On-chip phononic circuits tailor the transmission of elastic waves and couple to electronics and photonics to enable new signal manipulation capabilities. Phononic circuits rely on waveguides that transmit elastic waves within desired frequency passbands, which are typically designed based on the Bloch modes of the constitutive unit cell of the waveguide, assuming periodicity. Acoustic microelectromechanical system waveguides composed of coupled drumhead resonators offer megahertz operation frequencies for applications in acoustic switching. Here, we construct a reduced-order model (ROM) to demonstrate the mechanism of transmission switching in coupled drumhead-resonator waveguides. The ROM considers the mechanics of buckling under the effect of temperature variation. Each unit cell has two degrees of freedom: translation to capture the symmetric bending modes and angular motion to capture the asymmetric bending modes of the membranes. We show that thermoelastic buckling induces a phase transition triggered by temperature variation, causing the localization of the first-passband modes, similar to Anderson localization caused by disorders. The proposed ROM is essential to understanding these phenomena since Bloch mode analysis fails for weakly disordered (<5%) finite waveguides due to the disorder amplification caused by the thermoelastic buckling. The illustrated transmission control can be extended to two-dimensional circuits in the future.

Phononic circuits are attracting increased interest because they tailor the propagation of elastic and acoustic waves, which is advantageous for signal manipulation. For example, phononic circuits are useful for cellular phone duplexers by serving as acoustic isolators and mirrors.1–3 In medical ultrasound applications and acoustic nondestructive tests, phononic circuits promise to miniaturize the imaging aperture,4–6 decouple the electro-acoustic transduction,7,8 and slow the signal for smaller delay lines.9–11 Moreover, nanostructural phononics operating in the hypersonic (gigahertz to terahertz) frequencies enable thermal management,12–14 photonic-phononic interactions,15,16 and quantum information control.17,18 Phononic structures offer readily achievable nonlinearities allowing for strong optomechanical nonlinearities,19,20 targeted-energy transfer,21,22 and passive structural nonreciprocity.23–25 

Phononic circuits require accurately designed and fabricated waveguides to spatially constrain the acoustic transmission within a specific frequency range referred to as the passband (or the transmission band). In the passband, the temporal frequencies are linked to the spatial frequencies (i.e., the wavenumbers) through the dispersion relationship of the medium, providing additional control over the acoustic transmission.1,26 This temporal and spatial selectiveness stems from the dynamic characteristics of the unit cells whose periodic repetition forms the waveguide. Therefore, the unit cell design is directly linked to the waveguide characteristics via the Bloch modes of the unit cell. The Bloch modes are the vibrational modes that the unit cell exhibits under Floquet boundary conditions with a wavenumber spanning the irreducible Brillouin zone (IBZ).26 This approach calculates the possible wavenumber-frequency relationship known as the band structure of the phononic crystal (i.e., the unit cell). This band structure matches the transmission in an infinite periodic waveguide of the same repeated unit cell.26 

Bloch modes predict the transmission of sufficiently long and weakly disordered waveguides,5,6,12,16,27,28 although fabricated waveguides are neither infinite nor perfectly periodic. In these cases, the finite-structure modal frequencies lie within (or close to) the Bloch-mode passbands.26 For example, such a waveguide of N cells possesses at most N finite-structure modes for every passband; increasing N makes the N modes more densely packed within the passband, leading to the continuous Bloch-mode band structure as N + . The dense packing of modes originates from the structural periodicity whose absence (i.e., aperiodicity) generates frequency-distant modes that cannot approximate the passbands. In addition, the periodicity causes (spatially) extended mode shapes that permit the transmission of a signal between the ends of the structure.26 These features—the approximate passband and the extended mode shapes—are acoustically attractive and enable a finite periodic structure to operate as a waveguide. The Bloch-mode approach is computationally efficient in linear periodic systems because it enables the tailoring of a single unit cell to estimate the behavior of the entire waveguide. On the other hand, it significantly deviates from experimental results when the number of unit cells is limited, when there is aperiodicity (structural asymmetry) in the devices (whether intentional or uncontrolled), and when nonlinearities are profound.

Considering repetitive arrays of drumhead resonators composed of coupled flexible micro-membranes, we have recently shown that thermoelastic buckling of the membranes can switch the acoustic transmission.29 Waveguides made from coupled drumhead resonators were first proposed in 2013 by Hatanaka et al.,30 who showed that these waveguides sustain megahertz-to-gigahertz mechanical vibrations with high quality factors (high Q’s) and optical finesse, features that are valuable in mechanical, electrical, and optical applications.31–33 For instance, optomechanical interactions favor large surface-area structures (like the drumhead resonators) over beams/cantilevers.32–34 Another advantage of the drumhead resonators is their manufacturability via conventional micro/nanofabrication,32,33 while allowing for in situ structural tunability and actuation via piezoelectric,30,34 electrostatic,35 and thermal control.29 Therefore, drumhead resonators were applied in tunable optical cavities36 and low-loss nonlinear optomechanical coupling.37 Moreover, coupling drumhead resonators in the form of arrays, like the devices studied in this article, served in realizing phononic transistors,30 tunable one-dimensional (1D) phononic waveguides,35 cavity-switchable waveguides,38 and on-chip two-dimensional (2D) topological insulators.39 

In this work, we study the mechanism of transmission switching in the drumhead-resonator waveguides reported in Ref. 29, a phenomenon that has previously been attributed to buckling-induced aperiodicity. Specifically, we develop a reduced-order model (ROM) that mimics the experiments observed in Ref. 29 (Sec. II). The ROM accounts for out-of-plane translation, rotation, and coupling to accurately predict the first and second passbands of the waveguides as functions of the buckling state. The ROM uses the concept of the von Mises truss40 to capture the effect of buckling on drumhead-resonator waveguides, as illustrated in the electro-thermoelastic tunability of individual drumhead resonators in Ref. 41. In turn, the von Mises trusses permit modeling and predictive analysis of the drumhead-resonator waveguides via lumped springs and rigid masses, presenting simpler models that are amenable to analytical studies compared to finite element models (FEMs) (e.g., continuous beams on elastic foundations).

With the von Mises ROM, we calculate the Bloch modes (Sec. III) and compare them to the transmission of (60-cell) finite waveguides in cases of perfect periodicity and (<5%) weak disorder (Secs. IV and V). We investigate the acoustics of the finite waveguides by subjecting their first cell to nonzero initial velocities and monitoring the resulting free responses in the time and frequency domains as functions of the spatial propagation of wave packets in the waveguide. We find that when the weakly disordered finite waveguides are close to their critical buckling state, the transmission through the first passband vanishes. Stronger disorder results in a larger range of temperatures where the first passband does not transmit elastic waves. This contrasts with the corresponding perfectly periodic finite waveguide (i.e., with no disorder), where the first passband transmits elastic waves at all considered temperatures, even at the onsite of critical buckling. As for the second passband, the acoustic transmission persists for all considered disorders and temperatures.

To thoroughly explain the effect of buckling on the transmission, we inspect the dependences of the mode shapes of the considered waveguides on temperature (Sec. VI). The results show that the transmission switching is associated with converting the mode shapes from extended over the entire waveguide to localized at some cells. This localization of mode shapes with disorders conforms to Anderson's localization originally discovered in electromagnetic waves42–44 and then applied in elastic settings.27,28,45 Moreover, the ROM allows us to test the validity of certain assumptions (cf. Sec. VII) and reveal the physical mechanisms governing the acoustics of the considered waveguides (e.g., revealing the necessity of a narrowing of a passband in addition to disorder to affect transmission loss with buckling). Finally, we present an evaluation of this buckling-switchable transmission on a FEM of the experimental waveguide studied in Ref. 29 with 5% disorder far from or close to critical buckling (Mm. ). The FEM simulations agree with the predictions of the ROM, thus, conclusively proving that weak disorder leads to loss of transmission in the repetitive array of drumhead resonators due to buckling.

In this work, we study the phononic waveguides shown in Fig. 1(a). This waveguide consists of repetitive cells capable of transmitting flexural acoustic waves.29,38,46,47 This waveguide was studied in Ref. 29, where the cells are drumhead-like membranes composed of silicon nitride (SiNx) suspended by an etched silicon oxide (SiO2) layer on top of a silicon (Si) substrate. The involved materials and fabrication methods induce residual stresses in the waveguide, whose cells buckle as depicted in Fig. 1(b) by atomic force microscopy (AFM) conducted in Ref. 29.

FIG. 1.

(Color online) Thermally buckled elastic waveguide. (a) Schematic drawing of a microelectromechanical system (MEMS) phononic waveguide made of coupled drumhead resonators (Ref. 29); (b) three-dimensional (3D) topography map of three cells of the waveguide (Ref. 29) measured using AFM at room temperature—the colormap shows the out-of-plane deflections resulting from buckling in the structure; (c) measured transmission in the waveguide (Ref. 29) as a function of temperature and frequency of excitation applied to the first cell in the waveguide—the colormap depicts the amplitude of oscillations at the middle of the waveguide, which shows that the temperature change eliminates the transmission in passband I and detunes the frequency of passbands I, II, and III; (d) simulated transmission in Ref. 29 as a function of the imposed temperature change Δ T (to control buckling) via a COMSOL finite element model (FEM)—the scatter points depict the eigenfrequencies of the transmitting (large blue points) and non-transmitting (small black points) eigenmodes in passband I for Δ T = −160, −140, −120, −100, −80, and −20 K, whose transmission is estimated by sinusoidal forcing response (requiring more than 30 h of COMSOL simulations for every Δ T); the background shaded regions correspond to passbands I (blue) and II (green) calculated by Bloch analysis; schematic drawings of the proposed ROM of the waveguide undergoing thermal buckling showing it in (e) an undeformed and (f) a deformed state.

FIG. 1.

(Color online) Thermally buckled elastic waveguide. (a) Schematic drawing of a microelectromechanical system (MEMS) phononic waveguide made of coupled drumhead resonators (Ref. 29); (b) three-dimensional (3D) topography map of three cells of the waveguide (Ref. 29) measured using AFM at room temperature—the colormap shows the out-of-plane deflections resulting from buckling in the structure; (c) measured transmission in the waveguide (Ref. 29) as a function of temperature and frequency of excitation applied to the first cell in the waveguide—the colormap depicts the amplitude of oscillations at the middle of the waveguide, which shows that the temperature change eliminates the transmission in passband I and detunes the frequency of passbands I, II, and III; (d) simulated transmission in Ref. 29 as a function of the imposed temperature change Δ T (to control buckling) via a COMSOL finite element model (FEM)—the scatter points depict the eigenfrequencies of the transmitting (large blue points) and non-transmitting (small black points) eigenmodes in passband I for Δ T = −160, −140, −120, −100, −80, and −20 K, whose transmission is estimated by sinusoidal forcing response (requiring more than 30 h of COMSOL simulations for every Δ T); the background shaded regions correspond to passbands I (blue) and II (green) calculated by Bloch analysis; schematic drawings of the proposed ROM of the waveguide undergoing thermal buckling showing it in (e) an undeformed and (f) a deformed state.

Close modal

To better convey the buckling and acoustic mechanisms, we want to clarify some terminologies used in Refs. 29 and 41 and this paper. We refer to the point of critical buckling as the thermoelastic state leading to the minimum stiffness of the waveguide membranes, corresponding to the lowest natural frequency attained as a function of the stress state [e.g., at 230 K in Fig. 1(c) and in Fig. 4 in Ref. 41]. The pre-buckling regime corresponds to the thermoelastic state where the compressive effects (stresses/forces/strains/…) in the membrane are lower than the ones required for critical buckling; the post-buckling regime corresponds to the thermoelastic state with compressive effects larger than the ones at the point of critical buckling. Hence, the point of critical buckling can be detected by locating the state where the membrane attains minimum stiffness (i.e., minimum natural frequency).

For an ideal membrane with a symmetric cross section, increasing the compressive effects induces no static bending-induced deformation in the pre-buckled regime; once the ideal membrane attains the point of critical buckling, the static deformation drastically increases with further compression in the post-buckling regime.48 However, for a non-ideal membrane with an asymmetric cross section, the membrane exhibits some bending deflection in the pre-buckled regime, as demonstrated by Fig. 1(b) deflections of measured at room temperature (i.e., pre-buckled regime). The rate of static deflection starts to increase as a function of increasing compression, as the membrane reaches critical buckling and then transitions to post-buckling. We observe this behavior in the experiments and simulations of Refs. 29 and 41 [also exhibited later in this work by the ROM static simulations of Fig. 2(a)]. Near the point of critical buckling, the membrane's elasticity is strongly sensitive to the thermo-elastic conditions dictated by the residual stresses, temperature, geometry parameters (e.g., thickness), etc. We experimentally observed this strong sensitivity in the asymmetric single resonators of Ref. 41.

FIG. 2.

(Color online) Thermal buckling of the infinite perfectly periodic waveguide (i.e., Bloch modes). (a) Static equilibrium of a single cell as a function of temperature based on (4); (b) frequency dispersion curves as a function of the nondimensional wavenumber of the Bloch modes at a temperature of 390, 370, and 350 K; (c) Bloch-mode frequency extrema as a function of temperature illustrating the transmission detuning in an infinite perfectly periodic waveguide—we depict the Bloch-mode frequency extrema with k x / L close to 0 rad by the filled circles, whereas the extrema with k x / L close to π rad are shown by open circles; also the blue and green colors in (b) and (c) represent the Bloch-mode passbands I and II, respectively.

FIG. 2.

(Color online) Thermal buckling of the infinite perfectly periodic waveguide (i.e., Bloch modes). (a) Static equilibrium of a single cell as a function of temperature based on (4); (b) frequency dispersion curves as a function of the nondimensional wavenumber of the Bloch modes at a temperature of 390, 370, and 350 K; (c) Bloch-mode frequency extrema as a function of temperature illustrating the transmission detuning in an infinite perfectly periodic waveguide—we depict the Bloch-mode frequency extrema with k x / L close to 0 rad by the filled circles, whereas the extrema with k x / L close to π rad are shown by open circles; also the blue and green colors in (b) and (c) represent the Bloch-mode passbands I and II, respectively.

Close modal

In Fig. 1(c), we show the effect of buckling on the elastic transmission of the waveguide of Fig. 1(a).29 The temperature in Fig. 1(c) controls the state of buckling in the waveguide, where lower temperature increases compressions between cells to provoke stronger buckling. At each temperature, the colormap in Fig. 1(c) corresponds to the frequency response measured at the middle cell of the waveguide due to the electrostatic actuation of the gold (Au) pad covering the first cell [cf. Fig. 1(a)]. At high temperatures in Fig. 1(c) (i.e., above ∼230 K), the waveguide exhibits three frequency regimes of effective transmission corresponding to the first three passbands (labeled as I, II, and III). A decrease in temperature from 280 K down to ∼230 K decreases the mean frequency of all the passbands, indicating a softening behavior. During this softening, passband I diminishes its bandwidth until collapsing at ∼230 K, whereas passbands II and III maintain an almost constant bandwidth. Cooling the waveguide to below ∼230 K eliminates the transmission in passband I and increases the mean frequencies of passbands II and III while possessing almost constant bandwidths. The observed temperature-dependent changes in the frequency passbands and the switch in frequency detuning imply that the waveguide at ∼230 K is in a critical buckling state associated with the softest structural configuration (since buckling indicates a minimum linearized stiffness). Accordingly, the waveguide is pre-buckled for temperatures ≳230 K and post-buckled for temperatures ≲230 K, based on Ref. 29. In both buckling regimes, the frequency detunings of passbands II and III are direct consequences of the buckling state of the waveguide. However, the frequency detuning of passband I and its transmission loss in the post-buckled regime necessitates both buckling and disorder in the waveguide.29 

The requirement of disorder for inducing transmission loss was proven in Ref. 29 via COMSOL simulations, whose results are depicted in Fig. 1(d). The color-shaded regions in Fig. 1(d) illustrate the transmission predicted by Bloch modes as a function of buckling controlled by the imposed temperature change Δ T in COMSOL (cf. Ref. 29 for detailed info). The COMSOL simulated Bloch modes demonstrate the effect of buckling on detuning the frequencies of passbands I and II, yet the Bloch modes cannot predict transmission loss of passband I because Bloch modes assume perfect periodicity with the defined Floquet boundary conditions. Therefore, to capture the transmission loss of passband I depicted by the scatter points of Fig. 1(d), we had to simulate the sinusoidal response of a finite weakly disordered (<5%) waveguide in COMSOL.29 As illustrated by the results of Fig. 1(d), passband I loses transmission around critical buckling, indicating the possibility of transmission loss over a range of temperatures in experiments. We should note that the simulations differ from the experiments in Ref. 29 by showing a certain degree of wave transmission in passband I at the post-buckled regime [i.e., at Δ T = −160 and −140 K in Fig. 1(d)]. This transmission past the critical buckling point was not observed in experiments [cf. Fig. 1(c) in Ref. 29], which might be due to unmodeled effects at low temperatures.

Although the finite waveguide simulations in COMSOL depict the transmission loss and the necessity of disorder, these simulations are computationally expensive and impractical for design or in-depth computational investigations. For instance, determining the transmission of a 60-cell waveguide at each temperature of Fig. 1(d) requires more than 20 h of simulations on COMSOL (cf. Ref. 29 for detailed information). Hence, to perform further parametric investigation of the relationship between disorder, buckling, and elastic transmission in the drumhead waveguides, we propose the ROM depicted in Figs. 1(e) and 1(f). This ROM captures the thermally mediated elastic buckling based on the ROM of a single cell introduced in Ref. 41 exhibiting very good predictive capacity. Here, we extend the ROM of Ref. 41 to account for the coupling between the cells in the waveguide and model the acoustics of the entire phononic waveguide. Accordingly, we allocate to each cell a translational degree-of-freedom (DoF) (as in Ref. 41) and a rotational DoF to capture passbands I and II, respectively.

As shown in Fig. 1(e), each cell of index i in the waveguide consists of a rigid mass m i with a moment of inertia J i. Cell i undergoes the motion illustrated in Fig. 1(f) with translational coordinate u i and rotation angle θ i. The translation deforms the grounding springs of stiffnesses k i B and k i S representing the restoring forces for bending and stretching, respectively. As in Ref. 41, these translational bending and stretching springs are confined at distances d B and d S [see Fig. 1(f)] while possessing free (undeformed) lengths L B and L S, respectively; clearly, a free length larger than the confinement distance (i.e., L B > d B and L S > d S) introduces compressive strains (precompression) in the cell. We assume that the remaining springs in the ROM are undeformed at the initial undeformed configuration of the waveguide. For example, the springs with stiffnesses T i B, k i 1 C, k i C, T i 1 C, and T i C attached to the cell of index i do not apply any forces or torques in Fig. 1(e). The grounding torsional spring of stiffness T i B lumps the bending effects that oppose the rotation θ i due to the grounded boundary of the drumhead. The coupling springs with stiffnesses k i C and T i C account for the force and torque, respectively, applied by cell i to cell ( i + 1) due to the deformations illustrated in Fig. 1(f). Last, we represent the lattice length separating two successive cells by the length L [see Figs. 1(e) and 1(f)].

In a previous article,41 we described the static equilibrium and the equations of motion of a single drumhead resonator and identified its system parameters, which we refer to as the reference cell parameters. These parameters are the translating mass m Ref and springs k Ref B and k Ref S (we use the subscript “ Ref” to label the reference cell), which are reproduced in Table I. We start by studying the Bloch modes of an infinite waveguide based on a repetition of this reference unit cell, as shown in Figs. 1(e) and 1(f). The grounding translating springs exert the force F i Buck at the cell i { 1 , 2 , } and the temperature T expressed for any translational displacement u i as
F i Buck u i ; T = k i B d B u i d B δ B T + k i T u i 1 1 + δ S T 1 + u i d S 2 .
(1)
TABLE I.

Parameters of the reference cell (Ref. 41).

δ B δ S d ¯ S κ Ref T 1 2 π Λ Ref B (MHz) χ
β 0 β 1 (K−1) β 2 (K−2) γ 0 γ 1 (K−1)
7.65  −3.47 × 10−2  3.81 × 10−5  1.9  −4.07 × 10−3  9.40  1/12 
δ B δ S d ¯ S κ Ref T 1 2 π Λ Ref B (MHz) χ
β 0 β 1 (K−1) β 2 (K−2) γ 0 γ 1 (K−1)
7.65  −3.47 × 10−2  3.81 × 10−5  1.9  −4.07 × 10−3  9.40  1/12 
In (1), δ B T and δ S T are the temperature-dependent bending and stretching strains, respectively, stemming from the thermal expansion and the fabrication-residual stresses in the cells. We characterize these strains by the following temperature dependences (as discussed in Ref. 41),
(2)
δ B T = def d B L B L B = β 0 + β 1 T + β 2 T 2 ,
(2a)
δ S T = def d S L S L S = γ 0 + γ 1 T ,
(2b)
with the values of β 0, β 1, β 2, γ 0, and γ 1 listed in Table I. We assume that these temperature dependences govern the strains of all the cells in the waveguide. Moreover, we nondimensionalize the forces in this work by k Ref B d B, leading to the following nondimensional buckling force:
F ¯ i Buck u ¯ i ; T = u ¯ i δ B T + κ i T u ¯ i 1 1 + δ S T 1 + u ¯ i d ¯ S 2 ,
(3)
where κ i T = def k i T / k i B, d ¯ S = def d S / d B, and u ¯ i = def u i / d B with the overbar denoting a nondimensionalized entity.
Focusing on the reference cell that undergoes only translation while connected to k Ref B and k Ref S, we find its equilibrium displacement u ¯ Ref Eqm T by solving the following equation:
F ¯ Ref Buck u ¯ Ref Eqm ; T = 0 with the largest d F ¯ Ref Buck d u ¯ Ref u ¯ Ref ; T | u ¯ Ref = u ¯ Ref Eqm > 0.
(4)
The maximum condition in (4) ensures u ¯ Ref Eqm to be the most stable equilibrium solution, which should be favored experimentally. In Fig. 2(a), we plot the values of u ¯ Ref Eqm as a function of temperature based on (2–4), and the reference parameters in Table I. We observe that the single cell translates upward due to cooling, which increases the internal compressions leading to buckling of the cell.41 

To study the effect of buckling on wave transmission, we evaluate the Bloch modes of the cell at each temperature shown in Fig. 2(a). The Bloch modes correspond to the infinite waveguide of Figs. 1(e) and 1(f) made of cells whose parameters are identical to the considered single cell. In this perfectly periodic infinite waveguide, all the cells attain at T the equilibrium state of u ¯ i Eqm = u ¯ Ref Eqm T and θ i Eqm = 0 rad for all i 1 , 2 , 3 , , + . At every instant t, we track the oscillations of the i th cell about its equilibrium state via the perturbation coordinates,

(5)
v ¯ i t = def u ¯ i t u ¯ i Eqm ,
(5a)
h ¯ i t = def L ¯ θ i t L ¯ θ i Eqm , where L ¯ = def L d B .
(5b)

The above coordinates allow writing Newton's second law on any cell of index i > 1 in Figs. 1(e) and 1(f) as
μ i 1 0 0 χ d 2 v ¯ i d t 2 d 2 h ¯ i d t 2 + μ i 1 Λ i 1 C Λ i 1 C 2 Λ i 1 C 2 Λ i 1 C 4 Γ i 1 C · v ¯ i 1 h ¯ i 1 + ( μ i 1 Λ i 1 C Λ i 1 C 2 Λ i 1 C 2 Λ i 1 C 4 + Γ i 1 C + μ i Λ i C + Λ i Buck Λ i C 2 Λ i C 2 Λ i C 4 + Γ i C + Γ i B ) v ¯ i h ¯ i + μ i Λ i C Λ i C 2 Λ i C 2 Λ i C 4 Γ i C v ¯ i + 1 h ¯ i + 1 = 0 0 ,
(6)
where μ i = def m i / m Ref, χ = def J i / m i L 2, Λ i Buck T = def Λ i B ( d F ¯ i Buck / d u ¯ i ) u ¯ i = u ¯ i Eqm T, Λ i B = def k i B / m i, Λ i C = def k i C / m i, Γ i C = def T i C / m i L 2, and Γ i B = def T i B / m I L 2. Equation (6) only considers the linearized dynamics of the undamped i th cell. Note that the nondimensionalization in (6) results in (squared) frequency-like parameters (i.e., Λ i Buck, Λ i C, Γ i B, and Γ i C). This parameter conversion offers an advantage when comparing the model to experiments because frequencies are easier to identify than stiffnesses and directly affect the performance of the waveguides. For instance, we deduce the value of Λ Ref B listed in Table I from the experiments of a single cell in Ref. 41. For the remaining frequency-like parameters, we assume the following relationships for all i Ref , 1 , 2 , 3 , :
Λ i C T = 0.2 Λ i Buck T min 350 400  K Λ i Buck T ,
(7a)
Γ i B = χ Λ i B ,
(7b)
Γ i C T = χ 3 Λ i C T 3 4 Γ i B .
(7c)
To calculate the Bloch modes of a single cell, we apply the Floquet boundary conditions of v ¯ i / h ¯ i = p i e j ( k x / L ) i + ω t with a normalized wavenumber k x / L, a modal frequency of ω, a modal vector p i, and the imaginary number j 2 = 1. Additionally, we assume that all cells are identical to the single cell with parameters listed in Table I, transforming (6) into the following boundary value problem:
ω 2 1 0 0 χ + [ 2 1 cos k x L Λ Ref C + Λ Ref Buck Λ Ref C sin k x L Λ Ref C sin k x L 1 2 1 + cos k x L Λ Ref C + 2 1 cos k x L Γ Ref C + Γ Ref B ] p i = 0 0 .
(8)

For all k x / L 0 , π rad, the IBZ is defined by the respective pair of eigenfrequencies ω that zero the determinant of the matrix operating on p i in (8). These eigenfrequencies are the Bloch modes' frequencies forming the dispersion curves in Fig. 2(b) at 390, 370, and 350 K for the single cell. The lower (blue) and upper (green) curves in Fig. 2(b) correspond to passbands I and II of the transmission in the perfectly periodic infinite waveguide, respectively. We depict the transmission of this waveguide in Fig. 2(c), where we collect the extrema (maxima and minima) of passbands I and II [as in Fig. 2(b)] for the temperature T 350 , 400 K.

Figure 2(c) shows that cooling from 400 to ∼370 K reduces the mean frequencies of both passbands while narrowing the bandwidth of passband I. Cooling below ∼370 to 350 K increases again the mean frequencies of both passbands while widening the bandwidth of passband I. The first cooling phase from 400 to ∼370 K in Fig. 2(c) resembles the cooling phase in Fig. 1(c) between 280 and ∼230 K. However, the second cooling phase between ∼370 and 350 K in Fig. 2(c) diverges fundamentally from the experimental transmission in Fig. 1(c) between ∼230 and ∼80 K, where the transmission in passband I does not reemerge. Therefore, the ROM buckling cannot eliminate the transmission of passband I in a perfectly periodic infinite waveguide. This loss of transmission with buckling necessitates the consideration of disorder (i.e., the break of perfect periodicity) in the waveguide as previously established in Ref. 29.

Note that we adopt the relationships in (7) to emulate the experimental transmission in Fig. 1(c) between 280 and ∼230 K. For this reason, we select Λ i C T in (7a) to decrease until the transmission vanishes at the point of minimum frequency, min 350 400 K Λ i G T , leading to the shrinkage of passband I between 400 and ∼370 K in Fig. 2(c). In (7b), we assume that the rotation of the cell centerline (of length L) deflects an elastic foundation of stiffness density k i B / L. In (7c), we impose a temperature-constant bandwidth for passband II like the measurements in Fig. 1(c). The temperature-detuning of the mean frequencies of the passbands in Fig. 2(c) is considered for the identified parameters (i.e., Λ Ref B, β 0, β 1, β 2, γ 0, and γ 1) of the ROM in Ref. 41, which deviate from those in the devices used in Fig. 1(c) (extracted in Ref. 29).

We focus now on the waveguide disorder resulting from the thickness variation between the cells. We assume a thickness variation of the form
h i h Ref h Ref = σ h 4 2 N + 1 2 i N + 1 2 1 1 s i + r i 1 , 1
(9)
for i 1 , 2 , , N, where we denote by h i the thickness of the i th cell, by h Ref the thickness of the reference cell discussed in Sec. III, by σ h the level of thickness disorder, and by r i 1 , 1 a random rational number 1 , 1 generated at each i. We introduce the random number r i to account for the random errors of the fabrication process. The s i term in (9) represents the systematic errors resulting from wet etching that forms the waveguide cells as explained in Refs. 29, 38, 46, and 47.

The holes at the center of the cells in Figs. 1(a) and 1(b) are etching holes through which the etchant attacks the underlying layer and suspends the cells. Thus, there is a higher (linear) density of etching holes at the middle of the waveguide [of index ( N + 1 ) / 2] compared to the ends (of indices 1 and N). This higher etching-hole density increases the etching rate, leading to over-etching at the middle of the waveguide compared to its ends.29 We model this over-etching by s i in (9) as a linear distribution of the cell position from the middle of the waveguide. Figures 3(a) and 3(b) show two examples of thickness variation with σ h = 0 % (i.e., perfectly periodic waveguide) and σ h = 5%, respectively.

FIG. 3.

(Color online) Effect of weak thickness disorder on the static equilibrium of finite waveguides. [(a) and (b)] Thickness profile relative to the reference thickness h Ref; [(c) and (d)] static deflections at 400, 390, 380, 370, 360, and 350 K of the perfectly perdiodic (σh = 0 %) and weakly disordered (σh = 5 %) 60-cell waveguides of (a) and (b), respectively [refer to (9) for the definition of the disorder parameter σ h]. In (c) and (d), the thick segments represent the rigid masses of the cells in the ROM of Figs. 1(e) and 1(f) translating and rotating according to the computed equilibria from (11).

FIG. 3.

(Color online) Effect of weak thickness disorder on the static equilibrium of finite waveguides. [(a) and (b)] Thickness profile relative to the reference thickness h Ref; [(c) and (d)] static deflections at 400, 390, 380, 370, 360, and 350 K of the perfectly perdiodic (σh = 0 %) and weakly disordered (σh = 5 %) 60-cell waveguides of (a) and (b), respectively [refer to (9) for the definition of the disorder parameter σ h]. In (c) and (d), the thick segments represent the rigid masses of the cells in the ROM of Figs. 1(e) and 1(f) translating and rotating according to the computed equilibria from (11).

Close modal
The thickness variation implies a corresponding variation in the dynamical properties of the cells. Based on the theory of the mechanics circular plates49,50 and the assumption in Ref. 41, the thickness affects the parameters of the i th cell in Figs. 1(e) and 1(f) as follows:
κ i S κ Ref S = h i h Ref 2 ,
(10a)
Λ i B Λ Ref B = h i h Ref 2 .
(10b)
The scaling relationships (10) with the expressions in (7) quantify the effect of the cell's thickness on the ROM parameters.
With the ROM of Figs. 1(e) and 1(f), we apply Newton's first law to calculate the static equilibrium ( u ¯ i Eqm, L ¯ θ i Eqm) of each cell i 1 , 2 , , N in the N cell waveguide using
F ¯ 1 Buck u ¯ 1 Eqm ; T 0 F ¯ i Buck u ¯ i Eqm ; T 0 F ¯ N Buck u ¯ N Eqm ; T 0 Q ¯ Buck + K ¯ Stat u ¯ 1 Eqm L ¯ θ 1 Eqm u ¯ i Eqm L ¯ θ i Eqm u ¯ N Eqm L ¯ θ N Eqm q ¯ Eqm = 0 ,
(11)
where F ¯ i Buck u ¯ i Eqm ; T is the thermoelastic buckling force expressed in (3). In (11), we assume small angles of deformation allowing the approximation s in θ i Eqm θ i Eqm while neglecting the longitudinal displacement of the cells' ends. Under this approximation, we express the nondimensional static stiffness K ¯ Stat as
K ¯ Stat = K ¯ 1 0 2 × 2 N 2 0 2 × 2 i 2 K ¯ i 0 2 × 2 N i 1 0 2 × 2 N 2 K ¯ N ,
(12a)
with 0 M × P denoting a zero-filled matrix of M rows by P columns,
K ¯ 1 = μ 1 Λ 1 C μ 1 Λ 1 C 2 μ 1 Λ 1 C μ 1 Λ 1 C 2 μ 1 Λ 1 C 2 μ 1 Λ 1 C 4 + Γ 1 C + Γ 1 B μ 1 Λ 1 C 2 μ 1 Λ 1 C 4 Γ 1 C ,
(12b)
K ¯ = [ μ i 1 Λ i 1 C μ i 1 Λ i 1 C 2 μ i 1 Λ i 1 C + μ i Λ i C μ i 1 Λ i 1 C 2 μ i 1 Λ i 1 C 4 Γ i 1 C μ i 1 Λ i 1 C + μ i Λ i C 2 μ i 1 Λ i 1 C + μ i Λ i C 2 μ i Λ i C μ i Λ i C 2 μ i 1 Λ i 1 C 4 + Γ i 1 C + μ i Λ i C 4 + Γ i C + Γ i B μ i Λ i C 2 μ i Λ i C 4 Γ i C ] ,
(12c)
for 2 i N 1, and
K ¯ N = μ N 1 Λ N 1 C μ N 1 Λ N 1 C 2 μ N 1 Λ N 1 C 2 μ N 1 Λ N 1 C 4 Γ N 1 C μ N 1 Λ N 1 C μ N 1 Λ N 1 C 2 μ N 1 Λ N 1 C 2 μ N 1 Λ N 1 C 4 + Γ N 1 C + μ N Γ N B ,
(12d)
where Λ i C, Γ i B, and Γ i C are defined in (7).

We solve (11) via “fsolve” (gradient descent method) in matlab®. In the numerical solver, the starting guesses for u ¯ i Eqm in (11) correspond to the equilibria of individual cells in (1) [see Fig. 2(a)]. We assign the differences u ¯ i + 1 Eqm u ¯ i Eqm as starting guesses for L ¯ θ i Eqm for i 1 , 2 , , N 1 and L ¯ θ N 1 Eqm as the guess for L ¯ θ N Eqm. Figures 3(c) and 3(d) display the computed equilibria of (11) at different temperatures in the perfectly periodic and weakly disordered waveguides of Figs. 3(a) and 3(b), respectively. In Figs. 3(c) and 3(d), each segment (thick dashed line) represents a cell in the ROM of Figs. 1(e) and 1(f) translated and rotated according to the equilibrium of (11).

In Fig. 3(c), the cells at equilibrium undergo the same translational deflections without rotation, which results from the perfect periodicity of the waveguide of Fig. 3(a). For instance, the weakly disordered waveguide of Fig. 3(b) attains equilibrium with different cell translations and rotations, as shown in Fig. 3(d). For both waveguides of Figs. 3(c) and 3(d), the cooling increases the cells' baseline translational deflections, going from negative to positive values between 400 and 350 K [like the individual cell in Fig. 2(a)]. Moreover, each 10 K of cooling induces larger deflections at lower temperatures in Figs. 3(c) and 3(d), which mirrors the buckling susceptibility observed in Fig. 2(a).

At this point, we linearize the dynamics around the calculated equilibria to study the transmission in the finite waveguides for varying temperatures. Using the perturbation coordinates in (5), the linearized equations yield
M ¯ d 2 q ¯ d t 2 + K ¯ Buck + K ¯ Stat K ¯ q ¯ = 0 ,
(13)
where q ¯ = v ¯ 1 , h ¯ 1 , , v ¯ i , h ¯ i , , v ¯ N , h ¯ N T, K ¯ Stat is defined in (12), K ¯ Buck corresponds to the 2 N × 2 N matrix whose diagonal contains the linear stiffnesses of the buckling forces,
K ¯ Buck = diag μ 1 Λ 1 Buck , 0 , , μ i Λ i Buck , 0 , , μ N Λ N Buck , 0 ,
(14)
and M ¯ denotes the nondimensional mass matrix expressed as
M ¯ = M ¯ 1 0 2 × 2 N 1 0 2 × 2 i 1 M ¯ i 0 2 × 2 N i 0 2 × 2 N 1 M ¯ N ,
(15)
where M ¯ i = μ i 1 0 0 χ for i 1 , 2 , 3 , , N. In this work, we solve for the acoustics of the waveguides by direct integration of (13) using the matlab®ode45” function.
To this end, we consider the solution of (13) subject to a nonzero initial translational velocity at cell 1 and all other initial conditions set to zero,
q 0 = def q ¯ t = 0 = 0 and q ̇ 0 = def d q ¯ d t t = 0 = 1 0 3 μ 1 0 0 2 N 1 × 1 .
(16)
The initial conditions (16) induce a motion that propagates in the waveguides as depicted in Figs. 4 and 5 and Mm. . This motion enables the study of wave transmission in the considered finite waveguides. In Figs. 4 and 5 and Mm. , we consider the responses of the waveguides of Figs. 3(a) and 3(b) at 390 K, to address the effect of weak disorder at a fixed temperature corresponding to (relatively) broad passbands I and II based on the Bloch modes of Figs. 2(b) and 2(c).
FIG. 4.

(Color online) Elastic wave transmission through the perfectly periodic finite waveguide of Fig. 3(a) at 390 K: temporal responses in terms of (a) the translations v ¯ n and (b) the rotational h ¯ n perturbation coordinates of (from left to right) cell n { 1, 20, 40, 60  } due to the initial conditions in (16) {the rightmost plots in (a) and (b) show zoomed-in views of the responses of cell 60 in the duration [16, 20] μs}; (c) spatiotemporal evolution of the normalized mechanical energy E i Mech / E I n of (17) for i {1, 2,…, 60} in the ROM waveguide—the wavepacket labels highlight the instants at which the fast and slow wavepackets reach cell 60 and reflect.

FIG. 4.

(Color online) Elastic wave transmission through the perfectly periodic finite waveguide of Fig. 3(a) at 390 K: temporal responses in terms of (a) the translations v ¯ n and (b) the rotational h ¯ n perturbation coordinates of (from left to right) cell n { 1, 20, 40, 60  } due to the initial conditions in (16) {the rightmost plots in (a) and (b) show zoomed-in views of the responses of cell 60 in the duration [16, 20] μs}; (c) spatiotemporal evolution of the normalized mechanical energy E i Mech / E I n of (17) for i {1, 2,…, 60} in the ROM waveguide—the wavepacket labels highlight the instants at which the fast and slow wavepackets reach cell 60 and reflect.

Close modal
FIG. 5.

(Color online) Elastic wave transmission through the weakly disordered finite waveguide of Fig. 3(c) at 390 K: temporal responses in terms of (a) the translations v ¯ n and (b) the rotational h ¯ n perturbation coordinates of (from left to right) cell n { 1, 20, 40, 60  } due to the initial conditions in (16) {the rightmost plots in (a) and (b) show zoomed-in views of the responses of cell 60 in the period [16, 20] μs} and (c) spatiotemporal evolution of the normalized mechanical energy E i Mech / E I n of (17) for i {1, 2,…, 60} in the ROM waveguide—the “Fast wavepacket” label highlights the instant at which the fast wavepacket reaches cell 60 and reflects, and the “Slow wavepacket” label highlights the cells where the slow wavepacket is confined.

FIG. 5.

(Color online) Elastic wave transmission through the weakly disordered finite waveguide of Fig. 3(c) at 390 K: temporal responses in terms of (a) the translations v ¯ n and (b) the rotational h ¯ n perturbation coordinates of (from left to right) cell n { 1, 20, 40, 60  } due to the initial conditions in (16) {the rightmost plots in (a) and (b) show zoomed-in views of the responses of cell 60 in the period [16, 20] μs} and (c) spatiotemporal evolution of the normalized mechanical energy E i Mech / E I n of (17) for i {1, 2,…, 60} in the ROM waveguide—the “Fast wavepacket” label highlights the instant at which the fast wavepacket reaches cell 60 and reflects, and the “Slow wavepacket” label highlights the cells where the slow wavepacket is confined.

Close modal
Mm. 1.

Video of simulations of wave propagation in periodic and disorder waveguides at 390 K.

Mm. 1.

Video of simulations of wave propagation in periodic and disorder waveguides at 390 K.

Close modal

Figures 4(a) and 4(b) show the translational and rotational time-responses, respectively, for wavepackets propagating through the waveguide of Fig. 3(a) at a temperature of 390 K. We observe that the initial velocity imposed at cell 1 by (16) provokes a wavepacket that propagates to the last cell (with index 60) (cf. also Mm. ). The propagating front of this wave consists mainly of two distinct wavepackets with different wave speeds, namely, a “fast” wavepacket reaching cell 60 within ∼17  μs [cf. h ¯ 60 in Fig. 4(b)] and a “slow” wavepacket reaching that boundary cell within ∼38  μs [cf. v ¯ 60 in Fig. 4(a)]. Figure 4(a) shows the fast wavepacket characterized by low translational amplitudes compared to the slow wavepacket. Figure 4(b) shows both wavepackets possessing similar rotational amplitudes.

The slow and fast wavepackets in the finite waveguide are analogous to waves transmitted in passbands I and II, respectively, of the infinite perfectly periodic waveguide [cf. Figs. 2(b) and 2(c)]. For instance, the estimations of the group velocity of the passband structure in Fig. 2(b) at 350 K indicate that the fastest waves in passbands I and II reach the end of the waveguide within ∼36 and ∼16  μs, respectively, which validates the observed wavepacket propagations in Fig. 4. These time durations are estimated in the 60-cell waveguide as Δ t = min k x 0 , πL 59 L / c g k x , where c g = ( 1 / 2 π ) ( d k x / d f ) in each passband. We conclude that at the temperature considered, the finite waveguide supports the propagation of spatially extended wavepackets with frequency-wavenumber contents lying inside passbands predicted in the waveguide of infinite extent. To visualize the wave propagation over every cell, in Fig. 4(c), we plot the spatiotemporal normalized energy evolution in the corresponding perfectly periodic 60-cell waveguide. In particular, we depict the contour plot of the normalized mechanical energy E i Mech / E I n of each cell i 1 , 2 , , N defined by
E i Mech E I n = { μ i d v ¯ i d t 2 + μ i χ d h ¯ i d t 2 + Λ i Buck v ¯ i 2 + Γ i B v ¯ i 2 + 1 2 [ Λ i 1 C v ¯ i v ¯ i 1 2 + Λ i C v ¯ i + 1 v ¯ i 2 + Γ i 1 C h ¯ i h ¯ i 1 2 + Γ i C h ¯ i + 1 h ¯ i 2 ] } / q ̇ 0 T M ¯ q ̇ 0 + q 0 T K ¯ q 0 ,
(17)
with Λ 0 C = Λ N + 1 C = Γ 0 C = Γ N + 1 C = 0. Figure 4(c) demonstrates the two waves propagating in the primary front, reaching the waveguide boundary at cell 60. Moreover, Fig. 4(c) shows the slow wavepacket transmitting a significantly larger portion of the input energy since it mainly corresponds to the translational motion [cf. Fig. 4(a)] that is more effectively excited via the initial conditions in (16).

For comparison, in Fig. 5, we depict the corresponding wave transmission in the ( σ h = 5%) weakly disordered waveguide at 390 K, forced by the same excitation in (16). A drastically different acoustics are observed for the disordered system. We observe that only the early (fast) wavepacket propagates to cell 60 in the presence of disorder. Moreover, as Fig. 5(c) shows, the slow wavepacket (which carries the major portion of the available energy) becomes spatially localized in the first 30 cells, a result that indicates that the weakly disordered waveguide at 390 K cannot transmit the slow wavepacket corresponding to passband I of Figs. 2(b) and 2(c). This transmission loss for wavepackets in passband I is in full agreement with the experimental findings reported in Ref. 29 for a similar waveguide [and summarized in Figs. 1(a)–1(c)]. Therefore, the ROM developed in this work captures the experimental transmission loss mediated by buckling and provides conclusive proof regarding the important role that structural disorder plays for the transmission loss in buckled waveguides at certain temperature ranges.

This transmission loss mechanism is also observed by the FEM of the waveguide studied in Ref. 29, further illustrated in Mm. , presenting the time-series deformations of the centerlines of two waveguides based on the COMSOL simulations of Ref. 29. In particular, we consider two identical 20-cell weakly disordered waveguides at −20 K (i.e., far from critical buckling) and −120 K (close to critical buckling), respectively. In Mm. , we assess the capacity of waveguides to transmit propagating wavepackets with frequency contents inside passband I of their corresponding infinite waveguide (i.e., the passbands based on the Bloch modes of the constitutive unit cell). These FEM simulations show that an elastic wavepacket can propagate only in the weakly disordered waveguide far from critical buckling (cf. Ref. 29 for the FEM geometry and methods). Hence, with the ROM developed in this work, we confirm that buckling-induced transmission loss for waves in passband I is associated with weak disorder and thermoelastic effects, confirming the experimental and FEM results reported in Ref. 29.

Mm. 2.

Video of simulations of wave propagation in close to and far from the buckling temperature point.

Mm. 2.

Video of simulations of wave propagation in close to and far from the buckling temperature point.

Close modal
To further study the frequency transmission in the considered waveguides at 390 K, Fig. 6 presents the amplitude of the fast Fourier transforms (FFTs) of the displacements at selected cells subject to the initial conditions (16). For comparison, we overlay the FFT plots on top of the Bloch modes' passbands of the infinite waveguide [cf. Figs. 2(a)–2(c)] and the modal frequencies ω Mode = Λ Mode of the finite waveguide calculated by the eigenvalue problem,
Λ Mode M ¯ + K ¯ Φ = 0 ,
(18)
where Φ denotes the mode shape vector. For clarity, we do not show in Fig. 6 the modal frequencies that lie within the passbands.
FIG. 6.

(Color online) Effect of weak thickness disorder on the frequency responses of finite waveguides: FFTs of the [(a) and (c)] translational and [(b) and (d)] rotational temporal responses of cells 1 (red line) and 60 (purple line) at 390 K of the perfectly periodic 60-cell waveguide in (a) and (b) and the weakly disordered 60-cell waveguide in (c) and (d) [refer to the ROM of Figs. 3(a) and 3(c)]; the blue and green shaded background regions correspond to the frequencies of passbands I and II of the corresponding perfectly periodic waveguide at 390 K, respectively [cf. Figs. 2(b) and 2(c)], whereas the black vertical lines denote the modal frequencies of the finite waveguides calculated using (18) whose values are not inside the Bloch modes' passbands.

FIG. 6.

(Color online) Effect of weak thickness disorder on the frequency responses of finite waveguides: FFTs of the [(a) and (c)] translational and [(b) and (d)] rotational temporal responses of cells 1 (red line) and 60 (purple line) at 390 K of the perfectly periodic 60-cell waveguide in (a) and (b) and the weakly disordered 60-cell waveguide in (c) and (d) [refer to the ROM of Figs. 3(a) and 3(c)]; the blue and green shaded background regions correspond to the frequencies of passbands I and II of the corresponding perfectly periodic waveguide at 390 K, respectively [cf. Figs. 2(b) and 2(c)], whereas the black vertical lines denote the modal frequencies of the finite waveguides calculated using (18) whose values are not inside the Bloch modes' passbands.

Close modal

Figures 6(a) and 6(b) illustrate that all modal frequencies of the perfectly periodic finite waveguide are inside the passbands (there exist 60 modes in each passband), whereas Figs. 6(c) and 6(d) show certain modes of the ( σ h = 5%) weakly disordered finite waveguide are lying outside these passbands, i.e., in stopbands. Hence, the Bloch modes of the infinite waveguide constitute a perfect estimator of wave transmission only in the perfectly periodic finite waveguide.

In Figs. 6(a) and 6(b), the perfectly periodic finite waveguide corresponds to strong translational and rotational responses at cell 60, respectively, only when their frequency contents are inside the passbands. Outside of the passbands, however, the frequency responses at cell 60 are minimal compared to the response of the excited cell 1, indicating a lack of transmission throughout the waveguide (i.e., stop band). In Figs. 6(c) and 6(d), cell 60 of the ( σ h = 5%) weakly disordered finite waveguide admits weak responses compared to the response of cell 1 in the Bloch modes defining passband I. These frequency responses verify that the transmission loss shown in Fig. 5 corresponds to the transmission loss of wavepackets inside passband I. Notably, concerning wavepackets initiated in passband II [cf. Fig. 6(d)], the rotational response of cell 60 compares in magnitude to cell 1 due to the persistence of the transmission in this passband, as previously illustrated in Fig. 5.

Similar performance to the results of Figs. 4–6 is observed for different temperatures between 400 and 350 K, as shown in Fig. 7. In particular, the perfectly periodic finite waveguide of Fig. 3(a) does not lose transmission for the considered temperatures; the weakly disordered finite waveguide of Fig. 3(c) cannot transmit waves in passband I between 390 and 352 K.

FIG. 7.

(Color online) Effects of thickness disorder and thermal buckling on the frequency content of transmitted waves through finite waveguides: contour plots of normalized transmitted energy vs cell number and wave frequency at (a) 400 K, (b) 390 K, (c) 370 K, (d) 353 K, and (e) 350 K through the 60-cell waveguides with thickness disorders σ h of (i) 0%, (ii) 2.5%, and (iii) 5%.; black and yellow colors correspond to the limiting values of 0 and 1 for the normalized energy respectively.

FIG. 7.

(Color online) Effects of thickness disorder and thermal buckling on the frequency content of transmitted waves through finite waveguides: contour plots of normalized transmitted energy vs cell number and wave frequency at (a) 400 K, (b) 390 K, (c) 370 K, (d) 353 K, and (e) 350 K through the 60-cell waveguides with thickness disorders σ h of (i) 0%, (ii) 2.5%, and (iii) 5%.; black and yellow colors correspond to the limiting values of 0 and 1 for the normalized energy respectively.

Close modal
In Figs. 7 and 8, we summarize the results for all measured temperatures by considering the frequency contents of the wavepackets transmitting throughout the spatial extent of the perfectly periodic or weakly disordered waveguides. We relate the frequency transmission to the nondimensional kinetic energy E ¯ i Kin ( t ) attained by the cell of index i {1, 2, …, N} and numerically approximated by
E ¯ i Kin t m = 1 n FFT E ̃ i , m cos ω ̃ m t + θ ̃ m ,
(19a)
with
E ̃ i , m = 1 2 μ i ω ̃ m 2 v ̃ i , m 2 + χ h ̃ i , m 2 ,
(19b)
where we denote by n FFT the output number of FFT sampled frequencies ω ̃ m  0 rad/s with phase shifts θ ̃ m and by v ̃ i , m and h ̃ i , m the FFT amplitudes of v ¯ i t and h ¯ i t, respectively, at frequencies ω ̃ m for m {1, 2,…, n FFT}. Equation (19) assumes that the FFTs of both v ¯ i t and h ¯ i t have identical phase shifts θ ̃ m at ω ̃ m, leading to q ¯ i t m = 1 n FFT q ̃ i , m cos ω ̃ m t + θ ̃ m for q { v, h}.
FIG. 8.

(Color online) Effects of thickness disorder and thermal buckling on the frequency content of transmitted waves reaching 75% into the finite waveguides: contour plots of the transmitted energy to cell 45 vs the temperature and the wave frequency in the 60-cell waveguides with thickness disorders σ h of (a) 0%, (b) 2.5%, and (c) 5%; the displayed values are extracted from cell 45 of contour plots similar to the plots of Fig. 7 but over the temperature range of [350, 400] K with 1 K resolution; the black and yellow colors correspond to the limiting values of 0 and 1 for the normalized energy, respectively, the blue/green lines correspond to the frequency extrema of passbands I and II, respectively, cf. Fig. 2(c), and the solid/dashed lines to the filled and open circles in Fig. 2(c), respectively. Each of these plots requires less than 3.5 h of ROM simulations for all 51 temperatures of the displayed temperature range of [400, 350] K. These computations are significantly faster than COMSOL finite element modeling, which requires more than 30 h of simulations for each temperature [cf. Fig. 1(d)], indicating more than 400-fold savings in simulation time per temperature.

FIG. 8.

(Color online) Effects of thickness disorder and thermal buckling on the frequency content of transmitted waves reaching 75% into the finite waveguides: contour plots of the transmitted energy to cell 45 vs the temperature and the wave frequency in the 60-cell waveguides with thickness disorders σ h of (a) 0%, (b) 2.5%, and (c) 5%; the displayed values are extracted from cell 45 of contour plots similar to the plots of Fig. 7 but over the temperature range of [350, 400] K with 1 K resolution; the black and yellow colors correspond to the limiting values of 0 and 1 for the normalized energy, respectively, the blue/green lines correspond to the frequency extrema of passbands I and II, respectively, cf. Fig. 2(c), and the solid/dashed lines to the filled and open circles in Fig. 2(c), respectively. Each of these plots requires less than 3.5 h of ROM simulations for all 51 temperatures of the displayed temperature range of [400, 350] K. These computations are significantly faster than COMSOL finite element modeling, which requires more than 30 h of simulations for each temperature [cf. Fig. 1(d)], indicating more than 400-fold savings in simulation time per temperature.

Close modal

In Figs. 7 and 8, the contour plots depict the E ̃ i , m normalized by max E ̃ j , p over all j {2, 3,…, N} and all p {1, 2,…, n FFT} of the respective response. For better visualization, we coerce the contour values to 0 and 1 if they fall below the minimum threshold E Ths  2 × 10−3 and above the saturation limit E Sat  0.5, respectively. In essence, in Figs. 7 and 8, we assess the binary behavior of the considered waveguide by checking whether the energy can transmit to cell i at frequency ω ̃ m under the studied temperature and disorder conditions.

In particular, we investigate the acoustics of three waveguides with disorders σ h {0%, 2.5%, 5%} as defined in (9); the corresponding thickness profiles of the waveguides are provided in Fig. 3(a), Mm. , and Fig. 3(c), respectively. Moreover, to efficiently excite passband II in addition to passband I, we modify in this section the initial conditions in (16) into
q 0 = def q ¯ t = 0 s = 0 and q ̇ 0 = def d q ¯ d t t = 0 s = 1 0 3 μ 1 1 0 3 μ 1 χ 0 2 N 1 × 1 .
(20)
Note that by the new initial conditions (20), we deliver the same initial kinetic energy for both translational and rotational coordinates of cell 1. We use in (20) the same value of initial kinetic energy delivered in (16) to provide a fair analogy to the previous results.
Mm. 3.

Video of simulations of the propagation modes in periodic and disorder waveguides at 390 K within the first passband.

Mm. 3.

Video of simulations of the propagation modes in periodic and disorder waveguides at 390 K within the first passband.

Close modal

The plots of Figs. 7(a)–7(e) display the wave transmission in the three finite waveguides at 400, 390, 370, 353, and 350 K, respectively. For all these temperatures, the perfectly periodic finite waveguide ( σ h = 0%) transmits energy to the last cell 60, with frequency contents in both passbands I and II—cf. Figs. 7(a)(i)–7(e)(i). This transmission for all temperatures is unique for the perfectly periodic waveguide and does not occur in the weakly disordered finite waveguides considered in Figs. 7(a)(ii)–7(e)(ii) and Figs. 7(a)(iii)–7(e)(iii). For example, we observe transmission loss of waves with frequency content in passband I for the waveguide with σ h = 2.5% at 370 K in Fig. 7(c)(ii) and for the waveguide with σ h = 5% at 390, 370, and 353 K in Figs. 7(b)(iii)–7(d)(iii), respectively. Hence, larger disorders correspond to a more extended range of temperatures with transmission loss in passband I.

Moreover, we notice that the frequency width of passband I is smaller in the transmitting scenarios of the weakly disordered waveguides compared to the perfectly periodic waveguide. This width-narrowing of passband I accompanies a similar narrowing of passband II in the weakly disordered waveguides of Figs. 7(a)(ii)–7(e)(ii) and Figs. 7(a)(iii)–7(e)(iii). However, contrary to passband I, the weakly disordered waveguides keep transmitting energy to the last cell 60 in passband II at all temperatures considered, as depicted in Fig. 7. Therefore, we conclude that passband II is less susceptible to structural disorder than passband I, which fully agrees with what was experimentally witnessed in Ref. 29. Hence, the ROM developed herein accurately captures these acoustic aspects of the phononic lattice under investigation.

To further clarify the dependence of wave transmission on temperature, in Figs. 8(a)–8(c), we study the frequency content of transmitted waves reaching cell 45 in the three finite waveguides with disorder σ h {0%, 2.5%, 5%}, respectively. On top of the finite waveguides results, we overlay the corresponding passbands of the infinite waveguides [cf. Fig. 2(c)]. The results in Fig. 8(a) prove that, at all the considered temperatures, the wave transmission in the perfectly periodic finite waveguide has frequency contents solely inside the passbands. Thus, the passbands of the infinite waveguide perfectly estimate the wave transmission in the perfectly periodic finite waveguide at all temperatures, as concluded from Figs. 6(a) and 6(b) at 390 K.

However, this perfect estimation does not hold for the wave transmission in the weakly disordered finite waveguides considered in Figs. 8(b) and 8(c), especially around the critical-buckling temperature (i.e., 370 K). Although, as discussed previously, there is a band-narrowing of both passbands I and II, wave transmission loss is only realized for passband I [cf. Figs. 8(b) and 8(c)]. This behavior confirms that passband I is more susceptible to disorders than passband II, confirming the analogous conclusion illustrated in Fig. 1(c) of the experimental work in Ref. 29. Last, by comparing Fig. 8(b) to Fig. 8(c), we deduce that stronger disorders result in more severe wave transmission losses with thermal-mediated buckling.

To explain the causes of transmission loss due to disorder, we plot in Figs. 9 and 10 the spatial distributions (modeshapes) of two modes of the finite waveguides, specifically modes 30 and 90, at 400, 390, 370, 353, and 350 K. Thick dashed lines represent the normalized translational and rotational deformations of individual cells calculated by the eigenvector Φ of (18) at the considered temperature and waveguide. For better visualization, we join the cells with cubic splines depicted as thin lines to imitate the continuous deformation of the waveguide's centerline. Mode 30 considered in Fig. 9 is inside passband I of the infinite perfectly periodic waveguide (which contains also the first 60 modes of the perfectly periodic finite waveguide with no disorder), whereas mode 90 in Fig. 10 is located inside passband II (which also contains modes 61–120 of the corresponding perfectly periodic finite waveguide with no disorder).

FIG. 9.

(Color online) Effect of thickness disorder and thermal buckling on mode 30: modeshape at (a) 400 K, (b) 390 K, (c) 370 K, (d) 353 K, and (e) 350 K in (i) the perfectly periodic finite waveguide (where mode 30 is in passband I) and (ii) the ( σ h = 5%) weakly disordered finite waveguide.

FIG. 9.

(Color online) Effect of thickness disorder and thermal buckling on mode 30: modeshape at (a) 400 K, (b) 390 K, (c) 370 K, (d) 353 K, and (e) 350 K in (i) the perfectly periodic finite waveguide (where mode 30 is in passband I) and (ii) the ( σ h = 5%) weakly disordered finite waveguide.

Close modal
FIG. 10.

(Color online) Effect of thickness disorder and thermal buckling on mode 90: modeshape at (a) 400 K, (b) 390 K, (c) 370 K, (d) 353 K, and (e) 350 K in (i) the perfectly periodic finite waveguide (where mode 90 is in passband II) and (ii) the ( σ h = 5%) weakly disordered finite waveguide.

FIG. 10.

(Color online) Effect of thickness disorder and thermal buckling on mode 90: modeshape at (a) 400 K, (b) 390 K, (c) 370 K, (d) 353 K, and (e) 350 K in (i) the perfectly periodic finite waveguide (where mode 90 is in passband II) and (ii) the ( σ h = 5%) weakly disordered finite waveguide.

Close modal

Figures 9(a)(i)–9(e)(i) show that mode 30 of the perfectly periodic finite waveguide deforms with comparable amplitudes over the entire spatial extent of the system, i.e., from cell 1 (the excited cell) up to cell 60, at all the considered temperatures. Such modes with spatially extended amplitude distributions over the entire waveguide are called extended modes that are necessary for wave propagation in the finite waveguides. For instance, this type of extended modes enables energy initially applied to cell 1 to appreciably deform the remaining cells resulting in detectable mechanical energy propagation through the entire spatial length of the waveguide. The extended modes in Figs. 9(a)(i)–9(e)(i) verify the persistence of wave transmission via passband I of the perfectly periodic finite waveguide as depicted in Figs. 7(a)(i)–7(e)(i) and Fig. 8(a)  for all temperatures, even very close to the critical-buckling temperature.

Conversely, in the weakly disordered finite waveguides, not all modes are extended for all temperatures. For example, for disorder level σ h = 5%, mode 30 is an extended mode only at 400 and 350 K, respectively—cf. Figs. 9(a)(ii) and 9(e)(ii). This extended shape directly affects wave transmission through the waveguide: A nonzero initial deformation of cell 1 due to the applied excitation deforms the cells throughout the waveguide as shown by the modeshapes of Figs. 9(a)(ii) and 9(e)(ii), allowing energy to propagate throughout the waveguide [cf. Figs. 7(a)(iii), 7(e)(iii), and 8(c)]. At the other considered temperatures, the disorder shrinks passband I, and, in turn, mode 30 shifts to a stop band and becomes spatially localized. Nevertheless, cell 1 exhibits minimal vibrations in Figs. 9(b)(ii)–9(d)(ii) at 390, 370, and 353 K, respectively, justifying the inability to transmit energy by mode 30 in Figs. 7(b)(iii)–7(d)(iii) and 8(c). The modeshapes in Figs. 9(b)(ii)–9(d)(ii) are “localized modes” where large deformations are confined only locally without extending over the entire waveguide (as in extended modes).

Localized modes characterize aperiodic/disordered structures because extended modes necessitate the periodicity between the constitutive cells in the waveguides. In other words, cells of similar geometry and material configurations form a periodic structure with cells of similar (isolated) modal frequencies, which we refer to as modal periodicity. This modal periodicity is necessary to form the extended modes that enable transmission throughout the structure. In this work, we observe this modal periodicity breaking in the weakly disordered waveguides under the effect of buckling because the cells in the waveguides develop different (isolated) modal frequencies in passband I due to buckling-induced changes in the grounding and coupling stiffnesses (cf. Ref. 41) The buckling-induced differences in modal frequencies lead to localized modes, as in Figs. 9(b)(ii)–9(d)(ii), inhibiting energy transmission throughout the waveguides.

As a general conclusion, buckling and thermoelastic effects lead to shifts of modeshapes in the frequency domain due to disorder, which, in turn, “transforms” certain modes from extended to localized. Therefore, thermal effects and buckling magnify the “modal” disorder in the disordered waveguides, leading to energy localization and confinement, similar to Anderson localization.42 Due to this effect, many modes between 1 and 60 (inside passband I of the perfectly periodic finite waveguide) become localized in the weakly disordered waveguides, as seen at 390 K in Mm. . Indeed, at 390 K, all the leading 60 modes are localized in the waveguide with σ h = 5%, prohibiting passband I transmission [cf. Figs. 7(b)(iii) and 8(c)]. In addition, Mm. demonstrates the existence of some extended modes between modes 1 and 60 in the waveguide with σ h = 2.5% at 390 K, which explains the observed passband I transmission at 390 K of Figs. 7(b)(ii) and 8(b). Last, in Mm. , all the leading 60 modes of the perfectly periodic waveguide are extended modes, resulting in broader passband I at 390 K than the waveguide with σ h = 2.5%—compare Fig. 7(b)(i) to Fig. 7(b)(ii). Note that all 60 leading modes of the perfectly periodic waveguide are extended modes even at the critical-buckling temperature of 370 K, as shown in Mm. .

Mm. 4.

Video of simulations of the propagation modes in periodic and disorder waveguides at 370 K within the first passband.

Mm. 4.

Video of simulations of the propagation modes in periodic and disorder waveguides at 370 K within the first passband.

Close modal

The buckling-induced localization does not occur in mode 90 of Fig. 10, not even in the weakly disordered waveguide of σ h = 5%. Therefore, mode 90 remains an extended mode at all temperatures and weak disorders, enabling the propagation of a wave at the modal frequency of mode 90. Most modes between 61 and 120 possess similar extended modeshapes, as illustrated in Mm. and Mm. at 390 K and the critical-buckling temperature of 370 K, respectively. In addition, as expected, all modes between 61 and 120 are extended in the perfectly periodic finite waveguide at all temperatures. In contrast, some modes away from the median mode 90 are localized in the weakly disordered finite waveguides, resulting in the narrowing of passband II in Figs. 8(b) and 8(c). Mm. and Mm. show fewer localized modes for weaker disorders (i.e., σ h of 2.5% vs 5%), verifying the wider passband II of Fig. 8(b) compared to Fig. 8(c).

Mm. 5.

Video of simulations of the propagation modes in periodic and disorder waveguides at 390 K within the second passband.

Mm. 5.

Video of simulations of the propagation modes in periodic and disorder waveguides at 390 K within the second passband.

Close modal
Mm. 6.

Video of simulations of the propagation modes in periodic and disorder waveguides at 370 K within the second passband.

Mm. 6.

Video of simulations of the propagation modes in periodic and disorder waveguides at 370 K within the second passband.

Close modal
This section investigates the transmission of the waveguide under scenarios that were not considered in the previous sections or in previous work.29 For instance, Fig. 11 examines the effect of the assumed thickness distribution in (9). Recall that so far in this work, the thickness distribution in (9) accounts for a linear variation (mimicking the wet etching effects) and a random variation (mimicking the fabrication random errors). Since (9) is a hypothetical thickness disorder model, we want to study transmission with different thickness distributions, like the ones in Figs. 11(a) and 11(b), representing a purely random disorder and a purely linear variation according to (21a) and (21b), respectively,
h i h Ref h Ref = σ h 4 r i 1 , 1 ,
(21a)
h i h Ref h Ref = σ h 4 2 N + 1 2 i N + 1 2 1 1 ,
(21b)
with σ h = 2.5% and the notations defined in (9).
FIG. 11.

(Color online) Effects of the disorder distribution: thickness profile relative to the reference thickness h ref for the considered (a) purely random disorder in (21a) and (b) purely linear disorder in (21b) with σ h = 2.5%; contour plots of the transmitted energy to cell 45 vs the temperature and the wave frequency in the 60-cell waveguide with (c) the purely random thickness disorder of (a) and (d) the purely linear thickness disorder of (b). Refer to Fig. 8 for details about the plots of (c) and (d).

FIG. 11.

(Color online) Effects of the disorder distribution: thickness profile relative to the reference thickness h ref for the considered (a) purely random disorder in (21a) and (b) purely linear disorder in (21b) with σ h = 2.5%; contour plots of the transmitted energy to cell 45 vs the temperature and the wave frequency in the 60-cell waveguide with (c) the purely random thickness disorder of (a) and (d) the purely linear thickness disorder of (b). Refer to Fig. 8 for details about the plots of (c) and (d).

Close modal

With these thickness distributions, Figs. 11(c) and 11(d) show similar transmission behavior under the effect of buckling [cf. Figs. 8(b) and 8(c)]: Passbands I and II detune depending on the waveguide temperature (i.e., the state of buckling); passband I does not transmit energy at temperatures close to the state of critical buckling, whereas passband II enables wave transmission over the entire temperature range considered. Therefore, the transmission loss and detuning due to buckling are qualitatively robust to the disorder distribution. By comparing the transmissions in Figs. 11(c) and 11(d), we conclude that the disorder distribution quantitatively affects the transmission loss, where the 2.5% disorder is more effective (i.e., results in a larger range of transmission loss) for a random distribution of Figs. 11(a) and 11(c) compared to the purely linear distribution of Figs. 11(b) and 11(d) [the transmission of the mixed 2.5% disorder of (9) in Fig. 8(b) is between the ones corresponding to the disorders of Fig. 11].

It is important to note that experimental disorders can take different forms than the computational ones investigated in this work. Furthermore, disorders can affect other parameters (in addition to the thickness of the cells), like the diameters, the residual stresses, and the material properties of the resonators (membranes). Thus, we do not claim (in this work and in Ref. 29) that the experimental disorders are identical to the theoretical models in (9) and (21); we only highlight the effect of structural disorders (as low as 5% in different forms) on the acoustics of the waveguides around buckling.

Additionally in this section, we are interested in testing the assumed models of the translational coupling Λ i C T, the torsional grounding Γ i B T, and the torsional coupling Γ i C T in (7). For this reason, we consider models 1 and 2 of Table II in two waveguides with the same thickness disorder, σ h = 5%, cf. Fig. 3(b). Figures 12(a) and 12(b) show the simulated transmission for the waveguides with models 1 and 2 of Table II, respectively. For instance, Fig. 12(a) shows that both passbands I and II keep on transmitting over [350, 400] K for model 1 of Table II. Regarding the transmission depicted in Fig. 12(a), the mean frequency of each passband detunes as a function of buckling, while exhibiting a constant bandwidth in both passbands. Alternatively, in Fig. 12(b), corresponding to the waveguide with the parameters of model 2 in Table II, the transmission via passband I persists over [350, 400] K but vanishes at passband II around critical buckling, where passband II narrows to a minimum bandwidth. By comparing the transmissions of Fig. 8(c) and Figs. 12(a) and 12(b) corresponding to the same thickness disorder σ h = 5% (see Fig. 3b), we can conjecture that the transmission loss necessitates a narrowing passband, in addition to the disorder.

TABLE II.

Models 0, 1, and 2 of Λ i C T, Γ i B T, and Γ i C T used in Figs. 4–11, 12(a), and 12(b), respectively.

Equations Corresponding figures
Model 0  Same as (7): Λ i C T = 0.2 Λ i Buck T min 350 400 K Λ i Buck T  Figs. 4–11  
Γ i B = χ Λ i B 
Γ i C T = χ 3 Λ i C T 3 4 Γ i B 
Model 1  Λ i C = 0.2 max 350 400 K Λ i Buck T min 350 400 K Λ i Buck T  Fig. 12(a)  
Γ i B T = χ Λ i B × Λ i Buck T max 350 400 K Λ i Buck T 
Γ i C = 1 4 χ Λ i C 
Model 2  Λ i C = 0.2 max 350 400 K Λ i Buck T min 350 400 K Λ i Buck T  Fig. 12(b)  
Γ i B T = χ Λ i B × Λ i Buck T max 350 400 K Λ i Buck T 
Γ i C T = 0.047 max 350 400 K Λ i Buck T Λ i Buck T 
Equations Corresponding figures
Model 0  Same as (7): Λ i C T = 0.2 Λ i Buck T min 350 400 K Λ i Buck T  Figs. 4–11  
Γ i B = χ Λ i B 
Γ i C T = χ 3 Λ i C T 3 4 Γ i B 
Model 1  Λ i C = 0.2 max 350 400 K Λ i Buck T min 350 400 K Λ i Buck T  Fig. 12(a)  
Γ i B T = χ Λ i B × Λ i Buck T max 350 400 K Λ i Buck T 
Γ i C = 1 4 χ Λ i C 
Model 2  Λ i C = 0.2 max 350 400 K Λ i Buck T min 350 400 K Λ i Buck T  Fig. 12(b)  
Γ i B T = χ Λ i B × Λ i Buck T max 350 400 K Λ i Buck T 
Γ i C T = 0.047 max 350 400 K Λ i Buck T Λ i Buck T 
FIG. 12.

(Color online) Different types of transmission: contour plots of the transmitted energy to cell 45 vs the temperature and the wave frequency in the 60-cell waveguide of Fig. 3(b) (i.e., σ h = 5% thickness disorder) with Λ i C, Γ i B, and Γ i C modeled as (a) model 1 and (b) model 2 of Table II. Refer to Fig. 8 for details about the plots of (a) and (b).

FIG. 12.

(Color online) Different types of transmission: contour plots of the transmitted energy to cell 45 vs the temperature and the wave frequency in the 60-cell waveguide of Fig. 3(b) (i.e., σ h = 5% thickness disorder) with Λ i C, Γ i B, and Γ i C modeled as (a) model 1 and (b) model 2 of Table II. Refer to Fig. 8 for details about the plots of (a) and (b).

Close modal
We note that the effect on the passbands guided our selection of the models of Λ i C T, Γ i B, and Γ i C in Table II. We aimed to investigate three scenarios, namely, a narrowing passband I and an almost constant bandwidth of passband II [model 0 of Table II, qualitatively mimicking the experimental transmission of Fig. 1(c)]; constant bandwidths of passbands I and II [model 1 of Table II and Fig. 12(a)]; and a constant bandwidth of passband I and a narrowing passband II [model 2 of Table II and Fig. 12(b)]. The Bloch-mode analysis of Sec. III helps relate the effect of Λ i C T, Γ i B T, and Γ i C T on the transmission characteristics of the considered waveguides. Accordingly, the eigenvalue problem in (8) results in the following eigenvalues:
k x = 0 : { Λ T = Λ Ref Buck T , Γ T = Γ Ref B T + Λ Ref C T χ ,
(22a)
k x = π L : { Λ T = Λ Ref Buck T + 4 Λ Ref C T , Γ T = Γ Ref B T + 4 Γ Ref C T χ ,
(22b)
where Λ and Γ denote the first and second eigenvalues of (8), respectively. In our considered cases, the band structure {i.e., the eigenfrequencies of (8) for values of k x [ 0 , π L ]} admits monotonic eigenvalues as functions of k x in each passband, as illustrated by the eigenfrequencies of Fig. 2(b) (i.e., they possess no local extrema within ] 0 , π L [). This monotonic feature enables the deduction from (22) of the bandwidth-like metrics Δ Λ and Δ Γ for passbands I and II, respectively,
Δ Λ = 4 Λ Ref C T ,
(23a)
Δ Γ = Λ Ref C T 4 Γ Ref C T χ .
(23b)
Based on (23a) and (23b), we select the models of Λ i C T, Γ i B T, and Γ i C T to represent different types of transmission. Such expressions relating the effective parameters (e.g., Λ i C, Γ i B, and Γ i C) to the waveguide transmission properties are made possible due to the ROM developed in this work. Moreover, the ROM allows the performance of parametric studies (e.g., the studies conducted in this work and specifically in this section), which would be computationally expensive and unfeasible to conduct with finite element modeling or experimental testing. For example, it takes more than 180 h to simulate the 60-cell waveguide transmission of Fig. 1(d) via COMSOL at six temperatures, compared to less than 3.5 h with the ROM at 51 temperatures (cf. Figs. 8, 11, and 12), thus, providing more than 400-fold savings in computational duration.

We developed a ROM for on-chip phononic waveguides made from coupled drumhead resonators. In particular, we investigated the effect of thermal-induced buckling on eliminating transmission over a low-frequency passband in weakly disordered waveguides. The considered disorders are very small and typically result from fabrication errors. We show that buckling magnifies the effect of weak geometric aperiodicity by amplifying the modal disorders between constitutive cells of the waveguide. The resulting effective aperiodicity yields transmission loss in the narrowing passband of the waveguide due to spatial localization of subsets of modes (which is similar to Anderson localization in disordered waveguides). The effectiveness of buckling on localizing certain modes depends on the stiffness (i.e., modal) detuning with temperature. For instance, localization of modes due to buckling is effective in the passbands that narrow near critical buckling, leading to transmission loss [Figs. 8(b) and 8(c) and Fig. 12(b)].

Notably, the developed ROM can capture the dynamics of a two-passband waveguide under thermal buckling. We adopt the buckling model from the experimental results in Ref. 41 by introducing the thermal expansion of the plate microstructure of the individual cells (undergoing stretching and bending) and the fabrication-residual stresses. Moreover, we control the level of buckling in the ROM waveguide by assigning different temperatures. We study the transmission as a function of temperature by considering the Bloch modes and the free response of the perfectly periodic finite waveguides. Hence, we present a method to relate the free response to the frequency content of transmitted waves in the waveguide, which saves computational effort compared to simulating the frequency response of the ROM.

The present study highlights the important role of validated ROMs in the design of phononic or acoustic waveguides that undergo buckling phase transitions. In these cases, Bloch mode analysis fails to capture the experimental results even for weakly disordered finite waveguides. We highlight the fact that transmission in finite waveguides is achieved via extended modes of the waveguides, whereas the existence of localized modes inhibits wave transmission, as energy becomes spatially confined and does not transmit throughout the extent of the waveguide. The developed ROM fully captures these results, which offers a reliable and robust alternative predictive design tool for on-chip phononic waveguides, compared to experimental and/or finite element computational methods, which are not as versatile or computationally inexpensive. Particularly, the ROM paves the way for nonlinear studies of phononic waveguides, with the promise of advancing our physics-based understanding of the acoustics of nonlinear micro/nano waveguides, as well as acoustic filtering of nonlinear weakly disordered waveguides susceptible (and tunable) to buckling effects.

This work was supported in part by National Science Foundation (NSF) Emerging Frontiers in Research and Innovation (EFRI) Grant No. 1741565. This support is gratefully acknowledged by the authors.

1.
R. H.
Olsson
and
I.
El-Kady
, “
Microfabricated phononic crystal devices and applications
,”
Meas. Sci. Technol.
20
(
1
),
012002
(
2009
).
2.
R.
Aigner
, “
Volume manufacturing of BAW-filters in a CMOS fab
,” in
Proceedings of the Second International Symposium on Acoustic Wave Devices for Future Mobile Communication Systems
, Chiba, Japan (March 3–5,
2004
), pp.
3
5
.
3.
K. M.
Lakin
, “
Thin film resonator technology
,”
IEEE Trans. Ultrason. Ferroelectr. Freq. Control
52
(
5
),
707
716
(
2005
).
4.
S.
Mohammadi
and
A.
Adibi
, “
On chip complex signal processing devices using coupled phononic crystal slab resonators and waveguides
,”
AIP Adv.
1
(
4
),
041903
(
2011
).
5.
M. B.
Zanjani
,
A. R.
Davoyan
,
A. M.
Mahmoud
,
N.
Engheta
, and
J. R.
Lukes
, “
One-way phonon isolation in acoustic waveguides
,”
Appl. Phys. Lett.
104
(
8
),
081905
(
2014
).
6.
H.
Zhu
and
F.
Semperlotti
, “
Double-zero-index structural phononic waveguides
,”
Phys. Rev. Appl.
8
(
6
),
064031
(
2017
).
7.
A.
Buhrdorf
,
O.
Ahrens
, and
J.
Binder
, “
Capacitive micromachined ultrasonic transducers and their application
,” in
2001 IEEE Ultrasonics Symposium. Proceedings. An International Symposium (Cat. No. 01CH37263)
, Atlanta, GA (October 7-10, 2021) (
IEEE
,
New York
,
2001
), Vol.
2
, pp.
933
940
.
8.
K.
Shung
and
M.
Zippuro
, “
Ultrasonic transducers and arrays
,”
IEEE Eng. Med. Biol. Mag.
15
(
6
),
20
30
(
1996
).
9.
R.
Lu
,
T.
Manzaneque
,
Y.
Yang
, and
S.
Gong
, “
Lithium niobate phononic crystals for tailoring performance of RF laterally vibrating devices
,”
IEEE Trans. Ultrason. Ferroelectr. Freq. Control
65
(
6
),
934
944
(
2018
).
10.
L.
Binci
,
C.
Tu
,
H.
Zhu
, and
J.-Y.
Lee
, “
Planar ring-shaped phononic crystal anchoring boundaries for enhancing the quality factor of Lamb mode resonators
,”
Appl. Phys. Lett.
109
(
20
),
203501
(
2016
).
11.
L.
Haofeng
,
J.
Rui
,
C.
Chen
, and
L.
Xinyu
, “
Surface acoustic wave sensors of delay lines based on MEMS
,”
J. Nanosci. Nanotechnol.
10
(
11
),
7258
7261
(
2010
).
12.
T.
Gorishnyy
,
C. K.
Ullal
,
M.
Maldovan
,
G.
Fytas
, and
E.
Thomas
, “
Hypersonic phononic crystals
,”
Phys. Rev. Lett.
94
(
11
),
115501
(
2005
).
13.
A.
Cleland
,
D.
Schmidt
, and
C. S.
Yung
, “
Thermal conductance of nanostructured phononic crystals
,”
Phys. Rev. B
64
(
17
),
172301
(
2001
).
14.
M.
Sledzinska
,
B.
Graczykowski
,
J.
Maire
,
E.
Chavez-Angel
,
C. M.
Sotomayor-Torres
, and
F.
Alzina
, “
2D phononic crystals: Progress and prospects in hypersound and thermal transport engineering
,”
Adv. Funct. Mater.
30
(
8
),
1904434
(
2020
).
15.
M.
Merklein
,
B.
Stiller
,
K.
Vu
,
S. J.
Madden
, and
B. J.
Eggleton
, “
A chip-integrated coherent photonic-phononic memory
,”
Nat. Commun.
8
(
1
),
574
(
2017
).
16.
H.
Shin
,
J. A.
Cox
,
R.
Jarecki
,
A.
Starbuck
,
Z.
Wang
, and
P. T.
Rakich
, “
Control of coherent information via on-chip photonic–phononic emitter–receivers
,”
Nat. Commun.
6
(
1
),
6427
(
2015
).
17.
E.
Verhagen
,
S.
Deléglise
,
S.
Weis
,
A.
Schliesser
, and
T. J.
Kippenberg
, “
Quantum-coherent coupling of a mechanical oscillator to an optical cavity mode
,”
Nature
482
(
7383
),
63
67
(
2012
).
18.
R. W.
Andrews
,
R. W.
Peterson
,
T. P.
Purdy
,
K.
Cicak
,
R. W.
Simmonds
,
C. A.
Regal
, and
K. W.
Lehnert
, “
Bidirectional and efficient conversion between microwave and optical light
,”
Nat. Phys.
10
(
4
),
321
326
(
2014
).
19.
H.
Shin
,
W.
Qiu
,
R.
Jarecki
,
J. A.
Cox
,
R. H.
Olsson
III
,
A.
Starbuck
,
Z.
Wang
, and
P. T.
Rakich
, “
Tailorable stimulated Brillouin scattering in nanoscale silicon waveguides
,”
Nat. Commun.
4
(
1
),
1944
(
2013
).
20.
S.
Gertler
,
P.
Kharel
,
E. A.
Kittlaus
,
N. T.
Otterstrom
, and
P. T.
Rakich
, “
Shaping nonlinear optical response using nonlocal forward Brillouin interactions
,”
New J. Phys.
22
(
4
),
043017
(
2020
).
21.
A. F.
Vakakis
,
O. V.
Gendelman
,
L. A.
Bergman
,
D. M.
McFarland
,
G.
Kerschen
, and
Y. S.
Lee
,
Nonlinear Targeted Energy Transfer in Mechanical and Structural Systems
(
Springer
,
New York
,
2008
).
22.
Y.
Lee
,
A. F.
Vakakis
,
L.
Bergman
,
D.
McFarland
,
G.
Kerschen
,
F.
Nucera
,
S.
Tsakirtzis
, and
P.
Panagopoulos
, “
Passive non-linear targeted energy transfer and its applications to vibration absorption: A review
,”
Proc. Inst. Mech. Eng. K
222
(
2
),
77
134
(
2008
).
23.
C.
Wang
,
A.
Kanj
,
A.
Mojahed
,
S.
Tawfick
, and
A. F.
Vakakis
, “
Experimental Landau–Zener tunneling for wave redirection in nonlinear waveguides
,”
Phys. Rev. Appl.
14
(
3
),
034053
(
2020
).
24.
C.
Wang
,
A.
Kanj
,
A.
Mojahed
,
S.
Tawfick
, and
A.
Vakakis
, “
Wave redirection, localization, and non-reciprocity in a dissipative nonlinear lattice by macroscopic Landau–Zener tunneling: Theoretical results
,”
J. Appl. Phys.
129
(
9
),
095105
(
2021
).
25.
A.
Kanj
,
C.
Wang
,
A.
Mojahed
,
A.
Vakakis
, and
S.
Tawfick
, “
Wave redirection, localization, and non-reciprocity in a dissipative nonlinear lattice by macroscopic Landau–Zener tunneling: Experimental results
,”
AIP Adv.
11
(
6
),
065328
(
2021
).
26.
M. I.
Hussein
,
M. J.
Leamy
, and
M.
Ruzzene
, “
Dynamics of phononic materials and structures: Historical origins, recent progress, and future outlook
,”
Appl. Mech. Rev.
66
(
4
),
040802
(
2014
).
27.
P. D.
Garcia
,
R.
Bericat-Vadell
,
G.
Arregui
,
D.
Navarro-Urrios
,
M.
Colombano
,
F.
Alzina
, and
C. M.
Sotomayor-Torres
, “
Optomechanical coupling in the Anderson-localization regime
,”
Phys. Rev. B
95
(
11
),
115129
(
2017
).
28.
J. C.
Ángel
,
J. C. T.
Guzman
, and
A. D.
de Anda
, “
Anderson localization of flexural waves in disordered elastic beams
,”
Sci. Rep.
9
(
1
),
3572
(
2019
).
29.
S.
Kim
,
J.
Bunyan
,
P. F.
Ferrari
,
A.
Kanj
,
A. F.
Vakakis
,
A. M.
Van Der Zande
, and
S.
Tawfick
, “
Buckling-mediated phase transitions in nano-electromechanical phononic waveguides
,”
Nano Lett.
21
(
15
),
6416
6424
(
2021
).
30.
D.
Hatanaka
,
I.
Mahboob
,
K.
Onomitsu
, and
H.
Yamaguchi
, “
A phonon transistor in an electromechanical resonator array
,”
Appl. Phys. Lett.
102
(
21
),
213102
(
2013
).
31.
I.
Wilson-Rae
,
R.
Barton
,
S.
Verbridge
,
D.
Southworth
,
B.
Ilic
,
H. G.
Craighead
, and
J.
Parpia
, “
High-Q nanomechanics via destructive interference of elastic waves
,”
Phys. Rev. Lett.
106
(
4
),
047205
(
2011
).
32.
J.
Liu
,
K.
Usami
,
A.
Naesby
,
T.
Bagci
,
E. S.
Polzik
,
P.
Lodahl
, and
S.
Stobbe
, “
High-Q optomechanical GaAs nanomembranes
,”
Appl. Phys. Lett.
99
(
24
),
243102
(
2011
).
33.
B.
Zwickl
,
W.
Shanks
,
A.
Jayich
,
C.
Yang
,
A.
Bleszynski Jayich
,
J.
Thompson
, and
J.
Harris
, “
High quality mechanical and optical properties of commercial silicon nitride membranes
,”
Appl. Phys. Lett.
92
(
10
),
103125
(
2008
).
34.
D.
Hatanaka
,
I.
Mahboob
,
H.
Okamoto
,
K.
Onomitsu
, and
H.
Yamaguchi
, “
An electromechanical membrane resonator
,”
Appl. Phys. Lett.
101
(
6
),
063102
(
2012
).
35.
J.
Cha
and
C.
Daraio
, “
Electrical tuning of elastic wave propagation in nanomechanical lattices at MHz frequencies
,”
Nat. Nanotechnol.
13
(
11
),
1016
1020
(
2018
).
36.
J.
Thompson
,
B.
Zwickl
,
A.
Jayich
,
F.
Marquardt
,
S.
Girvin
, and
J.
Harris
, “
Strong dispersive coupling of a high-finesse cavity to a micromechanical membrane
,”
Nature
452
(
7183
),
72
75
(
2008
).
37.
J. C.
Sankey
,
C.
Yang
,
B. M.
Zwickl
,
A. M.
Jayich
, and
J. G.
Harris
, “
Strong and tunable nonlinear optomechanical coupling in a low-loss system
,”
Nat. Phys.
6
(
9
),
707
712
(
2010
).
38.
D.
Hatanaka
,
I.
Mahboob
,
K.
Onomitsu
, and
H.
Yamaguchi
, “
Phonon waveguides for electromechanical circuits
,”
Nat. Nanotechnol.
9
(
7
),
520
524
(
2014
).
39.
J.
Cha
,
K. W.
Kim
, and
C.
Daraio
, “
Experimental realization of on-chip topological nanoelectromechanical metamaterials
,”
Nature
564
(
7735
),
229
233
(
2018
).
40.
Z.
Bazant
and
L.
Cedolin
, “
von Mises truss
,” in
Stability of Structures: Elastic, Inelastic, Fracture and Damage Theories
(
World Scientific
,
Singapore
,
2010
), pp.
228
231
.
41.
A.
Kanj
,
P. F.
Ferrari
,
A. M.
van der Zande
,
A. F.
Vakakis
, and
S.
Tawfick
, “
Ultra-tuning of nonlinear drumhead MEMS resonators by electro-thermoelastic buckling
,”
Mech. Syst. Signal Process.
196
,
110331
(
2023
).
42.
P. W.
Anderson
, “
Absence of diffusion in certain random lattices
,”
Phys. Rev.
109
(
5
),
1492
1505
(
1958
).
43.
A. P.
Mosk
,
A.
Lagendijk
,
G.
Lerosey
, and
M.
Fink
, “
Controlling waves in space and time for imaging and focusing in complex media
,”
Nat. Photonics
6
(
5
),
283
292
(
2012
).
44.
M.
Segev
,
Y.
Silberberg
, and
D. N.
Christodoulides
, “
Anderson localization of light
,”
Nat. Photonics
7
(
3
),
197
204
(
2013
).
45.
H.
Hu
,
A.
Strybulevych
,
J.
Page
,
S. E.
Skipetrov
, and
B. A.
van Tiggelen
, “
Localization of ultrasound in a three-dimensional elastic network
,”
Nat. Phys.
4
(
12
),
945
948
(
2008
).
46.
M.
Kurosu
,
D.
Hatanaka
,
H.
Okamoto
, and
H.
Yamaguchi
, “
Buckling-induced quadratic nonlinearity in silicon phonon waveguide structures
,”
Jpn. J. Appl. Phys.
61
(
SD
),
SD1025
(
2022
).
47.
K. C.
Balram
,
M. I.
Davanço
,
J. D.
Song
, and
K.
Srinivasan
, “
Coherent coupling between radiofrequency, optical and acoustic waves in piezo-optomechanical circuits
,”
Nat. Photonics
10
(
5
),
346
352
(
2016
).
48.
Y.
Zhang
,
S.
Hosono
,
N.
Nagai
, and
K.
Hirakawa
, “
Effect of buckling on the thermal response of microelectromechanical beam resonators
,”
Appl. Phys. Lett.
111
(
2
),
023504
(
2017
).
49.
K.-T.
Wan
,
S.
Guo
, and
D. A.
Dillard
, “
A theoretical and numerical study of a thin clamped circular film under an external load in the presence of a tensile residual stress
,”
Thin Solid Films
425
(
1–2
),
150
162
(
2003
).
50.
G. W.
Vogl
and
A. H.
Nayfeh
, “
A reduced-order model for electrically actuated clamped circular plates
,”
J. Micromech. Microeng.
15
,
684
690
(
2005
).