Fact-checked by Grok 2 weeks ago

Reflectance

Reflectance is the ratio of the radiant flux reflected from a surface to the incident radiant flux upon it, defined as a dimensionless quantity that quantifies the proportion of electromagnetic radiation, such as light, bounced back by a material. For non-fluorescent materials, this ratio ranges from 0 (complete absorption) to 1 (perfect reflection), adhering to energy conservation principles. In and physics, reflectance is wavelength-dependent and influenced by factors including the material's , , angle of incidence, and surface geometry. It manifests in two primary forms: specular reflectance, where light reflects mirror-like from smooth surfaces at equal angles of incidence and , and diffuse reflectance, where light in multiple directions from rough or irregular surfaces. Advanced quantities, such as the (BRDF), describe angular dependencies for precise modeling of scattering behavior. Reflectance measurements are fundamental across scientific and fields, enabling characterization of material properties in , such as coatings and thin films. In , spectral reflectance data from satellites like NASA's instrument assess land cover, vegetation health, and Earth's for climate modeling. Biomedical applications use to detect tissue abnormalities, including ischemia and neoplasia, through non-invasive optical probes. Additionally, in , it aids in analyzing surface compositions of extraterrestrial bodies via spectral signatures.

Fundamentals

Definition and Basic Principles

Reflectance is defined as the fraction of incident electromagnetic radiant power, or , that is reflected by a surface at the boundary between two media. It is a , typically denoted by ρ, and ranges from 0, indicating perfect with no , to 1, representing perfect of all incident power. The basic equation for total reflectance is given by \rho = \frac{P_\text{reflected}}{P_\text{incident}}, where P_\text{reflected} is the reflected radiant power and P_\text{incident} is the incident radiant power. This formulation arises from principles in and applies to optical across the . The value of reflectance depends on several key factors, including the of the incident , the state of the , the angle of incidence, and the physical and chemical properties of the surface, such as its roughness, composition, and microstructure. For instance, at interfaces involving dielectrics or metals, these dependencies lead to variations that can be described through more specialized functions, though the core concept remains the ratio of reflected to incident power. The term reflectance has been formalized in international standards, such as ISO 80000-7, which specifies quantities for and optical , including the spectral variant ρ(λ) as the ratio of reflected to incident spectral . Early studies on laws date back to Pierre Bouguer's 1729 work Essai d'optique sur la gradation de la lumière, where he conducted the first goniophotometric measurements of reflectance at varying angles of incidence for surfaces like and , laying foundational principles for quantitative photometry. Spectral reflectance curves illustrate this wavelength dependence for common materials used in mirrors. For example, evaporated aluminum coatings show high reflectance, often exceeding 90%, from the ultraviolet through the visible spectrum up to about 2 μm in the near-infrared, making it suitable for broadband applications. Silver coatings achieve even higher reflectance, typically above 95% in the visible range (400–700 nm) and extending into the near-infrared, though they oxidize more readily. In contrast, gold mirrors exhibit lower reflectance in the visible (around 40–50% at 500 nm) but increase sharply to over 98% in the mid-infrared beyond 2 μm, due to the material's electronic structure. These behaviors are critical for selecting materials in optical systems, as documented in standard reference measurements.

Reflectance versus Reflectivity

Reflectivity refers to the intrinsic property of a to reflect , defined as the reflectance of a layer thick enough that further increases in thickness do not change the value. In contrast, reflectance is the measured ratio of the reflected by a specific sample to the incident , incorporating effects such as sample thickness, surface imperfections, and . For homogeneous, thick samples, the two terms are equivalent, but they diverge for thin films or layered structures where and multiple internal influence the overall reflection. The (CIE) formalizes this distinction in its International Lighting Vocabulary, adopting a convention where terms ending in "-ivity" (e.g., reflectivity) describe generic material properties, while those ending in "-ance" (e.g., ) apply to specific samples under defined conditions of spectral composition, , and . This usage addresses real-world measurements that account for deviations from ideal behavior, such as or in non-ideal surfaces. Reflectivity at a is fundamentally determined by the Fresnel reflection . For incidence, the reflection r is r = \left| \frac{n_2 - n_1}{n_2 + n_1} \right| where n_1 and n_2 are the refractive indices of the incident and second medium, respectively; the corresponding power reflectivity is then R = |r|^2. For a glass-air (n_1 \approx 1.5, n_2 = 1), this yields R \approx 4\% at incidence, representing the single- reflectivity. Measured reflectance for an actual plate, however, would be higher due to contributions from both air- interfaces and partial internal reflections. Standards like ISO 80000-7 define reflectance within radiometric quantities as the of reflected to incident radiant power or , emphasizing its role in and measurements. The International Union of Pure and Applied Chemistry (IUPAC), in its Gold Book, equates reflectivity with reflectance, describing both as the fraction of incident reflected by a surface or discontinuity.

Mathematical Formulations

Hemispherical Reflectance

Hemispherical reflectance, also known as hemispherical-hemispherical reflectance, is defined as the ratio of the total reflected by a surface to the total incident upon it from all directions within the above the surface, assuming diffuse or integrated illumination over that . This quantity is dimensionless, as both fluxes are measured in watts ( unit for ), and it represents the overall reflective efficiency of a surface under broad, uniform incident . The key equation for hemispherical reflectance R is given by R = \frac{\Phi_e^r}{\Phi_e^i}, where \Phi_e^r is the total reflected and \Phi_e^i is the total incident . This formulation assumes uniform illumination across the incident hemisphere, and for ideal conservative surfaces that do not absorb energy, R has an upper limit of 1, meaning all incident flux is reflected without loss. The derivation of involves integrating the reflected radiant power over the entire reflected , which spans a \Omega = 2\pi steradians. Specifically, the total reflected \Phi_e^r is obtained by integrating the reflected radiance L_r weighted by the cosine of the polar angle and the differential : \Phi_e^r = \int_{\Omega} L_r(\theta_r, \phi_r) \cos \theta_r \, d\omega_r, normalized by the incident to yield R. This accounts for the of reflected light across all outgoing directions, providing a comprehensive measure for surfaces under isotropic incident conditions. In applications such as systems, hemispherical reflectance is crucial for assessing the performance of reflective surfaces in concentrating (CSP) collectors, where it directly influences the of redirecting to receivers. For example, typical white paints used in solar reflectors exhibit hemispherical reflectance values of 0.8 to 0.9 across visible and near-infrared wavelengths, enabling high energy capture while minimizing absorption.

Directional-Hemispherical Reflectance

Directional-hemispherical reflectance, denoted as \rho(\theta_i, \phi_i), is defined as the ratio of the radiant flux reflected by an opaque surface into the entire outgoing hemisphere to the radiant flux incident on the surface from a specific direction specified by polar angle \theta_i and azimuthal angle \phi_i, assuming a collimated or directional beam illumination. This measure captures the total reflected energy for a given incident direction, integrating over all possible reflection directions in the hemisphere above the surface. The quantity is mathematically formulated in terms of radiance as \rho(\theta_i, \phi_i) = \frac{1}{L_e^i \cos \theta_i} \int_{\Omega_r} L_e^r (\theta_r, \phi_r) \cos \theta_r \, d\Omega_r, where L_e^i is the incident radiance from direction (\theta_i, \phi_i), L_e^r (\theta_r, \phi_r) is the reflected radiance in direction (\theta_r, \phi_r), \cos \theta_i normalizes for the projected incident area, \cos \theta_r accounts for the projected reflected area, and the integral is over the reflected hemisphere \Omega_r. It relates directly to the (BRDF), f_r(\theta_i, \phi_i; \theta_r, \phi_r), through \rho(\theta_i, \phi_i) = \int_{\Omega_r} f_r (\theta_i, \phi_i; \theta_r, \phi_r) \cos \theta_r \, d\Omega_r, since the reflected radiance follows L_e^r = f_r L_e^i \cos \theta_i. This integration of the BRDF over the reflection hemisphere provides a directional analog to broader reflectance metrics, emphasizing angle-specific surface response without deriving the full BRDF here. The value of \rho(\theta_i, \phi_i) varies with the incident angles, depending on the material's scattering properties, and is bounded between 0 and 1 to satisfy , with a maximum of 1 for perfectly reflecting surfaces. For specular surfaces, it remains near 1 across incident directions, though the reflected flux distribution peaks sharply at the angle, concentrating energy in a narrow lobe. In , directional-hemispherical reflectance is essential for modeling the of terrestrial surfaces as viewed by satellites under specific solar illumination geometries.

Spectral Reflectance

Spectral reflectance describes the fraction of incident that is reflected from a surface as a function of \lambda or \nu, providing a wavelength-resolved measure essential for characterizing or dispersive materials where properties vary across the . This spectral dependence arises from material-specific interactions, such as by electronic transitions in the visible range or in the , enabling precise optical analysis in fields like and material science. The spectral hemispherical reflectance, which integrates reflections over the entire outgoing , is defined as R_\lambda(\lambda) = \frac{\Phi_{e,\lambda}^r (\lambda)}{\Phi_{e,\lambda}^i (\lambda)}, where \Phi_{e,\lambda}^r (\lambda) is the spectral reflected and \Phi_{e,\lambda}^i (\lambda) is the spectral incident at \lambda. Similarly, the spectral directional-hemispherical reflectance accounts for incident direction and integrates over the reflection : \rho(\theta_i, \phi_i; \lambda) = \frac{1}{L_{e,\lambda}^i(\theta_i, \phi_i) \cos \theta_i} \int_{\Omega_r} L_{e,\lambda}^r (\theta_r, \phi_r; \lambda) \cos \theta_r \, d\Omega_r, where L_{e,\lambda}^i(\theta_i, \phi_i) is the incident spectral radiance from direction (\theta_i, \phi_i), L_{e,\lambda}^r (\theta_r, \phi_r; \lambda) is the reflected spectral radiance in direction (\theta_r, \phi_r), \cos \theta_i normalizes for the projected incident area, \cos \theta_r accounts for the projected reflected area, and the integral is over the reflected hemisphere \Omega_r = 2\pi sr. It relates to the spectral bidirectional reflectance distribution function (BRDF), f_{r,\lambda}(\theta_i, \phi_i; \theta_r, \phi_r; \lambda), through \rho(\theta_i, \phi_i; \lambda) = \int_{\Omega_r} f_{r,\lambda} (\theta_i, \phi_i; \theta_r, \phi_r; \lambda) \cos \theta_r \, d\Omega_r, since the reflected spectral radiance follows L_{e,\lambda}^r = f_{r,\lambda} L_{e,\lambda}^i \cos \theta_i. These formulations use the subscript \lambda for per-unit-wavelength quantities; equivalent expressions apply using frequency \nu for per-unit-frequency measures. Broadband or total reflectance R can be derived from spectral reflectance by integrating over the spectrum weighted by the incident light : R = \frac{\int R_\lambda(\lambda) S(\lambda) \, d\lambda}{\int S(\lambda) \, d\lambda}, where S(\lambda) represents the of the source illumination. This integration is crucial for applications like color rendering, where the perceived reflectance depends on both material properties and lighting conditions. Spectral reflectance reveals material-specific behaviors, such as the high near-infrared reflectance of , typically ranging from 0.5 to 0.8 in the 0.7–1.1 \mum band, attributed to light scattering by the internal mesophyll cell structure of healthy leaves. This contrast with low visible reflectance enables indices like the (NDVI) in , defined as NDVI = (R_{\rm NIR} - R_{\rm red}) / (R_{\rm NIR} + R_{\rm red}), to quantify vegetation density and by exploiting differences. In metals, variations are pronounced; for example, exhibits low reflectance in the (around 0.2–0.4) due to strong but approaches near-unity reflectance (>0.95) in the , stemming from free-electron behavior that enhances at longer wavelengths.

Surface and Material Characteristics

Specular and Diffuse Reflectance

Specular reflectance occurs when reflects from a smooth surface in a mirror-like manner, where the angle of incidence equals of reflection, producing a clear of the source. This behavior is characteristic of polished or flat interfaces, such as or metals, and is fundamentally described by the , which quantify the fraction of incident reflected based on the refractive indices of the materials and of incidence. For example, at normal incidence on ( ≈1.5) in the , the reflectance is approximately 4%, increasing to around 10% at grazing angles near 60 degrees. In contrast, diffuse reflectance arises from surfaces with microscopic roughness that scatter incident light in many directions, rather than concentrating it in a single specular direction, resulting in no distinct image formation. Ideal diffuse reflectors, known as Lambertian surfaces, exhibit radiance that is independent of the viewing angle. The radiant intensity observed from the surface follows Lambert's cosine law: the intensity I is proportional to the cosine of the angle \theta between the surface normal and the line of sight, expressed as I = I_0 \cos \theta, where I_0 is the intensity at normal incidence. A representative example is matte paper, which achieves nearly uniform diffuse reflectance of about 0.8 across visible wavelengths, appearing equally bright from all angles due to this scattering. Real surfaces often exhibit a transition between specular and diffuse behaviors, influenced by surface microstructure and roughness. Microfacet models treat the surface as composed of tiny mirror-like facets with varying orientations, where the overall reflectance is an average over these facets' contributions. The seminal Torrance-Sparrow model, developed in , incorporates a roughness \sigma (the standard deviation of facet angles) to predict the broadening of the specular peak; as \sigma increases, the reflection shifts toward more diffuse scattering. For instance, a polished metal surface with low \sigma (e.g., <0.1 radians) shows predominantly specular reflectance exceeding 70% in the specular direction, while higher roughness values produce intermediate effects blending both components.

Reflectance of Liquids and Water

The reflectance of liquids, including water, primarily occurs at the air-liquid interface and is governed by the , which describe the fraction of incident light reflected based on the refractive index and angle of incidence. For smooth liquid surfaces, this reflection is specular, akin to principles observed in polished solids, but it is dynamically altered by fluid motion and surface perturbations. In calm conditions, the reflectance of water in the visible spectrum ranges from approximately 0.02 at normal incidence to 0.1 at steeper angles, due to its refractive index of 1.33. Water exhibits particularly low surface reflectance compared to many materials, with about 2% at normal incidence arising from the modest refractive index contrast with air, though hemispherical reflectance under diffuse illumination averages around 5%. Spectral variation is minimal across the visible wavelengths (400–700 nm), maintaining near-constant low values, but reflectance increases in the infrared due to higher absorption. On wind-roughened surfaces, capillary waves and larger undulations scatter light, elevating the diffuse reflectance component to up to 0.2, depending on wind speed. To account for these effects, the effective reflectance is modeled as the sum of the Fresnel term for the mean interface and an additional scattering term from capillary waves:
R = R_{\text{Fresnel}} + R_{\text{scattering}}
where R_{\text{Fresnel}} is the specular reflection for a flat surface, and R_{\text{scattering}} integrates contributions from wave slopes. This formulation correlates with wind speed via models like Cox-Munk, which statistically distribute surface slopes (e.g., Gaussian for low winds, increasing variance with speed up to 10 m/s), enabling predictions of enhanced backscattering. In ocean color remote sensing, such adjustments correct for surface reflectance to isolate water-leaving radiance, facilitating chlorophyll detection by revealing subsurface absorption features around 443 nm.
A distinctive feature for water is Brewster's angle, approximately 53° from the normal, where p-polarized (parallel) reflectance drops to zero, as the reflected and refracted rays become perpendicular, eliminating reflection for that polarization. This polarization selectivity is exploited in applications like glint reduction in remote sensing.

Specialized Contexts

Grating Efficiency

In the context of diffraction gratings, grating efficiency generalizes the concept of reflectance to periodic structures, representing the fraction of incident optical power that is reflected into a specific diffraction order rather than a simple total reflection. This efficiency, denoted as \eta_m for the m-th order, is defined as \eta_m = \frac{P_m}{P_{\text{incident}}}, where P_m is the power diffracted into order m and P_{\text{incident}} is the incident power. The angular positions of these diffracted orders are determined by the grating equation, \sin \theta_m = \sin \theta_i + m \frac{\lambda}{d}, where \theta_i is the angle of incidence, \theta_m is the diffraction angle for order m, \lambda is the wavelength, and d is the grating groove spacing. This framework applies particularly to reflection gratings, where light is reflected and dispersed into multiple orders, contrasting with transmission gratings that allow light to pass through the structure. Blazed gratings, a common type of reflection grating, feature asymmetric groove profiles (often sawtooth-shaped) designed to concentrate diffracted power into a single desired order, typically the first order (m=1), achieving efficiencies of 70-90% under optimal conditions. These gratings outperform non-blazed designs by mimicking a specular reflection from the sloped facet of each groove, thereby enhancing energy transfer to the target order while minimizing losses to other orders or zero-order reflection. In contrast, transmission gratings distribute efficiency across orders more evenly but are less common in high-efficiency reflective applications due to material absorption constraints. Grating efficiency is inherently polarization-dependent, with transverse electric (TE, or s-polarization) modes often exhibiting higher efficiency than transverse magnetic (TM, or p-polarization) modes, especially at oblique incidence angles common in spectroscopic setups. This dependence arises from the interaction of the electric field with the grating's periodic microstructure. Reflection gratings with high efficiency are essential in spectrometers, where they enable precise wavelength dispersion for applications ranging from astronomical observations to laser spectroscopy. A notable example is the echelle grating, which operates in high orders (e.g., m > 30) to provide exceptional (often >10,000) over broad bandwidths, with efficiencies maintained above 50% despite the high-order operation, making it ideal for high-throughput instruments. In radiometry, reflectance is interconnected with other coefficients that describe the interaction of with matter, providing a complete of incident . These include , , and , which together obey conservation principles for energy balance. Kirchhoff's law of thermal radiation states that, for an opaque body in thermal equilibrium, the emissivity ε equals the absorptance α, and both are related to reflectance ρ by ε = α = 1 - ρ at each wavelength. This law ensures detailed balance between emission and absorption processes. Absorptance α is defined as the ratio of absorbed radiant flux Φ_e^a to incident radiant flux Φ_e^i, given by α = Φ_e^a / Φ_e^i. Transmittance τ, relevant for transparent or semi-transparent media, is the ratio of transmitted radiant flux Φ_e^t to incident flux, τ = Φ_e^t / Φ_e^i. For opaque materials where τ = 0, conservation requires ρ + α = 1. The following table summarizes key radiometric coefficients related to reflectance, all of which are dimensionless ratios between 0 and 1:
CoefficientSymbolDefinitionKey RelationExample
ReflectanceρReflected / incident For opaque: ρ + α = 1Mirror: ρ ≈ 1, α ≈ 0
AbsorptanceαAbsorbed / incident α = 1 - ρ (opaque): α = 1, ρ = 0
TransmittanceτTransmitted / incident ρ + α + τ = 1 (non-scattering)Clear glass: τ ≈ 0.9, ρ ≈ 0.08
EmissivityεEmitted / ε = α (Kirchhoff's law): ε = 1
These relations highlight , with examples illustrating ideal cases. For volumetric media, such as participating atmospheres or tissues, the μ quantifies the fractional loss of per unit path length, defined as μ = - (dI / dx) / I, where I is the . This leads to Beer's for non-scattering : I(x) = I_0 e^{-\mu x}. In contexts involving , μ = κ + σ ( plus coefficients) influences the overall reflectance by governing paths and multiple reflections within the medium; reflectance is often quantified via the single-scattering (ratio of to total ).

References

  1. [1]
    Reflectance - an overview | ScienceDirect Topics
    6.2.​​ Reflectance is a dimensionless number defined by the ratio of the radiant flux Φ, expressed in Watts, of the reflected (index r) and the incident (index i ...
  2. [2]
    [PDF] Reflectance quantities in optical remote sensing—definitions and ...
    Bidirectional reflectance properties can be derived from ground reflectance measurements by deducing the contribution of the hemispherical diffuse ...
  3. [3]
    Specular vs. Diffuse Reflection - The Physics Classroom
    Reflection off of smooth surfaces such as mirrors or a calm body of water leads to a type of reflection known as specular reflection.
  4. [4]
    Optical Properties of Materials | NIST
    Feb 26, 2021 · Such measurements include reflectance, transmittance, emittance, absorptance, and index of refraction.Missing: definition | Show results with:definition
  5. [5]
    Reflectance Spectroscopy - PMC - NIH
    Reflectance spectroscopy systems for biomedical applications contain three key components: a light source, a fiber optic probe, and a spectrometer.
  6. [6]
    [PDF] Reflectance Spectroscopy Planetary Science
    Measurement of transmission and reflectance spectra of frozen volatiles for application to outer solar system problems. Derivation of absorption coefficients.
  7. [7]
    [PDF] Geometrical considerations and nomenclature for reflectance
    Then, combining the basic definition of reflectance with eq (Bl), (B2), (B3), and (B4), the bidirectional reflectance is dpiei,<t>r,er,<}>r)=d<^r/d<Pi = dMr ...
  8. [8]
  9. [9]
    [PDF] Standard Reference Materials - NIST Technical Series Publications
    Figure 9 shows the average spectral specular reflec- tance of the aluminum and gold coated samples . The individual values for each sample will be reported in ...
  10. [10]
  11. [11]
    [PDF] Radiant properties of materials - INFRAMET
    The International Lighting Vocabulary of the CIE generally follows this school of thought and proposes following parameters: reflectance and reflectivity, ...
  12. [12]
    International Lighting Vocabulary | CIE
    The aim of this edition of the International Lighting Vocabulary is to promote international standardization in the use of quantities, units, symbols and ...
  13. [13]
    Fresnell's Equations: Reflection and Transmission
    ... Fresnel's equations give the reflection coefficients: = and, = The transmission coefficients are. = and. = Note that these coefficients are fractional ...
  14. [14]
    Fresnel Reflection - Fibercore - Humanetics
    Fresnel Reflection. Between a typical glass-air interface, such as that of a flat cleaved fiber (normal incidence), the end reflection is around 4%. Related ...
  15. [15]
    Return Loss of a Glass-Air Interface - Math Encounters Blog
    Apr 3, 2015 · Let's begin by precisely defining both Fresnel reflection and return loss. Note the 4% transmission loss (aka reflected power) figure ...
  16. [16]
    ISO 80000-7:2019 - Quantities and units — Part 7: Light and radiation
    This document gives names, symbols, definitions and units for quantities used for light and optical radiation in the wavelength range of approximately 1 nm ...Missing: reflectance | Show results with:reflectance
  17. [17]
    reflectance (R05235) - IUPAC Gold Book
    synonyms: reflectance factor, reflectivity. https://doi.org/10.1351/goldbook.R05235. Fraction of incident radiation reflected by a surface or discontinuity, ρ ...Missing: definition | Show results with:definition
  18. [18]
    Reflectance - RP Photonics
    Reflectance is defined as the ratio of reflected radiant flux (optical power) to the incident flux at a reflecting object – for example, an optical component ...
  19. [19]
    [PDF] Radiometry and Photometry
    Hemispherical reflectance. R. 1. Radiant flux reflected by a surface, divided by that received by that surface. Spectral hemispherical reflectance. R ν or. R λ.
  20. [20]
    Hemispherical Reflectance Results of the SolarPACES Reflectance ...
    The paper only reports on the simplest task among those of SRRR: the solar hemispherical reflectance measurement.
  21. [21]
    [PDF] A Guide to Reflectance Coatings and Materials - Pro-Lite Technology
    Most commercial white paints, however, are not particularly white, nor particularly stable. They typically have an integrated reflectance in the 85 - 88% range ...<|separator|>
  22. [22]
    Retrieval of Land Surface Albedo from Satellite Observations
    Inherent albedo is the directional-hemispherical surface reflectance integrated from surface BRDF over all viewing angles. It is completely independent of the ...
  23. [23]
    [PDF] Spectral reflectance - GovInfo
    wavelength; spectral reflectance is the fraction of the incident radiant flux that is reflected as a function of wavelength. A related quantity is the ...
  24. [24]
    Spectral Reflectance | NIST
    Mar 1, 1998 · This document describes the instrumentation, standards, and techniques used to measure spectral reflectance over the ultraviolet, visible, and near infrared ...
  25. [25]
    [PDF] Vegetation Index Pasture, Rangeland and Forage
    The cell structure in a plant leaf strongly reflects near- infrared light (from 0.7 to 1.1 µm). • The more leaves a plant has, the more these wavelengths of ...
  26. [26]
    Landsat Normalized Difference Vegetation Index - USGS.gov
    Landsat NDVI quantifies vegetation greenness, density, and plant health. It's calculated as a ratio of red and near-infrared values, (NIR - R) / (NIR + R).
  27. [27]
    [PDF] THE OPTICAL PROPERTIES OF PLATINUM AND GOLD IN THE ...
    Gold has a reflectance in the vacuum ultraviolet superior to many materials, and the stability of a gold coating after extended use and storage is excellent ...Missing: IR | Show results with:IR
  28. [28]
    8.2 Specular Reflection and Transmission
    The Fresnel equations describe the amount of light reflected from a surface; they are the solution to Maxwell's equations at smooth surfaces. Given the index of ...
  29. [29]
    Fresnel reflections - RP Photonics
    Fresnel reflections occur at material interfaces due to impedance mismatch, with reflectivity calculated using Fresnel equations.
  30. [30]
  31. [31]
    Lights - Diffuse and Lambertian Shading - Introduction to Shading
    Figure 2: The cosine law. The amount of light incident on the surface depends on the cosine of the angle between the light incident direction and the surface ...
  32. [32]
    Materials - Light Reflecting Factors - The Engineering ToolBox
    Light reflecting factor materials. ; Aluminum coatings, matte, 55 - 56 ; Aluminum, anodized, matte, 80 - 85 ; Aluminum, matte, 55 - 75 ; Aluminum, polished, 65 - 75.
  33. [33]
    8.4 Microfacet Models
    Microfacet-based BRDF models work by statistically modeling the scattering of light from a large collection of microfacets.
  34. [34]
    [PDF] ab initio rederivation of Fresnel equations confirms microscopic ...
    Dec 16, 2017 · We provide a systematic theoretical, experimental, and historical critique of the standard derivation of Fresnel's equations, which shows in ...
  35. [35]
    [PDF] Chapter 5 Within the Water - MISC Lab
    Quantities such as the Fresnel reflectance of the water surface, shown in Fig. 4.3, also must be revised to take into account the dependence of reflection ...
  36. [36]
    [PDF] Temperature-dependent optical constants of water in the thermal ...
    Apr 15, 2022 · Abstract: This paper describes a data archaeology and rescue of temperature-dependent thermal infrared (IR) optical constants for liquid ...
  37. [37]
    Airborne Spectral Measurements of Ocean Directional Reflectance in
    The algorithm models the water surface reflectance by the Cox–Munk model and assumes the water-leaving radiance to be Lambertian. The algorithm retrieves the ...
  38. [38]
    [PDF] part ii radiative transfer - MISC Lab
    The U = 0 curve is just the Fresnel reflectance seen in Fig. 4.3. Reflectances for capillary waves and distributed sources. The collimated-source reflectance r+ ...Missing: effective | Show results with:effective
  39. [39]
    Algorithm Descriptions - NASA Ocean Color
    Relative measure of water-leaving radiance associated with chlorophyll fluorescence ... ocean color remote sensing reflectance (Rrs) data (Dierssen et al. 2022) ...
  40. [40]
    Brewster Angle - an overview | ScienceDirect Topics
    The reflectivity at normal incidence (θ = 0°) is typically less than 30% (Figure 3A) and the p-polarized reflectivity goes to zero at the substrate's Brewster ...
  41. [41]
    Light Reflection from Water Waves: Suitable Setup for a Polarimetric ...
    The measured and theoretical Fresnel absolute reflectance components for a flat water surface. The theoretical curves assume a water refractive index of 1.34.
  42. [42]
    Efficiency Characteristics of Diffraction Gratings - Newport
    A grating with high efficiency may allow the reflectivity and transmissivity specifications for the other components in the spectrometer to be relaxed. Moreover ...
  43. [43]
    The Grating Equation - Newport
    The grating equation mλ= d (sinα + sinβ) (2-1) which governs the angular locations of the principal intensity maxima when light of wavelength λ is diffracted ...
  44. [44]
    Diffraction Gratings Tutorial - Thorlabs
    Sep 1, 2022 · Diffraction gratings, either transmissive or reflective, can separate different wavelengths of light using a repetitive structure embedded within the grating.<|control11|><|separator|>
  45. [45]
    Echelle Gratings - Thorlabs
    Echelle Gratings are special low period reflective gratings designed for use in the high orders. They are generally used with a second grating or prism.
  46. [46]
    [PDF] 2.1.2 Reflectance, transmittance, and absorption Reflectance ... - SPIE
    2.1.2 Reflectance, transmittance, and absorption. Reflectance is the amount of flux reflected by a surface, normalized by the amount of flux incident on it.
  47. [47]
    1.8 Reflection, Transmission and Absorption - Gigahertz-Optik
    Introduction into the concepts of reflection, transmission and absorption measurements.
  48. [48]
    [PDF] On Kirchhoff's law and its generalized application to absorption and ...
    Kirchhoff's Law states that at a point on the surface of a thermal radiator at any temperature and wavelength, the spectral directional emittance is equal to ...
  49. [49]
    emissivity, absorption, radiant flux, light - RP Photonics
    According to Gustav Kirchhoff's law of thermal radiation, the absorptance of a medium is identical to its emissivity: the more an object can absorb ...
  50. [50]
    [PDF] CHAPTER 3 ABSORPTION, EMISSION, REFLECTION, AND ...
    If the source of radiation is in thermal equilibrium with the absorbing medium, then. Iλ = Bλ , so that ελ = aλ . This is often referred to as Kirchhoff's Law.
  51. [51]
    Remote estimation of the diffuse attenuation coefficient in a ...
    Solutions of the radiative transfer equation are used to derive relationships of water reflectance to the diffuse attenuation coefficient (K) in moderately ...
  52. [52]
    [PDF] Apparent optical properties and radiative transfer theory* - MISC Lab
    diffuse attenuation coefficients. Diffuse irradiance attenuation, K d. (m-1) ... Background reflectance just below surface. Reflectance of disk. Sea surface ...