Fact-checked by Grok 2 weeks ago

Attenuation coefficient

The attenuation coefficient is a quantitative measure of the fractional reduction in the intensity of a of or as it passes through an absorbing or medium per unit distance traveled. It quantifies the loss due to mechanisms such as , , and other dissipative processes, and is fundamental to understanding wave propagation in diverse physical contexts. In radiation physics, particularly for X-rays and gamma rays, the linear attenuation coefficient (denoted μ) represents the probability of photon interaction (removal from the beam) per unit length of the absorber, with units of inverse length (e.g., cm⁻¹). The (μ/ρ) normalizes this by the material's , yielding units of area per mass (e.g., cm²/g), which facilitates comparisons across different materials and is essential for calculating penetration and energy deposition. These coefficients depend on , of the material, and interaction type (e.g., , ). In acoustics, the attenuation coefficient (often α) describes the of wave or in a medium, arising from viscous , , and acoustic relaxation processes. It can be expressed in nepers per meter (/m) or decibels per meter (/m), with the relationship α (dB/m) = 8.686 α (Np/m), and is critical for modeling in air, , or . In and electromagnetic wave propagation, the attenuation coefficient relates to the imaginary part of the complex refractive index, governing the of via Beer's law: I = I₀ e^{-α x}, where x is the path length. For ocean , the beam attenuation coefficient c(λ) = a(λ) + b(λ) combines a(λ) and b(λ) coefficients, influencing penetration in water columns. Applications span ( optic losses), ( effects on ), and ( attenuation).

Introduction

Definition and Basic Concept

The attenuation coefficient, often denoted as \alpha, is a fundamental parameter in physics that quantifies the fractional decrease in the of a of —such as , X-rays, or other electromagnetic —per unit traveled through a homogeneous medium. It characterizes the medium's capacity to reduce the beam's energy by processes that remove photons from the primary path, thereby describing how easily the material can be penetrated by the . This is essential for understanding wave propagation in diverse materials, where higher values indicate greater attenuation and thus more rapid loss. The basic relationship governing this phenomenon is the exponential attenuation law, derived from Beer's law, which states that the transmitted intensity I after propagating a distance x through the medium is given by I = I_0 e^{-\alpha x}, where I_0 is the initial intensity of the beam. This equation assumes a monochromatic, collimated beam in a non-scattering or uniformly scattering medium, highlighting the linear dependence on distance and the coefficient's role in exponential decay. Attenuation arises primarily from absorption and scattering, which redirect or eliminate photons from the beam. The units of the attenuation coefficient are typically inverse length, such as m^{-1} or cm^{-1}, emphasizing its interpretation as a linear measure of attenuation per unit path length. This linear form distinguishes it from related quantities like the , which normalizes by . Examples of its application span various , including gases (e.g., atmospheric attenuation of ), liquids (e.g., in underwater ), and solids (e.g., in medical ), where the coefficient helps predict signal degradation in practical scenarios.

Historical Context and Applications

The concept of the attenuation coefficient originated in the 18th and 19th centuries through foundational work in on light absorption in homogeneous media. first described the gradation of in 1729, laying early groundwork for quantifying transmission losses. This was formalized by in 1760, who established the exponential relationship between and path length in his treatise Photometria, providing the mathematical basis for attenuation in non-absorbing media. August Beer extended this in 1852 by incorporating the effects of solute concentration on absorption in liquids, completing the Beer-Lambert law that underpins the modern attenuation coefficient in . In the early , the concept was extended to higher-energy radiation, particularly s, through pioneering experiments on and . , in his work from 1904 onward, quantified absorption in various elements, identifying characteristic absorption series (K, L, etc.) and demonstrating that absorbed energy is largely re-emitted as secondary radiation. His findings, detailed in the 1916 Bakerian Lecture, enabled the application of attenuation principles to , influencing fields beyond visible . The attenuation coefficient plays a critical role in diverse applications, enabling the prediction of penetration and material interactions. In optical fiber communications, it quantifies signal loss due to and , typically expressed in dB/km, which guides the design of low-loss fibers for transoceanic data transmission. In , such as computed (CT) scans, it measures attenuation in tissues relative to , forming the basis for Hounsfield units that differentiate healthy from pathological structures. employs it to model by aerosols, assessing impacts on budgets and forcing. In , it informs gamma-ray shielding designs, calculating material thicknesses needed to reduce in reactors and storage facilities. This parameter's importance lies in its utility for material characterization and engineering solutions, such as filters and detectors, where it predicts how radiation diminishes exponentially with distance—the core model introduced by and . Attenuation coefficients vary significantly by and medium; for instance, in pure for red visible light (around 676 nm), it is approximately 0.42 m⁻¹, highlighting 's relative transparency in the .

Fundamental Mathematical Definitions

Bulk Attenuation Coefficient

The bulk attenuation coefficient, denoted as \alpha, quantifies the rate of intensity reduction for a propagating in a uniform medium, assuming no angular dependence. It arises from the describing infinitesimal intensity loss along the direction: \frac{dI}{dx} = -\alpha I, where I is the at distance x. This posits that the fractional loss in over an infinitesimal path length dx is proportional to the local and the material's intrinsic attenuation properties. Integrating this differential form yields the law: I(x) = I_0 e^{-\alpha x}, where I_0 is the incident at x = 0. This integrated form, foundational to the Beer-Lambert law in and analogous expressions in other fields like acoustics and radiation , assumes a constant \alpha throughout the medium. The derivation relies on key assumptions about the medium and geometry. The medium must be homogeneous, with uniform properties that do not vary spatially, and isotropic, meaning is independent of direction within the bulk. The incident is taken as collimated, implying rays with negligible , and effects, such as those from coherent waves, are ignored to focus on incoherent or classical intensity transport. These conditions ensure the proportionality in the holds without additional terms for redirection or . In practice, the bulk attenuation coefficient is extracted experimentally from measurements of transmitted and incident intensities over a known path length x: \alpha = -\frac{1}{x} \ln\left(\frac{I}{I_0}\right). This logarithmic relation directly follows from rearranging the exponential law and is widely used in for materials like optical filters or biological tissues. However, the approach is limited to narrow-beam configurations, where scattered radiation escaping the beam path is negligible, as broader beams would incorporate buildup from multiple , inflating the apparent . Additionally, boundary effects, such as or at interfaces, are disregarded, restricting applicability to deep far from edges.

Directional Attenuation Coefficient

The directional attenuation coefficient, denoted as \alpha(\theta, \phi), quantifies the fractional loss of radiance per unit path length for radiation propagating in a specific direction specified by the polar angle \theta and azimuthal angle \phi. This coefficient arises in scenarios where the medium exhibits anisotropy, such as in aligned fibers, porous materials, or oriented particle distributions, leading to direction-dependent attenuation distinct from isotropic bulk cases. In the equation (RTE), the directional attenuation term governs the diminution of radiance L(\mathbf{r}, \Omega) along the path length s in direction \Omega = (\theta, \phi), formulated as \frac{dL}{ds} = -\alpha(\theta, \phi) L + S(\mathbf{r}, \Omega), where S(\mathbf{r}, \Omega) encompasses emission, in-scattering, and other source contributions. Here, \alpha(\theta, \phi) typically equals the sum of directional and coefficients, \alpha_a(\theta, \phi) + \alpha_s(\theta, \phi), and its angular variation stems from the geometric or orientation of medium components. This form extends the simpler law for collimated beams by incorporating directional specificity for diffuse or polychromatic fields. The directional attenuation coefficient finds critical application in simulations for non-collimated light, particularly in planetary atmospheres where multiple and varying incidence angles influence energy through layered, particle-laden . For instance, in modeling atmospheric radiative fluxes, \alpha(\theta, \phi) enables accurate of directional radiance fields affected by aerosol orientations or cloud structures. In -dominated , such as those with forward-peaked phase functions (asymmetry g > 0), the effective directional attenuation coefficient is reduced compared to isotropic assumptions, as forward-scattered photons remain aligned with the direction and contribute less to net loss. This effect is evident in the reduced coefficient \alpha_s' = \alpha_s (1 - [g](/page/G)), lowering the overall \alpha(\theta, \phi) for near-forward paths and altering penetration depths in turbid atmospheres. Quantitative assessments show that for [g](/page/G) approaching 1, the effective attenuation can drop significantly below the total extinction value, impacting and climate modeling.

Spectral and Hemispherical Variations

Spectral Attenuation Coefficient

The spectral attenuation coefficient, denoted as \alpha(\lambda), quantifies the wavelength-dependent loss of as it propagates through a medium, incorporating both and effects that vary with . This parameter is essential for understanding light transmission in dispersive materials such as optical fibers, atmospheric gases, and biological tissues, where attenuation is not uniform across the spectrum. For a monochromatic of \lambda, the transmitted I(\lambda, x) after traveling a distance x through the medium is given by the form derived from Beer's law: I(\lambda, x) = I_0(\lambda) \, e^{-\alpha(\lambda) x}, where I_0(\lambda) is the initial ; here, \alpha(\lambda) has units of inverse length, such as m^{-1} or cm^{-1}. This formulation captures how shorter or longer wavelengths may experience differing attenuation rates due to resonant interactions with the medium's molecular structure. To determine \alpha(\lambda), spectrophotometric techniques are employed, involving the measurement of spectra through samples of varying thicknesses using instruments like UV-Vis or FTIR spectrometers. The coefficient is then extracted by fitting the logarithmic , \ln(T(\lambda)) = -\alpha(\lambda) \, d, where d is the sample thickness, often correcting for losses at interfaces. Such measurements yield detailed \alpha(\lambda) spectra, enabling analysis of structures in materials. In , \alpha(\lambda) often displays sharp peaks at wavelengths corresponding to bands, reflecting or vibrational transitions; for example, exhibits strong attenuation in the spectrum around 2.7 \mum and beyond 4.5 \mum due to O-H stretching and bending modes, significantly impacting applications like and propagation through the atmosphere.

Hemispherical Attenuation Coefficient

The hemispherical coefficient, denoted as \alpha_h(\lambda), quantifies the of diffuse (hemispherical) radiation in a medium and is defined as \alpha_h(\lambda) = -\frac{1}{E_d(\lambda)} \frac{d E_d(\lambda)}{dz}, where E_d(\lambda) is the downward and z is the depth. In isotropic media, where is direction-independent, \alpha_h(\lambda) = \alpha(\lambda). This coefficient arises in theory for media where radiation is incident from multiple directions, providing an effective value for the decay of rather than intensity. This coefficient is particularly useful in calculations involving diffuse and , such as solar radiation propagating through . For instance, in atmospheric models, \alpha_h(\lambda) enables estimation of the fraction of diffuse reaching the ground under overcast conditions, accounting for the integrated effects of cloud layers on both and scattered components. In turbid , such as particle-suspended waters or dense cloud formations, the hemispherical attenuation coefficient is elevated compared to clear conditions due to multiple paths that prolong the effective for diffuse photons, enhancing overall energy loss through repeated absorption and redirection events. For example, in coastal ocean environments, values of the diffuse attenuation coefficient often exceed 0.5 m^{-1} in the spectrum.

Decomposition into Absorption and Scattering

Absorption Coefficient

The absorption coefficient, often denoted as \kappa or \alpha_\text{abs}, quantifies the portion of the total attenuation attributable to the irreversible dissipation of electromagnetic into other forms, such as or /molecular excitations, within a medium. This process occurs when photons are absorbed by the material, leading to a reduction in the wave's intensity without redirection of the energy. Unlike , which redirects without loss, fundamentally removes from the propagating wave. The absorption coefficient relates to the overall attenuation coefficient \alpha through the decomposition \alpha = \kappa + \sigma, where \sigma denotes the scattering coefficient; this separation highlights absorption as the energy-loss component distinct from mere deflection. In practical terms, for a plane wave propagating through a homogeneous medium, the intensity I after distance z follows I(z) = I_0 e^{-\alpha z} = I_0 e^{-(\kappa + \sigma) z}, with the absorptive term e^{-\kappa z} specifically accounting for dissipated power. This relation is foundational in fields like optics and radiative transfer, enabling targeted analysis of loss mechanisms. Quantum mechanically, the absorption coefficient arises from the probabilities of photon-induced transitions between discrete energy levels in atoms or molecules, as described by time-dependent perturbation theory and , which link \kappa to the matrix elements of the dipole operator and the density of final states. These transition probabilities determine the likelihood of an electron absorbing a photon to jump from a lower to a higher energy state, with stronger overlaps yielding higher \kappa values at resonant frequencies. This microscopic foundation explains the material-specific and wavelength-dependent nature of absorption. In semiconductors, the absorption coefficient \kappa exhibits a sharp dependence on the photon energy relative to the material's bandgap E_g, remaining negligible below E_g and rising rapidly above it due to allowed interband transitions. For instance, in with E_g \approx 1.12 eV (corresponding to \lambda \approx 1.1 \mum), \kappa approaches \sim 10^4 cm^{-1} for wavelengths slightly shorter than 1.1 \mum, where photon energies enable valence-to-conduction band excitations and establish the onset of significant optical . This behavior is critical for applications like photovoltaic devices, where efficient energy conversion hinges on \kappa values in this regime.

Scattering Coefficient

The scattering coefficient, often denoted as \sigma or \alpha_\text{scat}, quantifies the contribution to total attenuation arising from the redirection of radiation paths, where incident photons are deflected into different directions rather than continuing straight or being absorbed. It represents the effective cross-sectional area per unit volume available for scattering events, determining the probability per unit path length that a photon will undergo scattering in the medium. This coefficient is complementary to the absorption coefficient, together comprising the total attenuation as the sum of redirection and energy loss processes. The total scattering coefficient relates to its angular dependence through integration over the phase function, which describes the probability distribution of scattering directions: \sigma = \int_{4\pi} \sigma(\theta) \, d\Omega, where \sigma(\theta) denotes the differential scattering coefficient and the integral is over all solid angles d\Omega. This formulation links the overall scattering rate to the directional phase function p(\theta) = 4\pi \sigma(\theta) / \sigma, typically normalized such that the average over yields unity (\int_{4\pi} p(\theta) \, d\Omega / 4\pi = 1), enabling modeling of anisotropic effects in . Scattering mechanisms vary with particle size relative to the radiation wavelength. Rayleigh scattering governs interactions with small particles (much smaller than the wavelength), producing a wavelength-dependent intensity proportional to $1/\lambda^4 and a dipole-like angular pattern that favors side scattering. Mie scattering applies to particles comparable in size to the wavelength, offering an exact electromagnetic solution for spheres that often results in strong forward scattering lobes due to diffraction and refraction. For much larger particles, geometric scattering approximations treat the process via ray optics, emphasizing reflection, refraction, and shadowing effects akin to macroscopic optics. In fog, the coefficient dominates the attenuation process for visible , with typical values around 0.05–0.1 m^{-1} in dense conditions, primarily due to by water droplets of 5–20 μm diameter; this redirection scatters light out of the direct beam, drastically reducing visibility to 40–80 m according to the Koschmieder relation.

Microscopic and Mass-Based Expressions

Expression in Terms of Cross-Sections and Density

The macroscopic attenuation coefficient \alpha, which quantifies the of wave intensity through a medium, can be derived from the microscopic interactions of the propagating wave with individual particles in the medium. Specifically, \alpha is given by the product of the n (particles per unit volume) and the total cross-section \sigma_{\text{total}} per particle, expressed as \alpha = n \sigma_{\text{total}}, where \sigma_{\text{total}} represents the effective area for all processes. This relation bridges the bulk optical property to the fundamental particle-level interactions, ensuring \alpha has units of inverse length (e.g., m^{-1}) since n is in m^{-3} and \sigma_{\text{total}} in m^2. The total cross-section \sigma_{\text{total}} is commonly defined for spherical particles as \sigma_{\text{total}} = \pi r^2 Q_{\text{ext}}, where r is the particle and Q_{\text{ext}} is the dimensionless extinction efficiency factor. This efficiency Q_{\text{ext}} depends on the size parameter x = 2\pi r / \lambda (with \lambda the ) and the complex of the particle, often calculated via Mie for values ranging from near 0 for small particles to approximately 2 in the geometric limit. Absorption and cross-sections contribute additively to \sigma_{\text{total}}, providing a unified measure of . In heterogeneous media containing multiple species, the attenuation coefficient becomes a weighted sum over the components: \alpha = \sum_i n_i \sigma_i, where n_i and \sigma_i are the and total cross-section for the i-th , respectively. This additive form assumes incoherent superposition of interactions, valid for dilute or non-interacting mixtures. For dilute gases, where particle interactions are infrequent and multiple negligible, this microscopic expression directly supports the Beer-Lambert law, I = I_0 \exp(-\alpha L), with L the path length, enabling precise quantification of gaseous attenuation from molecular cross-sections.

Mass Attenuation Coefficient

The , denoted as \mu / \rho, is defined as the of the linear attenuation coefficient \mu to the mass density \rho of the , typically expressed in units of cm²/g. This quantity represents the probability of interaction per unit of the attenuating medium, independent of its physical state or density variations. One key advantage of the is its independence from the material's density, enabling straightforward comparisons of attenuation efficiency between substances like gases, liquids, and solids without accounting for packing or compression effects. Additionally, it scales directly with the atomic composition, as higher elements generally exhibit larger values due to increased interaction probabilities. For a pure material, the is given by \frac{\mu}{\rho} = \frac{N_A}{A} \sigma, where N_A is Avogadro's number ($6.022 \times 10^{23} ^{-1}), A is the in g/, and \sigma is the total cross-section for interactions in cm²/atom. This expression links macroscopic attenuation to microscopic cross-sections, facilitating calculations for and compound materials. Tabulated mass attenuation coefficients are essential for practical applications; for instance, lead has a value of 5.549 cm²/g at 100 keV energy, supporting its selection in shielding designs where high attenuation per unit mass is required. The volumetric attenuation coefficient \alpha relates simply as \alpha = \rho \cdot (\mu / \rho).

Logarithmic and Scaled Forms

Napierian Attenuation Coefficient

The Napierian attenuation coefficient, denoted as \alpha_N, quantifies the of through a medium using the natural logarithm base e, as expressed in the fundamental law I = I_0 e^{-\alpha_N x}, where I is the transmitted , I_0 is the incident , and x is the path length. This form arises directly from the governing , dI/dx = -\alpha_N I, which integrates naturally to the solution without requiring logarithmic conversions. It is the standard coefficient in theory, where it represents the total due to and in atmospheric and oceanic models. One key advantage of the Napierian form is its mathematical simplicity in theoretical derivations, as it aligns seamlessly with the eigensolutions of wave equations in homogeneous media, avoiding the scaling factors needed for other logarithmic bases. This direct linkage to differential forms facilitates precise modeling in fields like photon transport and . To convert to coefficients based on other logarithms, the relation \alpha_{\text{base}} = \alpha_N / \ln(\text{base}) applies; for instance, the decadic \alpha_{10} satisfies \alpha_{10} = \alpha_N / \ln(10) \approx \alpha_N / 2.302585. Such conversions ensure compatibility across and scientific applications while preserving the underlying physical . In , the Napierian attenuation coefficient relates to the complex \tilde{n} = n_r + i \kappa, where the imaginary part \kappa () is given by \kappa = \alpha_N \lambda / (4\pi), with \lambda the in ; this connection describes how material absorption attenuates propagating fields. This formulation is essential for analyzing light-matter interactions in dispersive media, linking macroscopic attenuation to microscopic quantum processes.

Decadic and Decibel Attenuation

The decadic attenuation coefficient, denoted \alpha_{10}, describes the attenuation of I through a medium over a x via the relation I = I_0 10^{-\alpha_{10} x}, where I_0 is the initial . This coefficient is related to the Napierian attenuation coefficient \alpha_N by \alpha_{10} = \alpha_N / \ln(10), providing a base-10 logarithmic measure convenient for experimental and practical calculations. Attenuation expressed in decibels (dB) follows from the power ratio as $10 \log_{10}(I_0 / I) = 10 \alpha_{10} x. Equivalently, in terms of the Napierian coefficient, this becomes $10 \alpha_N x \log_{10}(e) \approx 4.343 \alpha_N x, where \log_{10}(e) \approx 0.4343. This formulation arises directly from the logarithmic nature of the intensity decay, enabling straightforward scaling for path length in measurements. In telecommunications, particularly fiber optics, attenuation is routinely specified in dB per kilometer (dB/km) to quantify signal loss over long distances. For instance, standard single-mode silica fibers exhibit an attenuation of approximately 0.2 dB/km at the 1550 nm wavelength, the primary telecommunications band, due to minimized Rayleigh scattering and material absorption. This low-loss characteristic supports transcontinental data transmission with minimal repeaters. The scale's logarithmic basis aligns with human perceptual responses in both acoustics, where scales logarithmically with , and electromagnetics, where it accommodates the vast of signal strengths from noise floors to peak powers.

Transmission and Extinction Coefficients

The T, also known as , quantifies the fraction of incident that emerges from a medium after traversing a path length x. It is defined as T = \frac{I}{I_0} = e^{-\alpha x}, where I is the transmitted , I_0 is the incident , and \alpha is the attenuation coefficient. This expression directly inverts the attenuation law, illustrating how decreases exponentially with increasing path length or attenuation strength, assuming no or other effects. In , the often serves as a for the attenuation coefficient \alpha, particularly when describing total loss in a medium. More specifically, it refers to the imaginary part k of the complex refractive index \tilde{n} = n + i k, where n is the real part representing the shift. The relationship between these quantities is given by \alpha = \frac{4\pi k}{\lambda}, with \lambda denoting the , or equivalently k = \frac{\alpha \lambda}{4\pi}. This formulation arises from the wave equation in absorbing media, where the imaginary component introduces in the amplitude. The accounts for both and mechanisms contributing to overall , as the total is the sum of absorptive and scattering coefficients. In thin-film optics, it plays a in designs for applications such as mirrors and filters, where minimizing k ensures low losses in multilayer stacks to achieve desired reflectivity or transmissivity. For instance, materials like HfO₂ are selected for based on their low extinction coefficients to maintain high performance across spectral bands.

Comparisons with Other Optical Coefficients

The attenuation coefficient, often denoted as \alpha or c in , quantifies the total loss of through a medium due to both and , expressed as I(z) = I_0 e^{-\alpha z}, where I(z) is the after z. In , the reflectivity R measures the of incident reflected at a surface or interface, defined as R = I_r / I_i, where I_r and I_i are the reflected and incident intensities, respectively; this is primarily a surface property governed by and independent of propagation depth. While reflectivity describes immediate bounce-back at boundaries, the attenuation coefficient governs volumetric decay inside the material, with the two interacting in multilayer systems where reflected light may still undergo attenuation if re-entering the medium. The single-scattering \omega_0, a dimensionless between 0 and 1, represents the ratio of to total , given by \omega_0 = \sigma_s / (\sigma_a + \sigma_s) = b / c, where \sigma_s and \sigma_a are the and cross-sections, b is the , and c is the total (extinction) . This contrasts with the by focusing on the relative contribution of versus within the total extinction process; for example, in clear ocean water at 514 , \omega_0 \approx 0.25, indicating absorption dominance, whereas in -dominated like turbid waters, \omega_0 approaches 1. In models, high \omega_0 implies more light redirection than permanent loss, aiding applications like atmospheric studies. Emissivity \epsilon, the ratio of emitted by a surface to that of a blackbody at the same and , is linked to via Kirchhoff's law of thermal , which states that for a body in , \epsilon(\lambda) = \alpha(\lambda), where \alpha(\lambda) is the absorptivity (fraction of incident absorbed). For opaque bodies, absorptivity relates inversely to reflectivity and transmissivity, but the attenuation coefficient \alpha (absorption component) influences this through the of transmitted , T = e^{-\alpha d}, where d is thickness; thus, high attenuation implies strong and, by extension, high emissivity in equilibrium. This connection is fundamental in thermal optics, ensuring energy balance between and . In astronomy, attenuation coefficients model interstellar , where the \tau_\lambda = \int \alpha_\lambda \, ds describes the cumulative loss along a , often following empirical laws like A_\lambda = 1.086 \tau_\lambda in magnitudes, with \alpha_\lambda peaking in the UV due to dust grain properties. These are compared to bolometric corrections, which adjust observed magnitudes to total luminosities by accounting for ; for instance, tables of coefficients A_\lambda / E(B-V) (typically R_V \approx 3.1) enable corrections for reddening, ensuring accurate rate estimates from UV-to-IR fluxes reprocessed by .

References

  1. [1]
    [PDF] 10. GLOSSARY
    Attenuation Coefficient—The fractional reduction in the intensity of a beam of radiation as it passes through an absorbing medium. It may be expressed as ...
  2. [2]
    Transmitted Intensity and Linear Attenuation Coefficient - NDE-Ed.org
    The linear attenuation coefficient (µ) describes the fraction of a beam of x-rays or gamma rays that is absorbed or scattered per unit thickness of the absorber ...
  3. [3]
    NIST: X-Ray Mass Attenuation Coefficients - Section 2
    The attenuation coefficient, photon interaction cross sections and related quantities are functions of the photon energy.
  4. [4]
    NIST: X-Ray Mass Attenuation Coefficients - Introduction
    Mass attenuation coefficients (μ/ρ) and mass energy-absorption coefficients (μen/ρ) are used to calculate photon penetration and energy deposition in materials.
  5. [5]
    The AAPM/RSNA physics tutorial for residents. X-ray attenuation
    An attenuation coefficient is a measure of the quantity of radiation attenuation by a given thickness of absorber. Linear and mass attenuation coefficients are ...
  6. [6]
    [PDF] Chapter 8 – Absorption and Attenuation of Sound
    Attenuation includes absorption, which converts sound energy to heat, and scattering, which redirects energy. Absorption includes viscous and heat conduction ...
  7. [7]
    [PDF] Lab 4: Scattering, backscattering and beam attenuation 22 July 2021
    Jul 22, 2021 · Is the spectral particulate attenuation coefficient well fitted by a power-law function? Is the spectral particulate scattering coefficient ...
  8. [8]
    The Bouguer‐Beer‐Lambert Law: Shining Light on the Obscure - PMC
    The Beer‐Lambert law is unquestionably the most important law in optical spectroscopy and indispensable for the qualitative and quantitative interpretation ...
  9. [9]
    [PDF] Charles G. Barkla - Nobel Lecture
    Barkla's lecture covered his experiments on the Quantum Theory of Radiation, scattering of X-rays, and the discovery of a J series of characteristic radiations.
  10. [10]
    Fiber Attenuation Coefficient - an overview | ScienceDirect Topics
    Fiber attenuation coefficient is defined as a measure of how much optical power is lost per unit length of optical fiber, primarily due to factors such as ...
  11. [11]
    PET attenuation coefficients from CT images - PubMed
    The purpose of this study was to compare this transformation with experimental CT values and corresponding PET attenuation coefficients. In 14 patients, ...
  12. [12]
    Solar attenuation by aerosols: An overview - ScienceDirect.com
    This work provides an overview of the effect of aerosols on solar radiation budget by considering two common turbidity parameters.
  13. [13]
    Investigation of gamma-ray shielding parameters of bismuth ... - NIH
    Nov 22, 2022 · The mass attenuation coefficient measures the strength of interaction of gamma photons with the material medium and hence it depends on the ...
  14. [14]
  15. [15]
    [PDF] Photon cross sections, attenuation coefficients, and energy ...
    ... attenuation coefficient for 23elements (iH to 92 U) between 10 keV and 100 ... wave- lengths) include, among others, those by Birks. [1963],. Heinrich.
  16. [16]
    Spectral attenuation coefficients from measurements of light ... - TC
    Apr 21, 2021 · ... (such as glacier ice) that both absorbs and scatters light is specified by the spectral attenuation coefficient: katt(λ)=kabs(λ)+ksca(λ),. where ...
  17. [17]
    [PDF] Regular Spectral Transmittance
    where μ(λ) is the spectral attenuation coefficient, an intrinsic property of the medium, and d is the distance through the object, i.e., thickness of object ...
  18. [18]
    Experimental measurements of the spectral absorption coefficient of ...
    An experimental method for measuring the spectral absorption coefficient of infrared optical fibers using FTIR spectroscopy is presented.
  19. [19]
    [PDF] Absorption Spectrum of Water Vapor Between 4.5 and 13 Microns 1
    InvestiO'ations of the infrared spectrum of sunlight between 7 and 13 f.L have shown that essentially all the structure is due 1,0 absorptioll by polY'atomic.
  20. [20]
    Attenuation processes of solar radiation. Application to the ...
    The aim of this paper is to analyze the attenuation processes of the solar radiation and to review the scientific works in this field, specifically the ...
  21. [21]
    Remote sensing of normalized diffuse attenuation coefficient of ...
    Aug 25, 2016 · It determines the decrease in solar energy (required for photosynthesis and heating in the upper water column) with the increase in depth.
  22. [22]
    Absorption Coefficient - RP Photonics
    An absorption (or attenuation) coefficient is a logarithmic measure for the distributed absorption in a medium.
  23. [23]
    Absorption and Scattering - SPIE Digital Library
    Here, α(λ)=Aa+Sa α ( λ ) = A a + S a [1/km] is the extinction coefficient, where Aa is the absorption coefficient, and Sa is the scattering coefficient.
  24. [24]
    Optical Properties of Silicon - PVEducation
    The absorption depth is the inverse of the absorption coefficient. An absoption depth of, for example, 1 um means that the light intensity has fallen to 36% (1 ...
  25. [25]
    Definition and units of scattering coefficient - OMLC
    The scattering coefficient μ s [cm -1 ] describes a medium containing many scattering particles at a concentration described as a volume density s [cm 3 ].
  26. [26]
    The Volume Scattering Function (VSF) - Ocean Optics Web Book
    May 19, 2021 · Scattering Coefficients​​ Integrating β ( ψ , λ ) over all directions (solid angles) gives the total scattered power per unit incident irradiance ...
  27. [27]
    [PDF] Mie Scattering Theory - DESCANSO
    Stationary phase points for the total scattering integral near = xo. : (a) Nxo= +83p , (b) near Nxo = -83p , and (c) Nxo. = 830p. Points occur at ...
  28. [28]
    Mie Theory Approximations - Ocean Optics Web Book
    Nov 13, 2021 · This is to emphasize that “Rayleigh scattering,” “Rayleigh-Gans scattering, and Geometric optics are not physically different types of scattering ...Missing: mechanisms | Show results with:mechanisms
  29. [29]
    Review on Parameterization Schemes of Visibility in Fog and Brief ...
    Dec 11, 2021 · The VIS in fog is affected by the extinction of fog droplets [69,70,76,84], and based on Mie scattering theory, the extinction coefficient is ...
  30. [30]
    Cross Section (sigma = Mu/n) - Physics Resource - Tutor Hunt
    Nov 14, 2015 · Cross Section ... In the formula sigma=mu/n, n is the number density of the target particles (SI units: m-3) mu is the attenuation coefficient ...
  31. [31]
    Macroscopic Cross Section - an overview | ScienceDirect Topics
    ... cross-section, σ, for a particular reaction, and can be expressed as Σ = N σ. For mixtures of elements or compounds, it is calculated as Σ = N₁σ₁ + N₂σ₂ + ...
  32. [32]
    X-ray Absorption and Fluorescence - Ug11bm
    Sep 26, 2012 · The X-ray beam intensity I(x) at depth x in a material is a function of the attenuation coefficient mu, and can be calculated by the Beer-Lambert law.
  33. [33]
    [PDF] 32 - Shielding Radiation. - Nuclear Regulatory Commission
    The mass attenuation coefficient (u/D) is the probability of an interaction (PE, CS, PP) per unit density thickness. Its units are cm2/g. In other words, u/D is ...
  34. [34]
    Comparison of gamma spectrometric method and XCOM method in ...
    Since the μM value is independent of the physical state and density of the material, it is a much more useful expression in comparing the attenuation ...
  35. [35]
    NIST Atomic Form Factors: Form factors and standard definitions
    The mass attenuation coefficient is conventionally given by the symbol [ /ρ] = σ/uA, where σ is the cross-section in barns/atom, u is the atomic mass unit, and ...Missing: formula | Show results with:formula
  36. [36]
    X-Ray Mass Attenuation Coefficients - Lead - NIST
    Lead Z = 82. HTML table format. Energy, μ/ρ, μen/ρ. (MeV), (cm2/g), (cm2/g). 1.00000 ... 5.658E-02, 3.478E-02. 2.00000E+01, 6.206E-02, 3.595E-02. Lead Z = 82
  37. [37]
    MHz free electron laser x-ray diffraction and modeling of pulsed ...
    Sep 5, 2023 · for a constant (Napierian) attenuation coefficient μ k ⁠, which depends naturally on the radiation beam frequency content. Here, ...
  38. [38]
    Absorbance, absorption coefficient, and apparent quantum yield
    Absorbance (D or OD) is the log10 of the ratio of incident to transmitted light. Absorption coefficient (a) is the absorbed portion per unit pathlength.
  39. [39]
    Absorbance, Absorption Coefficient, and Apparent Quantum Yield
    Aug 10, 2025 · where, a [m −1 ] is referred to as the Napierian absorption coefficient, representing the absorbed portion of the infinitesimally thin unit ...
  40. [40]
  41. [41]
    31 The Origin of the Refractive Index - Feynman Lectures
    We see that the imaginary part n″ of a complex index of refraction represents an absorption (or “attenuation”) of the wave. In fact, n″ is sometimes referred ...
  42. [42]
    [PDF] SOLID STATE PHYSICS PART II Optical Properties of Solids - MIT
    ˜n is the index of refraction and ˜k is the extinction coefficient. (We use the tilde over the optical constants ˜n and ˜k to distinguish them from the ...
  43. [43]
    Absorbance – logarithmic, transmittance, coefficient, attenuance
    In some cases, one uses a decadic absorption coefficient, which is smaller by the factor ln ⁡ 10 , so that the absorbance is simply that coefficient times ...
  44. [44]
    Decibel – gain, loss, dB, dBm, dBc/Hz - RP Photonics
    The number of decibels is 10 times the logarithm (to base 10) of the power amplification factor or loss factor, or alternatively 20 times the logarithm of the ...
  45. [45]
    Low-Loss Optical Fiber - an overview | ScienceDirect Topics
    In 1976, fiber with loss of 0.47 dB/km at 1200 nm was demonstrated [3]. Within three years, the fiber attenuation reached 0.2 dB/km at 1550 nm [5], which is the ...
  46. [46]
    [PDF] The Decibel Scale
    Expressing sound levels in decibels is also appropriate because human loudness perception tends to be logarithmic: a 10dB increase in SPL for a pure tone ...
  47. [47]
    [PDF] Beam Transmission and Attenuation Coefficients: Instruments - IOCCG
    Beam transmittance 𝑇(𝜆, 𝑟&) over an optical path of length 𝑟& [m], and the beam attenuation coefficient 𝑐(𝜆) [𝑚-1], are the probably the most used optical ...<|separator|>
  48. [48]
    [PDF] Lecture 14 - Optical and Photonic Glasses
    Introducing the absorption coefficient, α and the dimentionless extinction coefficient, k, one has: n* = n –i k α = 4 π k / λ. (in units of cm-1). Also: ε r.<|separator|>
  49. [49]
    Light and Materials - J.A. Woollam
    Light loses intensity in an absorbing material according to Beer's Law: Thus, the extinction coefficient relates how quickly light vanishes in a material. ...
  50. [50]
    Extinction Coefficient - an overview | ScienceDirect Topics
    The atmospheric extinction coefficient (bext), which is the sum of scattering and absorption by particles and gases, is a measure of the alteration of radiant ...
  51. [51]
    Thin film optical coatings for the ultraviolet spectral region
    Nov 21, 2017 · The aim is to design and realize high-quality and high-reflectivity UV coatings. ... Thin film HfO2 refractive index and extinction coefficient.
  52. [52]
    [PDF] Ch.3: Optical Properties of Water - MISC Lab
    The downwelling diffuse attenuation coefficient Kd(z;8) is defined in terms of the decrease with depth of the ambient downwelling irradiance Ed(z;8), which.
  53. [53]
    [PDF] CHAPTER 3 ABSORPTION, EMISSION, REFLECTION, AND ...
    This is often referred to as Kirchhoff's Law. In qualitative terms, it states that materials that are strong absorbers at a given wavelength are also strong ...
  54. [54]
    [PDF] Lecture 3 : Modelling stellar populations
    Interstellar dust grains scatter and absorb photons. These two processes are combined to produce the interstellar extinction. Iλ = Iλ0 exp(-τλ), Iλ0 is the ...
  55. [55]
    [PDF] arXiv:0804.0498v1 [astro-ph] 3 Apr 2008
    Apr 3, 2008 · In this paper we present tables of bolometric corrections and interstellar extinction coefficients, as a function of Teff, log g, and [M/H] ...