Fact-checked by Grok 2 weeks ago

Extinction coefficient

The extinction coefficient, also known as the or molar attenuation coefficient, is a quantitative measure of the ability of a to absorb at a specific , normalized to its concentration and the path length of the through the sample. It serves as a key parameter in spectrophotometric analysis, enabling the determination of substance concentrations in solutions via the Beer-Lambert law, and is particularly valuable in fields such as biochemistry for characterizing proteins and dyes. Although the term "extinction coefficient" is now considered obsolete by the International Union of Pure and Applied Chemistry (IUPAC) in favor of "molar absorption coefficient," it remains widely used in to describe absorption-dominated processes in dilute solutions where scattering is negligible. In mathematical terms, the extinction coefficient ε (λ) at λ is defined as ε (λ) = A (λ) / (c · l), where A (λ) is the (dimensionless), c is the amount concentration (typically in dm⁻³), and l is the (in cm). The standard units are L ⁻¹ cm⁻¹ (or M⁻¹ cm⁻¹), reflecting its role in the decadic logarithm form of the Beer-Lambert law: A = ε · c · l. Values of ε vary widely depending on the and ; for example, strong chromophores like aromatic compounds can have ε exceeding 10⁴ L ⁻¹ cm⁻¹, while weaker absorbers are below 10² L ⁻¹ cm⁻¹. Beyond solution-phase , the term "extinction coefficient" also appears in and to describe the combined effects of and on , often denoted as the imaginary part k (dimensionless) of the complex refractive index ñ = n + i k, or as the linear extinction coefficient with units such as cm⁻¹, where it quantifies energy loss. Accurate determination of extinction coefficients typically involves measuring at known concentrations, with corrections for impurities or instrumental errors to ensure precision in applications like and pharmaceutical analysis.

Fundamentals

Definition

The extinction coefficient is a fundamental parameter that quantifies the of , particularly , as it propagates through a , accounting for both and effects. This arises from interactions between the and the medium's particles or molecules, leading to a reduction in along the propagation path. In its general form, the extinction coefficient \alpha describes the linear decay of I with distance x traveled in the medium, following the relation I = I_0 e^{-\alpha x}, where I_0 is the initial intensity and \alpha has units of inverse length (e.g., cm^{-1}). This formulation integrates the combined contributions of the absorption coefficient and the scattering coefficient, with the total extinction coefficient defined as \alpha = \alpha_\text{abs} + \sigma_\text{scat}. Although it incorporates scattering, the extinction coefficient often emphasizes absorption in contexts where scattering is minimal, such as in dilute or homogeneous media. The term "extinction coefficient" emerged in scientific literature in the early , specifically around 1902, within the domains of and , evolving from 18th-century concepts of light attenuation pioneered by and . In specialized applications, it may refer to the molar extinction coefficient in spectroscopic analysis of solutions or to the imaginary component of the complex in optical materials.

Units and Notation

The extinction coefficient is represented by various notations across scientific disciplines to distinguish its specific form and application. In molecular , the molar extinction coefficient is typically denoted by the Greek letter ε (epsilon). In optical contexts, the extinction coefficient is often symbolized as κ (kappa) or sometimes k, representing the imaginary component of the complex refractive index. The linear , which quantifies extinction per unit length, is commonly denoted by α (alpha). Units for the extinction coefficient vary by type, reflecting its physical . The linear extinction coefficient α has units of inverse , such as cm⁻¹ in traditional literature or m⁻¹ in units. The extinction coefficient ε is expressed in L ⁻¹ cm⁻¹ (liters per per centimeter) or equivalently M⁻¹ cm⁻¹ (inverse molarity per centimeter), though the unit is m² ⁻¹. In , the extinction coefficient κ is dimensionless, as it forms part of the unitless ñ = n + iκ, where n is the real . A key relation connects the linear and molar forms: the linear α equals the product of the extinction coefficient ε and the solute concentration c, given by \alpha = \varepsilon c where c is in L⁻¹, ensuring consistent units of inverse for α. This conversion is fundamental in applying spectroscopic measurements to bulk material properties. efforts favor SI units like m⁻¹ for α and m² ⁻¹ for ε, but non-SI conventions persist widely in experimental literature due to historical practices in . All forms of the extinction coefficient are inherently wavelength-dependent, often notated as ε(λ), κ(λ), or α(λ) to emphasize this variation in and behavior across the .

In Spectroscopy

Molar Extinction Coefficient

The extinction coefficient, denoted as \epsilon, quantifies the probability of a absorbing at a specific , serving as an intrinsic property of the absorbing species that remains independent of its concentration in solution. It reflects the strength of transitions within the , where higher values indicate more efficient due to favorable overlap between ground and excited states. Several factors influence the magnitude of \epsilon, primarily the nature of electronic transitions such as \pi \to \pi^* or n \to \pi^*, which are modulated by chromophores—structural units like conjugated double bonds or aromatic rings that enable in the UV-visible range. The , proportional to the square root of \epsilon, determines intensity, with extended conjugation typically increasing values. further alter \epsilon by stabilizing excited states differently from ground states, often causing shifts in intensity through interactions. For dyes, \epsilon commonly ranges from $10^4 to $10^5 L mol^{-1} cm^{-1}, reflecting strong suitable for spectroscopic applications. The extinction coefficient varies with (\epsilon(\lambda)), exhibiting peaks at absorption maxima (\epsilon_{\max}) where transitions are most probable, and dropping sharply elsewhere. This dependence arises from the energy matching between incident photons and molecular orbitals, with \epsilon_{\max} providing a key metric for identifying chromophores. Representative examples illustrate these properties: displays \epsilon \approx 210 L mol^{-1} cm^{-1} at 255 , corresponding to a \pi \to \pi^* transition in its aromatic system. In , oxyhemoglobin exhibits a much higher \epsilon_{\max} \approx 4.67 \times 10^5 L mol^{-1} cm^{-1} at 410 in the Soret band, driven by porphyrin chromophores and intense metal-to-ligand charge transfer.

Beer-Lambert Law

The Beer-Lambert law, also known as Beer's law, states that the A of a is directly proportional to the concentration c of the absorbing species and the path length l through which the light passes, expressed as A = \epsilon c l, where \epsilon is the molar . This relationship forms the foundation for quantitative , allowing the determination of concentrations in . The law derives from the exponential attenuation of as it propagates through an absorbing medium. The transmitted intensity I relates to the incident intensity I_0 by I = I_0 10^{-\epsilon c l}, assuming infinitesimal slices where is proportional to the number of absorbing molecules encountered. is then defined logarithmically as A = \log_{10}(I_0 / I), yielding the linear form A = \epsilon c l upon substitution and simplification. The law holds under specific assumptions, including the use of monochromatic , a non-scattering medium, and dilute solutions where interactions between absorbing species are negligible. Limitations arise at high concentrations, where deviations up to 30% can occur due to intermolecular interactions, effects, or non-linear ; additionally, polychromatic or invalidates the proportionality. In practice, the law enables by constructing curves from known standards, plotting against concentration to determine unknown sample concentrations via , typically valid over limited ranges to avoid deviations. This approach is widely applied in UV-Vis and for tasks such as reaction kinetics monitoring and purity assessment.

In Optics

Complex Refractive Index

In optics, the , denoted as , describes the propagation and attenuation of in a and is expressed as = n + iκ, where n is the real part that governs the of the wave, and κ is the imaginary part known as the extinction coefficient that quantifies . This arises from the need to account for both dispersive and absorptive effects in where interacts with electrons, leading to a modification of the wave's and . The extinction coefficient κ physically represents the rate at which the electromagnetic wave's diminishes as it penetrates the material due to processes. It directly influences the of , often characterized by the skin depth δ, defined as \delta = \frac{\lambda}{4\pi \kappa}, where λ is the in ; this distance marks the scale over which the drops to 1/e of its surface value. Higher values of κ result in shallower penetration, which is crucial for understanding opaque behaviors in materials like metals. The wavelength dependence of κ, denoted κ(λ), is intrinsically linked to the real refractive index n(λ) via the Kramers-Kronig relations, a set of transforms derived from the principle in linear response . These relations ensure that the real and imaginary parts of ñ are not independent but connected across the entire spectrum, allowing computation of n(λ) from measured κ(λ) (or vice versa) through principal value . This highlights how features at specific wavelengths induce anomalous dispersion in n nearby. Representative examples illustrate κ's material-specific behavior. In metals like , κ is notably high across the visible range, with values typically between 2 and 3 (e.g., κ ≈ 2.3 at 500 nm), enabling strong plasmonic effects and reflectivity while limiting light penetration to tens of nanometers. In semiconductors, such as or , κ remains low below the bandgap energy but rises abruptly near and above it due to direct or indirect interband transitions, transitioning the material from transparent to absorbing over a narrow spectral window.

Relation to Absorption

The extinction coefficient κ, as the imaginary part of the complex refractive index, directly governs the of in materials, leading to an of the electromagnetic wave's as it propagates. The coefficient α, which quantifies this decay, is related to κ by the formula \alpha = \frac{4\pi \kappa}{\lambda}, where λ is the of in . This relation stems from the equation in absorbing media, where the imaginary component introduces exponential damping. Consequently, the I of a beam traversing a distance x through the material follows Beer's law: I = I_0 e^{-\alpha x}, with I_0 as the initial intensity, describing how absorbed energy reduces the transmitted light. In homogeneous bulk materials, κ primarily reflects intrinsic absorption processes, such as electronic transitions or free-carrier effects, rather than scattering. However, the broader concept of optical extinction encompasses both absorption and scattering losses, particularly in heterogeneous systems like suspensions or aerosols. For instance, in Mie theory, which models scattering by spherical particles comparable in size to the wavelength, the total extinction cross-section σ_ext is the sum of the absorption cross-section σ_abs and the scattering cross-section σ_sca, such that σ_ext = σ_abs + σ_sca; here, κ contributes mainly to σ_abs, while scattering dominates for non-absorbing particles. Material-specific values of κ illustrate these absorption characteristics. In transparent dielectrics, such as fused silica, κ remains low (typically < 0.001 in the visible spectrum), enabling high transmission with minimal attenuation over distances. In contrast, semiconductors exhibit a sharp rise in κ near their electronic bandgap, where interband transitions occur; for example, in gallium arsenide (GaAs) at the absorption edge (870 nm, 1.42 eV bandgap), κ reaches values on the order of 0.01, corresponding to α ≈ 10^4 cm^{-1} and limiting penetration depths to microns. Such behavior underscores κ's role in defining absorption edges critical for optoelectronic devices. Variations in κ arise from temperature and composition, influencing material band structure and carrier dynamics. Elevated temperatures often narrow the bandgap, red-shifting the absorption onset and increasing κ at a fixed wavelength due to enhanced phonon-assisted transitions; for instance, in silicon, κ near the bandgap can increase by a factor of several from 300 K to 500 K due to bandgap narrowing. Compositionally, alloying or doping alters κ profoundly—adding impurities like oxygen in silicon can boost free-carrier absorption, elevating κ in the infrared, while precise stoichiometric control in compound semiconductors minimizes it in transparent regions. These dependencies are essential for tailoring materials in applications like solar cells and lasers.

Measurement and Applications

Experimental Determination

The molar extinction coefficient (ε) in spectroscopy is commonly determined using UV-Vis spectrophotometry by measuring the absorbance of solutions with known concentrations. The procedure involves preparing a series of standard solutions of the analyte in a suitable solvent, typically at concentrations spanning the expected linear range of the . Absorbance spectra are recorded across the relevant wavelength range, with baseline correction performed by subtracting the absorbance of a blank solvent sample to account for instrumental and solvent contributions. Multi-wavelength fitting, such as least-squares regression on absorbance data at several wavelengths, is applied to derive ε from the slope of the calibration curve (A = ε c l, where A is absorbance, c is concentration, and l is path length), enhancing accuracy for samples with overlapping spectra. Calibration with known standards, like certified reference materials, verifies the instrument's photometric accuracy before measurement. In optics, the extinction coefficient (κ) for thin films is measured using spectroscopic ellipsometry, which probes changes in the polarization state of light reflected from the sample. The technique involves directing broadband light onto the film at a fixed angle of incidence (typically near the ) and measuring the ellipsometric parameters Ψ (amplitude ratio) and Δ (phase difference) as functions of wavelength. These data are then fitted to an optical model of the film-substrate system using regression algorithms to extract the complex refractive index ñ = n + iκ, where κ quantifies absorption. This method is particularly suited for non-destructive characterization of thin films (e.g., semiconductors or dielectrics) with thicknesses from nanometers to micrometers. For bulk materials, transmission spectroscopy determines the absorption coefficient (α), from which κ is derived via κ = α λ / (4π), where λ is the wavelength. The procedure entails measuring the transmittance T through a sample of known thickness d, correcting for surface reflections using , and calculating α = -(1/d) ln(T / (1 - R)^2), where R is reflectivity. Calibration against reference materials with established optical constants ensures reliability. Common error sources in these measurements include instrumental limitations such as resolution and stray light, which can deviate absorbance readings by up to several percent, as well as sample-related issues like impurities or non-uniformity affecting purity and homogeneity. Accuracies typically range from ±5% to 10% under standard laboratory conditions, depending on calibration rigor and instrument quality. Validation using relations like between real and imaginary refractive index parts can further confirm results.

Practical Uses

In biochemistry, the extinction coefficient at 280 nm serves as a key parameter for quantifying protein concentrations, leveraging the strong absorption by aromatic amino acids such as tyrosine and tryptophan residues. This method enables direct spectrophotometric measurement without requiring additional reagents, providing rapid and accurate assessments in protein purification and characterization workflows. For instance, the molar extinction coefficient (ε280) is calculated from the protein's amino acid sequence, allowing precise concentration determination in complex biological samples. Similarly, the A260/A280 ratio, derived from nucleic acid extinction coefficients, assesses DNA and RNA purity; pure double-stranded DNA typically exhibits a ratio of approximately 1.8, while pure RNA shows around 2.0, indicating minimal protein contamination. In , the extinction coefficient informs the design of silicon-based solar cells by highlighting low absorption in the region, where the imaginary part of the (κ) approaches zero, enabling efficient thermal management and reduced energy loss. This property supports strategies that maintain cell performance under operational heat loads. For optical coatings, the extinction coefficient guides layer selection and thickness optimization to minimize absorption losses, ensuring high or across targeted wavelengths in antireflection structures for photovoltaic and applications. Environmental monitoring utilizes the extinction coefficient of , with a molar value of 85.02 L mmol-1 cm-1 at 663.6 nm in buffered 80% acetone, to estimate algal biomass and detect blooms through spectrophotometric analysis of water samples. This approach correlates data with chlorophyll concentrations, aiding in the assessment of and degradation. In , aerosol coefficients, combining and components, quantify particulate impacts on and ; for example, stratospheric measurements from instruments like OMPS-LP track long-term trends in aerosol loading to model climate effects. As of 2025, emerging applications include plasmonics, where the extinction coefficient of nanoparticles tunes resonances for sensitive detection in chemical and biosensors, achieving shifts in extinction peaks upon analyte binding. In quantum dot displays, high molar extinction coefficients (on the order of 106 M-1 cm-1) enhance color conversion efficiency and brightness in light-emitting diodes, enabling vibrant, wide-gamut visuals with minimal reabsorption losses.

References

  1. [1]
    molar absorption coefficient (M03972) - IUPAC
    molar absorption coefficient ... Absorbance divided by the absorption pathlength, l , and the amount concentration, c : ε ( λ ) = 1 c l l g ( P λ 0 P λ ) = A ( λ ) ...
  2. [2]
    attenuation coefficient (A00516) - IUPAC
    Analogous to absorption coefficient but taking into account also the effects due to scattering and luminescence. It was formerly called extinction coefficient.Missing: definition | Show results with:definition
  3. [3]
    [PDF] THE MOLAR EXTINCTION OF RHODOPSIN BY GEORGE WALD ...
    Mola~ Extinaion.--This quantity, otherwise called the molar or molecular extinction coefficient, is defined by the equation: logzo Io/! -- • • ¢ • l, in ...
  4. [4]
    Molar Absorption Coefficients - Spectra and Spectral Data
    Aug 25, 2025 · Many students and researchers still use obsolete terms like "extinction coefficient." Here are some definitions for clarity. Molar absorption ...
  5. [5]
    [PDF] SOLID STATE PHYSICS PART II Optical Properties of Solids - MIT
    ... definition ε(ω) = [˜n(ω) + i˜k(ω)]2 where ˜n(ω) is the index of refraction and ˜k(ω) is the extinction coefficient. We can then write for the real part of ...<|control11|><|separator|>
  6. [6]
    Glossary – Improve
    Extinction coefficient: a measure of the ability of particles or gases to absorb and scatter photons from a beam of light; a number that is proportional to the ...
  7. [7]
    Accurate Measurement of Molar Absorptivities - PMC - NIH
    The key to accurate measurement of molar absorptivities is a thorough understanding of the sources of error which appear throughout the measurement procedure.
  8. [8]
    [PDF] 1 CHAPTER 5 ABSORPTION, SCATTERING, EXTINCTION AND ...
    Radiation is weakened by absorption and scattering, which combined is called extinction. The equation of transfer deals with absorption, scattering, and ...
  9. [9]
    Extinction Coefficient - an overview | ScienceDirect Topics
    The extinction coefficient is a characteristic that determines how strongly a species absorbs or reflects radiation or light at a particular wavelength. It is ...
  10. [10]
    The Beer-Lambert Law - Chemistry LibreTexts
    Jan 29, 2023 · The constant ϵ is called molar absorptivity or molar extinction coefficient and is a measure of the probability of the electronic transition. On ...
  11. [11]
    Chapter 2 Absorption of light | Photochemistry and Photophysics
    The greater the overlap integral between the ground and excited state the higher the molar extinction coefficient and the more strongly coloured the molecule.
  12. [12]
    Electronic Spectroscopy - Interpretation - Chemistry LibreTexts
    Jan 29, 2023 · The molar extinction coefficients for these transitions are around 104. Examples of pi donor ligands are as follows: F-, Cl-, Br-, I-, H2 ...
  13. [13]
    [PDF] Photophysical requirements for absorption/emission
    The molar extinction coefficient (ε) is proportional to the square of the transition dipole moment. ε ∝ µge 2 Page 5 • Schematic of the electronic component of ...
  14. [14]
    Molar Absorptivity - an overview | ScienceDirect Topics
    Molar absorptivity is the absorbance per mole per liter of a chromogen in a cuvette, influenced by concentration, wavelength, temperature, and matrix ...Charge Transfer Complexes · Identification And... · Capillary Zone...
  15. [15]
    Benzene - PhotochemCAD
    Absorption Spectrum. Absorption Wavelength, 254.75. Absorption Epsilon, 210. Absorption Coefficient, 210 at 254.75 nm. Solvent, cyclohexane. Instrument, Cary 3.
  16. [16]
    Tabulated Molar Extinction Coefficient for Hemoglobin in Water
    using 64,500 as the gram molecular weight of hemoglobin. where x is the number of grams per liter. A typical value of x for whole blood is x=150 g Hb/liter.
  17. [17]
    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 ...
  18. [18]
    Misuse of Beer–Lambert Law and other calibration curves - PMC - NIH
    Feb 2, 2022 · An example is the Beer–Lambert Law, used to predict the concentration of a new sample from its absorbance obtained by spectrometry.
  19. [19]
    Light and Materials - J.A. Woollam
    They are generally represented as a complex number. The complex refractive index (ñ) consists of the index (n) and extinction coefficient (k):Missing: definition iκ
  20. [20]
    What is refractive index? - Omega Optical
    Oct 20, 2021 · In reality, refractive index is a complex number comprised of a real part (n) and an imaginary part (k). The real part, as described above, ...
  21. [21]
    [PDF] MIT Open Access Articles Detection of shorter-than-skin-depth ...
    Jun 4, 2013 · ... skin depth δskin = λ/(4πκ) = 16.5 nm (the complex index of refraction of gold is ˜n = 1.67 + 1.94i17 at probe wavelength λ = 393 nm). The ...
  22. [22]
    [PDF] Refractive Index. - Manx Precision Optics
    The extinction coefficient determines how much the intensity of light is reduced as it penetrates a material,. i.e. the absorptance of the material. Together ...Missing: ñ = iκ
  23. [23]
    Kramers–Kronig Relations - RP Photonics
    The Kramers–Kronig relations allow one to calculate the refractive index profile and thus also the chromatic dispersion of an optical material.
  24. [24]
    User:Antoine Moreau/Proposed/Kramers-Kronig relations
    Jul 2, 2012 · If the extinction coefficient of the medium is known then the refractive index can be calculated and vice versa.
  25. [25]
    Refractive index of Au (Gold) - Johnson - RefractiveIndex.INFO
    Extinction coefficient[ i ]. k = 2.9278. Created with Highcharts 5.0.14 ... Unlike many other metals, gold's optical properties are relatively consistent over a ...
  26. [26]
    Absorption Coefficient | PVEducation
    The absorption coefficient determines how far into a material light of a particular wavelength can penetrate before it is absorbed.
  27. [27]
    Mie Scattering Theory - an overview | ScienceDirect Topics
    Note that, the absorption cross section is the difference of the extinction and scattering cross sections. The variations of the extinction cross-section σe ...
  28. [28]
  29. [29]
    Optical Properties of Silicon - PVEducation
    Silicon's optical properties include absorption coefficient, absorption depth, real and imaginary refractive index components, and extinction coefficient. ...
  30. [30]
    Temperature Dependence of Absorption Coefficients - IOP Science
    The absorption continuum slightly decreases as the temperature decreases; the variations are less than 10% in the millimeter region at 24 K. More evident ...<|control11|><|separator|>
  31. [31]
    Self-consistent optical parameters of intrinsic silicon at 300 K ...
    The extinction coefficient, k, and α are directly related, α=4πk/λ, where λ is free-space wavelength (λ and energy, E, were converted using λ (μm)=1.23984190/E ...
  32. [32]
    None
    ### Summary: Experimental Determination of Extinction Coefficient Using UV-Vis Spectrophotometry
  33. [33]
    [PDF] Errors in spectrophotometry and calibration procedures to avoid them
    The individual sources of error and possible test methods are discussed in the following, with special consideration of two points of view: fundamental tests ...
  34. [34]
    Spectroscopic Ellipsometry: Advancements, Applications and Future ...
    Dec 6, 2023 · Ellipsometry provides a non-destructive means of determining a thin film's thickness, refractive index, and extinction coefficient [24].
  35. [35]
    Spectroscopic Ellipsometry: Basic Concepts - HORIBA
    It is important to note that spectroscopic ellipsometry is an indirect technique which does not measure thin film thickness and/or optical properties directly.
  36. [36]
    Sequence-specific determination of protein and peptide ... - NIH
    Thus, the molar absorptivity (extinction coefficient) for a protein at 280 nm (ε280) can be accurately estimated directly from its amino acid sequence.– ...
  37. [37]
    Determination of the protein content of complex samples by aromatic ...
    Mar 23, 2022 · In contrast, the absorbance at 280 nm is mainly caused by the aromatic amino acids tryptophan and tyrosine. Thus, the 280-nm absorbance ...
  38. [38]
    6. ANALYSIS OF PROTEINS
    Direct measurement at 280nm. Tryptophan and tyrosine absorb ultraviolet light strongly at 280 nm. The tryptophan and tyrosine content of many proteins ...
  39. [39]
    [PDF] A Practical Guide to Analyzing Nucleic Acid Concentration ... - NEB
    In buffered solutions, pure dsDNA has an A260/ A280 of 1.85–1.88 and pure RNA has a ratio of around 2.1. The A260/A230 is a sensitive indicator of contam- ...
  40. [40]
    [PDF] 260/280 and 260/230 Ratios
    260/280 Ratio. The ratio of absorbance at 260 nm and 280 nm is used to assess the purity of DNA and RNA. A ratio of ~1.8 is generally accepted as. “pure” for ...
  41. [41]
    Passive Radiative Cooling of Silicon Solar Modules with Photonic ...
    The silicon solar cell itself has zero absorptivity (extinction coefficient) throughout the IR wavelength range (see Supporting Information, Figure S2), and ...
  42. [42]
    Optimization of amorphous silicon solar cells through photonic ...
    May 13, 2025 · In addition, the extinction coefficient of is relatively to zero as the material is a dielectric with zero absorption in the visible spectrum ...
  43. [43]
    Refractive indices and extinction coefficients of polymers for the mid ...
    Refractive indices and extinction coefficients of several polymers have been measured for the wide infrared wavelength region. The polymers used as coating ...
  44. [44]
    [PDF] Determination of accurate extinction coefficients and simultaneous ...
    Each coefficient is the mean of three separate determinations. The percentage variation about the mean (m) is presented as lO0.o/m, where o represents the ...
  45. [45]
    [PDF] Algal chlorophylls: a synopsis of analytical methodologies - CIGLR
    Remarkably, a consistency among (estimates of) extinction coefficients for diverse chlorophylls exists, indicating a robust, quantitative quality to the ...
  46. [46]
    [PDF] OMPS-LP aerosol extinction coefficients and their applicability in ...
    Jan 16, 2025 · While SAGE III/ISS aerosol extinction co- efficient measurements have been extensively employed for validation, comparison, and long-term ...
  47. [47]
    [PDF] Scattering and Absorbing Aerosols in the Climate System
    aerosol scattering and absorption alters the radiation balance and atmospheric stability through. 80 perturbations to the vertical temperature ...
  48. [48]
    Size, Composition, and Phase-Tunable Plasmonic Extinction in Au ...
    Jun 9, 2025 · The synthesis of Au–Sn nanoparticles with tailorable localized surface plasmon resonances (LSPR) is explored, with a focus on size dependence, composition, and ...
  49. [49]
    Sensing of Organic Vapors with Plasmonic Distributed Bragg ...
    Apr 24, 2025 · The extinction coefficient exhibits a narrow LPSR peak characteristic of silver nanoparticles at 400 nm as well as the presence of a ...
  50. [50]
    Gram-scale synthesis of single-crystalline graphene quantum dots ...
    Oct 28, 2014 · ... absorption edge at ~470 nm. They have a large molar extinction coefficient (εm) of ~106 M−1 cm−1 within the visible region, even larger ...
  51. [51]
    Influence of Mn and Co ions co-doping on the photovoltaic ... - Nature
    Oct 14, 2025 · Also, the quantum dots have higher molar extinction coefficients, which absorbs more light per unit concentration falling on it. In the present ...