Nonlinear acoustics is the branch of acoustics that examines the behavior of sound waves with sufficiently large amplitudes propagating in nonlinear media, where the relationship between acoustic pressure and particle displacement deviates from linearity, leading to phenomena such as waveform distortion, harmonic generation, and shock wave formation.[1] This field contrasts with linear acoustics by incorporating amplitude-dependent effects that violate the superposition principle, requiring higher-order terms in the governing equations derived from fluid mechanics, including the continuity, momentum, and energy equations.[2] A key parameter quantifying nonlinearity is the coefficient of nonlinearity B/A, which measures the deviation from linear compressibility in fluids and typically ranges from 3 to 12 in liquids and biological tissues.[2]The theoretical foundation of nonlinear acoustics traces back to early contributions by Leonhard Euler in 1755 and George Gabriel Stokes in 1848, who analyzed finite-amplitude waves, with modern developments accelerating in the mid-20th century through studies of ultrasonic propagation.[1] Fundamental models for describing nonlinear wave propagation include the Fubini solution for lossless distortion, the Burgers equation accounting for absorption and nonlinearity, and the Khokhlov-Zabolotskaya-Kuznetsov (KZK) equation for focused beams in dissipative media.[1] Notable phenomena encompass harmonic generation, where a monochromatic wave produces higher-frequency components (e.g., second and third harmonics) due to steepening of the waveform; shock wave formation, occurring when the Gol'dberg number exceeds unity, resulting in sawtooth profiles; and self-demodulation, which produces low-frequency beams from high-frequency nonlinear interactions.[2] Additionally, cavitation arises from nonlinear pressure fluctuations inducing bubble oscillations, while parametric acoustic arrays exploit sum and difference frequency generation to create directive sound beams.[1]Applications of nonlinear acoustics span multiple domains, particularly in biomedical engineering, where tissue harmonic imaging enhances ultrasound resolution by detecting generated harmonics rather than fundamental frequencies, and high-intensity focused ultrasound (HIFU) leverages nonlinear focusing for therapeutic heating without invasive procedures.[3] In underwater acoustics, nonlinear effects enable advanced sonar systems and seabed profiling through parametric arrays that produce narrow, low-frequency beams from high-frequency sources.[1] Other uses include nondestructive testing of materials, where nonlinear harmonics detect microstructural damage like cracks or fatigue in solids, and aeroacoustics, modeling shock waves in high-speed flows.[3] These applications underscore the field's impact on diagnostics, therapy, and engineering, with ongoing research addressing saturation effects and nonlinear propagation in complex media.[2]
Fundamentals
Definition and Importance
Nonlinear acoustics is a branch of physics and acoustics that examines sound waves of sufficiently large amplitudes, where the propagation deviates from linear superposition principles due to the nonlinear response of the medium.[3] This field focuses on amplitude-dependent phenomena arising from finite amplitude effects, such as waveformdistortion and the generation of higher harmonics during propagation.[4]The importance of nonlinear acoustics lies in its ability to explain behaviors in high-intensity sound fields that linear theory cannot capture, including wave steepening leading to shock wave formation and increased energy dissipation through nonlinear mechanisms.[5] These effects are critical in applications like medical ultrasound imaging and therapy, where harmonic generation enhances resolution, and in aerospace engineering for mitigating sonic booms from supersonic aircraft.[6] Nonlinear acoustic theory has enabled advancements in nondestructive testing and material characterization by detecting microstructural damage through cumulative nonlinear distortions.[7]Central to the field is the nonlinearity parameter B/A for fluids, introduced by Beyer in 1960, which quantifies the medium's nonlinear compressibility via the equation of state and determines the strength of finite amplitude distortions.[8] Nonlinearity becomes significant when the acoustic Mach number exceeds approximately $10^{-3}, as in intense ultrasound fields with pressures above 0.1 MPa, or in high-Mach-number flows like sonic booms where shock formation occurs.[9] Mathematical models extending linear wave equations account for these processes, providing predictive tools for nonlinear propagation.[10]
Physical Mechanisms
Nonlinear behavior in acoustic waves arises primarily from the nonlinear stress-strain relations inherent in the propagation medium. In fluids, this nonlinearity stems from the equation of state that relates pressure and density beyond the linear approximation, leading to a dependence of the sound speed on the wave amplitude.[11] For liquids, the key parameter quantifying this effect is the nonlinearity parameter B/A, which measures the deviation from Hooke's law in the pressure-density relationship and is derived from thermodynamic properties.[12] In solids, similar nonlinearities occur due to anharmonic lattice vibrations and elastic moduli that vary with strain, often characterized by parameters like the third-order elastic constants.[5]A fundamental consequence of these nonlinear relations is the variation of the local wave speed with particle velocity, expressed as c = c_0 + \beta u, where c_0 is the linear sound speed, u is the particle velocity, and \beta is the coefficient of nonlinearity (typically \beta = 1 + B/(2A) for fluids).[13] This amplitude-dependent speed causes faster parts of the wave to overtake slower ones, resulting in wave steepening.[14] As steepening progresses, the waveform distorts, eventually forming a shock wave where the pressure gradient becomes discontinuous.[6] Without dissipative effects, this process would lead to infinite gradients, but in reality, shocks form with finite thickness determined by the medium's properties.[11]Absorption and viscosity play crucial roles in counteracting nonlinear steepening by dissipating energy, particularly at high frequencies generated during distortion.[15]Viscosity introduces shear stresses that smooth the waveform, preventing the formation of multivalued profiles as first recognized by Stokes in his analysis of finite-amplitude waves.[16] Thermoviscous absorption balances the nonlinear distortion, stabilizing shock fronts and limiting the energy transfer to higher harmonics that would otherwise cause infinite steepening.[6] This interplay ensures that nonlinear effects manifest as sawtooth-like waveforms in dissipative media rather than breaking entirely.[11]The manifestation of nonlinearity varies significantly across media due to differences in molecular structure and interactions. In gases, nonlinearity primarily arises from the adiabatic compression process, where finite amplitudes lead to entropy variations and a relatively modest \beta \approx (\gamma + 1)/2 (with \gamma the specific heat ratio), resulting in weaker effects at typical intensities.[13] Liquids exhibit much higher nonlinearity owing to strong intermolecular forces and cohesive energies, yielding B/A values often 5–10 times larger than in gases, which amplifies steepening and shock formation at lower amplitudes.[5] In solids, nonlinearities are dominated by microscopic anharmonicities in the crystal lattice, leading to both compressional and shear wave distortions, with parameters like the Murnaghan constants describing higher-order stress-strain couplings that can exceed those in fluids.[15]
Historical Development
The foundations of nonlinear acoustics trace back to the mid-18th century, when Leonhard Euler formulated the basic equations governing finite-amplitude sound waves in gases in 1759, establishing the framework for understanding wave propagation beyond infinitesimal amplitudes.[17] Early 19th-century contributions built on this, with Siméon Denis Poisson deriving an exact solution for progressive waves of finite amplitude in 1808, revealing waveform distortion though its implications were initially overlooked.[18] In 1848, George Gabriel Stokes provided a critical explanation for this distortion, postulating that waveforms steepen and form shocks, with viscosity preventing discontinuities, marking one of the first recognitions of nonlinear effects in sound propagation.[18] Samuel Earnshaw advanced the field in 1860 by generalizing propagation laws for adiabatic gases and solving the piston problem, highlighting the dependence of wave speed on amplitude.[17]By the early 20th century, Lord Rayleigh contributed significantly in 1910 with a paper on steady shocks in thermoviscous fluids, analyzing shock structure and dissipation, which concluded a formative era of theoretical exploration.[18] Further progress came in 1931 when Robert Fay developed a theory for nearly stable waveforms in viscous gases, incorporating harmonic generation, and in 1935 when Guido Fubini-Ghirga offered an explicit solution for piston-driven waves using Bessel functions, quantifying distortion and higher harmonics.[17] These works shifted focus from linear acoustics, dominant in the early 1900s, toward recognizing nonlinearity as essential for finite-amplitude phenomena.Post-World War II, nonlinear acoustics emerged as a distinct field in the 1950s, driven by practical needs in supersonic flight and ultrasonics; Robert T. Beyer pioneered studies on finite-amplitude wave distortion and absorption in liquids, publishing key reviews that synthesized experimental and theoretical advances.[19] K. Uno Ingard contributed concurrently with analyses of nonlinear interactions, including scattering of sound by sound and parametric effects in finite-amplitude waves.[20] The era's breakthroughs addressed sonic booms from high-speed aircraft and ultrasonic applications in medicine and materials, spurring dedicated research beyond linear approximations.[21]In the 1960s, the field formalized key models, with David T. Blackstock and others adapting the Burgers equation—originally from fluid dynamics in the 1940s—for acoustic propagation, enabling solutions for dissipative nonlinear waves and shock formation.[22] The 1970s saw the development of the KZK equation in 1969 by L. Khokhlov and E. Zabolotskaya for nonlinear beam propagation, extended by V. P. Kuznetsov in 1971 to include absorption, facilitating analysis of focused ultrasonic fields.[23]Influential figures expanded applications: Jan D. Achenbach advanced nonlinear wave theory in solids during the 1970s–1990s, deriving equations for finite-amplitude elastic waves and surface acoustics in anisotropic media.[24] Werner Lauterborn, from the 1970s onward, illuminated cavitation-related nonlinearity through numerical models of bubble oscillations, linking acoustic nonlinearity to chaotic dynamics in liquids.[25] These contributions solidified nonlinear acoustics as vital for phenomena like sonic booms and ultrasonic effects.[21]
Theoretical Models
Linear versus Nonlinear Wave Propagation
In linear acoustic wave propagation, sound waves are modeled as small-amplitude disturbances where the superposition principle holds, allowing the total wave to be the linear sum of individual waves without mutual interaction.[26] This approximation assumes that wave speed is independent of amplitude, so an initially sinusoidal waveform remains sinusoidal as it propagates, with no generation of new frequency components beyond the original.[26] Such behavior is valid for low-intensity sounds, like everyday speech or background noise, where nonlinear effects are negligible.[13]Nonlinear wave propagation deviates from this ideal when amplitudes become sufficiently large, introducing amplitude-dependent effects that violate superposition and cause waveform distortion.[26] As the wave travels, compression regions propagate faster than rarefaction regions due to the medium's nonlinear response, leading to steepening of the waveform front and the generation of higher harmonics—multiples of the fundamental frequency—that were absent in the source signal.[13] Over distance, this distortion can evolve into a sawtooth-like profile, where energy cascades to higher frequencies, accompanied by enhanced absorption as the waveform's spectral content shifts upward.[27]The onset of nonlinearity is quantified by dimensionless parameters such as the Gol'dberg number (Γ), which compares nonlinear distortion to linear dissipation, or the acoustic Reynolds number (Re_ac), which assesses the ratio of inertial to viscous forces in the acoustic field.[27] When Γ ≈ 1 or Re_ac ≫ 1, nonlinear effects become prominent, marking the threshold beyond which linear theory fails.[27]Qualitatively, low-amplitude waves, such as those from a soft whistle, propagate without distortion, maintaining their shape over long distances in air.[13] In contrast, high-amplitude waves, like intense ultrasound pulses or sonic booms, distort rapidly, forming steep fronts that can develop into shocks, altering the pressure profile into an asymmetric N-shape and increasing energy loss through cumulative absorption.[13]
Governing Equations and Derivations
The governing equations for nonlinear acoustics are derived from the fundamental laws of fluid dynamics for compressible fluids, specifically the Navier-Stokes equations, which account for conservation of mass, momentum, and energy. The continuity equation expresses mass conservation as \frac{\partial \rho}{\partial t} + \nabla \cdot (\rho \mathbf{v}) = 0, where \rho is density and \mathbf{v} is the particle velocity vector. The momentum equation, including viscous effects, is \rho \left( \frac{\partial \mathbf{v}}{\partial t} + (\mathbf{v} \cdot \nabla) \mathbf{v} \right) = -\nabla p + \eta \nabla^2 \mathbf{v} + \left( \zeta + \frac{1}{3} \eta \right) \nabla (\nabla \cdot \mathbf{v}), with p as pressure, \eta as shear viscosity, and \zeta as bulk viscosity. The energy equation incorporates thermal conduction and relates internal energy changes to heat flux, often simplified under isentropic conditions to dp = c^2 d\rho, where c is the speed of sound.[28][16]For acoustic perturbations, these equations are simplified by assuming small-amplitude disturbances around a quiescent equilibriumstate (\rho = \rho_0 + \rho', p = p_0 + p', \mathbf{v} = 0 + \mathbf{v}', with primes denoting small quantities). The flow is taken as irrotational, allowing introduction of the velocity potential \phi such that \mathbf{v} = \nabla \phi. Linearization yields the classical wave equation \frac{\partial^2 \phi}{\partial t^2} - c_0^2 \nabla^2 \phi = 0, where c_0 = \sqrt{(\partial p / \partial \rho)_s} at equilibrium. To capture nonlinear effects for finite but small amplitudes, a perturbationexpansion is employed: \phi = \epsilon \phi_1 + \epsilon^2 \phi_2 + \cdots, where \epsilon \ll 1 scales the amplitude. Substituting into the momentum and continuity equations introduces quadratic nonlinear terms from the convective acceleration (\mathbf{v} \cdot \nabla) \mathbf{v} and from the nonlinear equation of state p' = c_0^2 \rho' + \frac{B}{2 \rho_0} (\rho')^2 + \cdots, with B = \rho_0^2 (\partial^2 p / \partial \rho^2)_s as the nonlinearity parameter.[28][15][16]In the lossless case, neglecting viscosity and thermal conduction, the perturbation expansion to second order yields the basic nonlinear wave equation for the velocity potential:\frac{\partial^2 \phi}{\partial t^2} - c_0^2 \nabla^2 \phi = \frac{\partial}{\partial t} \left[ (\nabla \phi)^2 + \frac{B + 2}{2 c_0^2} \left( \frac{\partial \phi}{\partial t} \right)^2 \right],where the right-hand side arises from the nonlinear convective term and the equation-of-state nonlinearity, scaled by the coefficient \beta = 1 + B/(2A) with A = \rho_0 c_0^2. This equation describes wave distortion, such as steepening and shock formation, without dissipation balancing the nonlinearity. For plane progressive waves in one dimension, it simplifies further to highlight the role of B/A in speed variation with amplitude.[28][16][15]Loss terms are incorporated by retaining viscous and conductive effects in the perturbation expansion, leading to dissipative contributions on the right-hand side. Viscosity introduces terms proportional to \nabla^2 \mathbf{v}, contributing to classical absorption \alpha \propto \omega^2 (with \omega as angular frequency), while thermal conduction adds a similar term via heat flux \mathbf{q} = -\kappa \nabla T in the energy equation, where \kappa is thermal conductivity and T is temperature. These losses prevent infinite steepening in the lossless case, enabling derivations of dissipative models by balancing nonlinearity with attenuation; for instance, the absorption coefficient is \gamma = \frac{\omega^2}{2 \rho_0 c_0^3} \left[ \left( \frac{4}{3} \eta + \zeta \right) + \kappa \left( \frac{1}{C_v} - \frac{1}{C_p} \right) \right], linking to higher-order acoustic equations.[28][16]
Westervelt Equation
The Westervelt equation represents a key theoretical model in nonlinear acoustics for describing the propagation of finite-amplitude sound waves in dissipative fluids, capturing both quadratic nonlinearity and classical absorption effects under weakly nonlinear conditions. Derived originally in the context of parametric acoustic arrays, it extends the linear wave equation by incorporating second-order terms from fluid dynamics. This hyperbolic partial differential equation is particularly suited for modeling plane progressive waves where absorption is small compared to nonlinear distortion.The derivation for plane progressive waves begins from the fundamental governing equations of compressible fluid dynamics: the continuity equation \frac{\partial \rho}{\partial t} + \nabla \cdot (\rho \mathbf{v}) = 0 and the momentum equation \rho \left( \frac{\partial \mathbf{v}}{\partial t} + (\mathbf{v} \cdot \nabla) \mathbf{v} \right) = -\nabla p + \nabla \cdot \boldsymbol{\tau}, where \rho is density, \mathbf{v} is particle velocity, p is pressure, and \boldsymbol{\tau} is the viscous stress tensor for absorption. The equation of state is expanded as p = p_0 + c^2 (\rho - \rho_0) + \frac{B}{2A} \frac{(\rho - \rho_0)^2}{\rho_0} + \cdots, introducing the nonlinearity parameter \beta = 1 + \frac{B}{2A}, which quantifies the medium's nonlinear response based on the coefficients A and B from thermodynamic relations. Under assumptions of irrotational flow (\mathbf{v} = \nabla \phi), small Mach number (weakly nonlinear regime, where acoustic perturbations are much smaller than ambient values), and small absorption relative to nonlinearity, a perturbation expansion is applied to second order in acoustic quantities. The nonlinear convective term (\mathbf{v} \cdot \nabla) \mathbf{v} and the quadratic pressure term generate a source-like nonlinearity, while classical thermoviscous absorption is modeled via the stress tensor, leading to a diffusive term. For plane waves propagating in one dimension (e.g., along x), these yield the Westervelt equation in the form\frac{\partial^2 p}{\partial x^2} - \frac{1}{c^2} \frac{\partial^2 p}{\partial t^2} = \frac{\beta}{(1 + \alpha) \rho_0 c^4} \left( \frac{\partial p}{\partial t} \right)^2 + \delta \frac{\partial^3 p}{\partial x^2 \partial t},where \alpha accounts for a correction factor in the nonlinear term (often near unity for gases, related to specific heat ratios), \rho_0 is ambient density, c is sound speed, and \delta is the absorption coefficient derived from viscosity and thermal conductivity (typically \delta = \frac{2 \alpha_0 (2\pi f)^2}{\rho_0 c^3} for frequency-dependent classical absorption \alpha = \alpha_0 f^2). This form arises after combining the expanded equations and neglecting higher-order terms beyond the weakly nonlinear approximation.Solutions to the Westervelt equation are generally obtained via numerical methods, such as finite-difference time-domain schemes, due to its nonlinearity, which prevents closed-form analytical expressions for arbitrary initial conditions. For initial sinusoidal waves, perturbation methods expand the solution as a series in harmonics, p(x,t) = \Re \sum_{k=1}^\infty p_k(x) e^{i k \omega t}, where the fundamental p_1 satisfies the linear wave equation, and higher harmonics p_k (for k \geq 2) are driven by nonlinear coupling terms like \beta (k \omega p_k) \sum_{m=1}^{k-1} m \omega p_m (k-m) \omega p_{k-m}^*, enabling analysis of spectral evolution and shock formation. These approaches reveal progressive waveform steepening and harmonic generation, with absorption balancing nonlinearity to prevent singularities.
Burgers Equation
The Burgers equation provides a fundamental one-dimensional model for nonlinear acoustic wave propagation in dissipative media, where nonlinear effects cause waveform steepening that is counteracted by viscous and thermal diffusion. This equation captures the essential dynamics of finite-amplitude sound waves, particularly the formation of shocks, without the complexities of multidimensional geometry. It arises in contexts such as high-intensity ultrasound and sonic boom propagation, offering insights into energy dissipation and harmonicgeneration.[28]For unidirectional plane waves propagating along the z-direction, the Burgers equation can be derived from the Westervelt equation by assuming a slowly varying profile in the retarded time τ = t - z/c, where the pressure perturbation p(z, τ) satisfies\frac{\partial p}{\partial z} = \frac{\beta}{\rho c^3} p \frac{\partial p}{\partial \tau} + \frac{\delta}{2 c^3} \frac{\partial^2 p}{\partial \tau^2}.Here, β = 1 + B/(2A) is the coefficient of nonlinearity related to the equation of state (with B/A characterizing the medium's nonlinearity), ρ is the ambient density, c is the small-signal sound speed, and δ = (2π² α_0 f_0^2)/ω_0^2 is the diffusivity of sound incorporating classical absorption mechanisms, with α_0 the absorption coefficient at frequency f_0. This parabolic form approximates the hyperbolic Westervelt equation under the far-field, forward-propagating assumption.[29][28]In terms of particle velocity u (related to pressure via u ≈ p/(ρ c) for progressive waves), the equation adopts the standard Burgers form\frac{\partial u}{\partial t} + u \frac{\partial u}{\partial x} = \nu \frac{\partial^2 u}{\partial x^2},where x is the propagation coordinate, t is time, and ν is an effective kinematic viscosity incorporating acoustic absorption. This mirrors the original Burgers equation from fluid dynamics, which approximates the Navier-Stokes equations for turbulence by retaining only the nonlinear convective term and diffusion. The left side drives shock formation through steepening, while the right side smears discontinuities over a diffusion length scale.[28][29]Exact analytical solutions are enabled by the Cole-Hopf transformation, which maps the nonlinear PDE to the linear heat equation: let u = -2ν (∂φ/∂x)/φ, where φ satisfies ∂φ/∂t = ν ∂²φ/∂x². This yields solutions for stationary shock profiles, such as a tanh-shaped transition between two constant states, and for N-waves—bipolar pulses typical in acoustic shocks where an initial compression forms a sawtooth, evolves into a triangular profile, and eventually decays due to dissipation. These solutions highlight the balance between nonlinearity (quantified by a shock parameter σ = β u_0 / (c δ ω)) and diffusion.[29][28]The Burgers equation is valid for plane progressive waves in unbounded media, tubes, or collimated beams where diffraction is negligible, assuming the acoustic Reynolds number is high (indicating weak dissipation relative to nonlinearity) and the shock thickness exceeds the molecular mean free path. It breaks down in higher dimensions, where effects like beam spreading require extensions such as the KZK equation, and is limited to moderate amplitudes where higher-order nonlinearities remain small.[28][29]
KZK Equation
The Khokhlov–Zabolotskaya–Kuznetsov (KZK) equation provides a parabolic approximation for modeling the nonlinear propagation of finite-amplitude acoustic beams, incorporating effects of nonlinearity, thermoviscous absorption, and diffraction in three dimensions. Originally derived by Zabolotskaya and Khokhlov in 1969 for lossless media under the quasi-plane wave approximation, it was extended by Kuznetsov in 1970 to include dissipation, making it suitable for directed beams with slowly varying amplitude along the propagation axis. This equation extends the one-dimensional Burgers equation by adding a transverse diffraction term, essential for beams where lateral spreading influences wave evolution.[30][31]The derivation starts from the lossless Westervelt or Kuznetsov equation and applies a paraxial approximation, assuming high-frequency waves where the acoustic wavelength is much smaller than the beam width, small angular divergence (typically <10°), and often axisymmetric geometry for computational efficiency. The KZK equation takes the form:\frac{\partial p}{\partial z} = \frac{\beta}{\rho c^3} p \frac{\partial p}{\partial \tau} + \frac{\delta}{2 c^3} \frac{\partial^2 p}{\partial \tau^2} + \frac{1}{2 i k} \nabla_\perp^2 pwhere \beta is the nonlinearity parameter, \rho is density, c is sound speed, \delta is the diffusivity of sound, k = \omega / c is the wavenumber, \nabla_\perp^2 is the transverse Laplacian, and time derivatives are evaluated at the retarded time \tau = t - z/c. This form captures cumulative nonlinear distortion along the beam axis while treating diffraction as a slowly varying perturbation, often implemented by operator splitting in numerical solutions where nonlinear and absorption terms are handled in time domain and diffraction in frequency domain.[30][32]In applications to focused ultrasound, the KZK equation models harmonic generation and saturation, where nonlinear steepening produces higher harmonics that absorb more energy, limiting the peak pressure at the focus and broadening the beam; for example, measurements at around 2 MHz show fundamental amplitude reductions of about 20% beyond 50 MPa in water.[33] For parametric arrays, it predicts gain in directed low-frequency sound from nonlinear mixing of two high-frequency primaries, with beam patterns exhibiting sidelobe suppression up to 30 dB compared to linear sources.[34] Numerical solutions typically employ finite-difference time-domain (FDTD) methods, splitting the operator into nonlinear/absorption steps solved via fractional-step schemes and diffraction via angular spectrum propagation, enabling efficient 2D/3D simulations on grids with resolutions of 10–20 points per wavelength.[35]
Applications and Phenomena
Sonic Booms
Sonic booms arise from the supersonic flight of aircraft, where the pressure disturbances generated by the vehicle's volume and lift undergo nonlinear steepening during propagation, evolving into characteristic N-shaped waveforms known as N-waves.[36] This steepening occurs because higher-pressure regions in the initial pressure signature travel faster than lower-pressure ones due to the nonlinear dependence of sound speed on pressure, causing the waveform to compress at the front and expand at the rear until shocks form, typically resulting in a sharp rise, a linear decay, and a sharp fall in pressure.[36] Seminal work by Whitham in 1952 established the relationship between the aircraft's shape and the resulting N-wave overpressure and duration, providing a foundational framework for predicting these signatures in uniform media.[36]As N-waves propagate through the atmosphere, nonlinear effects lead to energy dissipation primarily through weak shocks, where viscosity and thermal conduction absorb energy, causing the shock amplitude to decay with distance traveled.[36] Atmospheric variations in temperature, wind, and density introduce additional complexities, such as caustics—regions of focused energy due to ray bending—and multipath propagation, where waves follow multiple trajectories to the ground, potentially amplifying local overpressures or distorting the waveform.[37] These effects are particularly pronounced over long distances, with overpressure typically decaying as the inverse square root of range in nonlinear regimes, though focusing can produce superbooms exceeding 10 pounds per square foot in extreme cases.[36]Mitigation of sonic booms focuses on shaping the near-field pressure signature through aircraft design to produce lower peak overpressures and smoother waveforms at the ground, such as flat-top or ramp-like profiles that reduce perceived loudness.[38] Numerical modeling employs Burgers-like equations, augmented with atmospheric absorption and nonlinearity terms, to simulate propagation and optimize designs; for instance, adjoint-based methods coupled with computational fluid dynamics have achieved reductions of up to 2.5 dB in A-weighted loudness by adjusting aircraft geometry.[38] These approaches, as detailed in Rallabhandi's 2011 work, enable inverse design for low-boom concepts by minimizing discrepancies between predicted and target ground signatures.[38] Recent efforts include NASA's X-59 QueSST experimental aircraft, which completed its maiden flight on October 28, 2025, and is designed to cruise at Mach 1.4 while producing a quiet "sonic thump" perceived as no louder than 75 dBA on the ground, potentially paving the way for regulatory changes to allow overland supersonic commercial flights.[39]The environmental impacts of sonic booms, particularly during the 1970s Concorde era, included structural vibrations and annoyance from overpressures around 1.94 pounds per square foot, leading to widespread complaints and restrictions on supersonic flight over land.[40] In response, the U.S. Federal Aviation Administration prohibited civil supersonic operations over populated areas in 1973 under 14 CFR §91.817, limiting flights to oceanic routes and imposing land use restrictions that effectively ended routine overland supersonic travel for commercial aircraft.[41] Tolerable overpressure levels were assessed at 1-1.5 pounds per square foot to avoid significant annoyance or minor damage, influencing global policies and ongoing research into quieter boom designs.[36]
Acoustic Levitation
Acoustic levitation relies on nonlinear acoustic radiation forces arising from the interaction of high-intensity sound waves with matter, enabling the suspension of small objects against gravity without physical contact. These forces emerge as second-order effects in the acoustic field, where the time-averaged momentum flux produces a net radiation pressure that counteracts gravitational forces. For small rigid or compressible particles much smaller than the acoustic wavelength, the force is described by the Gor'kov potential, which quantifies the potential energy based on the particle's compressibility and density relative to the surrounding medium, leading to stable trapping positions.[42] In the case of gas bubbles, the primary Bjerknes force governs the dynamics, resulting from the coupling between the bubble's radial oscillations and the pressure gradient in the sound field, often directing bubbles toward pressure antinodes.[43]The typical setup involves generating a standing wave between an ultrasonic transducer and a reflector, creating regions of pressure nodes and antinodes along the axis. Particles denser than the medium, such as solids in air, are trapped at pressure nodes where the radiation force balances gravity, while lighter objects like bubbles may position at antinodes. Nonlinear effects become prominent at high sound pressure levels (typically above 150 dB), where wave steepening generates higher harmonics that distort the waveform and enhance the effective radiation force by increasing the mean pressure and inducing acoustic streaming, which can stabilize or destabilize levitation depending on the configuration.[44] Ultrasonic frequencies (around 20-40 kHz) are commonly used to achieve the necessary intensity while minimizing audible noise and attenuation in air.[45]Experimental milestones in acoustic levitation began gaining traction in the 1970s with advancements in high-power ultrasonic transducers, enabling practical demonstrations for containerless processing. Early work at NASA's Jet Propulsion Laboratory demonstrated the levitation of liquid drops and solid particles in air using standing waves, paving the way for microgravity simulations on Earth.[46] By the late 1970s, systems had scaled to levitate millimeter-sized objects, such as polystyrene spheres and water droplets, with stable positions maintained for minutes, highlighting the role of nonlinear saturation to prevent excessive harmonic generation. Subsequent developments in the 1980s extended this to multiple particles and manipulation, but initial 1970s experiments established the feasibility for controlled environments.[47]Applications of acoustic levitation include contactless handling of sensitive materials in manufacturing, such as pharmaceuticals or electronics, where avoiding contamination is critical, and material processing in simulated microgravity conditions to study crystal growth or alloy formation without container-induced impurities. In space research, it facilitates non-invasive manipulation of samples aboard spacecraft, as demonstrated in orbital experiments. However, limitations arise from particle size constraints—the Gor'kov approximation holds only for diameters less than about 1/10th the wavelength (e.g., <0.3 mm at 40 kHz)—and acoustic intensity thresholds, where excessive levels (>160 dB) trigger unwanted nonlinear phenomena like cavitation in liquids or excessive heating, restricting levitation to small-scale, low-mass objects under 1 g.[48][49]
Ultrasonic Nonlinear Effects
Ultrasonic nonlinear effects occur in high-frequency sound waves typically ranging from 1 to 20 MHz, which are commonly used in diagnostic and therapeutic applications.[50] In this frequency domain, nonlinear propagation leads to waveformdistortion, where higher harmonics are generated due to the medium's nonlinear response, enhancing imaging contrast and resolution compared to linear methods.[51] The nonlinearity parameter B/A, which quantifies this distortion, varies across biological tissues, with values typically between 6 and 11; for example, liver tissue exhibits B/A ≈ 7.5, while fat has a higher value around 11.[52][53]Harmonic generation in tissue arises from cumulative nonlinear distortion as ultrasound waves propagate, converting fundamental frequency energy into higher harmonics that provide sharper images with reduced artifacts in medical imaging.[54] This process improves axial and lateral resolution by leveraging the harmonics produced primarily during propagation through attenuating media, rather than solely from transducer emissions.[51] In tissue harmonic imaging, the received signals at twice the fundamental frequency (second harmonics) suppress near-field clutter and side-lobe artifacts, enabling clearer visualization of structures like vessel walls.[55]Cavitation represents another key nonlinear effect, where ultrasound induces bubble formation and oscillation in liquids or tissues, categorized as stable or inertial based on bubble dynamics. Stable cavitation involves sustained, non-destructive oscillations around the bubble's resonant size, promoting gentle mixing or drug release without collapse.[56] In contrast, inertial cavitation features rapid bubble expansion followed by violent collapse due to inertial forces, generating localized high pressures and temperatures exceeding 5000 K.[57]Sonoluminescence emerges as a nonlinear emission during inertial collapse, where the extreme conditions inside the bubble produce broadband light pulses, serving as an indicator of cavitation intensity.[58] This phenomenon has been observed in single-bubble setups levitated in acoustic fields, highlighting the role of nonlinear bubble dynamics.[59]In inhomogeneous media such as biological tissues, nonlinear ultrasound propagation amplifies scattering and absorption, leading to enhanced waveform steepening and higher harmonic content.[60] Variations in tissue density and speed of sound cause focusing or defocusing, which interact with nonlinearity to increase attenuation at higher frequencies and alter beam profiles.[61] These effects are modeled using approaches like the Khokhlov–Zabolotskaya–Kuznetsov (KZK) equation for parabolic beam propagation in absorbing media.[60] Consequently, scattering from interfaces or inclusions is intensified, impacting imaging depth and contrast in heterogeneous environments like tumors or layered organs.[62]
Musical Acoustics
Nonlinear effects play a crucial role in the sound production of musical instruments, particularly in wind and string families, where deviations from linear wave propagation lead to distinctive timbres and dynamic behaviors. In reed instruments, such as clarinets and oboes, the reed's vibration exhibits strong nonlinearity due to its interaction with the air flow; as the blowing pressure increases, the reed's displacement becomes asymmetric, generating higher harmonics that enrich the spectrum beyond linear predictions. Similarly, in brass instruments like trumpets, shocks form in the air column at high amplitudes, producing sawtooth-like waveforms with steep pressure gradients; this nonlinear propagation steepens the wave profile, amplifying upper harmonics and contributing to the instrument's bright, aggressive timbre during loud playing. These effects arise from general wave distortion in nonlinear media, where finite amplitude waves accumulate energy in higher frequencies.[63]String instruments introduce nonlinearity through inharmonicity, where the stiffness of the string causes higher partials to deviate from integer multiples of the fundamental frequency, stretching them upward and altering the perceived pitch. This inharmonicity is particularly pronounced in pianos, where longitudinal string tension modulates transverse vibrations, generating additional partials via nonlinear mixing and enhancing the instrument's percussive quality.[64] In contrast, linear models assuming ideal harmonic series fail to capture these deviations, leading to inaccurate timbre simulations.[65]Organ pipes exemplify nonlinearities in flue instruments, where the air jet's interaction with the labium edge produces self-sustained oscillations influenced by cubic nonlinearities in the flow; linear models predict stable sinusoidal output, but nonlinear effects introduce mode interactions and bifurcations, resulting in richer spectra with even and odd harmonics depending on pipe geometry.[66] Historical studies from the 1980s on wind instruments, including clarinets and trumpets, highlighted these dynamics through early numerical models, revealing how nonlinear reed and lip vibrations limit stable oscillation ranges and affect playability at varying intensities.[67]In performance spaces, nonlinear absorption emerges at high sound pressure levels, where materials like wall finishes and seating exhibit amplitude-dependent damping, reducing higher harmonics more than fundamentals and altering reverberation tails. This leads to harmonic distortion in concert halls during forte passages, softening the overall brightness compared to linear absorption assumptions. Psychoacoustically, these nonlinear harmonics shape timbre perception, with listeners interpreting enriched spectra as warmth or brilliance, though excessive distortion at loud volumes can impose perceptual limits, causing fatigue or masking of subtle nuances in ensemble playing.[68][69]
Parametric Arrays
Parametric arrays exploit nonlinear acoustic interactions to generate highly directional beams of low-frequency audible sound from high-frequency ultrasonic carriers. The principle relies on self-demodulation, where two collimated ultrasound beams with closely spaced frequencies, say f_1 and f_2, interact through the medium's nonlinearity, producing a difference-frequency wave at f_d = |f_1 - f_2| that propagates as an audible signal. This process occurs because the nonlinear term in the wave equation mixes the primary waves, effectively creating a virtual end-fire array of sources along the beam axis, which enhances directivity far beyond what is achievable with conventional low-frequency transducers of similar aperture size.[70][71]Theoretical modeling of parametric arrays often draws from the Khokhlov–Zabolotskaya–Kuznetsov (KZK) equation, which accounts for nonlinearity, diffraction, and absorption in directive beams; solutions reveal that the audible secondary soundfield behaves as if radiated from a continuous virtual source distributed along the interaction length, yielding a directivity gain proportional to the beam's absorption length and carrier amplitude. Berktay's far-fieldapproximation further simplifies this by describing the demodulated pressure as an envelope-modulated pulse, enabling predictions of beam patterns and sidelobe levels for practical designs. This virtual source concept results in narrow beamwidths, on the order of degrees for audible frequencies, making parametric arrays ideal for applications requiring spatial audio control.[72]The concept was first proposed by Westervelt in 1963 for both transmitting and receiving applications, with Berktay extending it in 1965 to broadbandunderwater signaling through pulsed or modulated carriers. In air, commercial audio projectors like the Audio Spotlight, developed by F. Joseph Pompei in the late 1990s and commercialized by Holosonics in the early 2000s, use amplitude-modulated ultrasound to project intelligible speech or music over tens of meters with minimal spillover. Underwater, parametric arrays facilitate long-range communication and imaging due to reduced absorption relative to air, as demonstrated in naval experiments generating low-frequency signals from compact transducers.[70][71][73]Despite these advantages, parametric arrays face challenges including low acoustic-to-acoustic efficiency, typically below 1% in air owing to strong thermoviscous absorption of the ultrasonic carriers, which limits output sound pressure levels and requires high input powers. Distortion arises from unwanted higher-order harmonics and intermodulation products in the secondary field, necessitating preprocessing techniques like multitone modulation or equalization to achieve low total harmonic distortion (THD) below 10% for clear audio reproduction. Ongoing research addresses these through advanced array designs and materials to improve conversion efficiency and beam quality.[72][74][75]
Medical and Industrial Uses
In medical applications, nonlinear acoustics plays a pivotal role in high-intensity focused ultrasound (HIFU) therapy, where intense acoustic waves propagate nonlinearly to form shock fronts that enable precise tumor ablation without invasive surgery. Developed in the 1990s, early clinical trials demonstrated HIFU's efficacy for treating liver tumors by focusing ultrasound beams to achieve localized heating and mechanical disruption through nonlinear distortion, with initial human ablations reported using ultrasound guidance.[76] Similarly, extracorporeal shock wave lithotripsy (ESWL) for kidney stone fragmentation relies on nonlinear wave propagation to generate high-pressure shocks that induce cavitation bubbles, whose collapse erodes stone surfaces through mechanical stress.[77] These cavitation effects, stemming from nonlinear acoustic phenomena like bubble oscillation and jetting, enhance fragmentation efficiency while minimizing surrounding tissue damage.[78]Industrial uses of nonlinear acoustics include ultrasonic cleaning, where nonlinear bubble dynamics drive cavitation to remove contaminants from surfaces by generating microjets and shock waves from collapsing bubbles under high-amplitude fields.[79] In ultrasonic welding and soldering, nonlinear frictional heating at interfaces converts acoustic energy into localized thermal effects, enabling solid-state bonding of thermoplastics and metals without bulk melting; for instance, vibrations at 20-70 kHz produce interfacial temperatures exceeding 200°C through nonlinear energy dissipation.[80][81]Nonlinear enhancements further extend these applications, such as in ultrasound imaging where microbubble contrast agents exhibit nonlinear oscillations—producing harmonics and subharmonics—to improve vascular visualization and perfusion assessment over linear tissue echoes.[82] The acoustic nonlinearity parameter B/A, which quantifies a medium's propensity for waveformdistortion, aids tissuecharacterization by differentiating pathological states; for example, elevated B/A values in fatty liver tissue correlate with disease progression, enabling non-invasive diagnostics.[83][52]Safety standards mitigate risks from nonlinear effects, with the U.S. Food and Drug Administration (FDA) regulating diagnostic ultrasound via the mechanical index (MI), capped at 1.9 to limit cavitation-induced bioeffects from nonlinear bubble activity.[84] Therapeutic devices like HIFU systems must adhere to intensity thresholds, such as spatial-peak pulse-average intensity below 190 W/cm², to prevent excessive nonlinear propagation and unintended tissue damage.[85]