Fact-checked by Grok 2 weeks ago

Penetration depth

Penetration depth, also known as skin depth in , is the distance into a conducting or lossy material over which the amplitude of an electromagnetic wave or decreases to $1/e (approximately 37%) of its surface value. This phenomenon arises due to the skin effect, where alternating currents tend to concentrate near the surface of conductors, leading to of the inside the material. In electromagnetic theory, the penetration depth \delta for good conductors is given by the formula \delta \approx \sqrt{\frac{2}{\omega \mu \sigma}}, where \omega = 2\pi f is the , \mu is the magnetic permeability, and \sigma is the of the material. This depth decreases with increasing frequency, conductivity, and permeability, meaning high-frequency signals penetrate less deeply—typically on the order of micrometers for radio frequencies in metals like . Over one skin depth, the power density attenuates to about 13.5% ($1/e^2) of its surface value, which has critical implications for applications such as , where materials like aluminum effectively block high-frequency waves but are less efficient at low frequencies. The extends to other domains, notably , where the London penetration depth (\lambda_L) describes the over which an external penetrates into a superconductor before being expelled by the . In this context, \lambda_L = \sqrt{\frac{m}{\mu_0 n_s e^2}}, with m the , e the , and n_s the density of superconducting electrons, and \mu_0 as the ; typical values range from 10–100 nm for materials like aluminum and lead. This penetration arises from the screening currents induced at the surface, enabling perfect in type-I superconductors while influencing vortex structures in type-II ones. Beyond , penetration depth appears in fields like and , denoting the distance a thermal wave or diffusing substance travels before its intensity drops significantly, often analogous to the $1/[e](/page/E!) criterion in periodic heating scenarios. These varied applications underscore the term's role in characterizing in dissipative media across physics.

General Concept

Definition

Penetration depth, denoted as δ, is defined as the characteristic distance over which the amplitude of a propagating wave or in a medium attenuates to 1/ (approximately 37%) of its initial value, primarily due to mechanisms such as , , or dissipative losses. This measure quantifies how far a wave or field can effectively penetrate before significant reduction in occurs, making it a fundamental parameter in understanding wave behavior in various materials. The concept traces its roots to 19th-century investigations into electrical conduction, including Lord Kelvin's 1881 formulation of the law for economical conductor sizing in transmission lines, which accounts for energy losses due to resistance. The term "penetration depth" emerged in early 20th-century physics, building on 's 1873 derivation of the , which described nonuniform current distribution and field decay in conductors, and was formalized for sinusoidal wave propagation in subsequent works. Penetration depth is typically expressed in units of meters (m) and varies with medium-specific properties, including the frequency of the wave and material constants such as electrical , , or coefficient. For example, electromagnetic waves at radio frequencies in have penetration depths on the order of centimeters, limiting their utility for deep-water applications without very low frequencies.

Physical Interpretation

The penetration depth characterizes the exponential decay of wave amplitude as it propagates through a lossy medium, modeled by A(z) = A(0) e^{-z/\delta}, where A(z) is the amplitude at depth z and A(0) is the initial amplitude; intensity, proportional to the square of the amplitude, follows I(z) = I(0) e^{-2z/\delta}, falling to $1/e (approximately 37%) of its initial value at z = \delta/2. This decay illustrates the irreversible dissipation of energy, where the wave's propagating component diminishes progressively, limiting effective interaction depth. Physically, penetration depth emerges from fundamental energy loss mechanisms in the medium. Absorption converts into through resonant interactions with material oscillators, such as electrons or molecules, leading to and an imaginary component in the that drives exponential . redirects in random directions via interactions with inhomogeneities, reducing forward without net loss but effectively shortening the penetration distance. introduces phase variations due to frequency-dependent speeds, distorting the and affecting , but it does not contribute to . Collectively, these mechanisms determine the loss rate, with penetration depth serving as its measure—the smaller \delta, the faster the is depleted. At interfaces between media, penetration depth governs the behavior of evanescent fields under conditions like , where the wave does not propagate into the lower-index medium but extends a decaying field into it, qualitatively enabling effects such as tunneling without violating . This field decays exponentially over a distance on the order of the , contrasting with propagating inside uniform media. Visually, the physical process is represented by diagrams showing wave amplitude starting strong at the medium's surface and curving downward exponentially within it, symbolizing energy dissipation; in contrast, total reflection at boundaries depicts the rebounding fully outside while a faint, rapidly vanishing tail penetrates briefly, highlighting the non-propagating nature of evanescent components.

Mathematical Foundations

Attenuation Constant

The attenuation constant, denoted as \alpha, quantifies the exponential decay rate of a wave's amplitude as it propagates through a medium, serving as the reciprocal of the penetration depth \delta, such that \alpha = 1/\delta. This parameter measures how rapidly the wave's amplitude diminishes with distance, with units of nepers per meter (Np/m), where the neper is a dimensionless unit based on the natural logarithm. In general wave propagation, \alpha arises from energy losses due to absorption and scattering, directly linking the spatial decay to the medium's dissipative properties. The amplitude A(z) of a wave traveling a distance z into the medium follows the relation A(z) = A(0) \, e^{-\alpha z}, where A(0) is the initial amplitude at z = 0. This exponential form highlights \alpha's role in defining the characteristic scale over which the wave penetrates before significant attenuation occurs. For wave intensity I, which is proportional to the square of the amplitude in many contexts (e.g., electromagnetic or acoustic waves), the power attenuation coefficient is $2\alpha, yielding I(z) = I(0) \, e^{-2\alpha z}. This doubled rate for intensity underscores the quadratic dependence on amplitude, emphasizing \alpha's foundational impact on both field and energy propagation. Experimentally, \alpha is determined through transmission measurements, where the wave's or is recorded after passing through samples of varying thickness. Plotting \ln(I(z)/I(0)) versus z produces a straight line with -2\alpha for intensity data, from which \alpha is extracted as half the absolute value of the ; alternatively, direct amplitude measurements yield a of -\alpha. Such techniques, often employing transducers or detectors, provide precise values under controlled conditions like constant and . The magnitude of \alpha depends on extrinsic factors such as wave frequency, medium , and material composition, which modulate and mechanisms. Frequency dependence is particularly notable in viscous media, where classical predicts \alpha \propto \omega^2 (with \omega the ), arising from viscous as derived in Stokes' analysis of sound attenuation. influences \alpha by altering and molecular relaxation processes, typically reducing attenuation in gases as rises due to decreased molecular collision rates. Material composition further tunes \alpha through variations in , elasticity, and dissipative constituents, such as impurities or that enhance .

Derivation for Plane Waves

The derivation of penetration depth for plane waves begins with the , which governs the behavior of electromagnetic fields in source-free, linear, isotropic media. For time-harmonic fields, this scalar equation takes the form \nabla^2 \mathbf{E} + k^2 \mathbf{E} = 0, where k is the complex wavenumber that accounts for both and in dissipative media. The complex nature of k arises from the material's response, incorporating losses through the complex permittivity or permeability. For a propagating in the positive z-direction, the can be expressed in one dimension as \mathbf{E}(z, t) = \mathbf{E}(0) e^{i(\omega t - k z)}, assuming monochromatic waves at \omega. Here, k = \beta - i \alpha, with \beta as the real phase constant and \alpha > 0 as the constant. Substituting this form into the yields the k^2 = \omega^2 \mu \tilde{\epsilon}, where \tilde{\epsilon} is the complex and \mu is the permeability (assumed real for non-magnetic media). The field then becomes \mathbf{E}(z, t) = \mathbf{E}(0) e^{i \omega t} e^{-i \beta z} e^{-\alpha z}, showing along the propagation direction. In lossy dielectrics, the complex permittivity is \tilde{\epsilon} = \epsilon (1 - i \tan \delta), where \epsilon is the real permittivity and \tan \delta = \sigma / (\omega \epsilon) is the loss tangent, with \sigma as the conductivity. The wavenumber is thus k = \omega \sqrt{\mu \epsilon (1 - i \tan \delta)}, and the attenuation constant \alpha is the imaginary part of this expression (adjusted for sign convention), \alpha = -\operatorname{Im}(k). The penetration depth \delta, defined as the distance over which the wave amplitude decreases to $1/e of its initial value, is then \delta = 1 / \alpha. For low-loss media where \tan \delta \ll 1, \alpha \approx (\omega \sqrt{\mu \epsilon} \tan \delta)/2, providing an approximate form for \delta. This derivation assumes linear and isotropic media, monochromatic plane waves with no , and time-harmonic fields derived from in the . The attenuation constant \alpha directly relates to the imaginary part of the , establishing \alpha = 1 / \delta as the general solution for wave decay in dissipative environments.

Electromagnetic Applications

Skin Depth in Conductors

In conductors, the skin depth, often denoted as \delta, represents the distance over which the of an electromagnetic wave decreases to $1/e (approximately 37%) of its value at the surface, due to ohmic losses. For good conductors, this penetration depth is given by the formula \delta = \sqrt{\frac{2}{\omega \mu \sigma}}, where \omega is the of the wave, \mu is the magnetic permeability of the material, and \sigma is its . This expression arises from the attenuation of the wave's inside the , E(z) = E_0 e^{-z/\delta}, where z is the depth from the surface. The physical origin of the skin depth lies in the induction of eddy currents within the conductor, which, according to Lenz's law, generate a secondary magnetic field that opposes the incident wave's changing field. These opposing currents effectively cancel the field deeper within the material, confining the wave's energy to a thin surface layer. As frequency decreases, \delta increases proportionally to $1/\sqrt{\omega}, allowing greater penetration; conversely, higher frequencies result in shallower penetration, enhancing the skin effect. In practical applications, the skin effect significantly impacts (AC) systems by increasing the effective of conductors at higher frequencies, leading to elevated power losses in wires and cables. For instance, in lines and windings, this necessitates designs that mitigate losses, such as using stranded or to distribute more evenly across the cross-section. At 60 Hz, typical for , the skin depth in is approximately 8.5 mm, meaning in thicker wires flows primarily near the surface, raising resistive heating and concerns in high-power applications. The skin depth approximation holds under the condition that the conductor is "good," specifically when \sigma \gg \omega \epsilon, where \epsilon is the permittivity, ensuring conduction currents dominate over displacement currents. This breaks down in poor conductors or at very low frequencies, where the full complex propagation constant must be considered instead of the simplified form.

Penetration in Dielectrics and Plasmas

In s, electromagnetic wave penetration is governed by the material's complex , where losses arise from dielectric relaxation and mechanisms rather than free charge conduction. The penetration depth \delta, defined as the distance over which the wave decays to 1/e of its initial value, is approximated for low-loss dielectrics as \delta \approx \frac{2}{\omega \sqrt{\mu \epsilon} \tan \delta}, with \omega the , \mu the permeability, \epsilon the real , and \tan \delta the loss tangent representing the ratio of imaginary to real components. This formulation highlights how losses, such as those from molecular reorientation or relaxation, attenuate the wave exponentially, distinct from the conductivity-driven in metals. In plasmas, particularly collisional ones like those in the ionosphere, penetration depth is influenced by the collective motion of charged particles and electron-ion or electron-neutral collisions. For such media, an approximate expression in the low-frequency, highly collisional limit (\omega \ll \nu) is \delta \approx \frac{c}{\omega_p} \sqrt{\frac{2 \nu}{\omega}}, where c is the speed of light, \omega_p the plasma frequency (\sqrt{n e^2 / \epsilon_0 m_e} with n electron density, e charge, m_e mass, and \epsilon_0 vacuum permittivity), and \nu the collision rate. This arises from the complex dielectric function \epsilon \approx 1 - \omega_p^2 / (\omega (\omega + i \nu)), leading to evanescent waves below the plasma frequency with attenuation enhanced by collisions, critical for understanding signal propagation and absorption in partially ionized gases. Representative applications include signal absorption in the Earth's atmosphere, treated as a lossy due to molecular by and oxygen, where depths reach on the order of kilometers at frequencies around 10 GHz under clear conditions. In and treatments, biological tissues act as lossy s with depths of 2–4 cm at 915 MHz or 2.45 GHz, enabling targeted energy delivery while minimizing surface heating. The penetration depth in these media often decreases with increasing , as higher \omega amplifies the factor in the formulas, though resonant absorptions (e.g., at molecular or resonances) can introduce frequency-specific variations; for instance, in dielectrics with constant \tan \delta, \delta scales inversely with \omega, while in collisional plasmas, the \sqrt{\nu / \omega} term further reduces \delta at higher frequencies relative to collision rates.

Optical Applications

Beer-Lambert Law

The Beer-Lambert law provides the optical formulation for penetration depth in absorbing media, describing the of as it propagates through a homogeneous material. The T, defined as the ratio of transmitted I to incident I_0, is given by T = \frac{I}{I_0} = e^{-\alpha L}, where \alpha is the absorption coefficient (with units of inverse ) and L is the path length through the medium. The penetration depth \delta, representing the distance over which the falls to $1/e (approximately 37%) of its initial value, is then \delta = 1/\alpha. This coefficient \alpha relates to the material's via the imaginary part k of the complex \tilde{n} = n + i k, expressed as \alpha = \frac{4\pi k}{\lambda}, with \lambda denoting the wavelength of the light in vacuum. Historically, the foundational exponential absorption principle was established by Johann Heinrich Lambert in his 1760 work Photometria, which applied it to light transmission through atmospheres and turbid media. August Beer extended this in 1852 by incorporating the effects of solute concentration in liquid solutions, leading to the combined form now known as the Beer-Lambert law. The contemporary version for non-scattering absorbing media retains the pure exponential form above, serving as the basis for quantitative spectroscopy. The coefficient \alpha exhibits strong dependence, \alpha(\lambda), determined by the material's electronic and vibrational spectra. For instance, in clear ocean , \delta \approx 20–$50 m at wavelengths around 475 nm due to minimal , but it decreases to about 3 m at 700 nm in the near- region where vibrational modes begin to dominate , reaching centimeters further into the . In pure , the penetration at 475 nm is higher, around 60 m. This spectral variation underlies phenomena like the coloration of . Experimental validation of the Beer-Lambert law relies on , where light A = -\ln T = \alpha L is measured across varying path lengths or concentrations to confirm . setups using UV-Vis spectrophotometers on dilute solutions, such as metal ion complexes, routinely demonstrate adherence to the law within limits of low optical density (typically A < 2), enabling precise determination of \alpha from spectral data.

Absorption and Scattering Effects

In scattering media, light penetration deviates from the simple exponential decay described by the Beer-Lambert law, which assumes pure absorption without particle interactions. Scattering redirects photons along tortuous paths, increasing the effective optical path length and enhancing absorption, thereby reducing the overall penetration depth compared to non-scattering conditions. Within radiative transfer theory, the effective penetration depth in turbid media accounts for both absorption and scattering effects through the formula \delta_\text{eff} = \frac{1}{\alpha + \sigma_s}, where \alpha is the absorption coefficient and \sigma_s is the scattering coefficient; this represents the e-folding distance for the ballistic (unscattered) photon component before significant attenuation occurs. This approach models the combined extinction in semi-infinite or slab geometries, where multiple scattering events further complicate deeper propagation by diffusing light energy. For particulate scattering, Mie theory describes interactions with particles comparable in size to the wavelength, governed by the size parameter x = \frac{2\pi r}{\lambda}, where r is the particle radius and \lambda is the wavelength; larger x values lead to more forward-directed scattering but overall reduced penetration depth due to increased extinction efficiency. In biological tissues, Mie scattering from cellular components like mitochondria (radii ~0.5–1 μm) dominates near-infrared light attenuation, while in atmospheric aerosols (particle radii ~0.1–10 μm), it limits visibility and signal return in optical systems. Monte Carlo simulations model these complex trajectories by tracing individual paths through the medium, revealing that multiple events increase the via diffusive transport, which amplifies and effectively shortens the penetration depth beyond the ballistic limit. These simulations are essential for quantifying path length distributions in highly environments, where undergo thousands of before or escape. In applications, such as laser therapy, near-infrared light penetrates to depths of approximately 1–5 mm in scattering-dominated regimes, enabling targeted treatments while being limited by turbidity. Similarly, in for , aerosol reduces lidar penetration depths to kilometers or less, constraining the profiling of properties.

Other Contexts

Acoustic Waves

In acoustics, the penetration depth of sound waves in fluids and solids refers to the characteristic distance over which the wave's amplitude diminishes by a factor of e (approximately 8.7 dB) due to attenuation processes. This depth, denoted as \delta, is inversely related to the attenuation coefficient \alpha, where \delta = 1 / \alpha, and attenuation arises primarily from energy dissipation mechanisms in the medium. For plane acoustic waves, the intensity decays exponentially as I = I_0 e^{-2\alpha x}, with x being the propagation distance, highlighting how viscous and thermal effects limit wave propagation in dissipative media like air, water, or biological tissues. The primary mechanisms of acoustic attenuation include classical absorption due to viscous drag and thermal conduction, molecular relaxation processes, and scattering from inhomogeneities. Classical absorption, derived from the Navier-Stokes equations for compressible fluids, quantifies energy loss from shear viscosity and heat transfer, yielding the attenuation coefficient \alpha = \frac{\omega^2}{2 \rho c^3} \left( \frac{4}{3} \eta + \frac{(\gamma - 1) \kappa}{C_p} \right), where \omega is the angular frequency, \rho is the medium density, c is the sound speed, \eta is the shear viscosity, \kappa is the thermal conductivity, \gamma is the heat capacity ratio, and C_p is the specific heat at constant pressure; this is known as the Stokes-Kirchhoff formula. Relaxation absorption involves internal molecular rearrangements, such as vibrational or rotational modes in polyatomic gases, while scattering redirects wave energy due to spatial variations in the medium's properties, like grain boundaries in solids or bubbles in fluids. These mechanisms collectively determine \delta, with classical effects dominating in simple fluids at moderate frequencies. In practical applications, penetration depths vary significantly with , as \alpha \propto f^2 for classical , leading to shallower depths at higher frequencies. For in soft tissues at MHz frequencies (e.g., 1–10 MHz), \delta is typically on the order of centimeters, enabling up to about 10–20 cm deep but limiting resolution in deeper structures. In ocean acoustics, low-frequency sound (below 100 Hz) experiences minimal , resulting in \delta on the order of kilometers, which facilitates long-range propagation over thousands of kilometers in the for applications like detection. Attenuation and penetration depth in are commonly measured using -echo techniques, where a short ultrasonic is transmitted into the medium, and the of the reflected is analyzed to estimate \alpha from the decay rate along the path. This method is widely employed in to quantify tissue properties, compensating for frequency-dependent losses to improve diagnostic accuracy.

Particle Penetration in Materials

In the context of and , the penetration of charged particles into solids is primarily governed by their loss through interactions with atomic electrons, quantified by the \frac{dE}{dx}. The Bethe-Bloch formula provides the theoretical foundation for this , expressing the loss per unit path length for swift charged particles as -\frac{dE}{dx} = \frac{4\pi z^2 e^4 N Z}{m_e v^2} \left[ \ln \left( \frac{2 m_e v^2}{I (1 - \beta^2)} \right) - \beta^2 \right], where z is the of the incident particle, e and m_e are the charge and mass, v is the , \beta = v/c, N is the of target atoms, Z is the of the target , and I is the of the . This formula scales with z^2 (projectile charge squared) and Z (target atomic number), while the material's A influences the N Z / A, affecting per unit mass. Relativistic corrections to the Bethe-Bloch formula account for high-speed effects through the \beta^2 and (1 - \beta^2) terms, which modify the logarithmic factor and become prominent as \beta approaches 1, leading to a gradual rise in at energies above ~1 GeV. The total R of a , defined as the distance traveled until it loses most of its , can be approximated as R \approx E / \left( \frac{dE}{dx} \right) under the continuous slowing-down when \frac{dE}{dx} varies slowly with energy. A practical application of penetration occurs in for doping, where ions such as or at keV energies penetrate to depths of ~10 nm to 1 μm, enabling precise control of electrical properties without bulk diffusion. At higher energies, such as those of muons (typically GeV scale), penetration depths reach kilometers into the , as these minimally ionizing particles traverse dense rock with minimal , allowing detection in underground experiments. For neutral particles like neutrons, penetration in materials arises from probabilistic interactions including moderation (elastic scattering that reduces energy) and capture (absorption leading to nuclear reactions), characterized by the macroscopic total cross-section \Sigma_t = N \sigma_t, where N is the atomic number density and \sigma_t is the microscopic total cross-section. The penetration depth \delta, equivalent to the mean free path, is then \delta = 1 / \Sigma_t, representing the average distance a neutron travels before interacting; this varies with neutron energy and material composition, from millimeters in high-cross-section absorbers like cadmium to meters in light moderators like water. Unlike charged particles, neutron penetration lacks a Coulombic energy loss mechanism, emphasizing nuclear rather than electronic interactions.

References

  1. [1]
    3.12: Skin Depth - Physics LibreTexts
    May 9, 2020 · Skin depth is the distance over which the magnitude of the electric or magnetic field is reduced by a factor of 1 / e ≅ 0.368 . Since power is ...
  2. [2]
    Electromagnetic Signal Attenuation (Skin Depth) | US EPA
    Jan 24, 2025 · Skin depth is the depth at which the EM signal attenuates to 1/e or approximately a third. Skin depth is typically measured in meters.
  3. [3]
    Depth of Penetration and Current Density - NDE-Ed.org
    The depth at which eddy current density has decreased to 1/e, or about 37% of the surface density, is called the standard depth of penetration (d). The word ' ...
  4. [4]
    Penetration depth – Knowledge and References - Taylor & Francis
    Penetration depth is the depth an electromagnetic wave can penetrate into a material, where its power density falls to 36.8% of its surface value.Missing: physics | Show results with:physics
  5. [5]
    Characteristic Lengths in Superconductors - HyperPhysics
    The London penetration depth refers to the exponentially decaying magnetic field at the interior surface of a superconductor. It is related to the density of ...Missing: definition | Show results with:definition
  6. [6]
    London penetration depth – Knowledge and References
    London penetration depth refers to the distance that an external magnetic field or current can penetrate into a superconductor, which is only a thin surface ...Missing: definition | Show results with:definition
  7. [7]
    Penetration Depth - an overview | ScienceDirect Topics
    Penetration depth is the distance where induced eddy currents or light intensity fall to 1/e of their surface value, also known as skin depth.
  8. [8]
    Lord Kelvin | On the Sources of Energy in Nature...
    I say no more on this point at present, as the economy of copper for electric conduction will be the subject of a special communication to the Section. I ...
  9. [9]
    One Hundred and Fifty Years of Skin Effect - MDPI
    Nov 16, 2023 · According to Maxwell, the current density in a conductor increases towards the conductor surface. This phenomenon is called the skin effect. The ...Missing: origin | Show results with:origin
  10. [10]
    Superlensing enables radio communication and imaging underwater
    Oct 26, 2023 · Both radio communication and radio imaging through water may be enabled by superlensing of surface electromagnetic waves propagating along the water surface.
  11. [11]
    7.2: Attenuation and Dispersion - Physics LibreTexts
    Mar 5, 2022 · The dispersion and attenuation effects by considering a particular case of time evolution of the electric polarization P(t) of a dilute, non-polar medium.
  12. [12]
    Increasing the Penetration Depth of Microwave Radiation Using ...
    Jun 7, 2019 · (16) The penetration depth response is highly frequency-dependent, where the relationship is governed by exponential decay. Wave propagation ...
  13. [13]
    Attenuation of Waves - Nondestructive Evaluation Physics : Waves
    Attenuation is the decay rate of a wave as it propagates, caused by scattering and absorption, which diminish the wave's intensity.<|control11|><|separator|>
  14. [14]
    Electromagnetic Absorption - an overview | ScienceDirect Topics
    The absorbed waves appear as heat energy, which is the result of kinetic energy resulting from resonance. This process is the general mechanism of wave ...<|control11|><|separator|>
  15. [15]
    [PDF] Lecture 7: Propagation, Dispersion and Scattering
    scattering absorption. Interaction of EMR with the atmosphere. • EMR is attenuated by its passage through the atmosphere via scattering and absorption.
  16. [16]
    Evanescent Waves - RP Photonics
    A larger penetration depth is achieved only when the incidence angle is close to the critical angle for total internal reflection. Larger penetration depths can ...Example: Total Internal... · Evanescent Wave Coupling to...
  17. [17]
    Understanding the penetration of electromagnetic velocity fields into ...
    Nov 1, 1999 · The acceleration fields give rise to electromagnetic waves whose penetration into ohmic conductors is described by exponential damping with a ...
  18. [18]
  19. [19]
    Wave Attenuation - an overview | ScienceDirect Topics
    Wave attenuation is defined as the loss of energy of a wave as it travels through a medium, resulting from energy absorption, heat production, ...
  20. [20]
    [PDF] Chapter 8 – Absorption and Attenuation of Sound
    So, for classical attenuation, it is proportional to the frequency squared at low frequencies. As a consequence, in experimental measurements of absorption, ...Missing: ω² | Show results with:ω²
  21. [21]
    [PDF] Chapter 9: Electromagnetic Waves - MIT OpenCourseWare
    May 9, 2011 · When a plane wave impacts a conducting surface at an angle, a complex wave propagation vector kt is required to represent the resulting ...
  22. [22]
    None
    ### Summary of Plane Waves in Dissipative Media Derivation
  23. [23]
    None
    ### Summary: Plane Waves in Lossy Media (ECE 604, Lecture 8)
  24. [24]
    10.7 Skin Effect - MIT
    The skin depth, (5), is inversely proportional to the square root of . Thus, an order of magnitude variation in frequency or in conductivity only changes by ...Missing: formula electromagnetic
  25. [25]
    Electromagnetic Waves in Conductors
    The skin-depth is smaller at higher frequencies. This implies that high-frequency waves penetrate a shorter distance into a conductor than low-frequency waves.
  26. [26]
    [PDF] EDDY CURRENTS, DIFFUSION, AND SKIN EFFECT
    Thus, for the 60 Hz AC current in the power wires, the skin depth is 8.4 mm, but for the 700 MHz frequency used by many cellphones, the skin depth in a copper ...
  27. [27]
    [PDF] Chapter 25 Resistance and Current
    • For Copper at 60 Hz the “skin depth” is about 8.5 mm. • The current density J decreases exponentially. • For “good conductors” like metals. • A wire of ...
  28. [28]
    [PDF] ECE 329 – Fall 2021
    𝝈 ≫. For a good conductor, this is called “skin depth”. In general,. 1/𝜶 can be called “penetration depth”. In a good conductor the fields attenuate rapidly.
  29. [29]
    Dielectric Loss Tangent - an overview | ScienceDirect Topics
    ... dielectric loss and low magnetic loss, the corresponding penetration depth can be expressed as follows [72]: (3) D p = 2 C ω ɛ ′ ( 1 + tan 2 δ − 1 ) 1 / 2.
  30. [30]
    [PDF] ECE 507 – Lecture 13 - ECE507 - Plasma Physics and Applications
    TRANSVERSE ELECTROMAGNETIC WAVES IN PLASMAS(continuation). Defining m ne o o. 2. 2 p ε ω = oo. 2. 1 c μ ε. = Wave equation for transverse wave in a plasma. 0)t, ...
  31. [31]
    Penetration of waves into half-space collisional plasma - ADS
    Numerical calculations for the case of a plasma frequency equal to 5.5 x 10 to the 8th/sec, a collision frequency equal to the plasma frequency or 0.001 the ...Missing: rate | Show results with:rate
  32. [32]
    Potential of Dual-Frequency Radar and Microwave Radiometer ...
    Microwave attenuation is frequency dependent: liquid water and water vapor attenuation generally increase with frequency (continuum absorption), and attenuation ...
  33. [33]
    Microwave Tissue Ablation: Biophysics, Technology and Applications
    At 915 MHz and 2.45 GHz, wave penetration is 2–4 cm in most tissues, which is often commensurate with the treatment target (eg, 2–4 cm tumors).
  34. [34]
    2.3 The Lambert-Beer law and the absorption coefficient α
    Labs is called the absorption length or penetration depth. Note, that the latter expression is somewhat misleading: the light decays only to about 1/3 of ...
  35. [35]
    [PDF] π IN THE SKY7
    Solve for the blue light and red light absorption coefficients. α = (4πk) / λ blue light: α = (4π . 1.01E-09) / ( ...
  36. [36]
    Photometrie. Photometria, sive De mensura et gradibus luminis ...
    Feb 14, 2016 · Photometria, sive De mensura et gradibus luminis, colorum et umbrae (1760). by: Lambert, Johann Heinrich, 1728-1777; Anding, E. ... PDF download.
  37. [37]
    The Bouguer-Lambert-Beer Absorption Law and Non-Planar ...
    In 1852, August Beer published another version of the absorption law in German which included the effect of concentration, stemming from his investigation ...
  38. [38]
    Beer–Lambert law for optical tissue diagnostics: current state ... - NIH
    Oct 28, 2021 · Light penetration depth into the skin was calculated as inverse value of the attenuation coefficient. This simple form of BLL failed only at ...
  39. [39]
    Light in the Ocean | manoa.hawaii.edu/ExploringOurFluidEarth
    Because of this, the higher energy light with short wavelengths, such as blue, is able to penetrate more deeply. At 40 m, saltwater has absorbed nearly all the ...
  40. [40]
    Estimation of light penetration, and horizontal and vertical visibility in ...
    Jun 5, 2007 · This study opens the way to the development of simple semianalytical ocean color algorithms that make the best use of spectral information.
  41. [41]
    Teaching Beer's Law and Absorption Spectrophotometry with a ...
    Jan 29, 2016 · The Beer-Lambert law, also commonly referred to as Beer's law, enables the quantitation of the concn. of an absorbing analyte in soln. Results ...Abstract · Introduction · Figure 2
  42. [42]
    There's plenty of light at the bottom: statistics of photon penetration ...
    Jun 3, 2016 · We propose a comprehensive statistical approach describing the penetration depth of light in random media.
  43. [43]
    Investigation of depth selectivity of polarization gating for tissue ...
    Optical depth τ is defined as τ=(µ s +µ a )D where D is the geometrical thickness and µ s and µ a are the scattering and absorption coefficients, respectively.
  44. [44]
    Penetration of Light into Multiple Scattering Media - Sage Journals
    Aug 1, 2012 · They define the penetration depth as the depth in a semi-infinite layer where the radiation flux density is reduced to 1% of that at the ...
  45. [45]
    [PDF] Optical properties of biological tissues: a review - OMLC
    May 10, 2013 · Mie scattering is the generic name for scattering by a sphere of any size, both small and large, and the common term Rayleigh scattering refers.
  46. [46]
    [PDF] Monte Carlo Modeling of Light Transport in Multi-layered Tissues in ...
    Craig M. Gardner (University of Texas, Austin) provided us his Monte Carlo simulation and convolution results of a multi-layered medium, which are ...
  47. [47]
    Monte Carlo modeling of light propagation in highly scattering ...
    Monte Carlo modeling of light propagation in highly scattering tissues. I. Model predictions and comparison with diffusion theory.<|control11|><|separator|>
  48. [48]
    Effect of wavelength and beam width on penetration in light-tissue ...
    Sep 12, 2017 · In the penetration depth of 1%, the intensity is reached at 5378 μm (5.4 mm) with a wavelength of 750 nm. This is in agreement with literature ...
  49. [49]
    Lidar Penetration Depth Observations for Constraining Cloud ...
    Lidar Penetration Depth Observations for Constraining Cloud Longwave Feedbacks ... Abstract. Satellite-borne active remote sensing Cloud-Aerosol Lidar and ...
  50. [50]
    [PDF] 3d. Acoustic Properties of Gases - LEO L. BERANEK - MIT
    The attenuation caused by heat conduction and viscosity of the air a, is not known so accurately. The classical absorption due to these causes,' as discussed ...
  51. [51]
    [PDF] On the sound dispersion and attenuation in fluids due to thermal and ...
    Jul 19, 2021 · This formula completes the classical attenuation relation called Stokes-. Kirchhoff. We represent in a simplified manner the Kirchhoff ap-.
  52. [52]
    Ultrasound | Physics - Lumen Learning
    The accepted rule of thumb is that you can effectively scan to a depth of about 500λ into tissue. For 7 MHz, this penetration limit is 500 × 0.22 mm, which is ...
  53. [53]
    Chapter Seven Acoustical Oceanography - ScienceDirect.com
    Sound waves propagate through the ocean extraordinarily well, especially at lower frequencies. Acoustical signals are attenuated to a much smaller degree than ...
  54. [54]
    Pulse-echo ultrasound attenuation tomography - IOPscience
    May 20, 2024 · We present the first fully two-dimensional attenuation imaging technique developed for pulse-echo ultrasound systems.
  55. [55]
    [PDF] 34. Passage of Particles Through Matter
    Aug 11, 2022 · 34.4 shows R/M as a function of βγ (= p/Mc) for a variety of materials. The mass scaling of dE/dx and range is valid for the electronic losses ...
  56. [56]
    [PDF] Chapter 9: Ion Implantation
    Ion implantation energies range from several hundred to several million electron volts, resulting in ion distributions with average depths from < 10 nm to 10 μ.
  57. [57]
    Penetration of cosmic ray muons into the Earth - ScienceDirect.com
    The result predicts the underground muon vertical intensity over seven orders of magnitude (10 km depth), compares favorably with published Monte Carlo ...
  58. [58]
    Macroscopic Cross-section | Definition & Examples - Nuclear Power
    The macroscopic cross-section represents the effective target area of all of the nuclei contained in the volume of the material (such as fuel pellet).