Losses associated with nonlinear wave propagation and exhibited by acoustic wave distortion and shock formation compromise the efficiency of contactless acoustic energy transfer systems. As such, predicting the shock formation distance and its dependence on the amplitude of the excitation is essential for their efficiency, design, and operation. We present an analytical approach capable of predicting the shock formation distance of acoustic waves generated by a baffled disk with arbitrary deformation in a weakly viscous fluid medium. The lossless Westervelt equation, used to model the nonlinear wave propagation, is asymptotically expanded based on the amplitude of the excitation. Because the solutions of the first- and second-order equations decay at different rates, we implement the method of renormalization and introduce a coordinate transformation to identify and eliminate the secular terms. The approach yields two partial differential equations that can be solved to predict the formation distance either analytically or numerically much faster than time-domain numerical simulations. The analysis and results are validated with solutions obtained from a nonlinear finite element simulation and previous experimental measurements.

Acoustic energy transfer (AET) is a transformative contactless energy transfer (CET) technology that utilizes acoustic waves to transfer energy between piezoelectric transducers. AET systems have been shown to outperform conventional electromagnetic CET technologies1–6 in critical applications.7,8 In particular, AET has been proposed to recharge and communicate with low-power (e.g., 1μW10mW) implanted medical devices,9–11 which eliminates the need for surgery to replace batteries11–17 and to develop battery-free underwater sensing networks to observe ocean conditions, track migration and habitats of marine animals, and monitor oil spills.8,18–22 These applications present the need in developing mathematical and numerical models capable of assessing the efficiency of AET systems.23–37 In general, most of the current approaches neglect nonlinear effects associated with acoustic wave propagation38 and the electro-elastic response of the piezoelectric transmitters and receivers.39,40 On the other hand, these effects become significant as the source strength is increased to enable higher energy transfer. As such, there is a need to expand the analysis capabilities to models that account for nonlinear effects and investigate their impact on the efficiency of energy transfer systems. In our previous work, we investigated theoretically the effect of material nonlinearity of a piezoelectric receiving disk on the energy transfer. We showed that the material nonlinearity can shift the optimum load resistance and that the shift is a function of the source strength.29 We have also shown experimentally that the interplay of all the nonlinearities and the standing wave effects between the transmitter and receiver in an AET system can manifest themselves in a complex manner and have an impact on the energy transfer efficiency.36 

One consequence of nonlinear acoustic wave propagation is exhibited by the distortion of its waveform due to a difference in the traveling speeds of the compression and rarefaction parts of the wave. In the frequency domain, this distortion is interpreted as energy transfer from the fundamental wave frequency to its higher harmonics. The accumulation of the distortion effect as the wave travels results in a discontinuity, referred to as a shock.38 The occurrence of a shock is associated with a significant loss in energy that is proportional to the cube of the difference in pressure across the discontinuity. This loss compounds as the wave propagates further, resulting in further reduction of the acoustic power.41 It is relevant to point out here that in a weakly viscous fluid such as water and air, shocks occur before the attenuation is significant.42 In other words, attenuation effects can be neglected, and a lossless second-order wave equation can be used to analyze the nonlinear wave propagation. In the case of a finite amplitude plane wave, the amplitudes of the higher harmonics grow at the expense of the amplitude of the fundamental or excitation frequency up to the initial shock formation location x¯. Beyond x¯, all components decay due to the energy dissipated as a consequence of the shock propagation.42 In the context of AET, the component of pressure at the excitation frequency pω that is generated by the transmitting disk operating under a high excitation voltage decreases as the distance from the disk is increased due to diffraction and transfer of energy to higher harmonics. However, the total acoustic power remains conserved up to x¯. Beyond x¯, in addition to the transfer of energy and diffraction, the decrease in pω will be compounded by additional losses in energy due to the formation and propagation of shocks. In AET systems, the power transfer efficiency depends mostly on the pressure associated with the excitation frequency, pω, as the higher harmonics do not necessarily coincide with the higher modes of the receiver.37 As such, the power transfer efficiency will be significantly compromised beyond x¯, which renders this distance as an essential design parameter for high-intensity AET power transfer.

Analytical expressions for x¯ are readily available for plane waves, spherical waves, and on the axis of a focused Gaussian beam,42,43 which is not the case in the AET system where the disk undergoes transverse deformations. Several numerical and analytical studies investigated nonlinear wave propagation and shock characteristics of acoustic waves generated by disks.44–51 The numerical simulations provide an accurate description of the response. However, they are computationally expensive, especially when solved in the time-domain. The objective of this effort is to develop an analytical approach to predict shock formation associated with a propagating acoustic wave generated by a vibrating disk with an arbitrary transverse displacement. We consider an axisymmetric-baffled-vibrating piezoelectric disk and use the Westervelt equation to investigate the associated nonlinear wave propagation. In particular, we scale the governing equation and boundary conditions with ϵ and obtain analytical expressions for the ϵ-order and ϵ2-order solutions using the Rayleigh integral. Next, we follow the work of Kelly and Nayfeh52 with some modifications to eliminate the secular terms by implementing the method of renormalization53,54 and obtain a uniformly valid solution of the Westervelt equation. We validate the predictions of the analytical approach with higher fidelity finite element simulations and previously published experimental results.

An axisymmetric-baffled piezoelectric disk with thickness h and radius a in a semi-infinite fluid medium is considered to analyze the nonlinear wave propagation of its generated acoustic wave as shown schematically in Fig. 1. The schematic defines Cartesian (x,y,z), cylindrical (rp,ψ,z), and spherical coordinate (r,θ,ψ) systems with the origin at the center of the disk Oc. An additional spherical coordinate system (rs,θs,ψ) with origin, Os, at z=r0 is also defined and used in the analysis. The disk is actuated using a dynamic potential difference V(t) across its flat surfaces at a frequency near that of its thickness mode. The resulting axisymmetric radial and transverse displacements are represented by u^(rp,z,t), and w^(rp,z,t), respectively. The transverse displacement of the thickness mode is chosen because it has a non-zero mean and is, therefore, favorable for generating acoustic pressure field.34,55

FIG. 1.

Schematic of the baffled disk with thickness h and radius a in a semi-infinite fluid medium.

FIG. 1.

Schematic of the baffled disk with thickness h and radius a in a semi-infinite fluid medium.

Close modal

The normal velocity continuity condition at the piezo-medium interface yields the necessary radiation boundary condition written as

(1)

where w(rp,t)=w^(rp,z,t)|z=0 and ϕ(rp,z,t) is the velocity potential of the fluid. This boundary condition is an approximation because the displacement, written using the Lagrangian description, would have to be mapped to an equivalent Eulerian description to obtain the exact boundary condition. Still, this approximation is acceptable because rpu^(rp,z,t)|z=0, except at rp=0 where u^(rp,z,t)|z=0=0; consequently, the effects of the approximation are insignificant. Because the modal deformation of the disk can take a complicated shape and does not possess a closed-form expression,55 we write w(rp,t) as

(2)

where ϵ, c0, and ω are, respectively, the acoustic Mach number, velocity of sound in fluid, and excitation frequency, and G(rp) and (rp), are, respectively, the displacement amplitude and phase as a function of the distance rp.

The lossless form of the Wesetervelt equation is used to analyze the nonlinear and non-planar wave propagation of the acoustic pressure p(rp,z,t). It is written as

(3)

where ρ0 and β are, respectively, the density and the coefficient of nonlinearity of the fluid. It is relevant to note that the lossless form is valid if the analysis is restricted to distances smaller than 1/αx¯ϵ, where α is the absorption loss parameter of the fluid and x¯ϵ is the shock formation distance for a plane wave.42 To solve Eq. (3) subjected to the boundary condition given by Eq. (1), we asymptotically expand the pressure as

(4)

and the velocity potential as

(5)

Substituting the assumed expansions of p(rp,z,t) and ϕ(rp,z,t) into Eqs. (1) and (3) yields the ϵ-order equation,

(6a)

the corresponding boundary condition,

(6b)

and the ϵ2-order equation,

(6c)

From Rayleigh’s or the Kirchhoff–Helmholtz (KH) integral, the pressure generated due to a baffled-vibrating surface of surface area S0 and transverse velocity U0eiωt is given by56,57

(7)

where R is the distance between the observation point and an infinitesimal point on the surface (AP in Fig. 1) and k=ω/c0. As such, the ϵ-order solution is given by

(8)

The ϵ-order solution in Eq. (8) contains the independent variables rp and z inside the integral. Hence, a partial differential equation with an integral forcing function needs to be solved to determine the ϵ2-order solution. Such a solution is not straightforward. To simplify the analysis, we rewrite p1(rp,z,t) in terms of an infinite series as suggested by Hasegawa et al.,58 using the spherical coordinate system (rs,θs,ψ) with the origin at z=r0 (see Fig. 1), as

(9a)

where

(9b)

and

(9c)

Following Hasegawa et al.,58 Eq. (9a) is modified to

(9d)

where

(9e)

r12=rp2+r02, ra2=a2+r02, Pn is the Legendre function, and hn(2) is the spherical Hankel function of the second kind, which is defined by hn(2)=jniyn, where jn and yn are, respectively, the spherical Bessel functions of the first and second kind. For subsequent analysis, the ϵ-order solution is converted from an exponential to a trigonometric form and written as

(9f)

where

(9g)
(9h)

Substituting the ϵ-order solution in the ϵ2-order, Eq. (6c) yields

(10a)

where

(10b)

and

(10c)

To obtain the analytical expression for the ϵ2-order solution, the product of Legendre functions is rewritten as52 

(11)

where the coefficient κqnm is determined from the orthogonality condition of the Legendre functions as

(12)

Substituting Eq. (11) into Eq. (10a), we obtain

(13)

As detailed in the  Appendix, the solution of ϵ2-order equation (13), p2(rs,θs,t), is obtained using the separation of variables. The final solution of the pressure is then written as

(14)

The ϵ-order and ϵ2-order terms in Eq. (14) decay at different rates with respect to rs. If left untreated, the ϵ2-order terms can become significant and contradict the scaling condition that ϵ-order terms >>ϵ2-order terms. To solve this contradiction and eliminate terms leading to this contradiction, we implement the method of renormalization and introduce the coordinate transformation,52 

(15)

Substituting the above transformation into Eq. (14) and applying the Taylor expansion yields

(16)

From Eq. (16), the necessary condition to eliminate the secular terms is

(17)

The transformation f becomes singular unless there is a particular relation between the phases of p2 and p1/rs|rs=η. To examine the phase relation, we represent p2 and p1/rs|rs=η, respectively, as p2(η,θs,t)=p¯2(η,θs)cos(2ωt2kη+p2) and p1/rs|r=η=p¯1(η,θs)cos(ωtkη+p1), where p¯2(η,θs), p2(η,θs), p¯1(η,θs), and p1(η,θs) are, respectively, the amplitude and phase of p2 and the amplitude and phase of p1/rs|rs=η. From Eq. (17), f(η,θs,t) is then determined as

(18)

It is evident from Eq. (18) that singularities arise in f because of the tangent functions unless p1p2/2=π/4. However, p1 and p2 do not always hold such a relation. To eliminate this singularity, we rewrite p2(η,θs,t) as

(19)

Then, eliminating the first part on the right hand side of Eq. (19) and using the remaining time-independent term as a feedback error to the ϵ-order solution yields

(20)

where

(21)
(22)

From Eq. (22), it is noted that the transformation f is independent of ϵ. This is a powerful consequence of the method of renormalization as once f is determined from p1 and p2, which are again independent of ϵ, it can be used for any value of ϵ.

Equations (20)–(22) constitute the solution of the nonlinear wave propagation of the acoustic pressure generated by a vibrating disk with transverse excitation according to Eq. (2). For a given rs, θs, and t, one needs to first evaluate f from Eq. (22) and then solve for η using Eq. (21). The value of η can then be used to determine the nonlinear pressure from Eq. (20). As the wave propagates in the medium, Eq. (21) will eventually yield multiple solutions for η due to the cumulative nature of the nonlinearity.38,49 The first location is where multiple solutions occur or when p/rs= is the shock location.38,52,59–61 Beyond this location, the solution is not valid, and a shock fitting criteria such as an equal-area rule should be used.38,53

The efficacy of the analysis presented in Sec. II to predict the nonlinear wave propagation and shock formation is assessed by comparing its predictions with those from (a) higher fidelity Finite Element (FE) simulations and (b) previously published experimental results.

A piezoelectric disk with radius a=5 mm and thickness h=2 mm made of PZT-5H material and submerged in water was considered for validation using numerical simulations. The material and acoustic properties of the disk and fluid are presented in Table I. We performed a series of axisymmetric frequency and time-domain FE simulations in COMSOL Multiphysics to obtain the acoustic radiation characteristics of the baffled-disk configuration. A quarter-circular fluid domain of radius rf was considered in the simulations. A spherical wave radiation boundary condition at the outer boundary was used to simulate an infinitely propagating wave, and triangular mesh elements were used in all the domains.

TABLE I.

Material properties of the disk and fluid. CE, e, and ɛS are, respectively, the stiffness at constant electric field, coupling parameter, and relative permittivity at constant strain matrices.

Parameter (units)Value
Density of the disk, ρ (kg/m37500 
C11E (GPa) 127.21 
C12E (GPa) 80.21 
C13E (GPa) 84.67 
C33E (GPa) 117.44 
C44E (GPa) 22.99 
e31 (C/m2−6.63 
e33 (C/m223.24 
e24 (C/m217.03 
ε11S 1704.4 
ε33S 1433.6 
Density of the fluid, ρ0 (kg/m3999.6 
Velocity of sound in the fluid, c0 (m/s) 1481.44 
Nonlinear parameter of the fluid, β 3.6 
Parameter (units)Value
Density of the disk, ρ (kg/m37500 
C11E (GPa) 127.21 
C12E (GPa) 80.21 
C13E (GPa) 84.67 
C33E (GPa) 117.44 
C44E (GPa) 22.99 
e31 (C/m2−6.63 
e33 (C/m223.24 
e24 (C/m217.03 
ε11S 1704.4 
ε33S 1433.6 
Density of the fluid, ρ0 (kg/m3999.6 
Velocity of sound in the fluid, c0 (m/s) 1481.44 
Nonlinear parameter of the fluid, β 3.6 

1. Linear acoustic radiation

In both analysis and simulations, the material and geometric nonlinearities of the disk are neglected. Furthermore, because of the presence of a baffle, only the transverse displacement of the top surface of the disk is required to determine its acoustic radiation. In the simulations, we extracted this displacement from a linear FE simulation and use it as a boundary condition in the nonlinear FE simulation for computational efficiency, realized by eliminating the multiphysics coupling between the piezoelectric and fluid domains.

Considering a maximum mesh size of λf0/6 in the fluid domain and 20 elements each along the radius and thickness of the disk, we solved the linear problem using the frequency domain FE solver and determined that the thickness extensional mode resonant frequency of the disk is f0=1.005 MHz (see p. 108 in Guo62), where λf0=c0/f0. The corresponding Rayleigh far-field distance is rfar=12ka2=53.3 mm. The response and radiation characteristics of the disk for an excitation voltage amplitude corresponding to ϵ=105 and an excitation frequency of f0 are presented in Fig. 2. In particular, Figs. 2(a)2(d) show, respectively, the sound pressure level (pref=1 Pa) in the fluid domain, deformation of the disk, transverse displacement of the top surface of the disk, and the beam pattern at different radial distances from the origin [see schematic in Fig. 2(a)]. The abnormal higher deformation amplitude on the bottom surface compared to the top surface of the disk in Fig. 2(b) is due to the absence of radiation damping unlike the top surface. From Fig. 2(d), we note that except for r=0.125rfar, the pressure along the z-axis is maximum for any r value. Consequently, we expect the shock to occur first on the z-axis. We also note that r=0.125rfar is a local minimum similar to those observed in the near-field of axial pressure generated by a piston.46,57,63 Because we expect the shock to occur along the z-axis first, we limit further analysis only to the axial pressure distribution.

FIG. 2.

(a) The sound pressure level (pref=1 Pa) in the fluid domain, (b) real and imaginary components of the deformation of the disk, (c) transverse displacement of the top surface of the disk, and (d) the beam (directivity) pattern at different radial distances from the origin for ϵ=105 at an excitation frequency of 1.005 MHz.

FIG. 2.

(a) The sound pressure level (pref=1 Pa) in the fluid domain, (b) real and imaginary components of the deformation of the disk, (c) transverse displacement of the top surface of the disk, and (d) the beam (directivity) pattern at different radial distances from the origin for ϵ=105 at an excitation frequency of 1.005 MHz.

Close modal

2. Nonlinear axial acoustic wave propagation

Next, the transverse displacement presented in Fig. 2(c) is used to obtain axial p1 and p2 according to the solution in Sec. II. Alternatively, p1 and p2 could be determined by solving the ϵ-order and ϵ2-order equations and the ϵ-order boundary condition in a frequency domain FE solver. A pressure release boundary condition on the disk should be used to solve the ϵ2-order equations to ensure that p2 vanishes on the surface of the disk. We note that p2 evaluated using the two approaches should differ only by the non-secular terms (NSTs) that were neglected in the analysis. The axial p1 and p2 evaluated from the model for n=50,r0=50/k, labeled as Series n=50, and by the frequency domain FE simulations are presented, respectively, in Figs. 3(a) and 3(b). A mesh size of λf0/12 was used in the FE simulations. Figure 3(a) shows an excellent agreement between p1 evaluated by the analysis and numerical simulations except at close distances to the disk. The disagreement in this region is due to limiting the series and can be minimized by choosing higher values of n and r0. From Fig. 3(b), we note a close agreement in the amplitude and a small phase shift. We attribute this shift to errors induced by choosing a finite value of n and to the contribution of the NST. It is relevant to point out that p1 and p2 determined using FE are more accurate and that eliminating p2 that includes NST in the method of the renormalization will result in a transformation that will account for the complete particular solution of the ϵ2-order equation. As for computational cost, the evaluation of p1 and p2 from the analysis requires the analytical expressions of integrals presented in the  Appendix and storing, reusing, and manipulation of these expressions, which can become computationally expensive, especially for higher values of n. On the other hand, the linear frequency domain FE simulations to determine p1 and p2 is more efficient as it involves a matrix inversion. Moreover, the FE solver can be easily extended to complicated geometries.

FIG. 3.

Comparison of axial (a) ϵ-order solution p1 and (b) ϵ2-order solution p2, determined by using the analysis developed for n=50,r0=50/k and by a frequency domain FE solver. The analytical solution is labeled as Series, n=50, and the FE solution is labeled as Freq. FE. On the z-axis, rs=rr0 and θ=θs=π.

FIG. 3.

Comparison of axial (a) ϵ-order solution p1 and (b) ϵ2-order solution p2, determined by using the analysis developed for n=50,r0=50/k and by a frequency domain FE solver. The analytical solution is labeled as Series, n=50, and the FE solution is labeled as Freq. FE. On the z-axis, rs=rr0 and θ=θs=π.

Close modal

Having determined p1 and p2, we compare next the nonlinear wave propagation predicted using Eqs. (20)–(22) with that determined from a nonlinear time-domain FE simulation for three different excitation levels, namely, ϵ=5×103, ϵ=2.5×103, and ϵ=1.25×103. Toward that objective, we constructed a fluid domain of rf=2.25×x¯ϵ that is excited by the transverse displacement presented in Fig. 2(c). A maximum mesh size of λf0/36 was selected so that the response at 6f0 can be captured. For the sake of completeness, the Westervelt equation with dissipation was simulated in time-domain using COMSOL Multiphysics by choosing the diffusivity of the sound as δ=1.34×106m2/s. However, the absorption is not expected to have a significant impact on the response because the characteristic absorption distance [1/α=2c03/(δω2)] for f0 is about 120 m, which is significantly larger than 2.25×x¯1.25×103(0.1m). The acoustics (Westervelt) model along with the shock-capturing stabilization scheme with q=1.35 and κ=0.00015 was used to solve for the nonlinear wave propagation in this study.

The steady-state response characteristics for ϵ=5×103 obtained using time-domain FE simulation and from Eqs. (19)–(22) using p1 and p2 as determined from the analysis (perturbation and series, n=50) and frequency domain FE simulations (perturbation and freq. FE) are compared next. The plots presented in Figs. 4(a)4(d) show, respectively, the steady-state axial pressure waveform at time t=1/(2f0), the mean intensity, the amplitude of the pressure component with frequency ω=2πf0, and the amplitude of the pressure component with frequency 2ω. Figures 4(b) and 4(c) also show, respectively, the mean intensity and pressure when β=0, i.e., for the linear case. Based on the closeness of the responses predicted when p1 and p2 are determined from the analysis and from the FE simulations, it can be concluded that the contribution of NST is not significant. It can also be concluded that the transformation does an excellent job in predicting the pressure distributions at closer distances and that the discrepancies grow as the distance increases. The discrepancy is because the ϵ3-order secular terms become dominant as the distance increases to the point where one needs to carry the method of renormalization to ϵ3-order to maintain the accuracy. Figures 4(a)4(d) also show regions where the transformation yields multiple solutions, which invalidates the solution, and as such, no solution is shown. As noted in previous studies,38,53 these regions correspond to the shock formation, and a shock fitting criteria should be used to obtain the true waveform in these regions. From the results, the transformation yields multiple solutions in a small region around r/x¯5×103=0.5 and when r/x¯5×103>1.13. However, from Fig. 4(b), the mean intensity predicted by the time-domain FE simulation deviates from the linear case only beyond r/x¯5×103=1.13. This deviation is a consequence of loss of energy due to the formation and propagation of shock.38,46 As such, we hypothesize that r/x¯5×103=1.13 is the shock formation location for ϵ=5×103. As expected, an accelerated decrease in pomega is seen after the shock formation at r/x¯5×103=1.13 in Fig. 4(c). To confirm that a shock did not occur around r/x¯5×103=0.5, we present the time series and Fourier coefficients of the pressure at r/x¯5×103=0.48 as obtained from the time-domain FE simulation, respectively, in Figs. 5(a) and 5(b). The Fourier component of pressure at 2f0 is comparable with that at f0 in Fig. 5(b), and the corresponding time series presented in Fig. 5(a) does not indicate a shock, confirming that no shock occurs in the region around r/x¯5×103=0.5. We attribute this anomaly predicted by the model to the local minima in the amplitude of pressure, which is essentially a singularity. As such, multiple solutions are also associated with local minima of the pressure amplitude in the near field. To the knowledge of the authors, this conclusion has not been made in previous investigations as the pressure fields analyzed in those studies did not exhibit local minima.

FIG. 4.

Comparison of a steady-state axial (a) pressure waveform at time t=1/(2f0), (b) mean intensity, (c) component of the pressure at ω=2πf0, and (d) the component of the pressure at 2ω obtained from the time-domain FE simulation and from the model developed when p1 and p2 are obtained from the analysis (perturbation and series, n=50) and from frequency domain simulations (perturbation and freq. FE) for ϵ=5×103 and an excitation frequency f0=1.005 MHz. The axial distance is normalized with the corresponding characteristic shock formation distance x¯5×103=13.03 mm.

FIG. 4.

Comparison of a steady-state axial (a) pressure waveform at time t=1/(2f0), (b) mean intensity, (c) component of the pressure at ω=2πf0, and (d) the component of the pressure at 2ω obtained from the time-domain FE simulation and from the model developed when p1 and p2 are obtained from the analysis (perturbation and series, n=50) and from frequency domain simulations (perturbation and freq. FE) for ϵ=5×103 and an excitation frequency f0=1.005 MHz. The axial distance is normalized with the corresponding characteristic shock formation distance x¯5×103=13.03 mm.

Close modal
FIG. 5.

The (a) time series and (b) amplitude Fourier coefficients of the pressure at r/x¯5×103=0.48 on the z-axis for ϵ=5×103 and excitation frequency f0=1.005 MHz obtained from a time-domain FE simulation. Here, x¯5×103=13.03 mm.

FIG. 5.

The (a) time series and (b) amplitude Fourier coefficients of the pressure at r/x¯5×103=0.48 on the z-axis for ϵ=5×103 and excitation frequency f0=1.005 MHz obtained from a time-domain FE simulation. Here, x¯5×103=13.03 mm.

Close modal

The steady-state axial pressure waveform at time t=0, mean intensity, the component of pressure at ω, and the component of pressure at 2ω akin to Figs. 4(a)4(d) for the case of ϵ=2.5×103 are, respectively, presented in Figs. 6(a)6(d). The agreement of the results obtained from the analysis and the time-domain FE simulation is similar to that observed in Figs. 4(a)4(d). The only major difference is the absence of an anomaly, which suggests that the occurrence of anomaly also depends on the magnitude of ϵ. More importantly, the results predicted by the analysis suggest that the shock formation location on the z-axis is r/x¯2.5×103=1.04, which is also the point after which the mean intensity predicted by the time-domain FE simulation deviates from the linear case. The characteristic shock formation distance for ϵ=2.5×103 is x¯2.5×103=26.07 mm.

FIG. 6.

Comparison of the steady-state axial (a) pressure waveform at time t=0, (b) mean intensity, (c) the component of the pressure at ω=2πf0, and (d) the component of the pressure at 2ω obtained from the time-domain FE simulation and from the model developed when p1 and p2 are obtained from the analysis (perturbation and series, n=50) and from frequency domain simulations (perturbation and freq. FE) for ϵ=2.5×103 and excitation frequency f0=1.005 MHz. The axial distance is normalized with the corresponding characteristic shock formation distance x¯2.5×103=26.07 mm.

FIG. 6.

Comparison of the steady-state axial (a) pressure waveform at time t=0, (b) mean intensity, (c) the component of the pressure at ω=2πf0, and (d) the component of the pressure at 2ω obtained from the time-domain FE simulation and from the model developed when p1 and p2 are obtained from the analysis (perturbation and series, n=50) and from frequency domain simulations (perturbation and freq. FE) for ϵ=2.5×103 and excitation frequency f0=1.005 MHz. The axial distance is normalized with the corresponding characteristic shock formation distance x¯2.5×103=26.07 mm.

Close modal

The results for the case of excitation amplitude ϵ=1.25×103 are presented in Figs. 7(a)7(d). Inspecting the steady-state axial pressure waveform at time t=0 predicted by the time-domain FE simulation in Fig. 7(a), we note that a shock does not take place in the range of 0<r<2.25×x¯1.25×103, where x¯1.25×103=52.13 mm. This result can also be inferred from Fig. 7(b), which does not show any deviation of the mean intensity predicted by the time-domain FE simulation from the linear case. Consequently, the deviation of pω obtained from the time-domain FE simulation from the linear case in Fig. 7(c) is very small. The results predicted by the analysis suggest that the shock formation location on the z-axis is around r/x¯1.25×103=1.64, which falls in the far-field region of the disk. This inconsistency is again a consequence of implementing the method of renormalization only up to the O(ϵ2) order, as discussed earlier.

FIG. 7.

Comparison of the steady-state axial (a) pressure waveform at time t=0, (b) mean intensity, (c) the component of the pressure at ω=2πf0, and (d) the component of the pressure at 2ω obtained from the time-domain FE simulation and from the model developed when p1 and p2 are obtained from analysis (perturbation and series, n=50) and from frequency domain simulations (perturbation and freq. FE) for ϵ=1.25×103 and excitation frequency f0=1.005 MHz. The axial distance is normalized with the corresponding characteristic shock formation distance x¯1.25×103=52.13 mm.

FIG. 7.

Comparison of the steady-state axial (a) pressure waveform at time t=0, (b) mean intensity, (c) the component of the pressure at ω=2πf0, and (d) the component of the pressure at 2ω obtained from the time-domain FE simulation and from the model developed when p1 and p2 are obtained from analysis (perturbation and series, n=50) and from frequency domain simulations (perturbation and freq. FE) for ϵ=1.25×103 and excitation frequency f0=1.005 MHz. The axial distance is normalized with the corresponding characteristic shock formation distance x¯1.25×103=52.13 mm.

Close modal

The efficacy of the developed model in predicting the nonlinear wave propagation and the shock formation location is also validated by comparing its predictions with the experimental data of Nachef et al.63 Nachef et al. considered a piezoelectric ceramic disk having a radius of 25 mm and measured the pressure waveforms over axial distances between 50 mm and 400 mm and lateral distances between 0 mm and 50 mm for different source strengths at an excitation frequency of f0=1 MHz. These waveforms were later treated as a benchmark by Khokhlova et al.46 to validate their numerical models for p0=0.2,0.64, and 1.43 MPa where p0=ϵρ0c02 is the amplitude of the pressure at the source. They identified that the shock formation distances on the axis are, respectively, 0.1rfar and 0.28rfar for p0=1.43 MPa and p0=0.64 MPa where the Rayleigh far-field distance rfar=1156.63 mm. They concluded that a shock did not occur on the axis before 400 mm46 when p0=0.2 MPa. The relevant material and geometric properties were identified as ρ0=1000kg/m3, c0=1500 m/s, and β=3.5 and the equivalent aperture of the circular piston as a=23.5 mm.46 

From the discussion in Sec. III A 2, p1 and p2 can be evaluated either by the analysis or from appropriate frequency domain linear FE simulations. For the sake of simplicity, we evaluated p1 and p2 using the FE simulations to determine the nonlinear wave propagation generated by a baffled piston. Figures 8(a) and 8(b) show, respectively, the comparisons of Fourier components of pressure at ω and 2ω predicted using Eqs. (20)–(22) with the linear case (β=0) and corresponding experimental data when p0=0.2 MPa. The corresponding results for p0=0.64 MPa and p0=1.43 MPa are presented, respectively, in Figs. 8(c) and 8(d) and Figs. 8(e) and 8(f). From the results, we note that the shock formation distances predicted by the method of renormalization are r/rfar=0.25 and r/rfar=0.1, respectively, for p0=0.64 MPa and p0=1.43 MPa and that a shock did not occur in the range considered for p0=0.2 MPa. These findings are very close to the shock formation distances identified by Khokhlova et al.,46 which validate the analysis.

FIG. 8.

Comparisons of Fourier components of the axial pressure generated by a baffled piston at ω and 2ω predicted using Eqs. (20)–(22) (perturbation and freq. FE) with the linear case (β=0) and corresponding experimental data, respectively, when (a) and (b) p0=0.2 MPa, (c) and (d) p0=0.64 MPa, and (e) and (f) p0=1.43 MPa. The axial distance is normalized using the Rayleigh far-field distance rfar=1156.63 mm, and the experimental data are extracted from Khokhlova et al.46 

FIG. 8.

Comparisons of Fourier components of the axial pressure generated by a baffled piston at ω and 2ω predicted using Eqs. (20)–(22) (perturbation and freq. FE) with the linear case (β=0) and corresponding experimental data, respectively, when (a) and (b) p0=0.2 MPa, (c) and (d) p0=0.64 MPa, and (e) and (f) p0=1.43 MPa. The axial distance is normalized using the Rayleigh far-field distance rfar=1156.63 mm, and the experimental data are extracted from Khokhlova et al.46 

Close modal

Various effects determine the desired distance in AET systems that delivers the optimum power transfer. The output power fluctuates as the separation distance between the transmitter and receiver is increased due to the spatial acoustic resonances. These fluctuations are strong at closer distances and often become weak with an increase in the separation distance as the influence of the reflected waves diminishes with an increase in the separation distance. When these fluctuations are strong, constructive interference positions become the desired distances for AET. However, for a finite-sized transmitter and receiver disks, the general trend in the output power decreases with an increase in the separation distance. On the other hand, if the transmitter’s excitation amplitude is strong enough to trigger the wave propagation nonlinearities, we note that the formation and propagation of shock are associated with a loss in energy, thereby compromising the output power transfer beyond the initial shock formation distance. As such, the knowledge of the shock formation distance is crucial for AET systems.

The shock formation distance of a piezoelectric transmitter disk with arbitrary deformation depends on the excitation amplitude, the operation frequency, deformation of the disk, the disk’s diameter, and the material properties of the acoustic medium such as density, speed of sound, and the coefficient of nonlinearity. A closed-form expression that estimates the shock formation distance is not available in the literature. The analysis developed in this work provides a quick estimate that can aid the design and operation of efficient AET systems. The analysis is based on the asymptotic expansion of the governing equation. The method of renormalization is used to predict the nonlinear wave generated by a finite amplitude baffled circular disk for a specified deformation profile. The lossless form of the Westervelt equation and the normal velocity continuity boundary condition were scaled with a non-dimensional parameter related to the amplitude of the transverse displacement of the disk. They were expanded and solved by introducing a coordinate transformation to eliminate the secular terms, which resulted in singularities that required redefining the relationship between the pressure components. We validated the analysis by comparing its predictions of the shock formation distance under different excitation levels with higher fidelity nonlinear finite element simulations and previously published experimental results. We also demonstrated the versatility of the analysis by showing that the ϵ-order and ϵ2-order solutions can be determined either from analytical expressions or from linear frequency domain finite element simulations. We showed that the analysis accurately predicted the nonlinear wave propagation and shock formation distance in the near-field of the disk and argued that the accuracy level could be increased by extending the analysis to ϵ3 or higher orders. We found out that the transformation yields multiple solutions at the shock location and at local minima that occur in the near-field of the pressure. We also showed an accelerated decrease in the power of the excitation frequency beyond the shock formation. Of particular importance is the validity of the solution for any excitation level because the ϵ and ϵ2 order solutions are independent of that level.

This work was supported by the National Science Foundation (NSF) (Grant No. ECCS-1711139), which is gratefully acknowledged. The authors would also like to thank Professor Saad Ragab (Virginia Tech) for providing valuable feedback in developing the model.

The authors declare that they have no conflict of interest.

The data that support the findings of this study are available from the corresponding author upon reasonable request.

To determine the particular solution of Eq. (13), we use the separation of variable technique and assume the solution as

(A1)

Substitution of Eq. (A1) in Eq. (13) yields

(A2a)
(A2b)

where

(A2c)
(A2d)

The homogeneous solutions of Eqs. (A2a) and (A2b) are X1(rs)=Ψ1(rs)=jq(2krs) and X2(rs)=Ψ2(rs)=yq(2krs), which are now used to determine its particular solution by using the variation of the parameter technique as

(A3a)
(A3b)

where x is a dummy variable. Since the closed-form expression of spherical Bessel functions in terms of trigonometric functions is required to identify the secular terms in the ϵ2-order solution [i.e., Ψqnm(rs) and Xqnm(rs)], we use an identity of Rayleigh’s formula and represent the spherical Bessel function of the first kind as

(A4)

where nn1 and n1n2 are, respectively, the Stirling numbers of the first and second kind. From Eq. (A4), the coefficients of the trigonometric functions can be determined, which can then be used to determine yn(x). By using the so obtained analytical expressions of the spherical Bessel functions, it can be shown that

(A5)

Using the above relation, Eqs. (A3a) and (A3b) are rewritten as

(A6a)
(A6b)

Evaluation of integrals in Eqs. (A6a) and (A6b) shows that Ψqnm(rs) and Xqnm(rs) contain sine-integral, cosine-integral, 1/rs[], and logarithmic terms. Furthermore, it can also be shown that logarithmic terms grow indefinitely for large rs with respect to the ϵ-order solution and are the secular terms. Although the rest of the terms do not grow indefinitely for large r, some of them have a major contribution to the ϵ2-order response, especially at closer distances to the disk, and hence govern the nonlinear interaction. As such, we note that an asymptotic form (such as Kelly and Nayfeh52) of these expressions will not convey accurate physics. Furthermore, we noticed that sine-integral and cosine-integral terms decay monotonically at a faster rate than logarithmic and 1/rs[] terms, and hence, we regard them as non-secular terms (NSTs) and are neglected in the further analysis.

It is relevant to point out that the homogeneous solution of the ϵ2-order equation that is responsible for satisfying the ϵ2-order boundary condition is not considered in the analysis as it represents a freely propagating wave arising from excitation at the boundary and hence does not exhibit an unbounded growth (i.e., NST). Moreover, due to Eqs. (A3a) and (A3b), the ϵ2-order vanishes on the surface of the disk (rs=r0/cosθs). This condition ensures that any transformation defined in the method of renormalization vanishes on the surface of the disk, as is required to describe accurate underlying physics.52 Furthermore, from the definitions of Δn(c) and Δn(c), we note that they are the only parameters that vary for a different deformation profile and that they are constants of integration in determining Ψqnm(rs) and Xqnm(rs). As such, the analytical expressions for these integrals can be utilized to determine p2 for any deformation pattern.

1.
J. L.
Villa
,
J.
Sallán
,
A.
Llombart
, and
J. F.
Sanz
, “
Design of a high frequency inductively coupled power transfer system for electric vehicle battery charge
,”
Appl. Energy
86
(
3
),
355
363
(
2009
).
2.
Z.
Bi
,
T.
Kan
,
C. C.
Mi
,
Y.
Zhang
,
Z.
Zhao
, and
G. A.
Keoleian
, “
A review of wireless power transfer for electric vehicles: Prospects to enhance sustainable mobility
,”
Appl. Energy
179
,
413
425
(
2016
).
3.
P.
Maharjan
,
M.
Salauddin
,
H.
Cho
, and
J. Y.
Park
, “
An indoor power line based magnetic field energy harvester for self-powered wireless sensors in smart home applications
,”
Appl. Energy
232
,
398
408
(
2018
).
4.
M.
Jafari
,
Z.
Malekjamshidi
, and
J.
Zhu
, “
Design and development of a multi-winding high-frequency magnetic link for grid integration of residential renewable energy systems
,”
Appl. Energy
242
,
1209
1225
(
2019
).
5.
X.
Yu
,
S.
Sandhu
,
S.
Beiker
,
R.
Sassoon
, and
S.
Fan
, “
Wireless energy transfer with the presence of metallic planes
,”
Appl. Phys. Lett.
99
(
21
),
214102
(
2011
).
6.
S.
Kim
,
J. S.
Ho
,
L. Y.
Chen
, and
A. S.
Poon
, “
Wireless power transfer to a cardiac implant
,”
Appl. Phys. Lett.
101
(
7
),
073701
(
2012
).
7.
M. G.
Roes
,
J. L.
Duarte
,
M. A.
Hendrix
, and
E. A.
Lomonova
, “
Acoustic energy transfer: A review
,”
IEEE Trans. Ind. Electron.
60
(
1
),
242
248
(
2013
).
8.
J.
Jang
and
F.
Adib
, “Underwater backscatter networking,” in
Proceedings of the ACM Special Interest Group on Data Communication
(
ACM
, 2019), pp. 187–199.
9.
P.
Valdastri
,
E.
Susilo
,
T.
Forster
,
C.
Strohhofer
,
A.
Menciassi
, and
P.
Dario
, “
Wireless implantable electronic platform for chronic fluorescent-based biosensors
,”
IEEE Trans. Biomed. Eng.
58
(
6
),
1846
1854
(
2011
).
10.
Y.
Vaiarello
,
W.
Tatinian
,
Y.
Leduc
,
N.
Veau
, and
G.
Jacquemod
, “
Ultra-low-power radio microphone for cochlear implant application
,”
IEEE J. Emerg. Sel. Top. Circuits Syst.
1
(
4
),
622
630
(
2011
).
11.
T.
Maleki
,
N.
Cao
,
S. H.
Song
,
C.
Kao
,
S.-C.
Ko
, and
B.
Ziaie
, “
An ultrasonically powered implantable micro-oxygen generator (IMOG)
,”
IEEE Trans. Biomed. Eng.
58
(
11
),
3104
3111
(
2011
).
12.
K. L.
Lee
,
C. -P.
Lau
,
H. -F.
Tse
,
D. S.
Echt
,
D.
Heaven
,
W.
Smith
, and
M.
Hood
, “
First human demonstration of cardiac stimulation with transcutaneous ultrasound energy delivery: Implications for wireless pacing with implantable devices
,”
J. Am. Coll. Cardiol.
50
(
9
),
877
883
(
2007
).
13.
J.
Charthad
,
M. J.
Weber
,
T. C.
Chang
, and
A.
Arbabian
, “
A mm-sized implantable medical device (IMD) with ultrasonic power transfer and a hybrid bi-directional data link
,”
IEEE J. Solid-State Circuits
50
(
8
),
1741
1753
(
2015
).
14.
D.
Seo
,
J. M.
Carmena
,
J. M.
Rabaey
,
M. M.
Maharbiz
, and
E.
Alon
, “
Model validation of untethered, ultrasonic neural dust motes for cortical recording
,”
J. Neurosci. Methods
244
,
114
122
(
2015
).
15.
J.
Charthad
,
T. C.
Chang
,
Z.
Liu
,
A.
Sawaby
,
M. J.
Weber
,
S.
Baker
,
F.
Gore
,
S. A.
Felt
, and
A.
Arbabian
, “
A mm-sized wireless implantable device for electrical stimulation of peripheral nerves
,”
IEEE Trans. Biomed. Circuits Syst.
12
(
2
),
257
270
(
2018
).
16.
L.
Jiang
,
Y.
Yang
,
R.
Chen
,
G.
Lu
,
R.
Li
,
J.
Xing
,
K. K.
Shung
,
M. S.
Humayun
,
J.
Zhu
,
Y.
,
Chen
et al., “
Ultrasound-induced wireless energy harvesting for potential retinal electrical stimulation application
,”
Adv. Funct. Mater.
29
(
33
),
1902522
(
2019
).
17.
S.
Baltsavias
,
W.
Van Treuren
,
M. J.
Weber
,
J.
Charthad
,
S.
Baker
,
J. L.
Sonnenburg
, and
A.
Arbabian
, “In vivo wireless sensors for gut microbiome redox monitoring,” arXiv:1902.07386 (2019).
18.
N. E.
Leonard
,
D. A.
Paley
,
R. E.
Davis
,
D. M.
Fratantoni
,
F.
Lekien
, and
F.
Zhang
, “
Coordinated control of an underwater glider fleet in an adaptive ocean sampling field experiment in Monterey Bay
,”
J. Field Robot.
27
(
6
),
718
740
(
2010
).
19.
M. C.
Domingo
, “
An overview of the internet of underwater things
,”
J. Netw. Comput. Appl.
35
(
6
),
1879
1890
(
2012
).
20.
G.
Xu
,
W.
Shen
, and
X.
Wang
, “
Applications of wireless sensor networks in marine environment monitoring: A survey
,”
Sensors
14
(
9
),
16932
16954
(
2014
).
21.
I. F.
Akyildiz
,
P.
Wang
, and
S.-C.
Lin
, “
Softwater: Software-defined networking for next-generation underwater communication systems
,”
Ad Hoc Netw.
46
,
1
11
(
2016
).
22.
Z.
Li
,
S.
Desai
,
V. D.
Sudev
,
P.
Wang
,
J.
Han
, and
Z.
Sun
, “
Underwater cooperative MIMO communications using hybrid acoustic and magnetic induction technique
,”
Comput. Netw.
173
,
107191
(
2020
).
23.
S.
Ozeri
,
D.
Shmilovitz
,
S.
Singer
, and
C.-C.
Wang
, “
Ultrasonic transcutaneous energy transfer using a continuous wave 650 kHz Gaussian shaded transmitter
,”
Ultrasonics
50
(
7
),
666
674
(
2010
).
24.
S.
Shahab
and
A.
Erturk
, “
Contactless ultrasonic energy transfer for wireless systems: Acoustic-piezoelectric structure interaction modeling and performance enhancement
,”
Smart Mater. Struct.
23
(
12
),
125032
(
2014
).
25.
S.
Shahab
,
M.
Gray
, and
A.
Erturk
, “
Ultrasonic power transfer from a spherical acoustic wave source to a free-free piezoelectric receiver: Modeling and experiment
,”
J. Appl. Phys.
117
(
10
),
104903
(
2015
).
26.
M.
Gorostiaga
,
M.
Wapler
, and
U.
Wallrabe
, “
Analytic model for ultrasound energy receivers and their optimal electric loads
,”
Smart Mater. Struct.
26
(
8
),
085003
(
2017
).
27.
M.
Gorostiaga
,
M.
Wapler
, and
U.
Wallrabe
, “
Analytic model for ultrasound energy receivers and their optimal electric loads II: Experimental validation
,”
Smart Mater. Struct.
26
(
10
),
105021
(
2017
).
28.
M.
Bakhtiari-Nejad
,
A.
Elnahhas
,
M. R.
Hajj
, and
S.
Shahab
, “
Acoustic holograms in contactless ultrasonic power transfer systems: Modeling and experiment
,”
J. Appl. Phys.
124
(
24
),
244901
(
2018
).
29.
V. C.
Meesala
,
M. R.
Hajj
, and
S.
Shahab
, “
Modeling and identification of electro-elastic nonlinearities in ultrasonic power transfer systems
,”
Nonlinear Dyn.
99
,
249
268
(
2020
).
30.
H.
Basaeri
,
Y.
Yu
,
D.
Young
, and
S.
Roundy
, “
A MEMS-scale ultrasonic power receiver for biomedical implants
,”
IEEE Sens. Lett.
3
(
4
),
1
4
(
2019
).
31.
B.
Herrera
,
F.
Pop
,
C.
Cassella
, and
M.
Rinaldi
, “AlN PMUT-based ultrasonic power transfer links for implantable electronics,” in 2019 20th International Conference on Solid-State Sensors, Actuators and Microsystems and Eurosensors XXXIII (Transducers and Eurosensors XXXIII) (IEEE, 2019), pp. 861–864.
32.
H.
Basaeri
,
Y.
Yu
,
D.
Young
, and
S.
Roundy
, “
Acoustic power transfer for biomedical implants using piezoelectric receivers: Effects of misalignment and misorientation
,”
J. Micromech. Microeng.
29
(
8
),
084004
(
2019
).
33.
A.
Allam
,
K.
Sabra
, and
A.
Erturk
, “
Aspect ratio-dependent dynamics of piezoelectric transducers in wireless acoustic power transfer
,”
IEEE Trans. Ultrason. Ferroelectr. Freq. Control
67
(
5
),
984
996
(
2019
).
34.
V. C.
Meesala
,
S.
Ragab
,
M. R.
Hajj
, and
S.
Shahab
, “
Acoustic-electroelastic interactions in ultrasound energy transfer systems: Reduced-order modeling and experiment
,”
J. Sound Vib.
475
,
115255
(
2020
).
35.
M.
Bakhtiari-Nejad
,
M. R.
Hajj
, and
S.
Shahab
, “
Dynamics of acoustic impedance matching layers in contactless ultrasonic power transfer systems
,”
Smart Mater. Struct.
29
(
3
),
035037
(
2020
).
36.
A.
Bhargava
,
V. C.
Meesala
,
M. R.
Hajj
, and
S.
Shahab
, “
Nonlinear effects in high-intensity focused ultrasound power transfer systems
,”
Appl. Phys. Lett.
117
(
6
),
064101
(
2020
).
37.
A.
Bhargava
and
S.
Shahab
, “
Acoustic-electroelastic modeling of piezoelectric disks in high-intensity focused ultrasound power transfer systems
,”
J. Intel. Mater. Syst. Struct.
(published online).
38.
M. F.
Hamilton
,
D. T.
Blackstock
et al.,
Nonlinear Acoustics
(
Academic Press
,
San Diego, CA
,
1998
), Vol.
237
.
39.
S.
Leadenham
and
A.
Erturk
, “
Unified nonlinear electroelastic dynamics of a bimorph piezoelectric cantilever for energy harvesting, sensing, and actuation
,”
Nonlinear Dyn.
79
(
3
),
1727
1743
(
2015
).
40.
V. C.
Meesala
and
M. R.
Hajj
, “
Identification of nonlinear piezoelectric coefficients
,”
J. Appl. Phys.
124
(
6
),
065112
(
2018
).
41.
I.
Rudnick
, “
On the attenuation of a repeated sawtooth shock wave
,”
J. Acoust. Soc. Am.
25
(
5
),
1012
1013
(
1953
).
42.
T.
Muir
and
E.
Carstensen
, “
Prediction of nonlinear acoustic effects at biomedical frequencies and intensities
,”
Ultrasound Med. Biol.
6
(
4
),
345
357
(
1980
).
43.
D.
Dalecki
,
E. L.
Carstensen
,
K. J.
Parker
, and
D. R.
Bacon
, “
Absorption of finite amplitude focused ultrasound
,”
J. Acoust. Soc. Am.
89
(
5
),
2435
2447
(
1991
).
44.
S. I.
Aanonsen
,
T.
Barkve
,
J. N.
Tjo/tta
, and
S.
Tjo/tta
, “
Distortion and harmonic generation in the nearfield of a finite amplitude sound beam
,”
J. Acoust. Soc. Am.
75
(
3
),
749
768
(
1984
).
45.
Y.-S.
Lee
and
M. F.
Hamilton
, “
Time-domain modeling of pulsed finite-amplitude sound beams
,”
J. Acoust. Soc. Am.
97
(
2
),
906
917
(
1995
).
46.
V.
Khokhlova
,
R.
Souchon
,
J.
Tavakkoli
,
O.
Sapozhnikov
, and
D.
Cathignol
, “
Numerical modeling of finite-amplitude sound beams: Shock formation in the near field of a cw plane piston source
,”
J. Acoust. Soc. Am.
110
(
1
),
95
108
(
2001
).
47.
J. H.
Ginsberg
, “
Nonlinear king integral for arbitrary axisymmetric sound beams at finite amplitudes. I. Asymptotic evaluation of the velocity potential
,”
J. Acoust. Soc. Am.
76
(
4
),
1201
1207
(
1984
).
48.
J. H.
Ginsberg
, “
Nonlinear king integral for arbitrary axisymmetric sound beams at finite amplitude. II. Derivation of uniformly accurate expressions
,”
J. Acoust. Soc. Am.
76
(
4
),
1208
1214
(
1984
).
49.
F. Y.
Coulouvrat
, “
An analytical approximation of strong nonlinear effects in bounded sound beams
,”
J. Acoust. Soc. Am.
90
(
3
),
1592
1600
(
1991
).
50.
K.-E.
Fro/ysa
and
F.
Coulouvrat
, “
A renormalization method for nonlinear pulsed sound beams
,”
J. Acoust. Soc. Am.
99
(
6
),
3319
3328
(
1996
).
51.
M.
Foda
, “
Axial nonlinear field of a vibrating circular transducer
,”
Arch. Acoust.
22
(
1
),
59
76
(
1997
), see https://acoustics.ippt.pan.pl/index.php/aa/article/view/937
52.
S. G.
Kelly
and
A.
Nayfeh
, “
Non-linear propagation of directional spherical waves
,”
J. Sound Vib.
72
(
1
),
25
37
(
1980
).
53.
A. H.
Nayfeh
and
D. T.
Mook
,
Nonlinear Oscillations
(
John Wiley & Sons
,
2008
).
54.
A. H.
Nayfeh
,
Perturbation Methods
(
John Wiley & Sons
,
2008
).
55.
N.
Guo
,
P.
Cawley
, and
D.
Hitchings
, “
The finite element analysis of the vibration characteristics of piezoelectric discs
,”
J. Sound Vib.
159
(
1
),
115
138
(
1992
).
56.
Y.-H.
Kim
,
Sound Propagation: An Impedance Based Approach
(
John Wiley & Sons
,
2010
).
57.
L. E.
Kinsler
,
A. R.
Frey
,
A. B.
Coppens
, and
J. V.
Sanders
, Fundamentals of Acoustics, 4th ed. (Wiley-VCH, 1999), 560 pp., ISBN: 0-471-84789-5.
58.
T.
Hasegawa
,
N.
Inoue
, and
K.
Matsuzawa
, “
A new rigorous expansion for the velocity potential of a circular piston source
,”
J. Acoust. Soc. Am.
74
(
3
),
1044
1047
(
1983
).
59.
A.
Nayfeh
and
S. G.
Kelly
, “
Non-linear interactions of acoustic fields with plates under harmonic excitations
,”
J. Sound Vib.
60
(
3
),
371
377
(
1978
).
60.
J. H.
Ginsberg
, “
Propagation of nonlinear acoustic waves induced by a vibrating cylinder. I. The two-dimensional case
,”
J. Acoust. Soc. Am.
64
(
6
),
1671
1678
(
1978
).
61.
J. H.
Ginsberg
, “
Propagation of nonlinear acoustic waves induced by a vibrating cylinder. II. The three-dimensional case
,”
J. Acoust. Soc. Am.
64
(
6
),
1679
1687
(
1978
).
62.
N.
Guo
, “The vibration characteristics of piezoelectric discs,” Ph.D. thesis (University of London, 1989).
63.
S.
Nachef
,
D.
Cathignol
,
J. N.
Tjo/tta
,
A. M.
Berg
, and
S.
Tjo/tta
, “
Investigation of a high intensity sound beam from a plane transducer. Experimental and theoretical results
,”
J. Acoust. Soc. Am.
98
(
4
),
2303
2323
(
1995
).