Fact-checked by Grok 2 weeks ago

Radiometry

Radiometry is the science of measuring , specifically optical radiation spanning the , visible, and portions of the , corresponding to wavelengths from approximately 0.01 to 1000 µm or frequencies between 3×10¹¹ and 3×10¹⁶ Hz. This field quantifies the energy transfer of such between sources and receivers, forming the basis for precise of light propagation, sources, and materials. Central to radiometry are key quantities that describe in geometric terms, including radiant flux (Φ), the total emitted in watts (); irradiance (E), the per unit area in /m²; radiant intensity (I), the per unit in /sr; and radiance (L), the per unit area per unit in /m²·sr. These measures adhere to the () and account for factors such as , the , and cosine dependencies in propagation. Radiometry differs from photometry, which restricts analysis to the (approximately 360–830 nm) and weights measurements by human visual sensitivity, using units like lumens and . Historically, radiometry advanced through efforts at the National Bureau of Standards (now NIST), beginning in the 1920s with verifications of and irradiance standards achieving few percent accuracy by the 1950s, then evolving to sub-percent precision in the 1960s–1970s via electrically calibrated detectors and spectroradiometry spurred by space and energy needs. Today, it underpins diverse applications, including astronomical observations, for weather and environmental monitoring, assessment, lighting efficiency optimization (which accounted for about 14% of U.S. electricity consumption in 2020), phototherapy, ultraviolet hazard regulation, and defense systems.

Introduction

Definition and Scope

Radiometry is the science of detecting and measuring radiant electromagnetic energy in the optical portion of the , spanning , visible, and wavelengths, corresponding to wavelengths from approximately 100 to 1 or frequencies between 3×10¹¹ and 3×10¹⁶ Hz. This field quantifies the energy carried by electromagnetic waves in terms of objective physical quantities, distinguishing it from photometry, which weights measurements according to human visual sensitivity. The scope of radiometry encompasses non-ionizing portions of the spectrum, such as (UV), visible light, and (IR), where measurements are crucial for assessing energy transfer in various environments. It includes both scalar measurements, which capture total without directional information (e.g., in watts per square meter), and vector measurements that account for directionality (e.g., radiance in watts per square meter per ). These approaches are essential for scientific and engineering applications, enabling precise quantification of to support fields like and materials testing. Key applications include determining the temperature of sources and evaluating levels, such as the standard 1000 /m² for terrestrial simulations. Radiometric measurements employ units, with expressed in watts () and energy in joules (J), providing a standardized basis for comparing across diverse wavelengths and contexts.

Historical Development

The discovery of infrared radiation in 1800 by British William marked an early milestone in understanding beyond the visible spectrum. observed that a placed beyond the red end of the solar spectrum, dispersed by a , registered higher temperatures, indicating the presence of invisible "heating rays." This finding laid foundational groundwork for radiometry by expanding the conceptual scope of to include non-visible wavelengths. In the mid-19th century, German physicist Gustav Kirchhoff advanced the theoretical framework of thermal radiation through his 1859 law, which established that for a body in thermal equilibrium, the emissivity equals the absorptivity at each wavelength, enabling the concept of ideal blackbody radiators. Building on this, Austrian physicist Josef Stefan empirically derived in 1879 the relationship showing that the total energy radiated by a blackbody is proportional to the fourth power of its absolute temperature, later theoretically confirmed by Ludwig Boltzmann in 1884. These blackbody radiation laws provided essential principles for quantitative radiometric measurements, influencing subsequent detector developments. The late 19th and early 20th centuries saw practical advancements in instrumentation, notably American astrophysicist Samuel 's invention of the in 1880, a highly sensitive thermal detector capable of measuring minute temperature changes from radiant heat, which was a thousand times more precise than prior devices. The emergence of photoelectric detectors in the early 1900s further enabled direct electrical responses to radiation, facilitating more accurate spectral measurements. In 1948, the General Conference on Weights and Measures (CGPM) formalized key radiometric-related units, including the for based on a blackbody radiator at the platinum freezing point, setting the stage for the (SI) adopted in 1960. Post-World War II progress accelerated with space-based applications, as the Nimbus satellite series in the 1960s introduced advanced radiometers for , including high-resolution infrared instruments that measured global patterns from orbit. The National Institute of Standards and Technology (NIST), formerly the National Bureau of Standards, evolved its radiometry standards through cryogenic radiometers and trap detectors, achieving traceability for optical measurements with uncertainties below 0.1% by the 1980s and incorporating sources for UV and calibrations in the 1990s; as of May 2025, NIST's project provided lunar reflectance data for satellite radiometric calibration with 10 times greater accuracy than prior benchmarks. Key contributors included for instrumentation and ongoing work by (CIE) Division 2 committees, which standardized physical measurements of and through global expert collaborations. Significant milestones included the adoption of absolute radiometry in the 1970s, exemplified by NIST's development of cavity-based absolute radiometers that substituted electrical for radiative power with high precision, enabling primary standards independent of secondary calibrations. Post-2010 advancements introduced quantum-based calibrations, such as using to achieve quantum efficiency measurements with 0.5% , enhancing radiometric accuracy in low-light regimes.

Core Concepts

Electromagnetic Radiation Fundamentals

Electromagnetic radiation, the foundation of radiometry, exhibits wave-particle duality, manifesting properties of both classical waves and discrete particles known as photons. As a wave, it consists of oscillating electric and magnetic fields perpendicular to each other and to the direction of propagation, traveling through vacuum at the constant speed of light, c = 3 \times 10^8 m/s. The fundamental relationship between its wavelength \lambda and frequency \nu is given by \lambda \nu = c, which determines the energy and behavior of the radiation across different regimes. The spans a continuous range of wavelengths, conventionally divided into distinct bands: radio waves (longest wavelengths, lowest frequencies), microwaves, (IR), visible light, (UV), s, and gamma rays (shortest wavelengths, highest frequencies). Each band corresponds to specific energies, quantified by E = h \nu, where h is Planck's (h = 6.626 \times 10^{-34} J s), linking the particle-like nature of to its . This division is crucial for understanding how interacts with , as shorter wavelengths carry higher per and can ionize atoms in UV, , and gamma regions. Blackbody radiation represents the idealized emission from a perfect absorber, serving as a reference for in radiometry. A blackbody absorbs all incident and emits energy solely dependent on its T, without regard to the material composition. The B(\lambda, T), which describes the power per unit area, , and , follows : B(\lambda, T) = \frac{2 h c^2}{\lambda^5} \frac{1}{e^{h c / \lambda k T} - 1} where k is Boltzmann's constant (k = 1.381 \times 10^{-23} J/K). This equation resolves the ultraviolet catastrophe of classical theory by incorporating quantum effects. , derived from , specifies the wavelength \lambda_{\max} of peak spectral radiance via \lambda_{\max} T = 2898 μm·K, shifting the emission peak to shorter wavelengths as temperature increases—for instance, visible light peaks around 500 nm for temperatures near 5800 K. Polarization describes the orientation of the electric field vector in electromagnetic waves, which can be linear, circular, or elliptical, influencing , , and during measurements. Coherent maintains a fixed relationship across wavefronts, enabling effects essential for precise radiometric instrumentation, whereas incoherent sources like thermal emitters produce random phases that average out in detection. These properties must be accounted for in radiometric setups to avoid measurement biases, particularly in polarized or partially coherent sources such as lasers or atmospheric .

Distinction from Photometry

Photometry is the science of measuring visible light in a manner that accounts for the spectral sensitivity of the human eye, specifically weighting the radiation according to the photopic luminous efficiency function V(λ), which peaks at approximately 555 nm in the green region of the spectrum. This approach contrasts with radiometry by incorporating a psychophysical element, where photometric quantities such as luminous flux are expressed in lumens (lm) rather than the absolute energy unit of watts (W) used in radiometry. The V(λ) curve, standardized by the International Commission on Illumination (CIE), defines the eye's relative sensitivity across the visible range from about 360 nm to 830 nm, effectively filtering radiometric measurements to reflect perceived brightness. The primary distinction between radiometry and photometry lies in their measurement paradigms: radiometry provides objective, wavelength-independent quantification of optical radiation spanning the , , and portions of the , focusing solely on physical energy flux without regard to human perception. In contrast, photometry is inherently subjective and perceptual, applying the V(λ) to restrict analysis to the and scale values relative to the eye's peak sensitivity at 555 nm, where the maximum reaches 683 lm/W for monochromatic radiation. This makes photometry unsuitable for non-visible wavelengths, such as or , where radiometry excels in applications like thermal imaging. Conversion between radiometric and photometric quantities is possible but limited; for monochromatic sources at 555 nm, the factor is precisely 683 /W, but broadband sources require integration over the spectrum using V(λ), yielding no universal equivalence due to varying spectral distributions. In practice, radiometric spectral data often serves as the foundation for computing photometric values in fields like , bridging the two through established CIE protocols that define both systems. This overlap ensures consistency, yet highlights radiometry's broader applicability for objective assessments beyond human vision.

Radiometric Quantities

Primary Quantities

, denoted Q_e, is the total amount of energy emitted, transferred, or received in the form of , irrespective of its wavelength distribution, and is measured in joules (J). This quantity serves as the foundational measure in radiometry for the absolute energy content involved in radiative processes. Radiant flux, symbolized as \Phi_e, represents the rate of flow of radiant energy with respect to time, defined as \Phi_e = \frac{d Q_e}{d t}, and has units of watts (W). It quantifies the power radiated by a source or passing through a surface, providing a direct measure of energy transfer dynamics in radiometric systems. For instance, the total power output of an incandescent or the of a star is characterized by its radiant flux \Phi_e. Radiant intensity, denoted I_e, is the radiant flux per unit solid angle in a specified direction, given by I_e = \frac{d \Phi_e}{d \Omega}, with units of watts per steradian (W/sr). This quantity is essential for describing the directional distribution of power from sources, particularly point-like emitters. A key relation among these primaries is that the total radiant flux equals the integral of radiant intensity over the solid angle of the emitting : \Phi_e = \int_{2\pi} I_e \, d\Omega. This integration holds for the total, wavelength-integrated case without . These quantities form the basis for and can be extended to spectral variants for wavelength-dependent analyses.

Derived Quantities

Derived quantities in radiometry extend the primary concept of radiant flux by incorporating spatial and angular dependencies, enabling the description of radiation fields from extended sources and surfaces. These quantities are essential for analyzing how interacts with areas and directions in , particularly for non-point sources where uniformity cannot be assumed. Radiance, denoted L_e, quantifies the per unit projected area perpendicular to the direction of propagation and per unit . It is defined as L_e = \frac{d^2 \Phi_e}{dA \cos \theta \, d\Omega} with units of watts per square meter per (W/m²·sr). This measures the power density along a specific , and a key property is its invariance along a in lossless media, meaning the value remains constant as propagates through or homogeneous isotropic materials without or . Irradiance, denoted E_e, represents the radiant flux incident on a surface per unit area, integrating contributions from all directions over the incident hemisphere. It is given by E_e = \frac{d \Phi_e}{dA} in units of W/m². This quantity describes the total power density received at a point, independent of direction, and can be computed from radiance via hemispherical integration E_e = \int_{2\pi} L_e \cos \theta \, d\Omega. Radiant exitance, denoted M_e, is the emitted by a surface per unit area into the outward hemisphere. Its definition mirrors that of but applies to outgoing emission: M_e = \frac{d \Phi_e}{dA} also in W/m². For a blackbody, this follows the Stefan-Boltzmann law, M_e = \sigma T^4, where \sigma = 5.670 \times 10^{-8} W/m²·K⁴ is the Stefan-Boltzmann constant and T is the absolute temperature. Radiosity, denoted J_e, accounts for the total leaving a surface per area, comprising both emitted and reflected components: J_e = M_e + \rho E_e, where \rho is the . It is expressed in W/ and is particularly useful for diffuse surfaces in analyses, as it lumps all outgoing regardless of direction. Representative examples illustrate the scale of these quantities. The solar irradiance at Earth's surface under clear-sky conditions reaches approximately W/m² near noon at mid-latitudes, representing the power density from after atmospheric . For a blackbody at 300 K (), the is about 460 W/m², highlighting the role of thermal emission in everyday environments. Spectral versions of these derived quantities incorporate wavelength dependence, denoted with subscript \lambda or \nu, to describe distributions.

Spectral Radiometry

Spectral Quantity Definitions

In spectral radiometry, quantities describe the distribution of as a function of or frequency, enabling detailed analysis of across the optical spectrum. These spectral quantities are essential for applications such as and , where understanding the wavelength-dependent behavior of is critical. Unlike quantities that aggregate total , spectral forms provide the density of per unit spectral interval, allowing reconstruction of properties through . The , denoted \Phi_{e,\lambda} in watts per nanometer (W/) or \Phi_{e,\nu} in watts per hertz (W/Hz), quantifies the total power emitted, transmitted, or received by a source or system within an infinitesimal interval of \Delta\lambda or \Delta\nu. This fundamental quantity serves as the basis for all other radiometric measures, representing the power density without regard to spatial or directional . For instance, it is used to characterize the output of lamps or lasers across their spectra. Spectral radiance, L_{e,\lambda} in watts per square meter per steradian per nanometer (W/m²·sr·nm) or L_{e,\nu} in W/m²·sr·Hz, measures the per unit , per unit , and per unit spectral interval, capturing the directional of from a surface or . It is under in free and is pivotal for modeling transport in optical systems, such as in or illumination design. provides a complete description of how varies with and at a point. Spectral irradiance, E_{e,\lambda} in W/m²·nm or E_{e,\nu} in W/m²·Hz, represents the incident on a surface per unit area and per unit , integrating contributions from over the . This is key for assessing levels, such as in applications or performance, where the content of incoming radiation determines efficiency. Standard notation conventions employ subscripts \lambda to indicate per-unit-wavelength quantities and \nu for per-unit-frequency forms, with the distribution functions ensuring dimensional consistency across the . The corresponding total () quantity, such as total \Phi_e, is obtained by integrating the quantity over all wavelengths or frequencies: \Phi_e = \int_0^\infty \Phi_{e,\lambda} \, d\lambda or \Phi_e = \int_0^\infty \Phi_{e,\nu} \, d\nu. This integration links and radiometry, where broadband totals emerge as sums of components. A representative example is the solar spectral irradiance at Earth's surface, which exhibits a peak in the visible range around 500 nm, corresponding to light and delivering the majority of for biological and photovoltaic processes. This spectral profile underscores the concentration of in the 400–700 nm band, influencing applications from climate modeling to .

Spectral Distribution Equations

In spectral radiometry, radiometric quantities are often expressed as functions of either \lambda or \nu, requiring careful conversion to maintain the physical invariance of energy content within intervals. The fundamental relation ensures that the spectral radiant flux in wavelength form equals that in frequency form, such that \Phi_{e,\lambda} \, d\lambda = \Phi_{e,\nu} \, d\nu, accounting for the differential d\nu = -(c / \lambda^2) \, d\lambda where c is the . Thus, the spectral density transforms as \Phi_{e,\lambda} = (c / \lambda^2) \Phi_{e,\nu}, with \nu = c / \lambda. A cornerstone of spectral radiometry is the modeling of thermal emission from blackbodies using , which provides the L_{e,\lambda}(\lambda, T) as a function of and T: L_{e,\lambda}(\lambda, T) = \frac{2 h c^2}{\lambda^5} \frac{1}{e^{h c / \lambda k T} - 1}, where h is Planck's constant and k is Boltzmann's constant. This equation describes the maximum possible at equilibrium and serves as a reference for calibrating other sources. To obtain total radiometric quantities from their spectral counterparts, integration over the entire spectrum is required. The total radiant flux \Phi_e is thus given by \Phi_e = \int_0^\infty \Phi_{e,\lambda} \, d\lambda = \int_0^\infty \Phi_{e,\nu} \, d\nu, where the equality holds due to the conversion relation ensuring conservation of total energy. For blackbody emission, this integral yields the Stefan-Boltzmann law, \Phi_e = \sigma T^4 with \sigma = 2 \pi^5 k^4 / (15 c^2 h^3), but the spectral form emphasizes wavelength-dependent contributions. For band-limited spectra, such as those from lasers or filters, approximations simplify calculations by assuming the quantity is nearly constant over a small \Delta\lambda or \Delta\nu. The effective in the is then \Phi_{e,\text{band}} \approx \Phi_{e,\lambda} \Delta\lambda, enabling precise modeling of monochromatic-like sources. In detector applications, \eta(\lambda) modulates the response, with the detected proportional to \int \Phi_{e,\lambda} \eta(\lambda) \, d\lambda over the band's narrow range, often approximating \eta(\lambda_0) \Phi_{e,\lambda_0} \Delta\lambda at central \lambda_0. An illustrative example is the spectral radiance from a blackbody source at 300 , approximating room-temperature emission. At the peak of approximately 9.66 \mum (per , \lambda_{\max} T = 2898 \, \mum \cdot), the radiance reaches about 9.92 m^{-2} sr^{-1} \mum^{-1}, highlighting the dominance of mid- for such sources in applications like imaging.

Integral Radiometry

Integral Quantity Definitions

Integral radiometric quantities represent the total or power integrated over the entire , typically from through wavelengths, without resolving details. These quantities are essential for applications involving sources where the full energy content is relevant, such as solar radiation assessments or total thermal flux calculations. Unlike quantities, forms aggregate contributions across all wavelengths, simplifying analysis for non-dispersive measurements. One key integral quantity is the radiant exposure, denoted H_e, which measures the time-integrated irradiance on a surface. It is defined as H_e = \int E_e \, dt, where E_e is the irradiance in watts per square meter (W/m²) and t is time in seconds, yielding units of joules per square meter (J/m²). This quantity captures the cumulative energy deposited on a surface over a period, such as during an exposure test or environmental monitoring. For example, the annual solar radiant exposure on a horizontal surface in the United States averages approximately $5 \times 10^9 to $7 \times 10^9 J/m²/year, depending on location and atmospheric conditions. Integral radiosity, denoted J_e, quantifies the total leaving a surface per unit area, encompassing both emitted and reflected components integrated over the above the surface and the full . It has units of / and is particularly useful for describing the overall outgoing radiation from opaque or diffuse surfaces in or lighting scenarios. In practice, for sources like —which spans a broad from about 300 to 2500 —integral radiosity requires averaging over the source's emission profile, whereas near-monochromatic sources like LEDs (with bandwidths under 50 ) allow approximations treating them as effectively single-wavelength emitters for integral calculations. All integral quantities ultimately relate to the primary radiant flux \Phi_e, the total power in watts integrated over the complete ; for instance, integrating or exitance over area yields , and further spectral integration ensures the totals align with \Phi_e for fully characterized sources. These integrals derive from spectral counterparts by summing over without weighting, assuming the full spectrum is considered.

Integration and Broadband Calculations

In radiometry, integral quantities such as total radiant flux \Phi_e are obtained by numerically integrating distributions over , typically using methods like the or when data are provided at discrete intervals. The approximates the integral \int \Phi_{e,\lambda} \, d\lambda by summing trapezoidal areas under the curve, given by I \approx \delta\lambda \left( \frac{\Phi_{e,1}}{2} + \sum_{i=2}^{n-1} \Phi_{e,i} + \frac{\Phi_{e,n}}{2} \right) for evenly spaced points with interval \delta\lambda, providing a first-order accurate method suitable for measured bands. enhances accuracy by fitting quadratic polynomials between points, exact for up to second-degree polynomials, and is preferred for spectra with curvature, such as those from thermal sources; it weights the middle point as $4 \times the endpoint contributions over even intervals. For broadband calculations, approximations simplify integration when full spectral data are unavailable, particularly for thermal emitters. The effective wavelength \lambda_{\text{eff}} represents the single wavelength where monochromatic radiance equals the integrated broadband value, calculated as \lambda_{\text{eff}} = \frac{\int \lambda L_{e,\lambda} S(\lambda) \, d\lambda}{\int L_{e,\lambda} S(\lambda) \, d\lambda} with source radiance L_{e,\lambda} and detector responsivity S(\lambda), useful for narrowband approximations in filter radiometers. For thermal sources, Planck averaging integrates the Planck function B(\lambda, T) weighted by the instrument response, yielding an effective temperature via \int B(\lambda, T) S(\lambda) \, d\lambda = B(\lambda_{\text{eff}}, T_{\text{eff}}). In the full broadband limit for blackbodies, the Stefan-Boltzmann law provides total exitance as M_e = \sigma T^4, where \sigma = 5.670 \times 10^{-8} \, \text{W} \cdot \text{m}^{-2} \cdot \text{K}^{-4}, derived from integrating Planck's law over all wavelengths. Uncertainty in these integrations arises primarily from spectral resolution limits and extrapolation beyond measured ranges. Finite resolution introduces bandwidth errors, correctable via series expansions like E_{\text{corr}} = E_0 + A_1 E' + A_2 E'' + \cdots, where E_0 is the measured and primes denote derivatives; coarser increases random but can be reduced by averaging over more points, from ~4.5% at low to ~0.8% at higher. at spectral edges, often using linear or fits, correlates errors across points and dominates for tails in non-thermal spectra, contributing up to several percent in UV-visible integrations without baseline corrections. Overall relative combines systematic (e.g., ) and random components via u(I)/I = \sqrt{ \sum (u_{R_i}/I)^2 + (u_S/I)^2 }, where u_{R_i} are random variances. Filter-based methods compute partial integrals by convolving the spectral distribution with the bandpass function f(\lambda), yielding the effective quantity Q_{\text{eff}} = \int \Phi_{e,\lambda} f(\lambda) \, d\lambda / \int f(\lambda) \, d\lambda, which approximates band-limited responses in practical instruments like radiometers. This accounts for non-ideal shapes (e.g., Gaussian or rectangular), ensuring accurate partial radiant power within the while rejecting contributions. A representative example is calculating total irradiance from the AM1.5 global solar spectrum, defined under ASTM G173 conditions ( 1.5, ). Integrating the provided spectral E_{\lambda} (in W/m²/nm) from 280 nm to 4000 nm yields a total of 1000 W/m², normalizing terrestrial solar input for photovoltaic and atmospheric studies; this uses trapezoidal integration over ~2000 discrete points for precision.

Surface Interaction Properties

Absorption and Emission Coefficients

In radiometry, the absorptance \alpha_\lambda quantifies the fraction of incident spectral that is by a at a specific \lambda, defined as \alpha_\lambda = \frac{\Phi_{\lambda, \text{[absorbed](/page/Absorption)}}}{\Phi_{\lambda, \text{incident}}}, where $0 \leq \alpha_\lambda \leq 1. This coefficient applies to opaque bodies under and is fundamental for understanding energy uptake in radiative interactions. The emissivity \epsilon_\lambda describes the efficiency of a material in emitting spectral radiance at wavelength \lambda compared to a blackbody at the same temperature T, given by \epsilon_\lambda = \frac{L_{\lambda, \text{emitted}}}{L_{b,\lambda}(T)}, where L_{b,\lambda}(T) is the blackbody spectral radiance. According to Kirchhoff's law of thermal radiation, for an opaque body in local thermodynamic equilibrium, the spectral emissivity equals the spectral absorptance, \epsilon_\lambda = \alpha_\lambda, ensuring detailed balance between absorption and emission processes. This equality holds directionally and holds for materials like metals and dielectrics across the infrared spectrum. The total emissivity \epsilon integrates the spectral emissivity over all wavelengths, weighted by the blackbody spectral radiance: \epsilon = \frac{\int_0^\infty \epsilon_\lambda L_{b,\lambda}(T) \, d\lambda}{\int_0^\infty L_{b,\lambda}(T) \, d\lambda}. This formulation accounts for the temperature-dependent Planck distribution, providing a broadband measure essential for calculations in applications. In the gray body approximation, the emissivity is assumed constant (\epsilon_\lambda = \epsilon) across the , simplifying analyses for materials where spectral variations are negligible, such as in many models. This idealization treats the body as partially absorbing and emitting uniformly, contrasting with selective emitters that vary strongly with wavelength. A practical example is human skin, which exhibits a total emissivity of approximately 0.98 in the infrared range (8–14 \mum), enabling accurate for medical diagnostics.

Reflection and Transmission Factors

In radiometry, the reflection factor, or spectral reflectance \rho_\lambda, quantifies the fraction of incident spectral radiant flux that is reflected by a surface at wavelength \lambda. It is defined as \rho_\lambda = \frac{\Phi_{r,\lambda}}{\Phi_{i,\lambda}}, where \Phi_{r,\lambda} is the reflected spectral flux and \Phi_{i,\lambda} is the incident spectral flux. Reflectance can be decomposed into specular and diffuse components: specular reflection occurs in a single direction following the law of reflection, as in mirror-like surfaces, while diffuse reflection scatters light into multiple directions according to the surface's microstructure. To account for the angular dependence of , the (BRDF), denoted f_r(\theta_i, \phi_i, \theta_r, \phi_r), describes how incident radiance from (\theta_i, \phi_i) is reflected into (\theta_r, \phi_r). It is given by f_r(\theta_i, \phi_i, \theta_r, \phi_r) = \frac{dL_r(\theta_r, \phi_r)}{L_i(\theta_i, \phi_i) \cos \theta_i \, d\Omega_i}, where L_r and L_i are the reflected and incident radiances, respectively, \cos \theta_i is the cosine of the incident , and d\Omega_i is the of incidence; the units are steradians inverse (sr^{-1}). This function enables detailed modeling of surface for non-Lambertian materials. The transmission factor, or spectral transmittance \tau_\lambda, measures the fraction of incident spectral radiant flux that passes through a semi-transparent medium at wavelength \lambda, defined as \tau_\lambda = \frac{\Phi_{t,\lambda}}{\Phi_{i,\lambda}}, where \Phi_{t,\lambda} is the transmitted spectral flux. Transmittance applies to materials like or the atmosphere, where propagates without full absorption. Conservation of energy governs these factors: for opaque surfaces, \rho_\lambda + \alpha_\lambda = 1, where \alpha_\lambda is the spectral ; for transparent media, \rho_\lambda + \alpha_\lambda + \tau_\lambda = 1. Kirchhoff's law relates to under , but reflection and transmission remain independent of temperature for non-emitting contexts. Representative examples illustrate these factors: a high-quality silver mirror exhibits \rho > 0.97 across the (400–700 ), dominated by . In the atmosphere under clear-sky conditions, \tau \approx 0.75 for direct beam solar wavelengths (approximately 300–1100 ), accounting for molecular and minor .

Measurement and Instrumentation

Radiometric Instruments

Radiometric instruments are essential devices for quantifying optical across various wavelengths, enabling precise measurements of quantities such as and radiance. These instruments operate on or photoelectric principles, with designs tailored to specific ranges and resolutions. detectors absorb to produce a temperature-dependent signal, while photoelectric detectors convert photons directly into electrical charge. Selection of an instrument depends on the required , , and environmental conditions, often involving cryogenic cooling for enhanced performance in low-signal scenarios. Thermal detectors measure by detecting heat-induced changes in , offering broad response from to far-infrared. Bolometers, a primary type, utilize a resistive element whose varies with due to absorbed ; the basic components include an absorber, , weak , heatsink, and resistive heater for electrical . They achieve high sensitivity through operation at low temperatures. Pyroelectric sensors, another variant, generate voltage from temperature fluctuations in ferroelectric materials and are particularly suited for detecting chopped or modulated beams, where chopping enhances signal isolation using lock-in amplification. Cryogenic bolometers exemplify advanced detection in astronomy, employing superconducting transition-edge sensors cooled to millikelvin for submillimeter and millimeter-wave observations, as demonstrated in experiments with kilopixel arrays. Photoelectric detectors respond to individual photons via the photoeffect, providing fast response times and wavelength-selective sensitivity characterized by quantum efficiency η(λ), defined as the ratio of generated electron-hole pairs to incident photons. Photodiodes, such as silicon-based models, operate by generating proportional to and exhibit external quantum efficiencies up to 100% in the ultraviolet-visible range. Silicon photodiodes typically cover 200-1100 nm, with peak responsivity near 950 nm, making them ideal for UV-visible radiometry. Photomultipliers amplify photocathode-emitted electrons through secondary emission in a series of dynodes, enabling with gains exceeding 10^6 and quantum efficiencies around 20-30% in the . Spectroradiometers provide wavelength-resolved measurements by dispersing or interfering incoming radiation before detection, essential for spectral distribution analysis. Monochromator-based systems use diffraction gratings to isolate narrow wavelength bands, coupling with photoelectric detectors for high-resolution scans across the UV to near-IR. Interferometer designs, such as Fourier-transform infrared (FTIR) spectrometers, employ Michelson interferometry to encode spectral information in an interferogram, which is Fourier-transformed to yield the spectrum, offering advantages in throughput and signal-to-noise for mid- to far-infrared radiometry. Broadband radiometers integrate over wide spectral bands, suitable for total assessments without wavelength discrimination. Pyranometers measure global horizontal using sensors with a hemispherical , typically responding from nm to 3000 nm for resource evaluation. radiometers establish absolute scales through blackbody-like absorption in a conical , where incident is fully trapped and converted to heat, enabling traceability for shortwave with uncertainties below 0.3%.

Calibration and Uncertainty Analysis

Calibration in radiometry ensures the accuracy of measurements by establishing traceability to fundamental standards, primarily through the use of blackbody sources for radiance calibration and cryogenic radiometers for absolute determination. Blackbody sources, which approximate ideal thermal radiators, are calibrated at facilities like those at the National Institute of Standards and Technology (NIST) using cryogenic chambers maintained at temperatures as low as 20 K to minimize and achieve high precision in radiance temperature measurements. Cryogenic radiometers, such as absolute cryogenic radiometers (ACRs) of the electrical substitution type, serve as primary standards for radiant power by equating optical input to electrical heating, enabling direct SI-traceable flux calibrations with uncertainties typically below 0.02%. These standards are essential for calibrating and visible sources sent to NIST from industry and research laboratories. Transfer standards facilitate the dissemination of primary calibrations to working instruments and are often realized through integrating spheres or standardized lamps that maintain traceability. Integrating spheres provide uniform radiance fields for calibrating detectors and sources, with their output calibrated against NIST-traceable standards like quartz-tungsten lamps to ensure irradiance and radiance consistency across laboratories. Lamp standards, such as those used in spectroradiometer setups, are pre-calibrated for and employed to transfer radiance scales, achieving traceability via comparisons to blackbody or cryogenic references. This approach allows for reliable on-site calibrations while preserving the integrity of the system. Uncertainty analysis in radiometric measurements follows the ISO Guide to the Expression of Uncertainty in Measurement () framework, which categorizes uncertainties into Type A, evaluated through statistical methods from repeated observations, and Type B, assessed via non-statistical means such as manufacturer specifications or environmental factors. Type A uncertainties capture random variations, like detector noise, while Type B address systematic effects, including source instability or alignment errors, with propagation combining these components using the law of to yield a combined standard uncertainty. In radiometry, this framework ensures comprehensive error budgeting, often resulting in overall uncertainties of 0.5% to 2% for well-calibrated systems. Spectral calibration refines wavelength scales in instruments like spectroradiometers using discrete emission lines from gas discharge lamps, such as mercury or , or tunable sources to align spectral response accurately. These methods achieve uncertainties below 0.1 nm in the visible and near-infrared regions by fitting observed line positions to known atomic transitions, enabling precise determinations. For instance, low-pressure gas lamps provide sharp lines for , with corrections applied for instrumental broadening to maintain . Practical examples illustrate the application of these methods, such as the of the , which represents the at and is determined via satellite-based radiometers like those on NASA's and Solar Irradiance Sensor (TSIS-1), yielding a value of approximately 1361 W/m² with an of about 0.03% as maintained through ongoing cryogenic radiometer comparisons into 2025. Inter-laboratory comparisons, coordinated by bodies like the Consultative Committee for Photometry and Radiometry (CCPR), validate consistency across global facilities, highlighting the robustness of transfer standards. These efforts underscore the importance of periodic verifications to sustain measurement reliability in radiometry.

Applications

Astronomy and Astrophysics

In astronomy and , radiometry provides the quantitative framework for measuring the and distributions of objects, enabling inferences about their distances, temperatures, compositions, and evolutionary states. measurements typically begin with photometric , which are converted to radiometric quantities such as (in W m^{-2} Hz^{-1}) using established zero points; for instance, in the system, a of zero corresponds to 3631 Jy (1 Jy = 10^{-26} W m^{-2} Hz^{-1}), facilitating direct comparison with physical outputs. Bolometric corrections then adjust these monochromatic to total bolometric by integrating over the entire spectrum, accounting for unseen or contributions; for main-sequence , these corrections vary from about -0.1 mag for A-type stars to -1.5 mag for M dwarfs, derived from synthetic spectra or empirical calibrations against interferometric diameters. Spectral radiometry through reveals key diagnostics, such as the z = \Delta \lambda / \lambda, where \Delta \lambda is the observed shift in or lines relative to their rest wavelengths \lambda; this parameter quantifies radial velocities for nearby galaxies or cosmological expansion for distant ones, with precisions reaching \Delta z \sim 10^{-4} using high-resolution echelle spectrographs. Blackbody applies to broadband photometry or spectra, matching the observed radiance B(\lambda, T) = \frac{2hc^2}{\lambda^5} \frac{1}{e^{hc/\lambda kT} - 1} to estimate effective temperatures, as demonstrated in early calibrations where color indices align with temperatures from 3000 for cool giants to 20,000 for hot O stars. Dedicated instruments enhance radiometric precision across wavelengths. Ground-based facilities like the Atacama Large Millimeter/submillimeter Array (ALMA) deliver interferometric imaging at 0.3–9 mm wavelengths with sensitivities down to microJansky levels, measuring dust emission and molecular line fluxes in protoplanetary disks and high-redshift galaxies. Space telescopes circumvent terrestrial limitations: the () provides absolute flux calibrations in the 0.1–2.5 \mum range using onboard standards, achieving photometric accuracies of 1–2%, while the (JWST) extends this to 0.6–28.5 \mum with mid-infrared spectrometers like , enabling radiometric analysis of faint early-universe sources at flux levels below 1 nJy. Landmark radiometric results underscore these techniques' impact. The cosmic microwave background (CMB), observed as isotropic , has a measured temperature of 2.7255 K from differential radiometers on the COBE , confirming the universe's at z \approx 1100. Type Ia supernova light curves, standardized by decline rates \Delta m_{15}, yield distance moduli \mu = 5 \log_{10} (d/10 pc) through peak absolute magnitudes around -19.3 mag, revealing cosmic acceleration with distances to z > 1 and precisions of 5–10%. Radiometric challenges persist, particularly from atmospheric extinction, which attenuates flux by up to 0.3 mag/airmass at optical wavelengths due to molecular absorption and aerosol scattering; corrections rely on site-specific models, such as those for Cerro Paranal with wavelength-dependent coefficients k(\lambda) \approx 0.12 + 0.3/\lambda (in \mum), derived from spectrophotometric monitoring. Space-based radiometry addresses absolute scale uncertainties by avoiding these effects, as in proposed standards like the Absolute Radiometric Measurements in Space (ARMS) project, which aims for 0.1% accuracy traceable to cryogenic radiometers for calibrating extragalactic fluxes.

Remote Sensing and Environmental Monitoring

Remote sensing relies on radiometric measurements to observe Earth's surface and atmosphere from airborne and spaceborne platforms, enabling the quantification of reflected and emitted across various wavelengths for environmental analysis. Instruments capture or , which is processed to derive biophysical parameters such as , health, and productivity, supporting applications in monitoring and . Satellite radiometry plays a central role in through sensors like those on aboard NASA's and Aqua satellites, which measure reflectance in the visible-near infrared (VNIR) and shortwave infrared (SWIR) bands to map surface properties such as and mineral composition. Similarly, , including and 9, provide high-resolution multispectral imagery in VNIR-SWIR ranges, allowing for the detection of urban expansion and agricultural changes with spatial resolutions down to 30 meters. In the thermal infrared (TIR) spectrum, MODIS sensors estimate by measuring emitted radiance around 11-12 micrometers, achieving accuracies of about 0.5°C, which is crucial for tracking ocean currents and El Niño events. Atmospheric correction is essential in radiometric to isolate surface signals from overlying effects, involving the subtraction of contributions in shorter wavelengths, which scatters and reduces visibility of underlying features. Algorithms also account for by atmospheric constituents, such as in bands around 0.94 and 1.38 micrometers, using models like (Second Simulation of the Satellite Signal in the Solar Spectrum) to retrieve accurate top-of-atmosphere radiance. These corrections enable reliable derivation of surface , a key bidirectional property that describes how is reflected from natural surfaces. Vegetation indices derived from radiometric data provide non-destructive assessments of plant health and coverage; the (NDVI), calculated as \text{NDVI} = \frac{\text{NIR} - \text{Red}}{\text{NIR} + \text{Red}}, leverages the contrast in spectral reflectance between near-infrared (NIR, ~0.7-1.1 μm) where vegetation strongly reflects and red (~0.6-0.7 μm) where absorbs, yielding values from -1 to 1 that indicate sparse to dense green biomass. This index, first formalized in the , has been widely applied using data from sensors like (Advanced Very High Resolution Radiometer) to monitor phenological cycles and crop yields globally. For climate monitoring, the Clouds and the Earth's Radiant Energy System () instruments on satellites like measure Earth's budget by quantifying incoming shortwave irradiance and outgoing longwave thermal , revealing imbalances that drive ; for instance, CERES data from 2000-2020 showed an increase in absorbed by about 0.5 W/m² per decade due to reduced . These broadband radiometric observations, spanning 0.3-50 μm, support models of energy fluxes and feedback mechanisms in the . Recent advancements in the , including through GOES-19 operational by 2025, enhance real-time with the Advanced Baseline Imager (ABI), which provides multispectral radiometric data for mapping at 2-km resolution every 5 minutes, improving forecasts of and availability. Additionally, radiometric tracking of changes—increases in surface due to forest loss—has been used to quantify ; for example, MODIS-derived data indicated approximately a 2% increase in from 2000–2020, correlating with about 10% loss and associated carbon emissions.

Industrial and Biomedical Applications

In industrial manufacturing, non-contact infrared thermography plays a crucial role in processes, such as monitoring weld integrity during automotive assembly, where emissivity-corrected IR cameras adjust for surface properties to accurately measure distributions and detect defects like cracks or incomplete fusions without physical contact. Emissivity correction in these thermal applications ensures precise radiance-to-temperature conversions, enhancing reliability in high-heat environments like . Similarly, UV is essential for optimizing curing processes in coatings and adhesives, where radiometers measure and cumulative dose to ensure uniform and prevent under- or over-exposure, thereby maintaining product consistency in industries like and . In the energy sector, radiometry supports efficiency testing by simulating the AM1.5 global spectrum, a terrestrial of 1000 W/m² that represents average atmospheric conditions, allowing precise evaluation of photovoltaic performance under controlled conditions as defined by ASTM G173. For LED lighting systems, output involves measuring and to quantify and , enabling optimization of for applications like general illumination and displays. Biomedical applications leverage radiometry for diagnostic and therapeutic precision, as seen in (OCT) for retinal imaging, which uses near-infrared radiance in the 800–1300 nm range to achieve micrometer-resolution cross-sectional views of ocular tissues, aiding in the detection of conditions like . In phototherapy for , radiometric monitoring of (typically 30–40 µW/cm²/nm in the 460–490 nm band) ensures optimal reduction dosage, with devices calibrated to deliver effective exposure while minimizing risks like . Laser safety protocols further integrate radiometry through ANSI Z136.1 standards, which define maximum permissible exposure (MPE) limits—such as 1.8 J/cm² for near-IR pulses—to prevent ocular and skin damage by comparing measured irradiances against these thresholds. Recent advancements, as of 2025, include integrated with for non-destructive pharmaceutical purity checks, enabling rapid detection of contaminants or adulterants in tablets by analyzing signatures across hundreds of wavelengths, thus improving in .

References

  1. [1]
    Radiometry and photometry FAQ | - The University of Arizona
    Radiometry is the measurement of optical radiation, which is electromagnetic radiation within the frequency range between 3×10 11 and 3×10 16 Hz.
  2. [2]
    Radiometry and Applications | GTPE - Georgia Institute of Technology
    In this Radiometry and Applications course, you will gain knowledge about the transfer of optical radiant power between a source and a receiver.
  3. [3]
    [PDF] The National Measurement System for radiometry and photometry
    The history of radiometry and photometry at NBS may be divided into ... applications for radiometry did not exist; moreover, the technology required ...
  4. [4]
    [PDF] Photometry and Radiometry - UBC Computer Science
    It means that radiant flux density can be measured anywhere in three-dimensional space. This includes on the surface of physical objects, in the space between ...
  5. [5]
    [PDF] Self-study manual on optical radiation measurements : Part I
    THE NATIONAL MEASUREMENT LABORATORY provides the national system of. physical and chemical and materials measurement; coordinates the system with measurement. ...
  6. [6]
    [PDF] Radiometric Instrumentation and Measurements Guides for ... - NREL
    May 8, 1996 · Broadband Radiometry Applications to PV Performance Testing . ... History and Methods," Proceedings of 21st IEEE Photovoltaic Specialists ...
  7. [7]
    [PDF] symbols, terms, units and uncertainty analysis for radiometric sensor ...
    The scope includes the radiant properties of sources; the geometry of radiation transfer; the measurement equation used to predict sensor response; the ...
  8. [8]
    The Herschel Experiment | Cool Cosmos - Caltech
    Herschel discovered the existence of infrared light by passing sunlight through a glass prism in an experiment similar to the one we describe here.Missing: primary source
  9. [9]
    Herschel and the Puzzle of Infrared | American Scientist
    Most encyclopedias and physics books credit the great British astronomer Sir William Herschel with the discovery of infrared radiation in 1800.Missing: primary | Show results with:primary
  10. [10]
    (PDF) Kirchhoff's Law of Thermal Emission: 150 Years - ResearchGate
    Kirchhoff's law correctly outlines the equivalence between emission and absorption for an opaque object under thermal equilibrium.
  11. [11]
    Josef Stefan (1835 - 1893) - Biography - MacTutor
    Boltzmann, who was one of Stefan's students, showed in 1884 that this Stefan-Boltzmann law could be demonstrated mathematically. After this work, Stefan looked ...
  12. [12]
    [PDF] A Concise History of the Black-body Radiation Problem - arXiv
    Aug 12, 2022 · In 1879, Austrian physicist Josef Stefan gave an empirical law of temperature dependence of the black- body radiation based on Tyndall and ...<|separator|>
  13. [13]
    Samuel Pierpont Langley - Bolometer - NASA Earth Observatory
    May 3, 2000 · In 1878 he invented the bolometer, a radiant-heat detector that is sensitive to differences in temperature of one hundred-thousandth of a ...Missing: 1880s | Show results with:1880s
  14. [14]
    Defining the International System of Units (SI) | NIST
    Sep 29, 2023 · 1948: The candela is the luminance of a Planck radiator (a black body) at the temperature of freezing platinum. 1979: The candela is the ...
  15. [15]
    The Legacy of Nimbus - NASA Earth Observatory
    Oct 11, 2014 · Launched on August 28, 1964, Nimbus 1 sent back more than 27,000 images before its solar-power systems failed on September 22, 1964. It was the ...Missing: 1960s | Show results with:1960s
  16. [16]
    Optical Radiation Metrology and Standard Radiometers | NIST
    Feb 18, 2021 · Radiometric, radiance temperature, photometric, and color scales have been realized based on the spectral responsivity of standard detectors and radiometers.Missing: evolution | Show results with:evolution
  17. [17]
    Division 2 Technical Committees | CIE
    Division 1: Vision and Colour · Division 2: Physical Measurement of Light and Radiation · Division Officers · Official Division Members · Division 3: Interior ...Missing: key contributors
  18. [18]
    [PDF] A national measurement system for radiometry, photometry, and ...
    source can be referenced to absolute detectors. In 1979 the Conference Generale des Poids et Mesures (CGPM) adopted the 1977 CIPM recommendation for the ...
  19. [19]
    Detection of 15 dB Squeezed States of Light and their Application for ...
    Sep 6, 2016 · Squeezed states of light, with 15 dB squeezing, are used to calibrate photoelectric quantum efficiency, achieving 99.5% with 0.5% uncertainty.Missing: post- | Show results with:post-
  20. [20]
    Anatomy of an Electromagnetic Wave - NASA Science
    Aug 3, 2023 · This energy can be described by frequency, wavelength, or energy. All three are related mathematically such that if you know one, you can ...Missing: duality c
  21. [21]
    Electromagnetic Spectrum - Imagine the Universe! - NASA
    The amount of energy a photon has can cause it to behave more like a wave, or more like a particle. This is called the "wave-particle duality" of light. It ...
  22. [22]
    24.3 The Electromagnetic Spectrum – College Physics
    The relationship among the speed of propagation, wavelength, and frequency for any wave is given by $latex \boldsymbol{v_W = f \lambda} $, so that for ...
  23. [23]
    Electromagnetic Spectrum - Introduction - Imagine the Universe!
    The other types of EM radiation that make up the electromagnetic spectrum are microwaves, infrared light, ultraviolet light, X-rays and gamma-rays. You know ...
  24. [24]
    Kilogram: Mass and Planck's Constant | NIST
    May 14, 2018 · Planck's equation shows that energy, in turn, can be calculated in terms of the frequency ν of some entity such as a photon (a particle of light) ...
  25. [25]
    6.5 What is the origin of the Planck Function? | METEO 300
    Pe(λ)/π is called the Planck Distribution Function of Spectral Radiance and commonly has units of W steradian–1 m–2 nm–1 and is often denoted by the letter I. ...
  26. [26]
    6.6 Which wavelength has the greatest spectral irradiance?
    The result is the Wien Displacement Law: λmax=2898 μm KT. [6.5]. For the sun with a photospheric temperature of about 5780 K, λmax ~ 0.500 μm or 500 nm, which ...
  27. [27]
    Module 2 - Radiation Basics - met.nps.edu
    Jan 12, 2004 · ... EM radiation is said to be in phase or coherent radiation. Most radio or radar transmissions are coherent. Visible and IR radiation is ...
  28. [28]
    Realization of the candela | NIST
    The value of Km is 683 lm/W at 540 × 1012 Hz (555.016 nm in standard air). This frequency corresponds to the same frequency stated in the definition of spectral ...
  29. [29]
    CIE spectral luminous efficiency for photopic vision
    Values of spectral luminous efficiency for photopic vision, V(lambda), lambda in standard air, 1 nm wavelength steps, original source: CIE 018:2019.
  30. [30]
    Radiometry and Photometry: Review for Vision Optics | NIST
    Jan 1, 2000 · The difference between radiometry and photometry is that radiometry includes the entire optical radiation spectrum (and often involves ...
  31. [31]
    The basis of physical photometry, 2nd edition | CIE
    This publication describes the basic conventions and principles of modern physical photometry and explains how physical photometry relates to radiometry.
  32. [32]
    17-26-061 | CIE
    ... luminous efficacy (≈ 683 lm·W−1). Note 1 to entry: The term "luminous radiation" refers to optical radiation weighted with the CIE spectral luminous ...
  33. [33]
    Photometry | NIST - National Institute of Standards and Technology
    While radiometry measures light in all spectral regions, including ultraviolet and infrared, photometry only measures in the visible spectral region from 360 nm ...
  34. [34]
    [PDF] APPENDIX I THE SI SYSTEM AND SI UNITS FOR RADIOMETRY ...
    The following tables show radiometric and photometric quantities and symbols, definitions and units. RADIOMETRIC QUANTITIES. QUANTITY. SYMBOL DEFINITION. UNITS.
  35. [35]
    [PDF] 2. Radiometric Units
    The radiometric system of units describes the radiant energy emitted by a source or striking a receiver. The basic quantity in this system is radiant energy ...
  36. [36]
    Basic Radiance and Radiance Invariance - SPIE Digital Library
    If rays are traced along a lossless boundary between two media having different indices of refraction, the solid angle changes according to Snell's law.
  37. [37]
    Stefan-Boltzmann constant - CODATA Value
    Stefan-Boltzmann constant $\sigma$ ; Numerical value, 5.670 374 419... x 10-8 W m-2 K ; Standard uncertainty, (exact).Missing: law | Show results with:law
  38. [38]
    [PDF] Light, Radiometry, BRDF, and Radiosity
    Radiometry. The radiometric quantities that characterize the. distribution of light in the environment are: • Radiant Energy. • Radiance.
  39. [39]
    [PDF] Introduction to Radiosity
    The radiometric term radiosity means the rate at which energy leaves a surface, which is the sum of the rates at which the surface emits energy and reflects (or ...
  40. [40]
    Solar Radiation Basics | Department of Energy
    The amount of solar radiation that reaches any one spot on the Earth's surface varies according to: ... (W/m2). Radiation data for solar water heating and ...Basic Principles · Measurement · Distribution<|separator|>
  41. [41]
    [PDF] Radiometry and Photometry
    Radiometry is the detection and measurement of light waves in the optical portion of the electromagnetic spectrum which is further divided into ultraviolet, ...
  42. [42]
    [PDF] Spectral radiance calibrations - NIST Technical Series Publications
    • Spectral Radiometry. violet, visible, and infrared radiation; provides standards dissemination • Spectrophotometry. and measurement quality assurance ...
  43. [43]
    Irradiance – intensity, radiant flux, radiometry, measurement
    A related quantity is the spectral irradiance, which is the irradiance per unit frequency or wavelength interval. It has units of W / (m2 Hz) or W / (m2 nm), ...
  44. [44]
    Spectral Quantities - RP Photonics
    Spectral quantities in radiometry and photometry describe the distribution e.g. of a radiant flux over different optical frequencies or wavelengths.Missing: conventions | Show results with:conventions
  45. [45]
    Solar Iradiance - HyperPhysics
    By examining the 400-700 nm nominal range of visible light, you can see that the solar irradiance is similar to that of a blackbody radiator.
  46. [46]
    Natural and Simulated Solar Radiation - PubMed
    Oct 25, 2021 · Terrestrial solar spectral irradiance peaks at around 500 nm in the blue-green region, whereas the diffuse component peaks in the UVAI-blue region of the ...
  47. [47]
  48. [48]
    Invited Article: Advances in tunable laser-based radiometric ...
    Validation of a trap detector quantum efficiency model. The SIRCUS facility has broadly tunable narrow-band lasers that cover the full spectral range of the ...
  49. [49]
    Blackbody Spectral Radiance Calculator
    Radiant Emittance: 459.300 W·m⁻² · Total Radiance: 146.200 W·m⁻²·sr⁻¹ · Radiance at 10.000 μm: 9.924 W·m⁻²·sr⁻¹·μm⁻¹ · Band Radiance from 8.000 - 12.000 μm: 38.502 ...
  50. [50]
    Where solar is found - U.S. Energy Information Administration (EIA)
    Jul 12, 2024 · The two maps below show U.S. average annual solar radiation in kilowatthours (kWh) per square meter per day (kWh/m2/d) for direct normal ...Missing: exposure m²/
  51. [51]
    1.4 Basic Radiometric Quantities - Gigahertz-Optik
    Radiometry deals with the measurement of energy per time (= power, given in watts) emitted by light sources or impinging on a particular surface.Missing: NIST | Show results with:NIST
  52. [52]
    [PDF] Determining the uncertainty associated with integrals of spectral ...
    Abstract. : This report reviews the mathematical techniques required to evaluate a spectral integral, including understanding uncertainty associated with the ...Missing: extrapolation | Show results with:extrapolation
  53. [53]
    [PDF] User's Manual: Routines for Radiative Heat Transfer and Thermometry
    This routine can also be used to calculate the true temperature using the ratio method with the effective wavelength and the ratio temperature replacing the ...
  54. [54]
    A radiometric determination of the Stefan-Boltzmann constant and ...
    The total radiant exitance of a black body at the temperature of the triple point of water, Ttp (273.16 K), and at a series of other temperatures in the ...
  55. [55]
    [PDF] NPL REPORT The evaluation of the performance of bandpass filters ...
    These filter radiometers consist of an aperture which defines the area over which radiation is collected, a bandpass filter which restricts the range of.
  56. [56]
    Reference Air Mass 1.5 Spectra | Grid Modernization - NREL
    Mar 15, 2025 · ASTM G-173 spectra represent terrestrial solar spectral irradiance on a surface of specified orientation under one and only one set of specified atmospheric ...
  57. [57]
    [PDF] Untitled - MIT
    The mean horizontal intensity is the average intensity ... useful quantity; it is called radiant density. ... Radiant energy is the time integral of radiant flux.<|control11|><|separator|>
  58. [58]
    [PDF] SURVEY of the LITERATURE on the SOLAR CONSTANT and the ...
    The absorptance a is the ratio of the' energy absorbed to the energy incident; this definition is essentially similar to those of reflectance and transmittance ...
  59. [59]
    Spectral Emissivity Measurements - ScienceDirect.com
    As described in Chapter 2 , by Kirchhoff's law the spectral emissivity of a material is equal to its spectral absorptance under thermal equilibrium conditions.
  60. [60]
    [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 ...
  61. [61]
    [PDF] Lecture Radiative Transfer #.1 Kirchoff's law - CalTech GPS
    In the absence of Kirchoff's law of emissivity, the outgoing radiation would be a blackbody with a temperature such that the total amount of energy radiated ...
  62. [62]
    [PDF] Emissivity measurement for infrared thermography and radiative ...
    Aug 18, 2023 · Total emissivity is the emissivity calculated over all wavelengths ... The hemispherical emissivity is the integration of the directional ...
  63. [63]
    Understanding Classical Gray Body Radiation Theory | COMSOL Blog
    Nov 1, 2018 · A gray body is an imperfect black body; i.e., a physical object that partially absorbs incident electromagnetic radiation. The ratio of a gray ...
  64. [64]
    Radiation Heat Transfer - The Engineering ToolBox
    For the gray body the incident radiation (also called irradiation) is partly reflected, absorbed or transmitted. Heat radiation - incident reflected transmitted ...
  65. [65]
    Emissivity of Human Skin in the Waveband between 2µ and 6µ
    The emissivity of human skin ε(λ) in the range 2μ to 6μ has recently assumed considerable importance because of the increasing medical use of infrared scanners.
  66. [66]
    spectral reflectance - Illuminating Engineering Society
    The ratio of the reflected flux to the incident flux at a particular wavelength, λ, or within a small band of wavelengths, Δ λ, about λ.
  67. [67]
    [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 ...
  68. [68]
    [PDF] Geometrical considerations and nomenclature for reflectance
    general BSSRDF was introduced first and the BRDF was then developed as the important special case. Meanwhile, their publication [8] established the even ...
  69. [69]
    [PDF] 2.1.2 Reflectance, transmittance, and absorption Reflectance ... - SPIE
    Reflectance is the amount of flux reflected by a surface, normalized by the amount of flux incident on it. Transmittance is the amount of flux transmitted by a ...
  70. [70]
    Protected Silver Mirrors - Thorlabs
    Silver coated mirrors offer the highest reflectance in the visible-NIR spectrum of any metallic mirror, while also offering high reflectance in the IR.
  71. [71]
    Spectral integration of clear-sky atmospheric transmittance
    Broadband atmospheric transmittance is a key component in many broadband clear-sky solar radiation models, whose performance is inevitably bound to the ...
  72. [72]
    [PDF] NIST Technical Note 1621: Optical radiation measurements based ...
    The most accurate temperature determination of blackbody sources is based on the spectral irradiance and radiance responsivity determinations of the filter ( ...Missing: exitance radiosity
  73. [73]
    [PDF] Extremely broadband calibrated bolometers and microbolometer ...
    The basic components of a bolometer calibrated with electrical substitution include an absorber, thermometer, weak thermal link, heatsink, and resistive heater.
  74. [74]
    [PDF] Uncooled multimirror broad-band infrared microbolometers
    These suspended micromesh bolometers have achieved noise equivalent power (NEP) as low as 2.7. 10. W/Hz at 304 mK with a response time of. 24 ms [14]. Other ...
  75. [75]
    [PDF] Broadband Radiometric LED Measurements
    The measurements of the low-NEP pyroelectric detector were performed in AC mode using a lock-in amplifier and a chopping frequency of 10.5 Hz. A beam geometry ( ...
  76. [76]
    Calibration of Transition-edge Sensor (TES) Bolometer Arrays with ...
    Oct 7, 2022 · The current and future cosmic microwave background (CMB) experiments fielding kilopixel arrays of transition- edge sensor (TES) bolometers ...
  77. [77]
    Ultraviolet Photodetectors: From Photocathodes to Low-Dimensional ...
    The quantum efficiency η of a detector is defined by the number of e-h pairs generated by the photodetector per incident photon. Its standardized value is ...
  78. [78]
    [PDF] Optical Sensors for Planetary Radiant Energy (OSPREy)
    Jul 1, 2012 · uses a Sony linear silicon CCD array with 2,048 pix- els and covers the wavelength range 200–1,100 nm. • StellarNet (www.stellarnet-inc.com) ...
  79. [79]
    [PDF] Photomultiplier Handbook - Picosecond Timing Project
    A photomultiplier is a sensitive detector of radiant energy, using a photocathode and secondary-emission amplification for high sensitivity.
  80. [80]
    [PDF] Metrology of Optical Systems - SPIE
    Fourier-transform infrared (FTIR) spectrometer is an optical instrument used typically to measure the infrared radiometric properties (reflectivity or ...
  81. [81]
    [PDF] Optical Radiation Measurements for Photovoltaic Applications - NREL
    Broadband radiometers (pyranometers and pyrheliometers) are used to assess solar resources for renewable applications and develop and validate broadband solar ...
  82. [82]
    [PDF] The NOAA Climate Monitoring and Diagnostic Laboratory Solar ...
    A radiometric scale is required to convert broadband radiometer signals to irradiance units. ... scale from a cavity radiometer to a pyranometer is accomplished.
  83. [83]
    Infrared cryogenic blackbody broadband calibration | NIST
    Jan 28, 2010 · NIST calibrates blackbodies using cryogenic chambers at 20K, radiometers, and measurements of radiant power to provide radiometric temperature.
  84. [84]
    [PDF] Cryogenic blackbody calibrations attheNational Institute of ...
    An absolute cryo- genic radiometer (ACR) of the electrical substitu- tion type has been developed as the standard reference detector for the LBIR calibrations.
  85. [85]
    Cryogenic Blackbody Calibrations at the National Institute of ...
    The IR sources are sent to NIST by customers from industry, government, and university laboratories. An absolute cryogenic radiometer is used as the standard ...
  86. [86]
    Spectroradiometry of sources | NIST
    Jan 21, 2010 · NIST uses transfer lamps and integrating sphere sources for spectral radiance and irradiance calibration, using blackbodies and monochromators, ...
  87. [87]
    [PDF] Integrating Sphere Radiometry and Photometry | Labsphere
    In total luminous flux measurements, the spectral efficiency of the integrating sphere must be included in the spectral response analysis of the complete system ...
  88. [88]
    Spectral radiance standard transfer for an integrating sphere as a ...
    Nov 25, 2022 · A National Institute of Standards and Technology (NIST)-traceable filter radiometer was used as the transfer scale to calibrate the integrating ...
  89. [89]
    [PDF] JCGM 100:2008 (GUM 1995 with minor corrections - BIPM
    should be recognized that a Type B evaluation of standard uncertainty can be as reliable as a Type A evaluation, especially in a measurement situation where ...Missing: radiometry | Show results with:radiometry
  90. [90]
    6. Uncertainty Estimates in Radiometry - ScienceDirect.com
    The GUM defines type A uncertainties as those evaluated by statistical means and type B as those evaluated by other means. Systematic effects are most often ...
  91. [91]
    [PDF] Calibration and Measurement Uncertainty Estimation of Radiometric ...
    The GUM method employs two types of uncertainty estimates: Type A and Type. B. Type A uncertainty uses a statistical method of valuation, such as standard ...Missing: ISO framework
  92. [92]
    Spectroradiometer spectral calibration, ISRF shapes, and related ...
    The spectral calibration data are obtained by making measurements of light from a gas discharge lamp producing emission spectral lines [see Fig. 1 (a)]. Since ...
  93. [93]
    B.RCLab Wavelength scale calibration
    For spectrographs, the wavelength calibration corresponds to the identification of the central wavelength reaching each pixel of a detector. For both kind of ...3 -- Methodology · 3.1 -- Wavelength... · 4.1 -- Spectral Lamps
  94. [94]
    3.2 Parameters of a Spectroradiometer - Gigahertz-Optik
    The wavelength accuracy of a spectrometer is defined by a variety of quantities. In principle, the quality of the wavelength calibration presents the most ...
  95. [95]
    Solar Irradiance Science | Earth - NASA
    One of the major SORCE contributions was to establish a more accurate value at 1361 Wm-2, which leads to 340 W m-2 for the globally averaged solar input to ...Missing: annual average J/
  96. [96]
    Inter-comparison Example | NIST
    This paper overviews results from Consultative Committee on Photometry and Radiometry (CCPR) the Supplementary Comparison S2.
  97. [97]
    Laboratory Intercomparison of Radiometers Used for Satellite ...
    This paper describes the activities and results of the first two phases of LCE-2: the SI-traceable radiometric calibration and indoor intercomparison, the ...
  98. [98]
    [PDF] Observational Astrophysics 1. Astronomical Measurements
    To continue the tradition of confusing names, flux can also be called “flux density” (density in the sense that the measurement is per unit frequency or unit ...
  99. [99]
    a stellar bolometric corrections database with variable extinction ...
    Bolometric corrections (BCs) are usually applied to the absolute magnitude of a star to obtain its bolometric absolute magnitude or luminosity, or conversely, ...
  100. [100]
    Periodicity of quasar and galaxy redshift - Astronomy & Astrophysics
    The relative change in wavelength is termed the redshift z and is defined as z = Δλ/λ0. Initially, astronomers ascribed the redshift of the spectrum to the ...
  101. [101]
    The Stellar Temperature Scale - Astrophysics Data System
    According to Planck's law for black-body radiation, the logarithmic intensity difference for two light sources is a linear function of i/X, and the gradient is ...
  102. [102]
    Optical atmospheric extinction over Cerro Paranal
    The present study was conducted to determine the optical extinction curve for Cerro Paranal under typical clear-sky observing conditions.
  103. [103]
    Infrared Thermography for Temperature Measurement and Non ...
    Infrared thermography (IRT) is a science using infrared radiation to acquire thermal information, where the intensity of radiation is related to temperature.
  104. [104]
    Implementation of Non-Contact Temperature Distribution Monitoring ...
    It is based on low-cost infrared thermal cameras linked with a calculation unit in order to produce a corrected thermal map of the surveyed structure at a ...
  105. [105]
    [PDF] Measurement Uncertainty of Surface Temperature Distributions for ...
    Aug 10, 2021 · This paper describes advances in measuring the characteristic spatial distribution of surface temperature and emissivity during laser-.
  106. [106]
    [PDF] Printers' Guide Radiometry and Methods of UV Monitoring - RadTech
    Radiometry is a powerful analytical tool for UV curing and invaluable as a quality control (QC) tool for process monitoring. Radiometry provides the ...Missing: applications | Show results with:applications
  107. [107]
    [PDF] Radiometric Methods for UV Process Design and Process Monitoring
    Radiometry is a powerful analytical tool for UV curing process design and invaluable as a QC tool for process monitoring. It is important to identify the ...
  108. [108]
    [PDF] LIGHT EMITTING DIODES | Labsphere
    Almost everyone is familiar with light-emitting diodes (LEDs) from their use as indicator lights and numeric displays on consumer electronic devices.
  109. [109]
    Standard Solar Spectra - PVEducation
    The AM1. 5 Direct (+circumsolar) spectrum is defined for solar concentrator work. It includes the the direct beam from the sun plus the circumsolar component ...
  110. [110]
    Optical Coherence Tomography - Medical Imaging Systems - NCBI
    OCT is an interferometry based three-dimensional imaging modality that can be used on scattering media, including several types of body tissues. It provides ...
  111. [111]
    Optical Coherence Tomography (OCT): Principle and Technical ...
    Aug 14, 2019 · Optical coherence tomography (OCT) is a non-contact imaging technique which generates cross-sectional images of tissue with high resolution.
  112. [112]
    Evaluation and Treatment of Neonatal Hyperbilirubinemia - AAFP
    Jun 1, 2014 · Measurements should be made with a radiometer specified by the manufacturer of the phototherapy system. See Appendix 2 in the original ...
  113. [113]
    [PDF] Neonatal Phototherapy: Monitoring the Optimal Dose - FI-Admin
    Dec 15, 2022 · Phototherapy, as a treatment for neonatal hyper- bilirubinemia, depends on how far the skin is from the light source and how light penetrates ...
  114. [114]
    [PDF] American National Standard for Safe Use of Lasers
    For exposed skin areas exceeding 1000 cm2, the MPE is 10 mW∙cm-2. 9. Measurements and Calculations. Guidance on laser safety measurements can be found in ANSI ...
  115. [115]
    ANSI Z136.4-2021: Laser Safety Measurements
    Evaluation consists of comparing measured exposures with the maximum permissible exposure (MPE) values found in ANSI Z136. 1 that are based on the ability of ...
  116. [116]
    Hyperspectral Imaging Techniques for Lyophilization - NIH
    Jul 23, 2025 · This section covers various modeling strategies, including HSI data preprocessing methods, spectral unmixing modeling, classification and ...
  117. [117]
    A review of machine learning in hyperspectral imaging for food safety
    This study aims to differentiate the various ML models employed in food safety and discusses the challenges and future directions for effectively quantifying ...A Review Of Machine Learning... · 3. Hyperspectral Imaging... · 4. Essential Algorithms In...