1. Introduction
In nearly 30 years since the development of the first phononic crystal (PC) [
1], a plethora of interesting properties of PCs has been discovered, foremost of which are band gaps, frequency ranges over which elastic wave propagation is forbidden. Band gaps give PCs a diverse range of applications including vibration suppression [
2], waveguiding [
3], and energy harvesting [
4]. Band gaps are easily defined and identified for undamped PCs, which constitute the majority of published works, but are much more difficult to define and identify in damped PCs. As all real materials exhibit damping, understanding band gaps in damped PCs is thus critical to developing real-world applications for PCs. The purpose of this work is to develop a method for defining and identifying band gaps in damped PCs which is based on measurable physical quantities. Such a method is amenable to experimental verification, which allows the performance of physically realized PCs to be assessed relative to model predictions.
Damping in PCs, and the related structures known as acoustic metamaterials, has been an increased focus of research in the last decade. Damping creates phenomena not found in undamped PCs, such as “wavenumber band gaps” [
5], and conversely, PCs can create novel damping behaviors not found in the bulk material, such as anisotropic dissipation [
6]. Damping is also unintentionally but inevitably present in many PCs which rely on key properties of materials that also happen to be damped. For example, locally resonant PCs [
7], pattern transformation-tunable band gaps [
8], and magnetically-tunable band gaps [
9,
10,
11] were enabled by the low stiffness, ability to undergo large strains, and magnetic tunability of elastomers, respectively. The inclusion of damping strongly affects the PC band structure, causing pronounced smoothing of dispersion curves [
12] or a re-ordering of dispersion curves known as “branch overtaking” [
13,
14]. In damped band structures, either frequency [
13], wavenumber [
12], or both [
15] become complex-valued, which creates challenges for identifying band gaps. In some works, such as those previously cited on elastomeric PCs [
7,
8,
9,
10,
11], damping is neglected to simplify the interpretation of the band structures. However, a full understanding of wave propagation can only be obtained when damping is considered, due to the strong effect of damping on the band structure. In particular, the need to understand the damped band structure of the magneto-active metastructures studied in [
11] provided a strong initial impetus for the current work, as the elastomeric material from which they were fabricated is highly viscoelastic.
In works where damping is included, various methods have been developed to aid in interpreting band structures, which generally depend upon the class of waves that are studied: free waves or driven waves. Free waves can decay in time and in space [
15]. However, the most common analysis technique for free wave solutions is the “indirect” or
method, which proceeds by prescribing a real-valued Bloch wavenumber and solving the resulting eigenvalue problem for the complex frequency [
14,
16], yielding waves which decay in time but not in space. The damping ratio
has become a standard figure of merit to characterize the temporal decay of the wave and has been used, e.g., to study the phenomenon known as metadamping [
17,
18,
19] and to study the transition between viscously and viscoelasticly damped PCs [
20]. An alternative figure of merit that has found limited use is the quality factor [
21]. On the other hand, driven waves decay in space but not in time and are obtained using the “direct” or
method, by prescribing real frequencies and computing complex wavenumbers. Band structures for damped driven waves are much more difficult to interpret than band structures for undamped driven waves, because all wave modes are propagative (with spatial decay), and the cusps of the dispersion curves become highly rounded, even for weak damping [
12]. The rounding of the dispersion curves makes it difficult to identify band gap edge frequencies and has led several authors to use various figures of merit to quantify the wave decay: the minimum imaginary component of all wavenumbers at a given frequency [
22,
23], the minimum imaginary component normalized by wavenumber absolute value [
24], and the effective loss factor [
22,
25]. In [
22,
24], the respective figures of merit are visualized with a light-to-dark shading scheme to visually indicate regions of highly attenuated wave propagation. However, the lack of a standardized figure of merit illustrates the ongoing difficulties of identifying band gaps for damped driven wave propagation.
An additional difficulty lies in identifying polarized band gaps, which may also exist in PCs even when complete band gaps cannot be found. Polarized band gaps have long been studied in 2D PCs, where different band gaps are observed for modes with in-plane displacements and out-of-plane displacements, see, e.g., in [
26]. More recently, polarized band gaps with other polarizations (flexural, torsional, longitudinal) were discovered in lattice-based PCs [
11,
27,
28,
29]. Identification of polarized band gaps is more complicated than complete band gaps, because it requires a method to classify the polarization of each Bloch mode shape. In [
11,
27,
28,
29], manual inspection of the Bloch mode shapes was used to determine the subset of modes that align with the polarization of interest, which is problematic as it requires judgment calls of whether each mode is significant for a given direction. Additionally, the PCs in [
11,
27,
28,
29] were physically realized using polymeric and elastomeric materials which have significant damping. Including damping in the band structure computation necessitates simultaneous consideration of the polarization and rate of decay for each mode, and to the authors’ knowledge, no such method has yet been proposed.
Band gaps may also be studied by considering finite structures created by truncating infinite PCs. The distribution of the finite structure’s natural frequencies is dictated by the dispersion relation of the unit cell [
30,
31], which prevents natural frequencies from falling within the band gap, causing a deep trough in vibration transmission in the band gap frequencies. The transmission trough increases in depth with the number of unit cells in the truncated structure [
32], which is due to the larger number of poles in the frequency response function that lie below the band gap [
33]. It has further been shown that the dissipated power in a finite periodic structure is confined near the source for excitation frequencies within the band gap [
34], which has been proposed as an alternate means to identify band gaps from finite structures. The correspondence between the dynamics of infinite and finite PCs is especially significant as it allows for experimental verification of band gap existence. However, the existence of damping in all real materials causes smoothing of the transmission trough edges (see, e.g., in [
25]), which makes experimental identification of band gaps difficult for highly damped materials [
11].
The goal of this work is to develop a method by which polarized band gap-like frequency ranges can be systematically identified in damped PCs. As damped PCs may exhibit strong wave attenuation due to Bragg scattering, yet possess no true band gaps due to the effect of dissipation on the eigenmodes, we introduce the term “fuzzy band gaps” for frequency ranges with strong spatial attenuation of waves which may or may not be true band gaps. We define a fuzzy band gap as any frequency range where the transmitted wave amplitude through a finite periodic structure is less than a prescribed threshold. Though the threshold wave amplitude is arbitrary, it can be chosen based on physical grounds or engineering requirements, such as the maximum allowable vibration amplitude of a structure. Although “fuzziness” in PCs can arise from multiple sources, such as disorder or parasitic resonators, in this article we focus only on fuzziness arising from dissipation. We then select a measurable physical quantity (e.g., displacement or force) as a measure of wave amplitude, approximate the transmitted wave amplitude in a finite PC using a weighted superposition of the Bloch waves, and search for fuzzy band gaps by comparing the transmitted wave amplitude to the threshold amplitude. The use of a measurable physical quantity allows the fuzzy band gaps thus obtained to be directly verified by experiments. We first validate the evanescence indicator for a simple 1D viscoelastic diatomic lattice and show that the predicted fuzzy band gaps agree well with simulations of the vibration transmission through finite PCs. The “fuzzy” moniker arises from the fact that the evanescence indicator reveals smooth transitions between propagating and decaying wave behavior, which make the edges of a fuzzy band gap appear like an out-of-focus or “fuzzy” photo. We next validate the evanescence indicator on “magneto-active metastructures”, which are damped PCs with complex 3D geometry based on prior work [
11], using a combination of finite element method simulations and experiments. We show that the evanescence indicator is capable of identifying fuzzy band gaps with respect to arbitrary wave polarizations, including hybrid polarizations that consist of a superposition of two or more simple polarizations. Because of its applicability to damped PCs with complex geometries and its direct link to experimental results, the evanescence indicator is a powerful tool for interpreting damped band structures of PCs fabricated from real materials.
2. Theory
In this section, we derive an “evanescence indicator”, a figure of merit which quantifies how strongly waves are attenuated in a periodic structure at a given frequency. To compute this evanescence indicator only requires knowledge of the dispersion relation for an infinite periodic structure, the accompanying Bloch wave mode shapes, a threshold displacement amplitude of interest , and the number of unit cells N that defines the extents of a finite periodic structure of reference. This method is applicable for both undamped and damped structures with 1D periodicity.
2.1. Background
According to the Bloch theorem, as adapted to elastodynamics, waves propagating in periodic structures can be described by a basis of eigenfunctions having the form
where
and
t are the spatial and time coordinates, respectively;
is the displacement field;
is a function with the same periodicity as the periodic structure;
is the Bloch wavevector; and
is the angular frequency. Quantities in boldface denote vectors and
,
,
,
may all be complex-valued, depending on the frequency of excitation and the material properties of the structure.
For driven wave propagation, i.e., waves which are excited by a harmonic displacement or force and decay in space but not in time, the “direct” or
method is used to solve the dispersion relation. This method proceeds by prescribing a real-valued angular frequency
and solving the resulting eigenvalue problem for the Bloch wavevector
. Writing the Bloch wavevector in terms of its real and imaginary parts and substituting
and
into (
1) yields
It is evident that the imaginary-valued exponential function governs the spatial and temporal oscillation of the wave, while the real-valued exponent defines an “envelope” that governs the spatial growth or decay of the wave. If the wavevector is real-valued, i.e., , then the wave undergoes no change in amplitude as it propagates and is termed a “propagating” wave. However, for complex-valued , the wave amplitude will decrease with increasing and the wave is known as an “evanescent” wave. As an evanescent wave propagates through a medium of infinite extent, that is, , the wave amplitude decays to zero as .
In undamped periodic media, the dispersion relation can be partitioned into discrete “pass bands” (frequency ranges where at least one real-valued exists) and “stop bands” or “band gaps” (frequency ranges where all are complex, i.e., all waves are evanescent). Within a band gap, wave transmission in an unbounded medium is prohibited, because all wave modes decay to zero amplitude.
However, it is not so straightforward to define band gaps in finite structures or in damped periodic structures. In finite periodic structures created by truncating an infinite structure, a nonzero transmission is observed at all frequencies, even within the band gap, because for any finite , . In damped periodic structures, all are complex, so all wave modes are evanescent according to the traditional definition. In each of these cases, the rate of spatial decay of the wave amplitude (which depends on ) becomes important for characterizing the effectiveness of the phononic crystal at suppressing wave transmission. The developed evanescence indicator uses this basic concept of faster/slower wave decay and links it to measurable physical quantities such as displacement or force, so that band gaps in damped phononic crystals can be systematically defined.
2.2. Derivation of Evanescence Indicator
Our general approach to defining an evanescence indicator for damped systems is as follows. First, we select a finite structure of reference for which the wave attenuation is measured and choose a physical quantity (e.g., stress or displacement) as a measure of wave amplitude. For concreteness, we study two examples in which the displacement at a single point and the average displacement magnitude over several points are selected to characterize the wave amplitude. Next, we approximate the wave amplitude for all frequencies of interest using information directly extracted from the dispersion relation: a weighted summation of the Bloch mode shapes, so that the contributions of all wave modes are considered simultaneously. The amplitude is calculated at two locations (“input” and “output”) and the normalized transmission (output divided by input) is computed. Finally, we define “fuzzy band gaps” using a threshold value of normalized transmission as the criterion for fuzzy band gap existence: if the transmission is greater than the threshold transmission, there is no fuzzy band gap, and if the transmission is less than the threshold transmission, a fuzzy band gap is considered to exist. Two methods are discussed for determining the weights in the summation of Bloch modes, because the correct calculation of the weights is critical to accurately identifying the fuzzy band gaps.
Consider a structure which is infinitely periodic in the
x-direction and of finite extent in the
y- and
z-directions. The structure is excited by arbitrary forcing (prescribed forces and/or displacements) at a real frequency
. Bloch wave propagation occurs in the
x-direction only, so that
, where the scalar
k is the magnitude of the wavevector
. As the Bloch wave modes form a basis for the displacement field, the total displacement field of the wave can be expressed as a superposition of Bloch waves:
where
is the amplitude of the
jth Bloch wave having wavenumber
and mode shape
, and the summation is carried out over all Bloch modes existing at frequency
. As all Bloch wave modes occur in pairs (
,
) and the summation is over all Bloch wave modes, waves propagating in both the
and
directions are present in the summation, and the modal amplitudes
are chosen to give a physically realistic solution (i.e., with
for physically unrealistic wave modes). We define the “input displacement” as the displacement
of the structure at the point
and the “output displacement” as the displacement
of the structure at the point
. Note that the terms “input” and “output” are chosen merely to reflect the convention that transmission is output divided by input; the locations of
and
are arbitrary and are left to the user to be chosen as appropriate for a given application. As the dynamics of finite periodic structures are known to resemble the dynamics of the corresponding infinite structure [
32], we propose that the displacement at points
in a
finite periodic structure can be approximately expressed by Equation (
3). This approximation is made in order to link the information contained in the dispersion relation to the finite structure upon which the fuzzy band gap definition is based. The magnitudes of the displacements at the points
are
which forms the basis for our proposed evanescence indicator. Here,
n is independently taken as
to indicate that the displacement magnitude is evaluated at the points
and
.
To develop the definition of a fuzzy band gap, we first give an alternate definition of a band gap. In infinite undamped periodic structures, band gaps are identified by seeking frequencies
where all Bloch wavenumbers
are complex or imaginary. Placing the input point
at the origin and the point
an infinite distance away on the
x-axis, it is clear that the the transmitted wave amplitude
is 0 within the band gap, as
and all wavenumbers have nonzero
within the band gap. Therefore, a band gap can be equivalently defined as a range of frequencies
for which the displacement amplitude transmitted through an infinite structure is equal to a threshold amplitude, namely 0,
This definition is not amenable to damped structures because it would define a band gap to exist at all frequencies, as all wavenumbers are complex in damped structures and therefore decay to 0 as
. Therefore, the first adaptation we make to this definition is to place the output point
a
finite distance from the input point
, rather than an infinite distance. In particular, we place
on the right boundary of a finite structure with
N unit cells:
(
a denotes the lattice constant of the periodic structure). We refer to the finite structure with
N unit cells as the “finite structure of reference”. Relocating the output point necessitates a second adaptation because the criterion of 0 transmitted wave amplitude clearly cannot be satisfied for a finite structure, as
for all finite
. We therefore introduce a threshold value of the normalized transmission,
, as the criterion for fuzzy band gap existence. Rather than requiring the displacement amplitude to be
exactly zero, we require it to be
approximately zero, or more precisely, we require it to be less than or equal to a small but nonzero number. The chosen value of the threshold transmission is arbitrary, but it can be selected based on physical grounds, such as the noise floor of an experiment, or the maximum allowable vibration amplitude of a structure for a desired fatigue life. Fuzzy band gaps may now be determined in damped PCs using a definition similar to that of band gaps in undamped PCs, with
replacing 0 as the threshold amplitude for defining a fuzzy band gap, and with
and
taken on the left and right boundaries, respectively, of the finite structure of reference, rather than placing
at infinity. Fuzzy band gaps are therefore defined as all frequencies
satisfying
where
and
are computed using Equation (
4). From Equation (
7), we define the evanescence indicator
for a finite structure of reference with
N unit cells as
Note that for , a fuzzy band gap is considered to exist. In contrast to the traditional band gap definition (where the band gaps could be identified by analyzing each wave mode individually as the transmitted amplitude of every wave mode must be 0 for the total transmitted wave amplitude to be identically zero), the generalized fuzzy band gap definition requires the additive contributions of all wave modes to be considered together since nonzero transmissions are allowed within the fuzzy band gap. The influence of damping is implicitly captured in this evanescence indicator because the Bloch wavenumbers and mode shapes depend on the damping.
It is straightforward to compute , and it is assumed that is known. (Note that can usually be computed while solving the eigenvalue problem that gives the dispersion relation, e.g., by computing the eigenvectors of the transfer matrix or the finite element, dynamic stiffness matrix.) However, the value of , the excited amplitude of wave mode j, is not readily apparent. Furthermore, the value of cannot be selected arbitrarily, as determines the existence of polarized fuzzy band gaps. For example, if a longitudinally-polarized fuzzy band gap (with displacements primarily in the x-direction) is sought, the excited amplitude should be very small for flexural- or torsional-type Bloch waves, which effectively eliminates the contribution of the flexural or torsional modes to the evanescence indicator. Therefore, a method to determine appropriate values of is required.
2.3. Propagated Bloch Mode Expansion
In this section, we propose a method by which the unknown coefficients
can be estimated, which is based on the modal superposition method and inspired by the Reduced Bloch Mode Expansion Method (RMBE) [
16]. Consider an arbitrary finite metastructure composed of
N unit cells and having
M dofs, which can be a discrete system or a discretization of a continuous system obtained using, e.g., the finite element method. Assuming harmonic motion, the response of the structure is governed by
where
is the dynamic stiffness matrix which captures the effects of inertia, damping, and elastic forces;
is the vector of nodal displacements; and
is the vector of external forces. The system is subject to arbitrary excitation at frequency
: either a prescribed displacement or force applied to each node. The system can be condensed to remove the prescribed displacements, which yields the modified equation of motion:
where
is the condensed dynamic stiffness matrix and
is a vector that collects the effects of all prescribed forces and displacements.
In traditional modal superposition, the eigenmodes of the system are used to transform the problem from the physical coordinates to a set of “modal” coordinates. If all M eigenmodes are used, this transformation simply involves a change of basis to the orthonormal basis consisting of the normal mode shapes, and the exact displacements can be recovered. A good approximation may be obtained by using only a small subset of the eigenmodes, which greatly reduces the computational cost of solving the system response at many frequencies. The eigenmodes are typically chosen by considering those normal modes whose eigenfrequencies are close to the frequency range of interest, as normal modes with greatly differing frequencies will not be preferentially excited.
Here, we propose to modify the modal superposition approach, so that the set of modal basis vectors is chosen not from the normal modes of the finite system, but rather from the Bloch modes shapes determined from the dispersion relation at the frequency of interest. This choice is motivated by many previous works which show that the dynamics of a finite periodic system with a sufficient number of unit cells, closely approximates the dynamics of the infinite system (see, e.g., in [
32]). As the Bloch mode shapes are computed only for a single unit cell, we create modal basis vectors
(
) for the finite structure by “propagating” a set of
n Bloch mode shapes using the periodic extension of
and the Floquet propagator
. We call these modal basis vectors “propagated Bloch modes”. We then expand the displacement vector
in terms of the propagated Bloch modes:
where
is an
matrix of the propagated Bloch modes and
is a vector of amplitudes associated with each mode. Thus, the
jth element of
is
, the unknown coefficients required for the evanescence indicator. This “propagated Bloch mode expansion” (PBME) is so-named in reference to the “reduced Bloch mode expansion” (RMBE) [
16], which similarly uses modal superposition of Bloch modes, but for efficient computation of band structures. Substituting Equation (
11) into Equation (
10), and premultiplying by the transpose
of
, yields
where
is an
matrix and
is an
n-element vector. Solving Equation (
12) gives the modal amplitudes
, which are used to compute the evanescence indicator (Equation (
8)).
2.4. Modal Participation Factor
The modal participation factor (MPF) is commonly used in structural dynamics in conjunction with the modal superposition method to determine if a suitable set of eigenmodes have been included in the reduced system model. The MPF
measures the contribution or “participation” of mode
to state
[
35,
36]. The states are typically chosen as either rigid-body translations
along the coordinate axes or rigid-body rotations
about axes parallel to the coordinate axes. When the states are chosen in this way, the MPFs give an indication of the excited amplitude of mode
in response to translational or rotational excitations in the specified direction. As the MPFs measure the amplitude of a mode in response to a given polarization of excitation, they are a logical choice for the modal amplitudes
in the weighted summation of Bloch modes. They also allow us to study fuzzy band gaps with respect to hybrid polarizations, for example, by choosing
to study fuzzy band gaps with respect to
y-translation and
x-rotation.
2.4.1. MPF—Definition
The MPFs
(
) for translations in the
i direction are collected in the vector
and are defined as
Likewise, the MPFs
(
) for rotations about the
i axis are collected in the vector
and defined as
where
is the system mass matrix,
is the modal mass of mode
,
is a matrix of the rigid-body translation states, and
is a matrix of the rigid-body rotation states. It is convenient to partition the vectors
and
as
, where
are vectors containing the dofs associated with displacements in the
-directions, respectively. Using this partitioning, the rigid-body states can be defined as
where
represent unit displacements in the coordinate directions,
represent unit rotations about axes parallel to the coordinate axes and passing through the point
, and
and
are vectors of zeros and ones, respectively. For the magneto-active metastructures used in
Section 4 to demonstrate the use of the evanescence indicator for real and highly viscoelastic PCs, it is found that high-frequency modes have low MPFs due to their low effective mass, because they primarily involve motion of the exterior lattice structure. Thus, to identify the primary translational and rotational polarization of each mode, it is convenient to define the normalized MPF vectors
and
:
2.4.2. MPF Scaling—Translations and Rotations
From Equation (
13) and Equation (
14) and the definitions of the rigid-body states (Equation (
15) and Equation (16)), it is evident that the translational and rotational MPFs are inversely proportional to the amplitude of modes
and proportional to the amplitude of the rigid-body states
. In regard to the former observation, if a different mode normalization
is chosen, the modal mass becomes
, and the modal participation factors become
Upon substituting
for the modal amplitude
and
for the Bloch mode shape
in Equation (
4) (
being the continuous-system equivalent of
), the same displacement field is obtained as when
and
are used. This shows that the displacement field given by Equation (
4) is independent of the mode normalization when using MPFs as the modal amplitudes
, which is a desired result because the mode normalization is arbitrary.
On the other hand, the dependence of the MPFs on the rigid-body amplitudes initially seems undesirable, because the rigid-body amplitudes depend on the arbitrary parameters
. In addition, the amplitude of the rigid-body rotation states
also implicitly depends on the dimensions of the structure, because the terms
are on the order of the dimensions of the structure. This is especially significant when studying multi-polarized fuzzy band gaps (e.g.,
Section 4 where we study fuzzy band gaps with simultaneous
y-translations and
x-rotations), because the structure dimensions can introduce a scaling mismatch between
and
. For example, when using standard SI units, the unit displacements
could be chosen as 1 m and the unit rotations
as 1 rad. For a structure with dimensions on the order of tens of millimeters, the terms
in Equation (
14) are also of order 10 mm. As
are 100 times larger than
, it follows that
. The rotational contribution would thus be unintentionally obscured when choosing
. To ensure that all polarizations are properly accounted for, it is recommended to choose the rigid-body amplitudes
so that the displacements involved in the rigid-body rotations are on the same order of magnitude as the rigid-body translations. The rigid-body amplitudes can also be set equal to the “input” excitation measured in experiments, as discussed further in
Section 4.
3. Example 1: Diatomic Lattice
To validate the developed evanescence indicator, we apply it to two example systems which are discussed in the literature. First, to illustrate the basic features of the evanescence indicator, we apply it to a simple discrete 1D system: the diatomic lattice [
14,
20,
37], which we modify from its usual undamped configuration to include viscoelastic springs in order to illustrate the application of the evanescence indicator to damped systems. Later, in
Section 4, we apply it to a more complex case of magneto-active metastructures [
11], which provided some of the initial impetus for this work, and demonstrate that the evanescence indicator can correctly predict polarized fuzzy band gaps in damped PCs with 3D geometries.
As shown in
Figure 1, the 1D viscoelastic diatomic lattice is characterized by a unit cell with masses
and
, separated by distance
and connected by springs with stiffness
. Thus, the lattice constant
a is
. To validate the indicator for damped as well as undamped systems, the springs are assumed to be viscoelastic, with the spring stiffness characterized by a complex number
, where
is the magnitude of the spring stiffness and the loss angle
is varied to control the dissipation of the lattice without changing the spring stiffness.
corresponds to an undamped lattice.
We use the transfer matrix method to solve for the dispersion relation of the lattice. The transfer matrix links the left boundary forces and displacements from the
pth unit cell to the
unit cell according to
where
(
) and
(
) are the displacements and forces, respectively, on the left (right) side of the unit cell, and
is the transfer matrix. We prescribe real frequencies
and compute the complex wavenumber
k and mode shapes
from the eigenvalues and eigenvectors of the transfer matrix. We use the PBME method to determine the modal amplitudes and compute the evanescence indicator. For the purposes of illustration, we choose
,
,
,
, and an input displacement amplitude
. All results are presented in terms of the normalized angular frequency
, where
is the lower edge frequency of the first band gap for the undamped PC. We vary the parameters
,
N, and
to study their effect on the computed evanescence indicator and the fuzzy band gaps. All results are presented in terms of normalized transmission amplitudes
and
. To verify that the true displacement at unit cell
N is accurately captured by a summation of Bloch modes with the computed modal amplitudes, we also compute the true response of the finite structure by constructing the dynamic stiffness matrix including all degrees of freedom, applying appropriate boundary conditions, and inverting the dynamic stiffness matrix.
The evanescence indicator is shown in
Figure 2 for a finite structure of reference with
. In each plot (a–c), the effect of the threshold displacement amplitude
on the fuzzy band gaps is shown by varying
, which corresponds to threshold transmission amplitudes
, respectively. The
values are plotted as horizontal dashed lines and per the generalized fuzzy band gap definition, a fuzzy band gap is defined to exist at any frequency where the evanescence indicator falls below
. As in [
22,
24], we use a shading model to visualize the evanescence indicator. In our model, regions with transmission
are shaded white, indicating that no fuzzy band gap exists at that frequency. Regions with transmission
are shaded with a solid color, and intermediate values
are assigned shades between white and the solid color, so that shaded bars indicate the fuzzy band gaps. This luminance-based shading scheme is an intuitive choice for representing relative values of data [
38], and it enables ready visual identification of band gaps while also illustrating how the wave amplitude progresses through smooth transitions near the fuzzy band gap edges. If the threshold value is based on an engineering requirement, such as the maximum allowable vibration amplitude in a structure, the “darkness” of the shading can be interpreted as the factor of safety of the system against the given requirement. White regions indicate a factor of safety of 1 or less, dark regions indicate a factor of safety of 10 or more, and intermediate shades indicate a factor of safety between 1 and 10. The shading near the fuzzy band gap edges is reminiscent of out-of-focus photos due to the lack of sharp contrast. This gives rise to the moniker “fuzzy band gaps”, as out-of-focus photos are often colloquially described as “fuzzy”.
From
Figure 2, it is evident that the existence of fuzzy band gaps is dependent on the selected value of
. For example, in panel (a), a fuzzy band gap exists between 1.05–1.24
for
because the evanescence indicator (black dashed line) falls below
(blue dotted line) in that frequency range. However, when the threshold transmission is selected as
or
, there is no fuzzy band gap, because the trough in the evanescence indicator is not deep enough to reach
or
(orange and green dotted lines). The existence/nonexistence of fuzzy band gaps is also readily apparent from the shaded bars, as the light blue shaded region between 1.05–1.24
indicates a fuzzy band gap, while the lack of any orange or green shading in the same frequency range shows that no fuzzy band gap exists for those threshold amplitudes.
In addition to determining the existence of fuzzy band gaps, the threshold amplitude also affects the frequencies at which fuzzy band gaps occur. For example, in panel (a), the edge of the high-frequency fuzzy band gap increases in frequency from 1.74 to 1.77 to 1.83 as is varied from to to , as determined by locating the intersection of the evanescence indicator curve with the horizontal dotted lines that indicate the selected threshold amplitudes. This gives the shaded bars a stair-step appearance.
Figure 2 also reveals how the loss factor
affects the existence, position, and width of the fuzzy band gaps. Conferring with
Figure 2a for the undamped case, it can be seen that a fuzzy band gap exists between 1.05 and 1.24
. As the dissipation in the lattice is increased to
, the width of the fuzzy band gap increases to 0.97–1.33
. The trough is deeper than for the undamped case, so the structure of reference meets the threshold amplitude with a greater margin, and thus the shading is darker than for the undamped fuzzy band gap. However, the trough is still not sufficiently deep to meet the
threshold amplitudes, and so a fuzzy band gap is not formed for those cases. As the dissipation is further increased by increasing the loss factor to
(
Figure 2c), the two fuzzy band gaps which existed for the
case merge into a single high-frequency fuzzy band gap starting at 0.86
. This is the result of the disappearance of the optical pass band, which is much more strongly affected by damping than the acoustic pass band [
14]. The disappearance of the optical pass band causes a large decrease in the edge frequency of the single fuzzy band gap identified for
and
. Finally, it is apparent that the fuzzy band gaps become “fuzzier” with increasing dissipation, that is, the shaded transition from white to solid color becomes wider as the dissipation is increased. This is a result of the slope of the evanescence indicator curve near the fuzzy band gap edges becoming less steep with increasing dissipation. This, in turn, is caused by increasingly smooth transitions in the
band structure as
is increased.
Figure 3 shows how the number of unit cells
N in the finite structure of reference affects the fuzzy band gaps. For
(
Figure 3a), there are two fuzzy band gaps when
and only one fuzzy band gap when
and
, as discussed previously. However, when
N is increased to 10 (
Figure 3b), the two fuzzy band gaps observed for
merge into a single fuzzy band gap, while for
, a new fuzzy band gap opens between 1.0–1.37
. This behavior is expected, as the additional length of the
structure of reference allows for a greater loss of amplitude by the decaying waves as they propagate [
32,
33].
Figure 3b also illustrates an interesting feature of the shading approach: the ability to identify near-passbands. Though the
structure exhibits a single fuzzy band gap for all frequencies greater than 0.9
when
, the peak in the evanescence indicator at 1.47
nearly reaches the threshold amplitude. This is visualized by a region of lighter blue color centered around 1.47
. The same phenomenon is observed for
and
in
Figure 3c. Finally, it is observed that increasing
N tends to decrease the “fuzziness” of the fuzzy band gaps, because the wave attenuation is magnified by the increased number of unit cells, resulting in steeper rolloffs at the fuzzy band gap edges.
To confirm the accuracy of the propagated Bloch mode expansion method, we compare the evanescence indicator with the true displacement of the
N-cell finite structures, computed using the full set of dofs for the finite structure. For each case of
and
N in
Figure 2 and
Figure 3, the normalized displacement amplitude for mass
of cell
N is plotted with a black dashed line. In general, the evanescence indicator agrees with the true displacement within half a decade or less. The agreement is nearly perfect for frequencies less than 0.75
in all cases, with the largest deviation observed in the frequency range of the optical mode. The deviation decreases as the loss factor is increased, with a maximum deviation of 5.9 dB for
and
, compared to a maximum deviation of 21.2 dB for
and
. Interestingly, the deviation between the evanescence indicator and the true displacement is almost unaffected by changing the number of unit cells
N, with a maximum deviation of 10.1dB for each of the cases shown in
Figure 3. At first glance this seems counterintuitive, as the PBME method applied to this system essentially consists of reducing a
-dof structure to two dofs, and increasing
N therefore results in a greater reduction in the number of dofs, which would be expected to give a worse approximation. However, increasing the number of dofs in the finite structure causes its dynamics to more closely resemble the dynamics of the infinite system, making the Bloch modes more closely resemble the modes of the finite structure. These competing effects appear to cancel out, allowing the PBME method to model arbitrarily large structures with only two dofs.
Figure 4 illustrates how the evanescence indicator and fuzzy band gap shading can be used to simplify the presentation and interpretation of damped dispersion relations. For the undamped lattice (
Figure 4a), the band gaps can be easily identified using either of two methods: by locating frequency ranges where
is 0 or
, or by locating frequency ranges where
is nonzero. However, the introduction of damping changes the eigenmodes and causes the band gap to close (
Figure 4b), as the acoustic and optical modes for
are no longer distinct but instead meet at the edge of the first Brillouin zone, and
is nonzero for all frequencies. This has been noted before in the literature, see, e.g., in [
12]. In the damped case, it is thus necessary to present both the real and imaginary parts of the wavenumber to understand the wave propagation behavior.
In contrast with the complicated picture of wave propagation painted by
Figure 4a,b,
Figure 4c–e provides a much clearer picture of the wave behavior. The evanescence indicator is computed for
,
for three different loss factors, and the fuzzy band gap shading is plotted directly on the dispersion relation using the same shading model as in
Figure 2 and
Figure 3. The fuzzy band gaps are clearly visible from the shaded regions, even for the damped lattices where the fuzzy band gap edges cannot be inferred from the dispersion curves alone (
Figure 4d,e). In addition, it is not necessary to present the imaginary part of the wavenumber to give a clear picture of the wave propagation behavior because the essential information about the decay of the wave amplitude is captured by the fuzzy band gap shading, which reduces the complexity of the figure. Interestingly, the fuzzy band gap identified by the evanescence indicator for the undamped lattice (shaded region in
Figure 4c) is noticeably narrower than the band gap identified by the traditional means (which exists between
and
). This can be explained by referring to the imaginary part of the wavenumber shown in
Figure 4a. Near the edges of the band gap,
is close to 0, which gives a slow spatial decay that is not sufficient to cause the transmission to fall below
within
unit cells. The evanescence indicator thus provides a more practical assessment of the effectiveness of the lattice for suppressing vibrations than the traditional method of identifying band gaps.
5. Discussion
The proposed evanescence indicator provides a flexible and robust way to identify frequency ranges of strong wave attenuation in damped PCs which are termed “fuzzy” band gaps, including polarized fuzzy band gaps and hybrid-polarized fuzzy band gaps. In contrast to prior methods that heuristically identify regions of relatively greater or lesser evanescence in damped band structures, or rely on manual inspection and value judgments of mode shapes to classify polarizations, the proposed indicator quantitatively identifies fuzzy band gaps based on measurable physical quantities.
As a consequence of the generalized definition of fuzzy band gaps in damped PCs, fuzzy band gaps can not be defined unambiguously. In contrast to traditional band gaps which are unambiguously defined as any frequency where all wave modes are evanescent, the proposed method includes two arbitrary parameters (N and ) that affect the fuzzy band gaps. These parameters can be based on physical arguments, so that fuzzy band gaps identified using this method can be related to experiments or simulations of finite structures. Alternatively, a single value of N and can be selected to provide a consistent comparison between different geometries or polarizations, or fuzzy band gaps can be visualized for different combinations of N and to illustrate how the choice of these parameters affects the computed fuzzy band gaps. In all cases, the chosen values of N and should be reported.
To obtain the most accurate results, one must ensure that the computed dispersion relation is sufficiently complete, that is, that all relevant wavenumbers are computed by the eigenvalue search algorithm. In particular, FEM packages may utilize eigenvalue solvers that compute only a small fraction of the eigenvalues of the system. To accurately compute the evanescence indicator, it is important that all modes with small are computed by the eigenvalue solver, as they exhibit the slowest decay and therefore carry the most wave energy. Further analysis is needed to provide guidance on the minimum value of that should be included. One should also ensure that eigenvalues outside the first Brillouin zone, which may be returned by eigenvalue solvers, are not included in the computation of the evanescence indicator, as including them would result in double-counting the modes inside the first Brillouin zone.
Though the evanescence indicator was able to accurately identify the fuzzy band gaps when using the MPF method to compute the modal amplitudes, a puzzling result of this work was that the PBME method did not yield modal amplitudes which accurately captured the behavior of finite MAE metastructures. The reasons for this are not entirely clear. One possible explanation is that the normal modes of the finite structure do not resemble the propagated Bloch modes, because the Bloch modes are not compatible with the applied boundary condition of uniform longitudinal displacement. Further research is needed to understand if and how the PBME method can be applied to multiply-connected periodic media. Even still, the evanescence indicator accurately captures details of vibration transmission through finite MAE metastructures calculated through simulations and measured in vibration experiments.
There are a number of opportunities to extend and improve the current evanescence indicator. At present, the method is limited to structures with 1D periodicity. Generalizing this method to 2D and 3D periodicity would be of great value, but is more complicated as the flow of energy is not necessarily parallel to wave propagation direction. Thus, the wave transmission between the chosen input and output locations, and thus the value of the evanescence indicator, may depend more on the direction of the group velocity than the direction of the wavevector, and appropriate selection of the input and output locations takes on additional significance.