Fact-checked by Grok 2 weeks ago

Luminosity

In astronomy, luminosity is the total amount of electromagnetic emitted by a celestial object, such as a , , or other astronomical body, per unit time across all wavelengths. It represents an intrinsic property of the object, independent of the observer's distance, and is a fundamental characteristic that helps determine a 's evolutionary , lifespan, and energy output when combined with its . Luminosity is typically measured in watts (joules per second) or, more commonly in stellar contexts, in units of (L), where the Sun's luminosity is 3.828 × 1026 W. Unlike apparent brightness, which is the flux of energy received per unit area on and diminishes with distance according to the , luminosity remains constant regardless of how far the object is from the observer. The relationship between the two is given by the for flux F = L / (4πd2), where d is the to the object; for example, at twice the distance, the apparent brightness decreases to one-fourth. This distinction is crucial for astronomers, as measuring both luminosity and apparent brightness allows for accurate distance calculations via methods like the in the magnitude system. Stellar luminosities vary enormously: the Sun serves as a baseline at 1 L, while brighter stars like Sirius emit about 25 L, and supergiants can exceed thousands of solar luminosities, reflecting differences in surface temperature and radius via the Stefan-Boltzmann law (L ∝ 4πR2σT*4). For galaxies and other extended sources, luminosity quantifies overall energy output, aiding in studies of cosmic and . These measurements, often derived from spectroscopic data and models, underpin key astronomical tools like the Hertzsprung-Russell diagram, which plots luminosity against temperature to classify stars.

Core Concepts

Definition

Luminosity is defined as the total amount of radiated by an object per unit time, typically in the form of , though the concept extends to other types such as emission in certain physical contexts. In physics and astronomy, it represents an intrinsic property of the emitting source, independent of the observer's or perspective. The term "luminosity" originates from the Latin word , meaning "," and entered scientific discourse in the within the field of to describe the intrinsic brightness of light sources. It was later adopted in to quantify the output of bodies. For isotropic , luminosity L is simply the total radiated power, expressed in watts (W). A standard reference value is the , L_\odot = 3.828 \times 10^{26} W, as defined by the in 2015. Bolometric luminosity specifically refers to the total energy integrated over all wavelengths of the . Unlike observer-dependent quantities such as or , luminosity captures the source's inherent energy emission rate. Luminosity represents the total radiant power emitted by a source per unit time, serving as an intrinsic property that does not depend on the or of any observer. In astronomy, this makes it a fundamental measure for characterizing the output of and other objects, independent of how far away they are located. Flux, denoted as F, is the amount of received per unit area per unit time at a given point, making it inherently observer-dependent and scaling inversely with the square of the from the source. Unlike luminosity, flux diminishes as the observer moves farther away, reflecting the spreading of over a larger spherical surface. Brightness, often used interchangeably with apparent brightness in observational contexts, relates closely to and describes the perceived of from a source as seen by an observer; more technically, radiance (a form of brightness) quantifies the directional power per unit area per unit , which can vary with the angle of observation but remains tied to the observer's perspective for point-like sources. In contrast to luminosity's holistic emission, radiance emphasizes the angular distribution of . Illuminance measures the luminous flux incident on a surface per unit area, a quantity specific to photometry and lighting engineering where it accounts for how falls on a detector or , rather than the source's total output. The following table summarizes these distinctions conceptually:
TermDefinitionDependenceKey Characteristic
LuminosityTotal emitted by the source per unit timeIntrinsic (source-based)Fixed regardless of observer position
Flux received per unit area per unit timeExtrinsic (-based)Decreases with inverse square of
Brightness (Radiance)Directional intensity per unit area per unit Extrinsic (observer angle and position)Varies with viewing direction and
Illuminance incident on a surface per unit areaExtrinsic (surface and incident flux)Application-specific to illuminated areas
For example, a star's luminosity is a constant value determined by its physical properties, but its apparent brightness—equivalent to the observed from —dims according to the as distance increases, appearing four times fainter if twice as far away.

Measurement and Units

Units of Luminosity

The unit for luminosity is the watt (W), equivalent to joules per second (J/s), representing the rate of emission. In astronomical contexts, luminosities are often expressed in multiples of this unit, such as 10^{26} W, to accommodate the vast scales involved in stellar and galactic outputs. A widely used astronomical unit is the , denoted L_\odot, which normalizes comparisons for stars and other objects; for example, many stars have luminosities expressed as multiples or fractions of L_\odot. The (IAU) defines the nominal value as L_\odot = 3.828 \times 10^{26} W, based on precise measurements of the and the . This value evolved from 19th-century estimates derived from ground-based observations of the (the flux at Earth's orbit), such as Claude Pouillet's 1838 measurement of approximately 1220 W/m², to modern space-based determinations yielding the IAU 2015 nominal value of 1361 W/m² (as of 2025), enabling luminosity calculations via L_\odot = 4\pi (1 \mathrm{AU})^2 \times S, where S is the . Refinements continued through 20th-century satellite missions like SORCE, achieving current precision within 0.1%. In the centimeter-gram-second (CGS) system, common in older astronomical literature, luminosity is measured in erg/s, where 1 = 10^7 erg/s; the is thus approximately 3.828 \times 10^{33} erg/s. For frequency-specific contexts, such as spectral luminosity L_\nu (power per unit ), units are W/Hz, allowing analysis of emission across bandwidths without specifying total integration. Luminosity can also include non-electromagnetic contributions, such as emission from in stellar cores. For , standard solar models indicate that neutrino luminosity accounts for about 2% of the total energy output, with the remainder primarily in .

Techniques for Measurement

Direct methods for measuring luminosity rely on quantifying the total energy from an astronomical object across all wavelengths and scaling it by to obtain the intrinsic power output, given by L = 4\pi d^2 F_{\rm bol}, where d is the and F_{\rm bol} is the bolometric . Bolometers serve as key instruments for this purpose, detecting through the temperature-dependent resistance of a material that absorbs photons and converts them to , thereby enabling direct measurement of total incoming power without wavelength selectivity. These devices are particularly effective for broadband determination, though practical implementation requires precise and often cryogenic cooling to minimize thermal . Accurate estimates, typically derived from measurements or standard candles, are essential to transform observed into luminosity. Indirect methods predominate for most observations due to the limitations of direct flux measurements for faint or extended sources, primarily involving the integration of the (SED)—the flux density as a function of or —to compute bolometric luminosity. The SED is constructed from multi-wavelength photometry or , and the total luminosity is obtained by integrating the flux over all wavelengths and multiplying by $4\pi d^2, often using numerical techniques to handle gaps in coverage. This approach allows estimation of the full energy budget even when complete data are unavailable, by fitting models to observed segments of the SED. A related technique employs bolometric corrections to adjust magnitudes from a single bandpass to the total output; the correction {\rm BC} = M_{\rm bol} - M_V quantifies the difference between bolometric and visual absolute magnitudes, enabling from visible-band data via empirical tables or theoretical models calibrated against and spectral type. For example, polynomials fitted to effective temperature provide BC values, such as those ranging from -0.07 mag for hot stars to -2.5 mag for cooler ones. Several challenges complicate bolometric luminosity measurements, including atmospheric absorption by Earth's ozone and water vapor, which obscures ultraviolet and infrared wavelengths in ground-based observations, necessitating space-based platforms for full coverage. Incomplete wavelength sampling can lead to systematic under- or overestimation of the total flux, particularly if emission peaks in unobserved bands, while for cosmological distances, redshift stretches the SED and shifts flux, requiring corrections that amplify uncertainties. Additionally, deriving reliable bolometric corrections demands high-quality stellar atmosphere models and extensive multi-band data, as inconsistencies in zero-points or extinction handling can introduce errors up to 0.14 magnitudes. These issues often require substantial telescope time for comprehensive photometry across the electromagnetic spectrum. Advancements in modern instrumentation have mitigated some challenges, with space telescopes like the (JWST) providing unprecedented sensitivity in the from 0.6 to 28.3 μm, allowing better integration of dust-reprocessed emission into SEDs for more accurate bolometric estimates. JWST's capabilities extend to mid- luminosity functions, enabling precise constraints on obscured components that ground-based observatories cannot access, thus reducing reliance on extrapolated corrections. Results from such measurements are frequently normalized to units (L_\odot) for contextual comparison.

Applications in Astronomy

Stellar Luminosity

Stellar luminosity depends primarily on a star's R, effective surface T_\mathrm{eff}, and to a lesser extent on its , which influences transport and opacity. For stars approximated as blackbodies, the total luminosity L is given by the Stefan-Boltzmann law: L = 4\pi R^2 \sigma T_\mathrm{eff}^4, where \sigma = 5.670374419 \times 10^{-8} W m^{-2} K^{-4} is the Stefan-Boltzmann constant. This relation highlights how luminosity scales with the of and the square of , making hotter or larger far more luminous. Variations in composition, such as , subtly affect T_\mathrm{eff} and R through impacts on and in stellar interiors. For main-sequence stars, where hydrogen dominates the , luminosity correlates strongly with mass M via the mass-luminosity relation, approximated as L / L_\odot \approx (M / M_\odot)^{3.5} for stars near . This empirical power-law exponent of 3.5 arises from theoretical models balancing gravitational contraction against generation and , though it steepens to around 5 for more massive stars and flattens below 0.5 M_\odot. The relation varies across evolutionary stages, with post-main-sequence stars deviating due to changes in processes. Representative examples illustrate this range: , a , defines the unit at L = 1\, L_\odot = 3.828 \times 10^{26} W. In contrast, the A2 exhibits a luminosity of approximately 200,000 L_\odot, driven by its enormous radius exceeding 200 R_\odot despite a cooler T_\mathrm{eff} around 8,500 K. At the extreme, the Wolf-Rayet star in the has a luminosity exceeding 6,100,000 L_\odot (log L / L_\odot = 6.79), reflecting its initial mass over 200 M_\odot and intense core fusion of heavier elements. During stellar evolution, undergoes dramatic changes as progress through life cycles. Low- to intermediate-mass like experience a sharp increase upon leaving the , expanding into red giants where shell hydrogen boosts luminosity by factors of up to 1,000 or more compared to their main-sequence values, reaching peaks around 2,000–3,000 L_\odot at the red giant branch tip. These shifts result from core contraction and envelope expansion, altering R and T_\mathrm{eff}. Massive similarly brighten as supergiants before shedding . The Hertzsprung-Russell (HR) diagram plots stellar luminosity against T_\mathrm{eff} (or spectral type), enabling classification into luminosity classes (I for supergiants, II for bright giants, III for giants, IV for subgiants, and V for main-sequence dwarfs) based on vertical position. This framework reveals evolutionary tracks, with main-sequence stars forming a diagonal band where luminosity rises with temperature, while giants and supergiants occupy luminous branches above, distinguishing evolutionary stage and physical state./Cosmology/Astrophysics_(Richmond)/25%3A_Luminosity_Class_and_the_HR_Diagram)

Luminosity of Non-Stellar Objects

The luminosity of non-stellar objects in astronomy encompasses the total electromagnetic output from extended structures such as galaxies and active galactic nuclei (AGN), which often integrates contributions from multiple components including , gas, , and non-thermal processes, unlike the point-source dominated by . For galaxies, this luminosity represents the aggregate light from billions of , interstellar gas, and , typically measured in bolometric or band-specific units relative to the Sun's luminosity (L⊙). The , as a representative , has an estimated total optical luminosity of approximately 2–4 × 10^{10} L⊙, primarily from its stellar disk, with additional contributions from ionized gas and re-emission in the . Active galactic nuclei and quasars exhibit extreme luminosities powered by accretion onto supermassive black holes, where the emission arises from a compact and surrounding relativistic jets, far exceeding the integrated stellar output of their host galaxies. Quasars, the most luminous subset of AGN, typically range from 10^{12} to 10^{14} L⊙ in bolometric luminosity, making them observable across cosmic distances and up to 10,000 times brighter than the entire . This non-stellar dominance highlights the role of release in driving such outputs, contrasting with the thermonuclear in . In , luminosity is often expressed as spectral luminosity L_ν in units of W Hz^{-1}, accounting for the frequency-dependent nature of emission from relativistic electrons in jets or diffuse plasmas. The conversion from observed flux density S_{obs} to L_ν incorporates cosmological effects and the source's α (where flux density S_ν ∝ ν^{-α}) via the formula: L_\nu = \frac{4\pi D_L^2 S_{obs}}{(1+z)^{1+\alpha}} Here, D_L is the luminosity distance, and z is the . For example, an observed flux density of 1 Jy at z=1 (assuming α ≈ 0.7 and a standard ΛCDM cosmology) yields L_ν ≈ 6 × 10^{26} W Hz^{-1} at the rest-frame . Other wavelengths reveal additional non-stellar contributions: X-ray luminosity from galaxy clusters, primarily from hot intracluster medium heated by gravitational collapse and mergers, typically spans 10^{43} to 10^{45} erg s^{-1} (0.1–2.4 keV band), with luminous examples like those in the REXCESS survey exceeding 10^{44} erg s^{-1}. In infrared bands, star-forming regions within galaxies emit via dust heated by young stars, with total infrared luminosities L_{IR} often reaching 10^{10} L⊙ or more for luminous systems, tracing obscured star formation that optical measurements miss. Measuring aggregated luminosity poses challenges due to the extended nature of these sources, where resolving individual components (e.g., stars versus diffuse gas) is difficult, and intrinsic output must be corrected for dust extinction that absorbs shorter wavelengths and re-emits in the . Dust obscuration can reduce observed optical fluxes by factors of 2–10 in star-forming galaxies, necessitating multi-wavelength integration to estimate total bolometric values accurately.

Mathematical Formulations

Fundamental Equations

The luminosity L of a radiating source is defined as the total power radiated per unit time across all wavelengths and directions, representing the intrinsic output of the object. In general, for sources with arbitrary distributions, the luminosity is given by the integral of the specific I_\nu over \nu, surface area A, and \Omega: L = \iint_A \int_\Omega \int_0^\infty I_\nu(\Omega) \cos\theta \, d\nu \, d\Omega \, dA, where \theta is the angle between the normal to the surface and the direction of , and the integration accounts for the in non-isotropic emission. This formulation arises from principles, where specific intensity I_\nu describes the per unit area, frequency, and , ensuring in the emitted field. For an isotropic or a distant observer approximating , the observed F at distance d follows the , derived from over the surface of a : F = \frac{L}{4\pi d^2}, which links the intrinsic luminosity to the measurable flux by distributing the total power uniformly over the spherical area $4\pi d^2. This relation assumes no absorption or scattering along the line of sight and isotropic emission. In the blackbody approximation, common for thermal emitters like stellar surfaces, the luminosity simplifies using the Stefan-Boltzmann law. The surface flux from a blackbody at temperature T is \sigma T^4, where \sigma = 5.670 \times 10^{-8} \, \mathrm{W \, m^{-2} \, K^{-4}} is the Stefan-Boltzmann constant. Integrating this flux over the emitting surface area A yields the total luminosity L = A \sigma T^4. For a spherical source of radius R, A = 4\pi R^2, so L = 4\pi R^2 \sigma T^4. This expression derives from Stefan's empirical observation in 1879, relating radiated power to temperature via experimental data on heated bodies, and Boltzmann's theoretical derivation in 1884 using electromagnetic theory and thermodynamic equilibrium in a radiation cavity. The derivation integrates the blackbody energy density and radiation pressure over the surface, assuming perfect absorption and emission (emissivity \epsilon = 1) and Lambertian diffusion. In contexts, luminosity represents the net radiative output after accounting for internal absorption and re-emission processes, ensuring overall within the source; for instance, in optically thick media, the emergent luminosity equals the generation rate balanced by these interactions. These equations rely on key assumptions, including isotropic emission for the and blackbody behavior for the Stefan-Boltzmann relation, which may deviate in non-thermal or anisotropic sources like jets or dusty envelopes, requiring the general form for accuracy.

Relations to Magnitude and Distance

In astronomy, the absolute bolometric magnitude M_{\text{bol}} provides a standardized for an object's total luminosity across all wavelengths, defined by the relation M_{\text{bol}} = -2.5 \log_{10} \left( \frac{L}{L_\odot} \right) + 4.74, where L is the luminosity, L_\odot is the (\approx 3.828 \times 10^{26} W), and the constant sets the zero point such that the Sun has M_{\text{bol},\odot} = 4.74. This formulation allows direct comparison of intrinsic brightness independent of or observational bandpasses. The difference in absolute bolometric magnitudes between two objects reflects the ratio of their luminosities through the magnitude scale's logarithmic definition: M_{\text{bol},1} - M_{\text{bol},2} = -2.5 \log_{10} \left( \frac{L_1}{L_2} \right). This relation stems from the historical Pogson scale, where a magnitude difference of 5 corresponds to a luminosity of 100, enabling astronomers to quantify relative brightnesses efficiently. For example, a with M_{\text{bol}} = -1 is approximately times more luminous than . To connect observed brightness to intrinsic luminosity, the relates the m (brightness as seen from ) to the M (brightness at a standard distance of 10 parsecs) via m - M = 5 \log_{10} \left( \frac{d}{10} \right), where d is the distance in parsecs; this equation derives from the -luminosity , F = L / (4\pi d^2), with magnitudes logarithmic in flux. Rearranging allows luminosity derivation as L = 4\pi d^2 F, where flux F is inferred from m. In cosmological contexts, the is replaced by the d_L, defined to preserve the relation F = L / (4\pi d_L^2) amid , given by d_L = (1 + [z](/page/Z)) D_M, where z is the and D_M is the transverse comoving (D_M = \int_0^z c \, dz' / H(z') in a flat , with H(z) the Hubble parameter). This adjustment accounts for dimming, making high-[z](/page/Z) observations appear fainter than in static space, and is essential for probing distant structure. These relations enable practical luminosity estimates for standard candles like Type Ia supernovae, which have a consistent peak (M_B \approx -19.3) due to uniform progenitor properties. By measuring their and redshift-derived distance, astronomers compute d_L via the , yielding luminosities that calibrate the and reveal acceleration from .

Broader Contexts

Luminosity in Physics and Engineering

In and , luminosity is quantified as (Φ_v), which measures the total visible light output from a source, weighted by the eye's as described by the photopic luminosity function V(λ). This function peaks at approximately 555 nm, reflecting the eye's maximum response to green light under normal viewing conditions. The unit of is the (lm), defined such that 1 lm corresponds to the produced by 1/683 watts of radiant power at a of 540 THz (555 nm). This conversion factor ensures that photometric measurements account for perceptual efficiency rather than raw energy. In applications, helps evaluate beam visibility and safety, particularly for visible-wavelength lasers used in tasks like alignment and cutting. In engineering contexts, luminosity informs the design of light sources through , defined as the ratio of to electrical power input, expressed in lumens per watt (lm/W). This metric assesses efficiency in converting to visible , crucial for applications in systems. Incandescent bulbs historically achieved around 15 lm/W, while modern light-emitting diodes (LEDs) exceed 200 lm/W for high-quality white-light modules, enabling energy savings in and settings. For instance, non-directional LED lamps commercially available by the mid-2020s reach up to 200 lm/W, significantly outperforming traditional fluorescents at about 50–100 lm/W. Engineers use this to optimize fixtures for tasks like illumination in factories or displays, balancing output with thermal management. Thermal radiation provides another framework for luminosity in physics and , where the total power radiated by a hot body is governed by the Stefan-Boltzmann . For a blackbody, the radiant (power per unit area) is given by M = \sigma T^4, where \sigma = 5.670 \times 10^{-8} W/m²K⁴ is the Stefan-Boltzmann constant and T is the absolute temperature in . The total luminosity L then becomes L = 4\pi R^2 \sigma T^4 for a spherical emitter of radius R, applicable to any opaque hot surface assuming isotropic emission. In , this models heat loss from industrial furnaces or high-temperature components, where emissivity \epsilon (0 ≤ \epsilon ≤ 1) modifies the equation to M = \epsilon \sigma T^4 for real materials. Accurate predictions aid in designing efficient heaters or predicting material degradation under thermal stress. In , luminosity extends to , quantifying the energy flux of these weakly interacting particles from sources like nuclear reactors. Reactors act as prolific factories, producing electron antineutrinos via in processes, with typical luminosities on the order of 10²⁰ s per second per gigawatt of thermal power. This luminosity enables experiments to probe fundamental interactions, such as parameters, using detectors positioned near reactor cores. In contexts, luminosity relates to rates in semiconductors, where radiative recombination in p-n junctions generates in devices like LEDs. The rate R for s is proportional to the density and lifetime, R = Bn^2 for direct-bandgap materials (with B as the radiative and n the minority concentration), determining the device's luminous output. This process underpins , linking to practical sources.

Emerging Applications and Developments

Recent observations from the (JWST), as of 2025, have revealed obscured quasars in high-redshift galaxies (z > 6) with high intrinsic luminosities, bridging the gap between classical quasars and less luminous active galactic nuclei in the epoch of reionization. These findings, based on fitting and broad-line region analysis, indicate a higher density of such luminous objects than previously predicted by UV luminosity functions, potentially reshaping models of early black hole growth. In applications, luminosity concepts underpin metrics through integration of the AM1.5 global solar spectrum, which standardizes photovoltaic output calculations at 1000 W/m² to simulate Earth's average distribution. Recent advancements, as documented in efficiency tables as of 2025, show record multi-junction cell efficiencies exceeding 47% under these conditions, optimizing matching to enhance conversion from . Quantum technologies leverage luminosity control in single-photon sources, such as nitrogen-vacancy () centers in , where engineered emission rates enable deterministic generation for protocols like entanglement distribution. Breakthroughs in 2025 have achieved up to 80% room-temperature collection from these centers, improving for quantum networks by minimizing loss in luminous output from defect-based emitters. In climate science, Earth's serves as a measure of planetary luminosity, with an of approximately 240 W/m² representing the total thermal emission to space that balances incoming . This value, derived from observations, highlights the role of greenhouse gases in modulating luminosity escape, with recent data showing slight reductions due to atmospheric changes. As of 2025, detection has advanced through refined photometry combining TESS and JWST data, exploiting luminosity contrasts from planetary s to characterize atmospheres with precisions down to parts per million in depth. These techniques, applied to over 100 TESS objects of interest, enable detailed of habitable-zone worlds by isolating subtle luminosity variations against host star backgrounds.

References

  1. [1]
    17.1 The Brightness of Stars - Astronomy 2e | OpenStax
    Mar 9, 2022 · Perhaps the most important characteristic of a star is its luminosity—the total amount of energy at all wavelengths that it emits per second ...<|control11|><|separator|>
  2. [2]
    Luminosity and Apparent Brightness | ASTRO 801
    The luminosity of a star, on the other hand, is the amount of light it emits from its surface. The difference between luminosity and apparent brightness depends ...
  3. [3]
    Glossary term: Luminosity - IAU Office of Astronomy for Education
    Luminosity is the energy output per second from some emitting body such as a star. The most common use is for electromagnetic radiation.
  4. [4]
    Luminosity and Distance - University of Oregon
    The luminosity is how much energy is coming from the per second. The units are watts (W). Astronomers often use another measure, absolute magnitude .Missing: physics | Show results with:physics
  5. [5]
    Luminosity - Etymology, Origin & Meaning
    Originating in the 1630s from French luminosité or luminous + -ity, luminosity means brightness quality, intensity of light, or intrinsic brightness in ...Missing: physics optics
  6. [6]
    IAU 2015 Resolution B3 on Recommended Nominal Conversion ...
    Oct 26, 2015 · IAU 2015 Resolution B3 adopts five solar conversion constants (nominal solar radius, nominal total solar irradiance, nominal solar luminosity, ...
  7. [7]
    [PDF] Flux and luminosity • Brightness of stars • Spectrum of light
    Luminosity - A star produces light – the total amount of energy that a star puts out as light each second is called its Luminosity.
  8. [8]
    2 Radiation Fundamentals‣ Essential Radio Astronomy
    The specific intensity or brightness is an intrinsic property of a source, while the flux density of a source also depends on the distance between the source ...
  9. [9]
    Radiometry and photometry FAQ | - The University of Arizona
    Illuminance is another SI derived quantity which denotes luminous flux density . It has a special name, lux, and is lumens per square metre, or lm/m2. The ...Missing: luminosity | Show results with:luminosity<|control11|><|separator|>
  10. [10]
    lecture 2: distances and luminosities of stars
    Jan 11, 1997 · The apparent brightness (aka apparent flux) of a star depends on how far away it is. A star that is twice as far away appears four times fainter.
  11. [11]
    NOMINAL VALUES FOR SELECTED SOLAR AND PLANETARY ...
    The resolution recommends the use of nominal solar and planetary values, which are by definition exact and are expressed in SI units.ABSTRACT · INTRODUCTION · IAU 2015 RESOLUTION B3 · DISCUSSION AND...
  12. [12]
    Solar Physics Historical Timeline (1800-1999)
    1838: The solar constant​​ Pouillet's pyrheliometer. The solar constant is a measure of the sun's luminosity, and is defined as the amount of solar energy per ...
  13. [13]
    [PDF] Variations ofthe SolarConstant
    The question of the constancy of the so-called solar constant, or solar irradiance, has been addressed for nearly a century. However, for a variety of ...
  14. [14]
    Unit Conversions
    1 Joule = 2.39x10-1 calorie · 1 Joule = 107 ergs · 1 eV = 1.602177x10-12 erg = 1.602177x10-19 Joule · 1 Solar Luminosity = 3.826x1033 ergs/s = 3.826x1026 Joules/s ...
  15. [15]
    [nucl-ex/0602005] Solar Neutrinos - arXiv
    Feb 5, 2006 · A model-independent analysis of existing low-energy data shows good agreement between the neutrino and electromagnetic luminosities at the +/- ...
  16. [16]
    [PDF] Observational Astrophysics 1. Astronomical Measurements
    Luminosity is the rate at which energy is emitted or radiated from a source in all frequencies/wave- lengths (therefore, luminosity is, in other words, power).
  17. [17]
    The spectral energy distributions and the bolometric luminosities of ...
    Dec 21, 2023 · The comprehensive SEDs allowed us to measure accurate bolometric luminosities and derive robust bolometric corrections for the different tracers ...
  18. [18]
    ON THE USE OF EMPIRICAL BOLOMETRIC CORRECTIONS FOR ...
    1. INTRODUCTION. Bolometric corrections are widely used in astronomy to infer either luminosities or absolute magnitudes of stars. Empirical corrections in the ...ABSTRACT · COEFFICIENTS FOR THE... · THE USE OF EMPIRICAL...
  19. [19]
    [PDF] Fundamentals of Stars II: Revisiting Bolometric Corrections - arXiv
    Apr 8, 2025 · It is now practical to measure the heat received from a star by the use of a radiometer. Considerable sensitiveness in the method has been ...
  20. [20]
    Exploring the faintest end of mid-infrared luminosity functions up to z ...
    In this work, we report the first rest-frame MIR LFs at 7.7, 10, 12.8, 15, 18, and 21 μm as well as the total IR LF from the JWST Cosmic Evolution Early Release ...3.1 Sed Fitting · 3.3 Luminosity Function · 4.3 Luminosity Density...
  21. [21]
    [PDF] arXiv:1807.00378v2 [astro-ph.EP] 11 Jul 2018
    Jul 11, 2018 · The luminosity of the star is determined via the. Stefan-Boltzmann law applied to stars. L⋆ = 4πR2. ⋆σT4 eff. (3) where R⋆ is the stellar ...
  22. [22]
    Stefan-Boltzmann constant - CODATA Value
    Concise form, 5.670 374 419... x 10-8 W m-2 K ; Click here for correlation coefficient of this constant with other constants ; Source: 2022 CODATA
  23. [23]
    The Mass-Luminosity Relationship | ASTRO 801
    Stars on the Main Sequence must be using the energy generated via nuclear fusion in their cores to create hydrostatic equilibrium. The condition of hydrostatic ...Missing: seminal paper
  24. [24]
  25. [25]
    Deneb: August 8, 2025 - StarDate Online
    Aug 8, 2025 · So Deneb's true luminosity – the value when you add up all wavelengths of light – is somewhere between 50,000 and 200,000 times the Sun's. If ...
  26. [26]
    G. C. Sloan: Main Sequence Stars, Giants, and Supergiants
    Jan 11, 2021 · As it does this, the temperature of the core increases and the luminosity increases, in some cases by a factor of 1000 or more. This extra ...
  27. [27]
    On the mass-to-light ratio of the local Galactic disc and the optical ...
    ... luminosity should be larger in proportion than the I-band one. The total I-band luminosity of the Milky Way is thus ∼3.8 ± 0.6 × 1010 L⊙ (where the error ...
  28. [28]
    Lecture 32: Quasars
    Feb 28, 2025 · Quasars have luminosities which lie in the range L = 10 to 100,000 LMW, where LMW is the luminosity of the Milky Way galaxy, which in turn ...
  29. [29]
    What is radio luminosity?
    For example, consider a 1 Jy signal from a radio source at a redshift of 1, at a frequency of 1.4 GHz. Ned Wright's cosmology calculator calculates a ...
  30. [30]
    An HST Comparison of High and Low X-ray Luminosity Clusters
    Oct 15, 2001 · Eight of these clusters have low X-ray luminosities (<10^{44} erg/s), while the remaining nine clusters have L_x>10^{45} ergs/s. The clusters ...
  31. [31]
  32. [32]
    II. Dust extinction in galaxies: Luminosity functions and infrared excess
    We present the complete prescription of dust attenuation implemented in the new semi-analytical model called GAS.
  33. [33]
    Dust in Galaxies - Department of Physics and Astronomy
    Dust is a useful first proxy for phases of the interstellar medium that are harder to trace, since cool gas and dust are coupled gravitationally.
  34. [34]
    Most Luminous Object in the Universe Discovered
    Jun 27, 1991 · Astronomers define the luminosity of an object as the total amount of energy emitted at all wavelengths.
  35. [35]
    [PDF] Stefan 1879
    Über die Beziehung zwischen der Wärmestrahlung und der Temperatur. Von dem w. M. J. Stefan. 1. Über die Versuche von Dulong und Petit. Dulong ...
  36. [36]
    Ableitung des Stefan'schen Gesetzes, betreffend die Abhängigkeit ...
    Ableitung des Stefan'schen Gesetzes, betreffend die Abhängigkeit der Wärmestrahlung von der Temperatur aus der electromagnetischen Lichttheorie.
  37. [37]
    [PDF] General Introduction 1. Luminosity, Flux and Magnitude
    The astronomical unit is based on the mean distance of the Earth from the Sun, a light year is the distance light travels in one year, and the parsec is the ...
  38. [38]
    [PDF] Astrophysics Introductory Course
    Bolometric magnitudes mbol are a measure for the total luminosity integrated over all wavelengths. One defines: mbol = mV + B.C. where B.C. is called bolometric ...
  39. [39]
    [PDF] Brightness, Magnitudes, and Luminosity: A Tutorial (Prof. Harriet ...
    Formulas: The magnitude scale expresses a given ratio of brightness (say, between two stars) as a difference in magnitudes. This involves taking logarithms ...
  40. [40]
    Absolute Magnitude and Luminosity - Properties of Stars
    Jun 18, 2022 · The ratio of the two stars' luminosities is (Lum.*1)/(Lum.*2) = 10-0.4(abs mag*1 - abs mag*2) or in an approximate relation: Lum.*1/Lum.
  41. [41]
    What is distance modulus? - Las Cumbres Observatory
    The expression m - M is called the distance modulus and is a measure of distance to the object. An object with a distance modulus of 0 is exactly 10 parsecs ...
  42. [42]
  43. [43]
    [PDF] Distances in Cosmology - MPA Garching
    Luminosity Distance (standard candles). The luminosity distance dL is defined to satisfy the relation: where Fobs is the observed flux from an astronomical ...
  44. [44]
    Cepheid Variable Stars, Supernovae and Distance Measurement
    Once the period of a distant Cepheid has been measured, its luminosity can be determined from the known behavior of Cepheid variables. Then its absolute ...
  45. [45]
    Simulated host galaxy analogues of high-z quasars observed with ...
    For the EIGER quasars, the Eddington luminosities are 2.54 × 10 14 L ⊙ ⁠, 4.06 × 10 13 L ⊙ ⁠, and 3.88 × 10 13 for J0148+0600, J159-02, and J1120+0641, ...
  46. [46]
    JWST Unveils Obscured Quasars in the Most Luminous Galaxies at ...
    Jul 14, 2025 · Here we report the discovery of obscured quasars hosted by the most luminous galaxies at z > 6, possibly bridging the gap between the two AGN populations.
  47. [47]
    Solar Cell Efficiency Tables (Version 66) - Green - Wiley Online Library
    May 26, 2025 · Consolidated tables showing an extensive listing of the highest independently confirmed efficiencies for solar cells and modules are presented.
  48. [48]
    Recent progress in hybrid diamond photonics for quantum ... - Nature
    May 8, 2025 · For instance, NV center was created at the center of a photonic crystal cavity by combining nitrogen ion implantation and silicon masks, ...Missing: luminosity | Show results with:luminosity
  49. [49]
    Scientists unlock the quantum magic hidden in diamonds
    Oct 8, 2025 · Researchers have found a way to extract almost every photon from diamond color centers, a key obstacle in quantum technology.Missing: luminosity | Show results with:luminosity
  50. [50]
    [PDF] Balancing the Energy Budget - nasa ceres
    Earth's temperature has averaged 15 C (~59 F), with a balance of 240W/m2 of energy being absorbed from the Sun; and. 240 W/m2 being emitted to space as heat ...Missing: luminosity m²
  51. [51]
    Identification of the Top TESS Objects of Interest for Atmospheric ...
    Apr 23, 2024 · We present a sample of TESS planets and planet candidates that we identify as “best-in-class” for transmission and emission spectroscopy with ...