Fact-checked by Grok 2 weeks ago

Longitudinal wave

A longitudinal wave is a type of in which the particles of the medium oscillate parallel to the direction of wave propagation, resulting in alternating regions of (high particle ) and (low particle ). This particle motion distinguishes longitudinal waves from transverse waves, where oscillations occur perpendicular to the propagation direction. Unlike transverse waves, which require a medium capable of (such as solids), longitudinal waves can propagate through solids, liquids, and gases due to the medium's compressibility. Key characteristics of longitudinal waves include their dependence on the medium's elastic properties and density for propagation speed, typically expressed as v = f \lambda, where v is the wave speed, f is the frequency, and \lambda is the wavelength. The amplitude represents the maximum displacement from equilibrium, while the wavelength corresponds to the distance between consecutive compressions or rarefactions. Prominent examples include sound waves, which transmit pressure disturbances through air or other fluids as longitudinal pressure variations, and primary (P) waves generated during earthquakes, which travel fastest through Earth's interior by compressing and expanding rock and other materials. Other applications encompass ultrasound waves used in medical imaging and longitudinal vibrations in solids, such as those in coiled springs like a slinky.

Fundamentals

Definition

A is a in which the of particles in the medium occurs parallel to the direction of energy propagation. This contrasts with transverse waves, where is perpendicular to the propagation direction. In longitudinal waves, the disturbance creates a series of compressions and rarefactions along the path of travel, with compressions representing regions of increased particle density and rarefactions indicating decreased density. The deformation of the medium in a longitudinal wave can be visualized as alternating bunches and stretches in a coiled or as high- and low-pressure zones in a , where particles momentarily cluster tightly before spreading out. These patterns allow the wave to transmit through the medium without net of the particles from their original positions over time. The term "longitudinal wave" originated in the early 19th century, with its first known use recorded in 1813 amid developing theories of wave motion in physics. Physicists such as John William Strutt, 3rd Baron Rayleigh, furthered the conceptual framework in the late 19th century through studies of sound propagation, as detailed in his influential two-volume work The Theory of Sound (1877–1878), which analyzed longitudinal vibrations in various media. Understanding longitudinal waves requires familiarity with fundamental wave properties, including (the distance between consecutive compressions or rarefactions), (the number of wave cycles per unit time), (the maximum displacement of particles from equilibrium), and (the position of a point within the wave cycle).

Nomenclature

The term "longitudinal" originates from the Latin longitudo, meaning "length," reflecting the parallel alignment of particle oscillations with the direction of wave propagation. In the context of wave mechanics, key terminology includes , denoting regions where medium particles are densely packed, increasing local or , and , referring to regions of expanded spacing with reduced or . Longitudinal describes the fractional change in length along the propagation direction, quantifying the deformation in the medium. Standard symbols in physics for longitudinal include \xi(x, t) to represent the longitudinal of particles from their , where x is the position coordinate and t is time. The associated longitudinal is conventionally expressed as \partial \xi / \partial x, capturing the spatial of . In fluid media, the pressure variation induced by the wave is denoted by p, often related to the and . Conventions for wave parameters in longitudinal contexts follow general wave physics, with k as the wave number (k = 2\pi / \lambda, where \lambda is ) describing spatial periodicity, and \omega as the (\omega = 2\pi f, where f is ) characterizing temporal . Longitudinal waves differ from vector-based transverse waves by being irrotational and representable via a in formulations, where the field derives from the of a scalar .

Mathematical Formulation

Displacement and Particle Motion

In a longitudinal wave, the of particles from their positions is to the of wave . The particle \xi(x, t) for a traveling in the positive x-direction can be expressed as \xi(x, t) = A \cos(kx - \omega t), where A is the , k = 2\pi / \lambda is the wave number, and \omega = 2\pi f is the . This oscillatory motion results in particles moving back and forth along the axis, creating regions of and without any transverse component. The variation in density arises directly from the spatial gradient of the displacement. The local density \rho is related to the equilibrium density \rho_0 by \rho = \rho_0 \left(1 - \frac{\partial \xi}{\partial x}\right), where the term \frac{\partial \xi}{\partial x} represents . Regions of occur when \frac{\partial \xi}{\partial x} < 0, leading to increased density, while rarefaction happens when \frac{\partial \xi}{\partial x} > 0, resulting in decreased density. For the displacement function above, \frac{\partial \xi}{\partial x} = -k A \sin(kx - \omega t), which oscillates between positive and negative values, driving these density fluctuations. The velocity of the particles is the time derivative of the displacement, given by u(x, t) = \frac{\partial \xi}{\partial t} = -\omega A \sin(kx - \omega t). The acceleration follows as a(x, t) = \frac{\partial^2 \xi}{\partial t^2} = -\omega^2 A \cos(kx - \omega t). This acceleration links to the initiation and propagation of the wave through Newton's second law, where the net force on a particle element, typically from pressure gradients, equals mass times acceleration, enabling the wave to sustain oscillatory motion. In fluids, longitudinal waves involve purely irrotational motion, meaning the velocity field satisfies \nabla \times \mathbf{u} = 0, with no vorticity generated due to the absence of shear stresses. In solids, however, longitudinal waves correspond to shear-free compression, where the displacement is dilatational without shear deformation, propagating via the material's bulk modulus and density, distinct from transverse shear waves that solids also support.

Wave Equation

The wave equation for longitudinal waves describes the propagation of particle displacements or pressure variations in a medium. In one dimension, it is derived by considering the dynamics of a small element of the medium, combining Newton's second law with the constitutive relations from elasticity theory. For solids, such as a thin rod, the longitudinal \xi(x, t) along the propagation x produces a \partial \xi / \partial x. By , the resulting is \sigma = E \partial \xi / \partial x, where E is the . The net on an infinitesimal element of length \Delta x and cross-sectional area A is then A (\partial \sigma / \partial x) \Delta x = A E (\partial^2 \xi / \partial x^2) \Delta x. Applying Newton's second law to this element of mass \rho A \Delta x (with \rho the ) yields \rho A \Delta x (\partial^2 \xi / \partial t^2) = A E (\partial^2 \xi / \partial x^2) \Delta x, simplifying in the limit \Delta x \to 0 to the one-dimensional : \frac{\partial^2 \xi}{\partial t^2} = c^2 \frac{\partial^2 \xi}{\partial x^2}, where the longitudinal wave speed is c = \sqrt{E / \rho}. For fluids, the derivation similarly relies on mass conservation () and the linearized , with the B relating perturbations to changes via dp = (B / \rho) d\rho. In one dimension, the for small perturbations is \partial \tilde{\rho} / \partial t + \rho_0 \partial v / \partial x = 0, where \tilde{\rho} is the perturbation, \rho_0 the equilibrium , and v the . The linearized Euler equation gives \rho_0 \partial v / \partial t = -\partial \tilde{p} / \partial x, with perturbation \tilde{p} = (B / \rho_0) \tilde{\rho}. Differentiating the with respect to time and substituting from the Euler , followed by using the relation, leads to the wave for : \frac{\partial^2 \tilde{p}}{\partial t^2} = c^2 \frac{\partial^2 \tilde{p}}{\partial x^2}, with wave speed c = \sqrt{B / \rho_0}. This one-dimensional form assumes propagation along a single axis and neglects transverse effects. To extend to three dimensions, longitudinal waves are modeled as irrotational, with \mathbf{v} = \nabla \phi using a \phi(\mathbf{r}, t), and \boldsymbol{\xi} = \nabla \phi. The potential then satisfies the three-dimensional \partial^2 \phi / \partial t^2 = c^2 \nabla^2 \phi, or equivalently for in fluids, \partial^2 \tilde{p} / \partial t^2 = c^2 \nabla^2 \tilde{p}. For plane waves in an unbounded medium, boundary conditions typically involve no incoming waves from , ensuring outgoing or solutions. The general solution in one dimension consists of traveling waves, such as \xi(x, t) = f(x - c t) + g(x + c t), where f and g are arbitrary functions representing right- and left-propagating components, respectively. In three dimensions, solutions take the form \phi = A e^{i(\mathbf{k} \cdot \mathbf{r} - \omega t)}, with \omega = c k for dispersionless .

Propagation Characteristics

Speed in Media

The speed of longitudinal waves in isotropic elastic solids is given by the formula c_L = \sqrt{\frac{\lambda + 2\mu}{\rho}}, where \lambda and \mu are the Lamé constants representing the material's and rigidity, respectively, and \rho is the . This expression arises from the wave equation for dilatational motion, where is parallel to the , involving no deformation; the effective is thus \lambda + 2\mu, combining volumetric and deviatoric responses. In contrast, the speed of transverse waves in the same medium is c_T = \sqrt{\frac{\mu}{\rho}}, which depends solely on and is typically lower than c_L since \lambda + 2\mu > \mu. In liquids, the speed is given by c = \sqrt{\frac{B}{\rho}}, where B is the adiabatic and \rho is the . For example, in at 20°C, this speed is approximately 1480 m/s. In ideal gases, the longitudinal wave speed, equivalent to the , is c = \sqrt{\frac{\gamma P}{\rho}}, where \gamma is the adiabatic index (ratio of specific heats), P is the , and \rho is the ; this derives from adiabatic compression assumptions in the fluid's ./Book%3A_University_Physics_I_-Mechanics_Sound_Oscillations_and_Waves(OpenStax)/17%3A_Sound/17.03%3A_Speed_of_Sound) The propagation speed is influenced by material properties such as (which affects molecular spacing and elasticity), (inversely proportional in the formulas), and elastic moduli (directly scaling the square root term); for instance, higher generally increase speed in gases due to enhanced molecular activity, while in , variations from dominate./Book%3A_University_Physics_I_-Mechanics_Sound_Oscillations_and_Waves(OpenStax)/17%3A_Sound/17.03%3A_Speed_of_Sound) Representative values illustrate these differences: in air at 20°C and standard , the speed is approximately 343 m/s, reflecting low and , whereas in , it reaches about 5960 m/s due to high elastic moduli and moderate ./Book%3A_University_Physics_I_-Mechanics_Sound_Oscillations_and_Waves(OpenStax)/17%3A_Sound/17.03%3A_Speed_of_Sound) In anisotropic media, such as crystals, the longitudinal wave speed varies with propagation direction because elastic properties are tensorial, leading to direction-dependent velocities without a simple scalar formula.

Attenuation Mechanisms

Attenuation in longitudinal waves describes the progressive reduction in wave amplitude and energy as it propagates through a medium, primarily due to energy dissipation into other forms such as heat. The displacement amplitude A decays exponentially with propagation distance x according to the relation A = A_0 e^{-\alpha x}, where \alpha is the attenuation coefficient with units of inverse length (nepers per meter). Consequently, the wave intensity I, which is proportional to the square of the amplitude, follows I = I_0 e^{-2\alpha x}, reflecting the quadratic dependence of energy on displacement. This exponential decay quantifies how longitudinal waves, such as sound or seismic P-waves, weaken over distance in real media, distinguishing them from ideal lossless propagation. Attenuation mechanisms for longitudinal waves are categorized into classical and quantum types, each dominating under specific conditions like or material state. Classical mechanisms encompass viscous drag, stemming from molecular friction that opposes particle motion, and , which arises from heat diffusion across compressions and rarefactions in the wave. These processes, prevalent in fluids and higher-temperature solids, irreversibly convert acoustic energy into via irreversible . In contrast, quantum mechanisms, particularly relevant in crystalline solids at low temperatures, involve , where quantized vibrations (s) collide, leading to anharmonic interactions that redistribute and dissipate wave energy among phonon modes. Phonon-phonon scattering, for instance, follows and becomes the primary pathway when classical effects are negligible. The magnitude of \alpha typically varies with the \omega of the wave, influencing differently across media. In viscous fluids, classical mechanisms yield \alpha \propto \omega^2, as derived from the Stokes-Kirchhoff formula, where higher frequencies amplify the relative motion and energy loss due to and conduction. This dependence is evident in liquids like , where increases sharply with frequency, limiting high-frequency penetration. In certain , especially those exhibiting viscoelastic or anelastic behavior, \alpha \propto \omega, reflecting linear frequency scaling from mechanisms like internal friction or specific interactions. Such dependence arises in non-linear viscoelastic media, where wave distortion enhances dissipation proportionally to frequency. To experimentally determine \alpha, techniques like the pulse-echo method are employed, involving the transmission of short ultrasonic pulses into the medium and analysis of echoed signals from boundaries. In this approach, the amplitude ratio between successive echoes, after correcting for reflection coefficients and diffraction effects, yields \alpha via logarithmic fitting of the decay. This method is widely used for longitudinal waves in both fluids and solids, providing frequency-resolved measurements that validate theoretical models of attenuation.

Examples in Acoustics and Mechanics

Sound Waves

Sound waves represent a primary example of longitudinal waves in gaseous and liquid media, manifesting as propagating disturbances that cause alternating regions of and , thereby producing oscillations in and . In air, these waves typically involve small-amplitude variations superimposed on the ambient , enabling the transmission of audible frequencies from approximately 20 Hz to 20 kHz. The mathematical description of the pressure variation in a plane progressive sound wave is given by p(x,t) = p_0 + \Delta p \cos(kx - \omega t), where p_0 is the equilibrium pressure, \Delta p is the amplitude of the pressure fluctuation, k = 2\pi / \lambda is the wavenumber, and \omega = 2\pi f is the angular frequency. The pressure amplitude \Delta p relates to the maximum particle displacement A through \Delta p = \rho c \omega A, with \rho denoting the medium's density and c the speed of sound; this connection stems from the acoustic impedance Z = \rho c, which quantifies the medium's resistance to the oscillatory flow induced by the wave. The in air derives from the adiabatic nature of compressions and rarefactions in the wave, where air parcels exchange no heat with their surroundings. For an ideal diatomic gas like air, the speed is expressed as c = \sqrt{\frac{\gamma R T}{M}}, with \gamma = 1.4 as the ratio of specific heats, R the , T the , and M the (approximately 0.029 kg/ for dry air). This formulation arises by equating the restoring force from gradients to the inertial response of elements, yielding the wave speed as the square root of the adiabatic divided by . strongly influences c, increasing it by roughly 0.6 m/s per ; at 20°C (293 K), c \approx 343 m/s for dry air. modifies this slightly by lowering the effective M due to vapor's lower molecular weight (18 g/mol versus 29 g/mol for dry air), raising c by 0.1% to 0.6% depending on relative levels up to 100% at standard conditions. In air, sound waves experience attenuation mainly from classical mechanisms involving and , which dissipate energy as , particularly prominent at higher frequencies. The attenuation coefficient \alpha (in nepers per meter) approximates \alpha \approx \frac{\omega^2}{2 \rho c^3} \left( \frac{4}{3} \eta + \frac{(\gamma - 1) \kappa}{C_p} \right), where \eta is the dynamic , \kappa the thermal conductivity, and C_p the at constant pressure; the intensity decays as e^{-2\alpha x}. This quadratic frequency dependence makes negligible below 1 kHz but significant above 5 kHz, where it limits over distances beyond a few hundred meters for typical audible sounds. Nonlinear effects emerge in high-amplitude sound waves due to the pressure-dependent variation in local propagation speed, causing compressive regions to accelerate relative to rarefactions and leading to steepening. Nonlinear steepening becomes evident in intense acoustic sources like explosions or loudspeakers at over 140 SPL (corresponding to peak particle velocities of about 0.7 m/s in air). Full shock waves form when the propagation distance exceeds the shock formation distance; for typical audible waves, this requires particle velocities around c/20 (approximately 17 m/s, or about 168 SPL) for distances comparable to one , resulting in a discontinuous sawtooth profile characterized by a sharp rise and linear decay, accompanied by higher generation and increased .

Seismic P-Waves

Seismic P-waves, also known as primary waves, are the fastest type of body wave generated during earthquakes, traveling through the Earth's interior at speeds typically ranging from 5 to 7.5 km/s in the crust. These propagate by alternating and of the medium, causing particles to oscillate to the direction of wave travel, and they can traverse solids, liquids, and gases alike. Unlike shear waves, P-waves do not require rigidity in the medium, allowing them to penetrate the outer core while refracting at boundaries such as the core-mantle . The velocity of P-waves varies with depth due to changes in the Earth's material properties, generally increasing from the crust into as enhances the moduli relative to . In the , speeds reach 8 to 13 km/s, reflecting higher and moduli under greater lithostatic . This relationship is captured by the formula v_p = \sqrt{\frac{K + \frac{4}{3} G}{\rho}} where v_p is the P-wave speed, K is the (measuring to compression), G is the (measuring to shear deformation), and \rho is the of the medium. P-waves are detected by seismographs, which record ground vibrations as differential motion between a stationary mass and the moving , often capturing or displacement components to identify the initial compressional arrivals. Arrival time differences from multiple stations enable location and, through , the construction of three-dimensional velocity models that map discontinuities like the Mohorovičić (Moho) boundary and mantle transitions. Attenuation of P-waves in the Earth arises primarily from anelasticity, where energy is dissipated as through mechanisms such as viscoelastic relaxation, particularly in regions of that reduce the quality factor Q. The Q factor, defined as Q = 2\pi E / \Delta E (with E as peak stored energy and \Delta E as energy lost per cycle), quantifies this low-loss propagation; values are low (e.g., Q \approx 20-90) in the beneath mid-ocean ridges due to , increasing to hundreds in the deeper . This frequency-dependent dissipation provides insights into , volatile content, and melt distribution within the and .

Specialized Contexts

Pressure Waves in Fluids

Pressure waves in fluids are longitudinal disturbances propagated by variations in that induce corresponding changes, driving particle motion parallel to the direction of wave propagation. These waves arise in both gases and liquids, often as transient phenomena rather than periodic oscillations, and are governed by the principles of compressible . In irrotational flow, the velocity field can be expressed as \mathbf{v} = \nabla \phi, where \phi is the , leading to the Euler equation \nabla p = -\rho \frac{\partial \mathbf{v}}{\partial t} that describes the balancing the fluid's . Such waves are commonly generated by sudden piston motions in confined fluids or by explosive events that rapidly compress surrounding media. For instance, a abruptly advancing into a column creates a compression front that travels outward, satisfying the irrotational Euler equation and producing a propagating pulse. Explosions, similarly, initiate strong gradients through rapid release, forming blast waves where the initial shock front evolves according to the same foundational equations. The propagation speed c of these pressure waves in a compressible fluid is derived from the equation of state and given by c = \sqrt{\frac{dp}{d\rho}}, where dp and d\rho represent infinitesimal changes in and density, respectively. This speed reflects the medium's ; in ideal incompressible fluids, where density remains constant (d\rho = 0), the wave speed approaches , implying instantaneous transmission. waves in fluids constitute a specific, often subset of these disturbances, propagating at the same characteristic speed under small-amplitude conditions. In practical applications, pressure waves manifest as hydraulic shocks in pipelines, known as water hammer, where sudden valve closures generate pressure surges calculated by the Joukowsky equation \Delta p = \rho c \Delta v, with \rho as fluid density, c as wave speed, and \Delta v as velocity change. These surges can reach magnitudes sufficient to damage infrastructure, emphasizing the need for gradual flow control. Blast waves from explosions represent another key application, featuring strong shocks where post-shock pressures and densities jump discontinuously across the front, leading to rapid energy dissipation in the fluid. At interfaces between dissimilar fluids, pressure waves exhibit reflection and transmission governed by continuity of pressure and the mismatch in acoustic impedance Z = \rho c. Pressure remains continuous across the boundary, ensuring no abrupt discontinuity, while the particle velocity experiences a jump proportional to the impedance difference, with the reflection coefficient for pressure amplitude given by R = \frac{Z_2 - Z_1}{Z_2 + Z_1}. This behavior determines the partitioning of wave energy, with significant reflection occurring for large impedance contrasts, such as air-water boundaries.

Longitudinal Waves in Electromagnetics

Longitudinal electromagnetic waves, characterized by an component parallel to the wave \mathbf{k}, are rare compared to the ubiquitous transverse electromagnetic waves. In free space and non-dispersive media, such propagating longitudinal waves are forbidden by , \nabla \cdot \mathbf{E} = 0 in charge-free regions, as a longitudinal \mathbf{E}-field would imply nonzero divergence, violating the condition for far-field radiation. This constraint arises from in , where plane-wave solutions require \mathbf{E} \perp \mathbf{k} and \mathbf{B} \perp \mathbf{k} for energy propagation via the . Longitudinal components can nevertheless appear in specialized settings. In near-field regions close to sources, such as antennas, the reactive fields include a significant longitudinal \mathbf{E}-component parallel to the propagation direction, decaying rapidly with distance and not contributing to far-field radiation. Within waveguides, transverse magnetic (TM) modes support a longitudinal electric field E_z along the guide's axis (direction of \mathbf{k}), enabling guided propagation with partial longitudinal character, though the overall mode remains hybrid. Engineered metamaterials further enable true propagating longitudinal waves by breaking natural symmetries, supporting them across broad frequency bands up to Bragg resonances through tailored permittivity and permeability. A key realm for longitudinal electromagnetic waves is in plasmas, where Langmuir waves—electrostatic oscillations of —exhibit purely longitudinal with \mathbf{E} \parallel \mathbf{k}. These arise from charge bunching, consistent with allowing \nabla \cdot \mathbf{E} = \rho / \epsilon_0 in the presence of plasma density perturbations. For warm electrons, the is \omega^2 = \omega_p^2 + 3 v_{\mathrm{th}}^2 k^2, where \omega_p = \sqrt{n_e e^2 / \epsilon_0 m_e} is the electron plasma frequency, v_{\mathrm{th}} = \sqrt{k_B T_e / m_e} is the thermal velocity, and the factor of 3 accounts for three-dimensional isotropy. This relation shows weak dispersion, with phase velocities much greater than the thermal velocity (superluminal for sufficiently small k), distinguishing these waves from non-propagating cold-plasma oscillations at \omega = \omega_p. Langmuir waves find critical applications in plasma diagnostics and particle acceleration. In diagnostics, Langmuir probes excite these waves, and measuring their frequency directly yields the local electron density n_e via \omega_p, providing a standard tool for characterizing plasma parameters without invasive perturbations. In wakefield acceleration, intense laser or particle beams drive large-amplitude Langmuir-like wakes in underdense plasmas, generating gigavolt-per-meter electric fields for compact electron acceleration to GeV energies, as pioneered in the laser-wakefield concept. As of 2025, experiments have achieved acceleration of high-charge electron bunches to energies exceeding 10 GeV in compact, meter-scale setups. These waves propagate at speeds near c, enabling relativistic particle injection and staging for future high-energy colliders.

References

  1. [1]
    Transverse and Longitudinal Waves - HyperPhysics
    Transverse waves have displacement perpendicular to propagation, while longitudinal waves have displacement parallel to propagation. Sound waves in air are  ...
  2. [2]
    Longitudinal Wave Demo - UCSC Physics Demonstration Room
    Sound waves are a great example of a longitudinal wave, as they move through a medium of particles creating compressions and rarefactions. The reason there is ...
  3. [3]
    Waves - Physics
    Mar 1, 1999 · In a longitudinal wave, such as a sound wave, the particles oscillate along the direction of motion of the wave. Surface waves, such as water ...
  4. [4]
    16.1 What Are Waves? - Front Matter
    A longitudinal wave is a wave in which the displacement of the medium is parallel to the direction that the disturbance travels. The figure below shows a wave ...
  5. [5]
    Wave Motion in Mechanical Medium - Graduate Program in Acoustics
    In a longitudinal wave the particle displacement is parallel to the direction of wave propagation. The animation at right shows a one-dimensional ...Missing: definition | Show results with:definition
  6. [6]
    [PDF] The pre-history of 20th century acoustics: the legacy of Lord Rayleigh
    Jan 20, 2024 · The talk will review Rayleigh's life and key contri- butions, which covered structural vibration as well as air- borne acoustics.
  7. [7]
    CHEM 101 - Electromagnetic radiation and waves
    May 10, 2024 · All propagating (traveling) waves have four basic characteristics: speed, frequency, wavelength, and amplitude.
  8. [8]
    Longitudinal - Etymology, Origin & Meaning
    longitudinal(adj.) ... late 14c., from Medieval Latin longitudinalis, from Latin longitudo (see longitude). also from late 14c.
  9. [9]
    Longitudinal Wave - an overview | ScienceDirect Topics
    A longitudinal wave is defined as a type of wave in which the displacement of particles occurs in the same direction as the wave motion, characterized by ...Missing: origin | Show results with:origin
  10. [10]
  11. [11]
    [PDF] Chapter 5 – The Acoustic Wave Equation and Simple Solutions
    Notice that the condensation, , is the fractional change in density. This equation tells us that if the displacement, , varies with x then there is a density ...
  12. [12]
    Fluid Dynamics Course Notes - AstroBaki - CASPER
    Apr 25, 2017 · Sound Waves\. Sound waves are important in astrophysics because they provide the principal mechanism by which disturbances propagate in fluids.<|control11|><|separator|>
  13. [13]
    [PDF] Waves in an Isotropic Elastic Solid - Columbia University
    In an isotropic elastic solid, P-waves (longitudinal) travel at c = (λ+2µ)/ρ, and S-waves (transverse) at c = µ/ρ.
  14. [14]
    [PDF] 3. Waves in solids
    Mar 30, 2021 · • Hooke's law: −Kx. • Forces shall balance each other: • Mẍ ... Derivation of longitudinal wave equations (I). •. No details provided ...
  15. [15]
    Sound Waves in Fluids
    ... fluid velocity produced by the wave, and. $\displaystyle K = \rho\,\frac{dp}{d, (7.194). is the bulk modulus. The bulk modulus is a quantity with the units of ...
  16. [16]
    Simple problems - 4.4 Dynamic elasticity
    So there are two types of plane wave in an isotropic solid. The S-wave travels at speed , and material particles are displaced perpendicular to the direction of ...
  17. [17]
    Solids and Metals - Speed of Sound - The Engineering ToolBox
    Speed of sound in normal air is 343 m/s. In water the speed of sound is 1433 m/s. Speed of Sound in some common solids are indicated in the table below. Note ...
  18. [18]
    Longitudinal and transverse waves in anisotropic elastic materials
    Aug 9, 2025 · In cubic crystals, longitudinal velocities are mainly determined by C 11 and the combination (C 12 + 2C 44 ), and only have a slight ...
  19. [19]
    Attenuation of Waves - Nondestructive Evaluation Physics : Waves
    The quantity α \alpha is the attenuation coefficient of the wave traveling in the z-direction. The dimensions of α \alpha are nepers/length, where a neper is a ...
  20. [20]
    Bulk viscosity and compressibility measurement using acoustic ...
    Mar 31, 2009 · Viscosity and thermal conduction are the two mechanisms that cause attenuation when ultrasound propagates through a homogeneous media and ...
  21. [21]
    Acoustic Attenuation in Dielectric Solids | Phys. Rev.
    This treatment is extended to the case when the phonon gas is driven by an acoustic wave. An expression for the acoustic attenuation 𝛤 is obtained in the form 𝛤 ...
  22. [22]
    [PDF] Chapter 8 – Absorption and Attenuation of Sound
    The first two absorption processes (viscosity and thermal conductivity) follow classical absorption processes, which we will consider in a general form ...Missing: longitudinal drag
  23. [23]
    The attenuation of longitudinal waves in non-linear viscoelastic media
    It is shown that the degree of attenuation of a longitudinal wave in such media is a linear function of the frequency.
  24. [24]
    Measurement of ultrasound speed and attenuation coefficient of ...
    Measurement of ultrasound speed and attenuation coefficient of brain phantom using pulse echo and through transmission method. Abstract: This paper presents ...
  25. [25]
    Pulse-echo Technique to Compensate for Laminate Membrane ...
    May 25, 2022 · Accurately measuring the attenuation coefficient (AC) of reference phantoms is critical in clinical applications of quantitative ultrasound.
  26. [26]
    Sound Waves - Galileo
    From F = ma to the Wave Equation​​ Having found how the local pressure variation relates to s(x,t), we're ready to derive the wave equation from F=ma for a thin ...
  27. [27]
    17.2 Speed of Sound – University Physics Volume 1
    The equation for the speed of sound in air v = γ R T M can be simplified to give the equation for the speed of sound in air as a function of absolute ...
  28. [28]
    Acoustics Chapter One: Speed of Sound
    The speed of sound is directly influenced by both the medium through which it travels and the factors affecting the medium, primarily temperature and humidity ...
  29. [29]
    Absorption and Attenuation of Sound in Air
    Mar 22, 2016 · This thermal-viscous classical absorption depends on the square root of temperature, and the square of the frequency. relaxation processes --- ...
  30. [30]
    [PDF] Nonlinear Acoustic Propagation
    Consider a 1 kHz source in air at 1 atmosphere and 20˚C at a SPL = 100 dB. At what distance is a fully-developed shock wave formed? What about SPL = 120 dB and ...
  31. [31]
    Seismic Waves and Earth's Interior
    They typically travel at speeds between ~1 and ~14 km/sec. The slower values corresponds to a P-wave traveling in water, the higher number represents the P-wave ...Using P And S-Waves To... · Seismic Wave Propagation · P-Waves In Earth
  32. [32]
    Seismic Shadow Zone: Basic Introduction - IRIS
    The seismic shadows are the effect of seismic waves striking the core-mantle boundary. P and S waves radiate spherically away from an earthquake's hypocenter.
  33. [33]
    9.1 Understanding Earth Through Seismology – Physical Geology
    Mantle rock is generally denser and stronger than crustal rock and both P- and S-waves travel faster through the mantle than they do through the crust.
  34. [34]
    How fast do seismic waves travel and what controls this? - Geometrics
    The higher the shear modulus, the higher the s-wave velocity. Mathematically,. where. K = bulk modulus. µ = shear modulus. ρ = density. Note that Vp depends on ...
  35. [35]
    Seismographs - Keeping Track of Earthquakes - USGS.gov
    This combination of instruments tells a seismologist the general direction of the seismic wave source, the magnitude at its source, and the character of the ...Missing: tomography | Show results with:tomography
  36. [36]
    Seismic Tomography - EarthScope Consortium
    Jul 23, 2025 · One powerful technique is called Seismic Tomography, which works much like a CT scan of the Earth. Just as doctors take X-rays from different ...Missing: seismographs | Show results with:seismographs
  37. [37]
    [PDF] 1.25 Deep Earth Structure: Q of the Earth from Crust to Core
    Because of its strong dependence on temperature, partial melting, and water content, mapping anelastic attenuation in the Earth has the potential to provide ...
  38. [38]
    Euler Equations
    On this slide we have two versions of the Euler Equations which describe how the velocity, pressure and density of a moving fluid are related.
  39. [39]
    [PDF] A Physical Introduction to Fluid Mechanics
    Dec 19, 2013 · pressure disturbances caused by the piston motion are not small, shock waves will appear. A shock wave is a very thin region where the ...
  40. [40]
    Blast Wave - an overview | ScienceDirect Topics
    A blast wave is defined as a shock wave that propagates through an explosive mixture at supersonic velocity, resulting from the almost instantaneous ...
  41. [41]
    Speed of Sound - Equations - The Engineering ToolBox
    c = (dp / dρ)1/2 (1). where. c = speed of sound (m/s, ft/s). dp = change in pressure (Pa, psi). dρ = change in density (kg/m3, lb/ft3). Speed of Sound in Gases ...
  42. [42]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics
    Dec 15, 2021 · (b) The speed of sound in air (k = 1.4,R = 287 J kg 1 K 1) at standard conditions (T = 288 K) is 340 m s 1 (= 1115 ft s 1 ⇡ 1/5 mi s 1).
  43. [43]
    Water Hammer Impulse Surge Pressure - LMNO Engineering
    ΔP = ρ c ΔV. The equation for wave speed, c, during water hammer is based on mass conservation and allows the pipe wall material to expand (Wylie, 1993, p. 6 ...
  44. [44]
    Instantaneous Water Hammer Equation - Waterhammer.com
    The instantaneous water hammer equation is defined as: ΔP = -ρaΔV. Where. ΔP= pressure rise at the location of the velocity change. ρ = fluid density. a = wave ...
  45. [45]
    Blast wave kinematics: theory, experiments, and applications
    Jul 25, 2022 · In a later stage, a blast wave will decay and the strong-shock approximation will no longer be valid (and in the early stages of a chemical ...
  46. [46]
    Acoustic Impedance - an overview | ScienceDirect Topics
    To be more exact, reflection occurs where there is a difference of acoustic impedance (Z) between the media (Fig. 1.12A). The impedance is a measure of how ...
  47. [47]
    [PDF] Lecture 9: Reflection, Transmission and Impedance
    Matching the impedances of two different wave carrying media is of critical importance in electrical engineering. Say one wishes to drive an antenna, such ...
  48. [48]
    [PDF] Electromagnetic waves in plasmas and conductors
    Nov 11, 2024 · This note describes electromagnetic waves in plasmas ... The longitudinal waves are also known as plasma oscillations or Langmuir waves.
  49. [49]
    Experimental estimation of the longitudinal component of a highly ...
    Sep 9, 2021 · This paper describes a method that allows us to estimate and visualize the longitudinal component of the field in a relatively simple way.
  50. [50]
    Longitudinal electromagnetic waves with extremely short wavelength
    It is shown that the metamaterial supports longitudinal waves at an extremely wide frequency band from very low frequencies up to the Bragg resonances of the ...Abstract · Article Text
  51. [51]
    [PDF] Chapter 5 Electromagnetic Waves in Plasmas
    These have just the same shape except the electron plasma waves have much larger vertical scale: On the EM wave scale, the plasma wave curve is nearly vertical.
  52. [52]
    Laser Electron Accelerator | Phys. Rev. Lett.
    Jul 23, 1979 · Laser Electron Accelerator. T. Tajima and J. M. Dawson. Department of Physics, University of California, Los Angeles, California 90024. PDF ...Missing: original | Show results with:original