Fact-checked by Grok 2 weeks ago

Light intensity

The term "light intensity" is often used ambiguously in physics and . In general contexts, it refers to the amount of light energy or delivered per unit area to a surface ( or ), quantifying how brightly an area is illuminated or how much light power is concentrated in a given space. In radiometry, however, intensity specifically denotes , the power per unit . This distinction is important, as (power per unit area) from a follows the , decreasing with the square of the distance from the source. The concept encompasses both radiometric measures, which deal with physical without regard to human perception, and photometric measures, which weight the according to the human eye's sensitivity. is defined as the per unit , expressed in watts per (W/sr), representing the directional distribution of light power from a source. In contrast, is the photometric equivalent, measured in (cd), which accounts for the eye's spectral response peaking at 555 nanometers (corresponding to at 540 × 10¹² ); it is defined such that a source emitting at this frequency with a of 1/683 W/sr has a of 1 cd. These distinctions are crucial in fields like astronomy and , where unweighted physical energy flux is needed, versus , where perceived matters. A related quantity, (or illumination), specifically measures the photometric on a surface and is the per unit area, with the SI unit of (lx), where 1 lx equals 1 per square meter. The itself derives from the , equaling the emitted into a unit by a source of 1 . Measurement typically involves photometers or meters calibrated to standard illuminants that mimic daylight or incandescent sources, ensuring accuracy for applications such as , where affects , or plant growth, where optimal levels support . In , (W/m²) serves a similar for total electromagnetic power, independent of visibility.

Fundamentals

Definition

Light is a form of , consisting of oscillating electric and magnetic fields that propagate through space as waves or packets of energy known as photons, with the spanning s from about 400 to 700 nanometers. This radiation carries energy, and its characteristics, such as and , determine properties like color and . In physics, light intensity refers to the per unit area on a surface to the direction of , known as in , quantifying how much is delivered to a given area. It differs from total , which measures overall output per unit time regardless of direction, and from , which accounts for through a surface without directional specificity; intensity emphasizes the density of on the receiving surface. A related quantity, , is the per unit , typically modeled for s. For a emitting P uniformly over a \Omega (measured in steradians), the I is given by I = \frac{P}{\Omega}, where this relation establishes as a directional measure essential for understanding light and distribution. The concept of light intensity originated in 19th-century , initially within photometry to describe the apparent brightness of sources like candles, driven by practical needs for standardizing illumination in emerging industrial and scientific applications. By the late 19th and early 20th centuries, as electromagnetic theory advanced through works like those of James Clerk Maxwell, the term evolved into a precise radiometric quantity, integrating wave and particle models of while distinguishing it from related metrics like . This shift provided a foundational framework for modern , enabling quantitative analysis of 's energy flow without reliance on human perception.

Relation to Other Optical Quantities

Light intensity, particularly in the context of , describes the radiated by a source per unit , distinguishing it from , which is the total emitted regardless of direction. , denoted as Φ and measured in watts, represents the overall flow from a source, while radiant intensity I, in watts per (W/sr), captures the directional concentration of that , essential for understanding how light propagates in specific directions. In contrast, irradiance E quantifies the power incident on a surface per unit area, in watts per square meter (W/), focusing on the density of received rather than emitted directionality. Radiance L further refines this by specifying power per unit per unit , often per unit , in units of W/··Hz, providing a comprehensive measure for extended sources where both position and direction matter. These distinctions position intensity hierarchically between total and the more detailed radiance, emphasizing over areal or total integration. For extended sources, relates directly to radiance through integration over the source's surface. The key equation is: I(\omega) = \int L(\omega, \theta, \phi) \cos \theta \, dA where I(\omega) is the in \omega, L(\omega, \theta, \phi) is the radiance at position on the surface, \theta is the angle between the surface and \omega, and the covers the source area dA. This simplified form for point-like approximations aggregates radiance contributions, accounting for the foreshortening effect via \cos \theta. A conceptual illustration of intensity's behavior involves its falloff with under the for point sources: the same spreads over progressively larger spherical surfaces, reducing as $1/r^2, where r is . Consider a versus a nearby —both may have similar , but the lamp delivers higher due to proximity, demonstrating geometric spreading without altering the source's inherent directional power. A common misconception arises from equating intensity with everyday "brightness," which actually aligns more closely with perceived radiance in photometry, influenced by human visual sensitivity rather than pure radiometric power per . This confusion often leads to overlooking intensity's strict directional focus in favor of subjective visual impressions.

Radiometry

Radiant Intensity

Radiant intensity, denoted I_e, is defined as the amount of radiant flux emitted per unit solid angle from a point source in a given direction, measured in watts per steradian (W/sr). This quantity quantifies the directional distribution of radiant power and applies to electromagnetic radiation across all wavelengths, independent of human visual perception. The radiant , I_{e,\lambda}(\lambda), extends this concept to specify the distribution per unit , with units of /(sr·μm) or similar, allowing analysis of wavelength-dependent emission. For sources, the total radiant is obtained by the form over all wavelengths: I_e = \int_0^\infty I_{e,\lambda}(\lambda) \, d\lambda. This is essential for sources like emitters where emission spans multiple bands. For point sources, the describes how the (radiant flux per unit area) decreases with due to over an expanding spherical . Consider an isotropic emitting total \Phi_e uniformly in all directions. The for such a source is I_e = \frac{\Phi_e}{4\pi}, constant with . At a distance r from the source, this spreads over the surface of a with area $4\pi r^2. The E_e at that distance, normal to the radial direction, is then E_e = \frac{\Phi_e}{4\pi r^2} = \frac{I_e}{r^2}, demonstrating E_e \propto 1/r^2. This derivation follows from the : the total \Phi_e remains constant, but the area illuminated increases quadratically with radius, leading to the dilution of . For non-isotropic sources, the on-axis follows a similar form, E_e(\theta=0) = I_e / r^2, where I_e is the in the radial direction. In infrared contexts, is critical for characterizing blackbody radiators used in , where the follows ; for cavity designs approximating ideal blackbodies, the is the multiplied by the effective aperture area. In ultraviolet applications, beams can achieve high in narrow beams due to low , enabling applications such as precise material .

Intensity in Wave and Particle Models

In the classical wave model of light, intensity is defined as the time-averaged power per unit area carried by the electromagnetic wave, which is proportional to the square of the electric field amplitude, I \propto |E|^2. This relationship arises because the energy density of the electromagnetic field scales with E^2 and B^2, and for plane waves where |B| = |E|/c, the dominant contribution is from the electric field. The instantaneous energy flux is given by the Poynting vector \mathbf{S} = \frac{1}{\mu_0} \mathbf{E} \times \mathbf{B}, and for monochromatic plane waves, the time-averaged intensity is I = \frac{1}{2} c \epsilon_0 E_0^2, where E_0 is the peak electric field amplitude, c is the speed of light, \epsilon_0 is the vacuum permittivity, and \mu_0 is the vacuum permeability. In the quantum particle model, is described as a stream of , each carrying \hbar \omega, where \hbar is the reduced Planck's constant and \omega is the . The then represents the due to the photon flux density N ( per unit area per unit time), yielding I = N \hbar \omega. This formulation equates the classical to the product of photon arrival rate and individual photon , bridging wave and particle descriptions for non-interacting fields. For sources, the spectrum of is integrated over frequencies, but the local at a given follows this photon-based scaling. Coherence plays a crucial role in the wave model, as patterns modulate the distribution. In Young's double-slit experiment, light from a source passes through two closely spaced slits, producing two wavefronts that overlap and , resulting in an pattern I(\theta) = 4 I_0 \cos^2(\delta/2), where I_0 is the from a single slit and \delta is the phase difference dependent on the path length difference. Constructive at maxima yields four times the single-slit , while destructive at minima reduces it to zero, demonstrating how wave superposition directly governs observable variations. This effect requires temporal and spatial to maintain stable phase relations between the slits. The transition from classical to quantum descriptions of intensity is exemplified by , where classical Rayleigh-Jeans theory predicted infinite intensity (the ""), incompatible with observations. resolved this by quantizing energy in discrete packets, leading to for the B(\nu, T) = \frac{2 h \nu^3}{c^2} \frac{1}{e^{h\nu / kT} - 1}, where h is Planck's constant, \nu is , T is , and k is Boltzmann's constant. This formula accurately describes the radiance spectrum peaking at a frequency that shifts with , providing the foundational quantum explanation for thermal emission and reconciling wave-particle duality in intensity distributions.

Photometry

Luminous Intensity

is the photometric quantity analogous to , but weighted according to the human eye's sensitivity to visible light wavelengths. It quantifies the emitted by a in a particular direction per unit , expressed in the unit of (cd), where 1 cd equals 1 per (lm/sr). The is defined by fixing the luminous efficacy of monochromatic radiation at a of 540 × 10¹² Hz (corresponding to 555 nm under standard conditions) to exactly 683 lm/W. For a general distribution, I_v is calculated as I_v = 683 \int_0^\infty I_e(\lambda) V(\lambda) \, d\lambda, where I_e(\lambda) is the in W/sr, and V(\lambda) is the CIE 1931 photopic luminosity function, which peaks at 1 for 555 nm and describes the eye's relative sensitivity across the . This weighting ensures that reflects perceived brightness rather than total energy. In low-light conditions, adaptations account for shifts in visual sensitivity. , occurring at intermediate levels (approximately 0.001 to 3 /m²), uses a blended luminosity function V'(\lambda) that interpolates between photopic and scotopic responses, as defined in CIE models like the 2010 Mesopic System. , dominant in very dim conditions below 0.01 /m², employs the CIE 1951 scotopic luminosity function V''(\lambda), which peaks near 507 nm to reflect dominance. These functions adjust the 683 /W efficacy constant accordingly for accurate photometry in such regimes. The standardization of luminous intensity traces to the 9th General Conference on Weights and Measures in 1948, which adopted the as an to replace earlier candle-based definitions. This was revised in the 2019 SI redefinition, which anchored the unit to the exact value of 683 lm/W for the specified monochromatic source, eliminating reliance on physical artifacts and ensuring stability based on fundamental constants.

Human Perception Factors

Human perception of light intensity is shaped by biological mechanisms in the , which adjust sensitivity to varying environmental conditions rather than directly mirroring physical measurements. The contains and photoreceptors, with s dominating in low-light conditions for and s in brighter , leading to nonlinear responses that influence how is subjectively experienced. The describes a shift in relative perceived of colors under dim illumination, where shorter-wavelength and greens appear brighter compared to longer-wavelength , due to the increased of photoreceptors to shorter wavelengths during dark . This , first observed in the , becomes evident at twilight or in low-light settings, such as when foliage stands out more vividly against a dim . Adaptation processes further modulate perceived intensity over time. In dark adaptation, the dilates to allow more entry, while photopigments regenerate, increasing sensitivity by up to 100,000-fold over 30-40 minutes, making faint lights appear progressively brighter. Conversely, adaptation occurs rapidly within seconds to minutes, involving constriction and photopigment bleaching to reduce sensitivity and prevent in bright environments, thus compressing the range of perceivable intensities. Psychophysical principles like the Weber-Fechner law quantify this perceptual scaling, stating that the perceived change in intensity is proportional to the logarithm of the physical intensity: I_{\text{perceived}} \propto \log(I_{\text{physical}}). This logarithmic relationship, derived from early experiments measuring just-noticeable differences in levels, explains why doubling physical intensity yields diminishing subjective increments, with a constant Weber fraction (around 0.02 for ) indicating proportional sensitivity across intensities. Color temperature also affects subjective intensity, with cooler lights (around 6500 K, resembling daylight) often perceived as brighter and more alerting than warmer lights (around 2700 K, like incandescent bulbs), even at equivalent physical output, due to greater stimulation of cone pathways sensitive to blue hues. Studies in controlled environments show that cool white lighting enhances perceived across various color contexts, influencing mood and visual comfort.

Measurement and Units

Radiometric Units

In , the primary unit for measuring , defined as the per unit , is the watt per (W/sr) for . This quantifies the power emitted by a source in a specific , independent of , and is fundamental for characterizing point-like or directional sources. For spectral radiant intensity, which describes the distribution across wavelengths or frequencies, the units are watt per per hertz (W/sr/Hz) for frequency-based spectra or watt per per meter (W/sr/m) for wavelength-based spectra. These allow precise quantification of broadband or polychromatic sources by integrating over the spectrum to yield total in W/sr. Prior to the adoption of the system, was often expressed in the CGS unit of erg per second per (erg/s/sr), where 1 erg/s/sr equals 10^{-7} W/sr due to the conversion factor between the erg (10^{-7} J) and the joule as the base for the watt. This historical unit facilitated early radiometric measurements in and but has been largely supplanted by SI units for international standardization. SI traceability for measurements is ensured through standards maintained by national institutes such as the National Institute of Standards and Technology (NIST) and the International Bureau of Weights and Measures (BIPM), which provide reference artifacts and protocols to achieve uncertainties below 1% for electro-optical instruments. These standards, often involving transfer lamps or integrating spheres, link laboratory instruments to the and second, enabling consistent global comparisons. A common conversion relates radiant intensity I (in W/sr) to irradiance E (in W/m²) at a distance r (in m) from an isotropic point source, given by E = I / r^2, which accounts for the dilution of flux over a spherical surface. This relation is essential for practical applications, such as estimating illumination from distant sources, though it assumes perpendicular incidence and neglects atmospheric effects.

Photometric Units

The (cd) is the (SI) base unit for , defined as the , in a given direction, of a source that emits of 540 × 10¹² hertz and that has a in that direction of ¹/₆₈₃ watt per . This definition, established in 2019, fixes the of at 555 nanometers to exactly 683 lumens per watt, providing a precise link between photometric and radiometric quantities tailored to human vision. Related photometric quantities derive from the to quantify light as perceived by the under standard visibility conditions. The () measures , representing the total amount of visible light emitted by a source across all directions, equivalent to the flux from a uniform source of one intensity subtending a of one . The () quantifies , or the incident per unit area on a surface, with one equal to one per square meter, commonly used to specify lighting levels for visibility in design standards. The conversion between radiometric power and luminous quantities relies on the luminous efficacy formula, which weights spectral power by the eye's : \Phi_v = 683 \int P_e(\lambda) V(\lambda) \, d\lambda Here, \Phi_v is the in lumens, P_e(\lambda) is the spectral radiant power in watts, and V(\lambda) is the spectral , with the constant 683 lm/W reflecting the maximum efficacy at 555 nm. This ensures photometric units account for the human visual system's peak in the green-yellow spectrum. The (CIE) established the foundational V(\lambda) function in 1931 as the photopic luminosity curve, based on 2° field-of-view experiments averaging sensitivity data from multiple observers, peaking at unity for 555 nm and dropping to near zero below 400 nm and above 700 nm. This standard enables consistent visibility assessments under well-lit (photopic) conditions, influencing global lighting regulations. In 2006, the CIE advanced mesopic photometry to address intermediate lighting levels (approximately 0.001 to 10 cd/m²), where both and responses contribute, by proposing a state-of-the-art system blending photopic and scotopic functions with an adaptation coefficient to better model perceived in low-light environments like nighttime roads. This update, formalized in CIE Publication 191 (2010), enhances accuracy for applications requiring visibility at transitional luminances without altering core photopic standards. Prior to the candela's adoption, photometric measurements used legacy units like , originally defined by the light output of a standardized oil burning at a specified rate, which suffered from reproducibility issues due to material variations. The 9th General Conference on Weights and Measures in ratified the to replace these inconsistent units, phasing out candlepower internationally by aligning photometry with the framework and the emerging V(λ) standard.

Applications

In Illumination and Lighting Design

In illumination and lighting design, achieving uniform light intensity distribution is essential for creating functional and comfortable indoor environments. The Illuminating Engineering Society (IES) provides standards for recommended levels, measured in (lx), to ensure adequate visibility without excessive energy use. For general office spaces, IES RP-1 recommends horizontal of 300-500 lx at workplane height to support tasks like reading and computer use, promoting uniformity ratios of no more than 3:1 across the room to minimize shadows and visual fatigue. These guidelines emphasize even intensity distribution through strategic fixture placement and reflectors, balancing productivity with in architectural planning. Light source selection plays a critical role in optimizing intensity output and efficacy, defined as luminous flux per unit power in lumens per watt (lm/W). Incandescent lamps typically achieve only 10-20 lm/W due to high thermal losses, whereas modern LEDs can exceed 200 lm/W, with some high-efficiency models reaching over 220 lm/W as of 2025, enabling brighter illumination with significantly lower power consumption. This disparity allows LED systems to deliver equivalent intensity—such as 1,000 lm for task lighting—with roughly 80-90% less energy than incandescents, making them a cornerstone of sustainable in commercial and residential applications. Controlling is vital to prevent discomfort from high-intensity sources, often managed through intensity gradients that gradually reduce light output toward viewing angles. The Unified Glare Rating (UGR) method, standardized by the (CIE), quantifies discomfort glare by comparing luminaire to background , incorporating factors like source size, , and angular spread in its calculation: \text{UGR} = 8 \log_{10} \left( \frac{0.25}{L_b} \sum \frac{L^2 \omega}{p^{1.6}} \right), where L_b is background (cd/m²), L is source (cd/m²), \omega is (sr), and p is Guth . Designers apply this by using diffusers, louvers, or indirect fixtures to create smooth intensity falloff, targeting UGR values below 19 for offices to ensure visual comfort without compromising overall . Smart lighting systems further enhance efficiency by adaptively adjusting based on , daylight, and user needs via sensors and controls. These technologies can reduce use in commercial spaces by up to 50% compared to static setups, primarily by dimming or switching fixtures in unoccupied areas and integrating with . For instance, sensors and daylight harvesting enable real-time , aligning with IES standards while minimizing waste in dynamic environments like offices and .

In Scientific Research and Astronomy

In , light intensity plays a crucial role in analyzing and molecular levels through or line strengths. The relative intensities of lines from a sample in follow the , where the intensity I of a line is proportional to the N of the upper , the degeneracy g, and the exponential factor e^{-E/kT}, expressed as I \propto N g e^{-E/kT}, with E as the of the level, k the Boltzmann constant, and T the temperature. This relationship allows researchers to determine temperature by measuring intensity ratios between lines from thermally coupled levels, as the ratio simplifies to \frac{R(T)}{R(T_0)} = \exp\left[-\frac{\Delta E_{21}}{k_B}\left(\frac{1}{T} - \frac{1}{T_0}\right)\right], where \Delta E_{21} is the energy gap between levels. For instance, in luminescence thermometry using Gd³⁺ ions, intensity ratios from crystal field splittings (e.g., 72 cm⁻¹ gap for low temperatures around 30–51 K) enable high-precision temperature mapping with sensitivities up to 11.6% K⁻¹, aiding studies in plasma physics and material science. In astronomy, measurements of stellar light intensity are essential for determining cosmic distances via the , which relates the m (observed brightness) to the M (intrinsic brightness at 10 parsecs). The fundamental relation is m - M = 5 \log_{10} d - 5, where d is the distance in parsecs, derived from the of light intensity dilution over distance. Corrections for interstellar extinction and are applied in practice, such as adding terms for dust absorption, to refine distance estimates for galaxies and clusters. This method underpins the , enabling the calibration of Hubble's constant and mapping the universe's expansion history through observations of Cepheid variables or Type Ia supernovae. Laser research utilizes light intensity to probe ultrafast phenomena, particularly in pulses where peak intensities reach terawatt per square centimeter (TW/cm²) scales, inducing nonlinear effects like formation. In interactions with matter, such as ultrathin liquid sheets, a critical around 50 TW/cm² triggers rapid decline due to and plume expansion at velocities up to 21,000 m/s. These intensities, achieved with few-cycle pulses, facilitate applications in high-harmonic generation and science, with clamped values near 55 TW/cm² in filamentation experiments limiting further escalation while enabling controlled energy deposition. Post-2020 advances in intensity mapping have enhanced cosmological probes of by statistically mapping large-scale structure through line emissions like neutral hydrogen's 21 cm signal, without resolving individual galaxies. The Five-hundred-meter Aperture Spherical radio Telescope (FAST) array, upgraded for wide-band observations (0 < z < 2.5), achieves baryon acoustic oscillation constraints to ~1% accuracy, improving and matter density parameter estimates when combined with surveys like SKA1-Mid. This technique traces dark matter-dominated filaments by correlating intensity fluctuations with underlying gravitational potentials, yielding equation-of-state constraints such as \sigma(w_0) = 0.09, and supports analyses of 21 cm maps to reveal .

References

  1. [1]
    Light Intensity: Units, Measurement, Examples & Uses - Vedantu
    Rating 4.2 (373,000) Light intensity is the rate at which light spreads over a surface, measured in lux, which is lumens per square meter. Candela is the fundamental unit.
  2. [2]
    Intensity | - The University of Arizona
    The dimensionality of intensity is (photometric or radiometric) flux per steradian; the luminous and radiant are merely adjectives.
  3. [3]
    Candela | NIST - National Institute of Standards and Technology
    May 16, 2019 · While “radiant intensity” is the intensity of light without adjustments for human vision, luminous intensity is a “photometric” unit that ...
  4. [4]
    [PDF] symbols, terms, units and uncertainty analysis for radiometric sensor ...
    6) 1.2.3 RADIANT INTENSITY/PHOTON INTENSITY. The average radiant intensity Imeis the ratio of the total flux [W] to the total solid angle o)s. about the source ...
  5. [5]
    Lux | Light Measurement, Photometry & Illumination - Britannica
    Sep 22, 2025 · One lux (Latin for “light”) is the amount of illumination provided when one lumen is evenly distributed over an area of one square metre.
  6. [6]
    Chapter 1: The Physics of Light and Color
    Qualitatively, brightness is the lightness or darkness of a color, and is a measure of the light's intensity. Unlike hue and saturation, however, brightness ...
  7. [7]
    The Basics of Light - William P. Blair
    Optical light runs from about 400 to 700 nanometers. It's the same way as we move throughout the electromagnetic spectrum. Each range of light we have defined ...
  8. [8]
    Power Per Unit Solid Angle - HyperPhysics
    Power per unit solid angle, also called pointance, expresses the directionality of radiated energy, measured in watts per steradian for radiant power.
  9. [9]
    Radiant Intensity – radiant flux per unit solid angle - RP Photonics
    Radiant intensity is radiant flux per unit solid angle, measured in W/sr, and applies to emitted, transmitted, reflected, or received radiation.
  10. [10]
    Radiant Intensity - ECE 532, 1. Radiometry - OMLC
    Radiant intensity is the power directed into a cone of a solid angle, divided by that solid angle, measured in W/sr. For isotropic sources, it's P/(4π) [W/sr].
  11. [11]
    A History of Light and Colour Measurement, Science in the Shadows
    ... origins from the measurement of light intensity and radiant heat. During its rise in the 19th century, the subject was dominated by utilitarian need and ...
  12. [12]
    [PDF] Physics, Chapter 36: Light and Its Measurement
    We may express the intensity of a source of light in terms of the rate at which it emits radiation, in units of power, or energy emitted per unit time. Since ...<|separator|>
  13. [13]
    [PDF] Light Radiometry and Photometry
    Definition: The radiant (luminous) intensity is the power per unit solid angle emanating from a point source. Definition: The irradiance (illuminance) is the ...
  14. [14]
    [PDF] Flux and Irradiance - SPIE
    The commonly used spatial quantities are flux, irradiance, intensity, and radiance. Flux, Φ, is the optical power or rate of flow of radiant energy. Irradiance ...
  15. [15]
    [PDF] and color - UMIACS
    ▫ Radiance L ( θ. 1. ) is power (energy flux) emitted per unit area into a ... ▫ Radiant intensity is integral of radiance with respect to dA over the ...
  16. [16]
    Understanding Radiance (Brightness), Irradiance and Radiant Flux
    Radiance is power per unit area per unit solid angle, irradiance is power per unit area incident on a surface, and radiant flux is radiant energy per unit time.
  17. [17]
    [PDF] Radiometry is the measurement of radiation in the electromag
    Radiant Intensity is the amount of flux emitted through a known solid angle. It is measured in watts/steradian. To measure radiant intensity, start with the ...
  18. [18]
    [PDF] Radiometry and Photometry
    Radiant Intensity (I = d d ). Radiant Intensity is the amount of flux emitted through a known solid angle. It is measured in watts/steradian. To measure ...
  19. [19]
    [PDF] Light, Radiometry, BRDF, and Radiosity
    • Radiant Intensity: Radiant power per solid angle of a point source ... Radiance at x from x'. So now rendering equation is (without emitter). Next ...
  20. [20]
    Inverse Square Law for Light - HyperPhysics Concepts
    The light from a point source can be put in the form where E is called illuminance and I is called pointance.<|control11|><|separator|>
  21. [21]
    [PDF] Radiometry Energy & Power
    Radiant intensity, I, is measured in W/sr. Intensity is power per unit solid angle. Radiance, L, is measured in W/m2-sr. Radiance is power per unit projected ...
  22. [22]
    Radiometric Measurement - Newport
    The radiant intensity (Ir) and radiance (Lr) are derivatives of power and irradiance that also account for the divergence or spreading of the light source ...
  23. [23]
    Poynting Vector - RP Photonics
    The Poynting vector can be shown to indicate the direction of energy flow of an electro-magnetic wave, and with its magnitude also the optical intensity.
  24. [24]
    Photon Flux - PVEducation
    The photon flux is defined as the number of photons per second per unit area: T he photon flux is important in determining the number of electrons which are ...
  25. [25]
    Photon Flux Density - an overview | ScienceDirect Topics
    Photon flux density (photon irradiance) is defined as the quotient of the photon flux [mol s −1] incident on an element of a surface by the area [m²] of that ...
  26. [26]
    [PDF] CIE Technical Note 004:2016
    “The candela is the luminous intensity, in a given direction, of a source that emits monochromatic radiation of frequency of 540 × 1012 hertz and that has a ...
  27. [27]
    Mesopic luminous-efficiency functions - Optica Publishing Group
    We expected the luminous-efficiency curve at the −2-log level to approach the rod's efficiency curve or the CIE scotopic efficiency function V′(λ). This is ...
  28. [28]
    candela - BIPM
    It was then ratified in 1948 by the 9th CGPM, which adopted a new international name for this unit, the candela, symbol cd; in 1954 the 10th CGPM ...
  29. [29]
    Light and Dark Adaptation - Webvision - NCBI Bookshelf - NIH
    May 1, 2005 · Dark adaptation refers to how the eye recovers its sensitivity in the dark after exposure to bright lights.
  30. [30]
    Dark Adaptation and Purkinje Shift: A Laboratory Exercise in ... - NIH
    This phenomenon is known as the Purkinje shift and is another principle discussed in many undergraduate textbooks (e.g., Blake and Sekuler, 2005).Missing: explanation | Show results with:explanation
  31. [31]
    Perception Lecture Notes: Color
    Because of this, the brightness of a blue object compared to a red one increases during dark adaptation, called the Purkinje shift. You may have noticed that ...Missing: effect | Show results with:effect
  32. [32]
    The Effect of Pupil Size on Visual Resolution - StatPearls - NCBI - NIH
    Feb 28, 2024 · Theoretically, larger pupils allow more light to reach the retina, potentially enhancing contrast sensitivity, especially in dim light.
  33. [33]
    Perception Lecture Notes: Psychophysics
    Perception Lecture Notes: Psychophysics. Professor David Heeger. What you ... Nowadays, nobody really believes Fechner's interpretation of Weber's law.
  34. [34]
    Psychophysical Methods – Introduction to Sensation and Perception
    Introduction to Sensation and Perception. Basics: neuroscience and psychophysics ... Know Weber's law (also called Weber-Fechner law). The sensitivity of a ...
  35. [35]
    Effect of warm/cool white lights on visual perception and mood in ...
    Cool white light reduced the warmth of color and increased brightness in all three color environments compared to warm light. The preference for cool or warm ...
  36. [36]
    Effect of warm/cool white lights on visual perception and mood in ...
    Aug 31, 2021 · Cool white light reduced the warmth of color and increased brightness in all three color environments compared to warm light. The preference for ...Missing: impact intensity
  37. [37]
    [PDF] APPENDIX I THE SI SYSTEM AND SI UNITS FOR RADIOMETRY ...
    Radiant intensity is another SI derived unit and is measured in W/sr. Intensity is power per unit solid angle. The symbol is I. Intensity is the derivative ...
  38. [38]
    [PDF] Guidelines for Radiometric Calibration of Electro-Optical Instruments ...
    Jan 21, 2011 · Calibration to achieve relative uncertainty of less than 1 % of the measured radiance is essential for the accurate retrieval of atmospheric ...
  39. [39]
    1.5 Calculation of Radiometric Quantities (Examples) - Gigahertz-Optik
    ... radiant intensity emitted in a direction enclosing the angle ϑ with the surface's normal. ... radiance or luminance standards. Here you will find ...
  40. [40]
    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.
  41. [41]
    [PDF] Lighting Design for Open-Plan Offices - à www.publications.gc.ca
    The IESNA recommends 300 lux if computer use is intensive and 500 lux if it is intermittent.<|separator|>
  42. [42]
    Lumileds Drives Efficacy to Over 200 lm/W in Popular 2835 Package
    Oct 19, 2018 · Lumileds today introduced the LUXEON 2835 HE, which achieves unprecedented efficacy of over 200 lm/W, while providing market defining color consistency.
  43. [43]
    [PDF] Lighting world first: Philips breaks 200 lumens per watt barrier
    “The new prototype lamp marks the first time that lighting engineers have been able to reach 200 lm/W efficiency in a real-life lamp without compromising on.
  44. [44]
    UGR (Unified Glare Rating) | ERCO Lighting knowledge
    The calculation rule totals the glare effect of all light sources based on their average luminance, spread and position in relation to the viewing direction. It ...
  45. [45]
    [PDF] Understanding Unified Glare Rating (UGR)
    The Unified Glare Rating (UGR) predicts the glare caused by an electric lighting system along a psychometric scale of discomfort. In simpler words, UGR ...
  46. [46]
    Lighting Systems | Better Buildings Initiative - Department of Energy
    Commercial lighting uses about 4.0 quads of primary energy annually – more than 20% of total commercial building energy use in the United States.
  47. [47]
    GSA provides guidance on procuring and using energy-efficient ...
    Feb 1, 2024 · Conversions to LED lighting typically save 50% of electricity over a fluorescent baseline, and lighting controls can save an additional 80% of ...
  48. [48]
    One ion to catch them all: Targeted high-precision Boltzmann ...
    Nov 22, 2021 · The intensity ratio of the emission lines originating from two thermally coupled excited levels within the same configuration follows simple ...
  49. [49]
    What is distance modulus? - Las Cumbres Observatory
    Apparent magnitude, absolute magnitude and distance are related by an equation:m - M = 5 log d - 5m is the apparent magnitude of the objectM is the absolute ...
  50. [50]
  51. [51]
    None
    ### Summary of Recent Advances in Intensity Mapping for Cosmology (Post-2020)