y direction [cf. inset of Figure A8a]. Although the 2-D FEM model is suitable for simulating the propagation of modes in the x direction and polarized in the sagittal plane, we conducted a full 3-D simulation, which grants the appearance of the shear horizontally polarized surface modes and allows for the study of propagating directions other than the x direction. Bloch–Floquet periodic boundary conditions were applied to the vertical boundaries of the unit cell. The dimension of the unit cell, denoted a, corresponds to the periodicity of the grating and confines the first Brillouin zone. The lateral dimension, b, was chosen to be smaller than a to prevent unwanted effects in the ω ( k ) region of interest related to periodicity in the y direction.
Figure A8. (a) Band structure of a one-dimensional surface phononic crystal composed of a periodic array of strips of period a calculated close to the boundary of the first Brillouin zone. The color coding of the solid lines indicates the attenuation of the waves. Dispersion curves of the Rayleigh waves (R) in the EL approximation as well as of the bulk transverse (T) waves are also shown. Displacement profiles of (b) the SAW at k a / π = 0.911 ( Ω = 2.65), (c) the leaky mode at k a / π = 0.745 ( Ω = 3.645) and (d) the accidental BIC mode at k a / π = 0.911 ( Ω = 3.17). The x- and z-components of the displacement field are shown for leaky mode, together with the total displacement. The inset shows the unit cell used in the FEM simulations.
Figure A8. (a) Band structure of a one-dimensional surface phononic crystal composed of a periodic array of strips of period a calculated close to the boundary of the first Brillouin zone. The color coding of the solid lines indicates the attenuation of the waves. Dispersion curves of the Rayleigh waves (R) in the EL approximation as well as of the bulk transverse (T) waves are also shown. Displacement profiles of (b) the SAW at k a / π = 0.911 ( Ω = 2.65), (c) the leaky mode at k a / π = 0.745 ( Ω = 3.645) and (d) the accidental BIC mode at k a / π = 0.911 ( Ω = 3.17). The x- and z-components of the displacement field are shown for leaky mode, together with the total displacement. The inset shows the unit cell used in the FEM simulations.
Crystals 15 01030 g0a8
Figure A8a shows the low-frequency part of the grating band structure for k = k x close to the boundary of the one-dimensional Brillouin zone. The color code characterizes the mode attenuation. The first sound line (T, dotted line) defines the threshold between the radiative and non-radiative modes. The branch that almost overlaps with the dotted line is formed by surface modes with pure shear horizontal polarization. The lower branch, which closely follows the dispersion of the Rayleigh modes in the EL approximation (R, dashed line), corresponds to the Rayleigh surface waves in the grating. The total displacement profile of the Rayleigh mode is illustrated in Figure A8b. As discussed in the main text, this branch is determined mainly by the n = 0 spatial harmonic for wavenumbers up to a region close to the Brillouin ZB at k a / π = 1 . Here, the phase-matching condition between the n = 0 and n = 1 harmonics, | k π a , 0 | = | k π a , 1 | results in the opening of a frequency gap at the zone center. The branch above the frequency gap (folded-back branch) is predominantly related to harmonic n = 1 . This corresponds to Rayleigh waves up to the critical frequency at the crossing of the first sound line. Above this frequency, the modes acquire a radiative transverse component related to the n = 0 harmonic, in addition to the evanescent components of the n = 1 harmonic. The attenuation increases from γ 1 × 10−16 for the non-radiative Rayleigh-like wave to γ 1 × 10−5 for radiative modes above the critical frequency. To clearly observe the features of radiative surface waves, the total displacement of the mode at k a / π = 0.745 with Ω = 3.645 and γ = 8 × 10−4 is plotted in Figure A8c], respectively. The wavefronts are visible, propagating within the material with a propagation direction forming an angle of approximately 50° from the surface, in agreement with Equation (26). The displacement near the surface of the leaky waves transforms according to representation A of the k -group C s . Because the x and z components form the basis of the representation A (see Table A2), the far field of the leaky modes below the second sound line consists of transverse bulk waves polarized in the sagittal plane only (see Equation (14)). At a certain value of k ( k a / π 0.911 and Ω 3.17), the radiative component is suppressed and the mode in the folded-back branch becomes a Rayleigh mode with virtually no attenuation. Figure A8d shows the total displacement of this mode, which can be referred to as an accidental BIC, showing a field profile localized on the surface without radiated field into the bulk.
Table A2. Character table and basis functions for the irrreducible representations of group C s .
Table A2. Character table and basis functions for the irrreducible representations of group C s .
C s E σ
A′11x,z
A″1−1y
At k x = 0 , the branches in the EL approximation related to n = 1 and n = 1 harmonics intersect following the phase-matching condition between these harmonics, | k 0 , 1 | = | k 0 , 1 | (dashed lines in Figure A9). All irreducible representations of the point group C 2 v are one-dimensional (see Table A3), thus no two-fold degenerate modes can occur. Therefore, the degeneracy is lifted when the surface is perturbed with a finite-sized grating (see Figure A9). The resulting two modes transform according to the B 1 and A 1 characters, and their displacement profiles are shown in Figure A9. Although the modes are predominantly related to the n = 1 and n = 1 harmonics with evanescent components, the n = 0 harmonic introduces radiative components. Because the frequencies are above the first and second sound lines of the space harmonic n = 0 , radiation to the bulk might be contributed by both transverse and longitudinal components. As expected, the displacement profiles inside the bulk material show wavefronts in the direction normal to the surface plane ( θ = 90 ). From the plots of the components of the displacement profiles, we see that the far field of modes B 1 and A 1 is x- and z-polarized, respectively, which is expected from the basis functions associated with these representations (see Table A3).
Table A3. Character table and basis functions for the irrreducible representations of group C 2 v .
Table A3. Character table and basis functions for the irrreducible representations of group C 2 v .
C 2 v E C 2 σ σ x
A11111z
A211−1−1
B11−11−1x
B21−1−11y
Figure A9. Dispersion relation around the center of zone of modes forming the second frequency gap. The color coding of the solid lines indicates the attenuation of the waves. The dashed lines show the dispersion curves in the EL approximation of the branches related to the n = 1 and n = 1 space harmonics. The total displacement profile together with the x- and z-components of the modes at the edge of the frequency gap and at the zone center are shown at the right side of the figure.
Figure A9. Dispersion relation around the center of zone of modes forming the second frequency gap. The color coding of the solid lines indicates the attenuation of the waves. The dashed lines show the dispersion curves in the EL approximation of the branches related to the n = 1 and n = 1 space harmonics. The total displacement profile together with the x- and z-components of the modes at the edge of the frequency gap and at the zone center are shown at the right side of the figure.
Crystals 15 01030 g0a9

References

  1. Destrade, M.; Saccomandi, G. Linear elastodynamics and waves. In Continuum Mechanics Volume II; Merodio, J., Saccomandi, G., Eds.; Encyclopedia of Life Support Systems Publishers: Oxford, UK, 2012; pp. 16–21. [Google Scholar]
  2. Schröder, C.T.; Scott, W.R., Jr. On the complex conjugate roots of the Rayleigh equation: The leaky surface wave. J. Acoust. Soc. Am. 2001, 110, 2867–2877. [Google Scholar] [CrossRef]
  3. Lim, T.C.; Farnell, G.W. Search for Forbidden Directions of Elastic Surface-Wave Propagation in Anisotropic Crystals. J. Appl. Phys. 1968, 39, 4319–4325. [Google Scholar] [CrossRef]
  4. Farnell, G.W. Properties of Elastic Surface Waves. In Physical Acoustics Volume VI; Mason, W.P., Thurston, R.N., Eds.; Academic Press: New York, NY, USA, 1970; pp. 120–149. [Google Scholar]
  5. Kuok, M.H.; Ng, S.C.; Zhang, V.L. Angular dispersion of surface acoustic waves on (001), (110), and (111) GaAs. J. Appl. Phys. 2001, 89, 7899–7902. [Google Scholar] [CrossRef]
  6. Tarasenko, A.; Čtvrtlík, R.; Kudělka, R. Theoretical and experimental revision of surface acoustic waves on the (100) plane of silicon. Sci. Rep. 2021, 11, 2845. [Google Scholar] [CrossRef]
  7. Stegeman, G.I. Normal-mode surface waves in the pseudobranch on the (001) plane of gallium arsenide. J. Appl. Phys. 1975, 47, 1712–1713. [Google Scholar] [CrossRef]
  8. Every, A.G. Supersonic surface acoustic waves on the 001 and 110 surfaces of cubic crystals. J. Acoust. Soc. Am. 2015, 148, 2937–2944. [Google Scholar] [CrossRef] [PubMed]
  9. Hsu, C.W.; Zhen, B.; Stone, A.D.; Joannopoulos, J.D.; Soljačić, M. Bound states in the continuum. Nat. Rev. Mater. 2016, 1, 16048. [Google Scholar] [CrossRef]
  10. Yamanouchi, K.; Shibayama, K. Propagation and amplification of Rayleigh waves and piezoelectric leaky surface waves in LiNbO3. J. Appl. Phys. 1971, 43, 856–862. [Google Scholar] [CrossRef]
  11. Shimizu, Y.; Tanaka, M. Leaky surface wave on quartz and a new cut with extremely small temperature coefficient. Jpn. J. Appl. Phys. 1984, 24, 139–141. [Google Scholar] [CrossRef]
  12. Hashimoto, K.-Y.; Yamaguchi, S.; Mineyoshi, S.; Kawachi, O.; Ueda, M.; Endoh, G. Optimum leaky-SAW cut of LiTaO3 for minimised insertion loss devices. In Proceedings of the 1997 IEEE Ultrasonics Symposium Proceedings. An International Symposium, Toronto, ON, Canada, 5–8 October 1997; pp. 245–254. [Google Scholar]
  13. Matsukura, F.; Uematsu, M.; Hosaka, K.; Kakio, S. Longitudinal-type leky surface acoustic wave on LiNbO3 with high-velocity thin film. In Proceedings of the 2013 IEEE International Ultrasonics Symposium (IUS), Prague, Czech Republic, 21–25 July 2013; pp. 1692–1695. [Google Scholar]
  14. Kakio, S.; Hosaka, K. Loss reduction of leaky surface acoustic wave by loading with high-velocity thin film. Jpn. J. Appl. Phys. 2016, 55, 07KD11. [Google Scholar] [CrossRef]
  15. Naumenko, N.F. High-velocity non-attenuated acoustic waves in LiTaO3/quartz layered substrates for high frequency resonators. Ultrasonics 2019, 95, 1–5. [Google Scholar] [CrossRef] [PubMed]
  16. Farnell, G.W.; Adler, E.L. Elastic Surface Wave Propagation in Thin Layers. In Physical Acoustics Volume IX; Mason, W.P., Thurston, R.N., Eds.; Academic Press: New York, NY, USA, 1972; pp. 35–127. [Google Scholar]
  17. Glass, N.E.; Maradudin, A.A. Leaky surface-elastic waves on both flat and strongly corrugated surfaces for isotropic, nondissipative media. J. Appl. Phys. 1983, 54, 796–805. [Google Scholar] [CrossRef]
  18. Maznev, A.A.; Every, A.G. Bound acoustic modes in the radiation continuum in isotropic layered systems without periodic structures. Phys. Rev. B 2018, 97, 014108. [Google Scholar] [CrossRef]
  19. Maznev, A.A. Band gaps and Brekhovskikh attenuation of laser-generated surface acoustic waves in a patterned thin film structure on silicon. Phys. Rev. B 2008, 78, 155323. [Google Scholar] [CrossRef]
  20. Maznev, A.A.; Every, A.G. Surface acoustic waves in a periodically patterned layered structure. J. Appl. Phys. 2009, 106, 113531. [Google Scholar] [CrossRef]
  21. Tong, H.; Liu, S.; Zhao, M.; Fang, K. Observation of phonon trapping in the continuum with topological charges. Nat. Commun. 2020, 11, 5216. [Google Scholar] [CrossRef]
  22. Muzar, E.; Stotz, J.A.H. Surface acoustic wave modes in two-dimensional shallow void inclusion phononic crystals on GaAs. J. Appl. Phys. 2019, 126, 025104. [Google Scholar] [CrossRef]
  23. Zhen, B.; Hsu, C.W.; Lu, L.; Stone, A.D.; Soljačić, M. Topological Nature of Optical Bound States in the Continuum. Phys. Rev. Lett. 2014, 113, 257401. [Google Scholar] [CrossRef]
  24. Wang, J.; Li, P.; Zhao, X.; Qian, Z.; Wang, X.; Wang, F.; Zhou, X.; Han, D.; Peng, C.; Shi, L.; et al. Optical bound states in the continuum in periodic structures: Mechanisms, effects, and applications. Photonics Insights 2024, 3, R01–R26. [Google Scholar] [CrossRef]
  25. Maradudin, A.A.; Zierau, W. Surface acoustic waves of sagittal and shear-horizontal polarizations on large-amplitude gratings. Geophys. J. Int. 1994, 118, 325–332. [Google Scholar]
  26. Hu, P.; Wang, J.; Jiang, Q.; Wang, J.; Shi, L.; Han, D.; Zhang, Z.Q.; Chan, C.T.; Zi, J. Global phase diagram of bound states in the continuum. Optica 2022, 9, 1353–1361. [Google Scholar] [CrossRef]
  27. Nardi, D.; Banfi, F.; Giannetti, C.; Revaz, B.; Ferrini, G.; Parmigiani, F. Pseudosurface acoustic waves in hypersonic surface phononic crystals. Phys. Rev. B 2009, 80, 104119. [Google Scholar]
  28. Westafer, R.S.; Mohammadi, S.; Adibi, A.; Hunt, W.D. Computing Surface Acoustic Wave Dispersion and Band Gaps. In Proceedings of the COMSOL Conference, Boston, MA, USA, 8–10 October 2009. [Google Scholar]
  29. Basu, U.; Chopra, K. Perfectly matched layers for time-harmonic elastodynamics of unbounded domains:theory and finite-element implementation. Comput. Methods Appl. Mech. Engrg. 2001, 192, 1337–1375. [Google Scholar]
  30. Hosten, B.; Castaings, M. Finite elements methods for modeling the guided waves propagation in structures with weak interfaces. J. Acoust. Soc. Am. 2005, 117, 1108–1113. [Google Scholar] [CrossRef]
  31. Graczykowski, B.; Alzina, F.; Gomis-Bresco, J.; Sotomayor Torres, C.M. Finite element analysis of true and pseudo surface acoustic waves in one-dimensional phononic crystals. J. Appl. Phys. 2008, 119, 025308. [Google Scholar] [CrossRef]
  32. Ke, W.; Yaacoubi, S.; Mckeon, P. On the comparison of absorbing regions methods. In Proceedings of the Acoustics 2012, Nantes, France, 23–27 April 2012. [Google Scholar]
  33. Gulyaev, Y.V.; Plesskii, V.P. “Slow” acoustic surface waves in solids. Sov. Tech. Phys. Lett. 1977, 3, 87–88. [Google Scholar]
  34. Born, M.; Wolf, E. Principles of Optics: Electromagnetic Theory of Propagation, Interference and Diffraction of Light, 7th ed.; Cambridge University Press: Cambridge, UK, 1999. [Google Scholar]
  35. Freund, I. Polarization singularity indices in Gaussian laser beams. Opt. Commun. 2002, 201, 251–270. [Google Scholar] [CrossRef]
  36. Senthilkumaran, P. Singularities in Physics and Engineering, 1st ed.; IOP Publishing: Bristol, UK, 2018. [Google Scholar]
  37. Rayleigh, L. Theory of Sound; Dover Publications: New York, NY, USA, 1945; Sec. 272a. [Google Scholar]
  38. Boulanger, P.; Hayes, M.A. Bivectors and Waves in Mechanics and Optics, 1st ed.; Chapman & Hall: London, UK; New York, NY, USA, 1993. [Google Scholar]
  39. Aono, T.; Tamura, S.T. Effects of isotropic overlayer on the focusing of surface phonons. Phys. Rev. B 1997, 55, 6754. [Google Scholar] [CrossRef]
  40. Yoda, T.; Notomi, M. Generation and annihilation of topologically protected bound states in the continuum and circularly polarized states by symmetry breaking. Phys. Rev. Lett. 2020, 125, 053902. [Google Scholar] [CrossRef]
Figure 1. (a) 3D unit cell used in the FEM simulation of the 2D surface phononic crystal, where a is the lattice constant, and d and h are the height and diameter of the cylinder, respectively. (b) The first Brillouin zone of a square lattice crystal.
Figure 1. (a) 3D unit cell used in the FEM simulation of the 2D surface phononic crystal, where a is the lattice constant, and d and h are the height and diameter of the cylinder, respectively. (b) The first Brillouin zone of a square lattice crystal.
Crystals 15 01030 g001
Figure 2. (a) Band structure of a surface phononic crystal composed of cylindrical pillars calculated along Γ -X direction of the first Brillouin zone. The lattice parameter is a, and the height and diameter of the cylinders are h = 0.05 a and d = 0.75 a , respectively. The color coding of the solid lines indicates wave attenuation. Dispersion curves of the Rayleigh waves (R) in the EL approximation, as well as those of the bulk transverse (T) and longitudinal (L) waves, are also shown. The band structure close to the boundary of the 1st Brillouin zone (b,c) at the zone center. Displacement profiles of (d) the accidental BIC ( k a / π = 0.877 and Ω = 3.2758), (e) a leaky mode ( k a / π = 0.738 and frequency Ω = 3.675), (f) a SAW ( k a / π = 0.877 and frequency Ω = 2.524), and (g) the symmetry-protected BIC ( k a / π = 0.0 and frequency Ω = 5.848).
Figure 2. (a) Band structure of a surface phononic crystal composed of cylindrical pillars calculated along Γ -X direction of the first Brillouin zone. The lattice parameter is a, and the height and diameter of the cylinders are h = 0.05 a and d = 0.75 a , respectively. The color coding of the solid lines indicates wave attenuation. Dispersion curves of the Rayleigh waves (R) in the EL approximation, as well as those of the bulk transverse (T) and longitudinal (L) waves, are also shown. The band structure close to the boundary of the 1st Brillouin zone (b,c) at the zone center. Displacement profiles of (d) the accidental BIC ( k a / π = 0.877 and Ω = 3.2758), (e) a leaky mode ( k a / π = 0.738 and frequency Ω = 3.675), (f) a SAW ( k a / π = 0.877 and frequency Ω = 2.524), and (g) the symmetry-protected BIC ( k a / π = 0.0 and frequency Ω = 5.848).
Crystals 15 01030 g002
Figure 3. Distribution of the far-field polarization state near the accidental BIC: (a) Intensity, (b) azimuth ( γ 0 ), and (c) ellipticity ( χ 0 ).
Figure 3. Distribution of the far-field polarization state near the accidental BIC: (a) Intensity, (b) azimuth ( γ 0 ), and (c) ellipticity ( χ 0 ).
Crystals 15 01030 g003
Figure 4. (a) Band-structure representation in the one-dimensional empty lattice approximation up to the second Brillouin zone. The black, dark blue, and light blue lines correspond to the Rayleigh modes, transverse modes, and longitudinal modes of velocities v S A W , v t and v l , respectively. n = 0 ( n = ± 1 ) space harmonic is represented by solid (dotted) lines. Slowness diagram for k x = π / a (b), k x / ω r > v t 1 (c), and ( k x k g ) / ω r > v t 1 (d). The slowness curves with black (red) lines indicate real-valued (imaginary-valued) k z / ω solutions. Slowness curves of the n = 0 ( n = ± 1 ) space harmonic are plotted as solid (dotted) lines.
Figure 4. (a) Band-structure representation in the one-dimensional empty lattice approximation up to the second Brillouin zone. The black, dark blue, and light blue lines correspond to the Rayleigh modes, transverse modes, and longitudinal modes of velocities v S A W , v t and v l , respectively. n = 0 ( n = ± 1 ) space harmonic is represented by solid (dotted) lines. Slowness diagram for k x = π / a (b), k x / ω r > v t 1 (c), and ( k x k g ) / ω r > v t 1 (d). The slowness curves with black (red) lines indicate real-valued (imaginary-valued) k z / ω solutions. Slowness curves of the n = 0 ( n = ± 1 ) space harmonic are plotted as solid (dotted) lines.
Crystals 15 01030 g004
Figure 5. Dimensionless frequency of the accidental BIC obtained from the FEM simulation (squares) of a 1D SPnC as a function of the ratio between the period and the width of the groove. The horizontal dashed lines mark the position of the BIC from the closing channel condition in the EL approximation ( Ω a u x ) and the value of the critical frequency ( Ω c ), respectively.
Figure 5. Dimensionless frequency of the accidental BIC obtained from the FEM simulation (squares) of a 1D SPnC as a function of the ratio between the period and the width of the groove. The horizontal dashed lines mark the position of the BIC from the closing channel condition in the EL approximation ( Ω a u x ) and the value of the critical frequency ( Ω c ), respectively.
Crystals 15 01030 g005
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

Alzina, F. Rayleigh Bound States in the Continuum in Shallow Surface Relief Phononic Crystals. Crystals 2025, 15, 1030. https://doi.org/10.3390/cryst15121030

AMA Style

Alzina F. Rayleigh Bound States in the Continuum in Shallow Surface Relief Phononic Crystals. Crystals. 2025; 15(12):1030. https://doi.org/10.3390/cryst15121030

Chicago/Turabian Style

Alzina, Francesc. 2025. "Rayleigh Bound States in the Continuum in Shallow Surface Relief Phononic Crystals" Crystals 15, no. 12: 1030. https://doi.org/10.3390/cryst15121030

APA Style

Alzina, F. (2025). Rayleigh Bound States in the Continuum in Shallow Surface Relief Phononic Crystals. Crystals, 15(12), 1030. https://doi.org/10.3390/cryst15121030

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop