Fact-checked by Grok 2 weeks ago

Effective temperature

The effective temperature of a celestial body, such as a star or planet, is the temperature of a black body of the same size that would radiate the same total amount of electromagnetic power as the body itself. This parameter provides an average measure of the body's surface temperature, approximating the conditions in its photosphere where most radiation originates, even though real bodies are not perfect black bodies with uniform emissivity across wavelengths. In stellar astrophysics, it is a fundamental property used to classify stars by spectral type and color, with hotter stars exhibiting higher effective temperatures and bluer light. The effective temperature T_{\text{eff}} is calculated from the body's luminosity L and radius R using the Stefan-Boltzmann law: L = 4\pi R^2 \sigma T_{\text{eff}}^4, where \sigma = 5.67 \times 10^{-8} W m^{-2} K^{-4} is the Stefan-Boltzmann constant. For the Sun, this yields an effective temperature of 5772 K (nominal value as of IAU 2015), serving as a for other . In , effective temperature estimates the temperature a would have without an atmosphere, influencing models of and climate; for , it is about 255 K, significantly cooler than the actual surface average due to the . Beyond astronomy, the term "effective temperature" also appears in as an index for human , originally defined as the temperature of still, saturated air that produces the same sensation of warmth as the actual combining air temperature, , and . This older formulation has evolved into the standard effective temperature (SET*), a more comprehensive metric standardized by organizations like to evaluate indoor climate conditions and design energy-efficient buildings.

Fundamentals

Blackbody Radiation

A blackbody is an idealized physical object that absorbs all incident electromagnetic radiation, regardless of frequency or angle of incidence, and re-emits energy solely in the form of thermal radiation. This perfect absorption implies that a blackbody also serves as a perfect emitter, with the emitted spectrum depending only on its temperature and independent of its composition or structure. In the late , efforts to describe encountered significant challenges, particularly with the Rayleigh-Jeans law, a classical that predicted an infinite at high frequencies, known as the . This discrepancy between theory and experimental observations, such as those from cavity experiments, highlighted the limitations of . In December 1900, Max Planck resolved this issue by introducing a revolutionary formula for the spectral radiance of blackbody radiation, marking the birth of quantum theory. Planck's derivation assumed that the energy of oscillators within the blackbody is quantized in discrete units of hf, where h is Planck's constant and f is the frequency, rather than continuous as in classical models. This led to Planck's law, which describes the spectral radiance B(\nu, T) as: B(\nu, T) = \frac{2 h \nu^3}{c^2} \frac{1}{e^{h\nu / kT} - 1} where \nu is frequency, T is temperature, c is the speed of light, and k is Boltzmann's constant. The law accurately fits experimental data across all frequencies, avoiding the ultraviolet catastrophe by suppressing high-frequency contributions at finite temperatures. From Planck's law, several key relations emerge for blackbody radiation. The Stefan-Boltzmann law states that the total power P radiated by a blackbody is proportional to the fourth power of its absolute temperature: P = \sigma A T^4 where \sigma = 5.670 \times 10^{-8} W m^{-2} K^{-4} is the Stefan-Boltzmann constant and A is the surface area. This law was first observed empirically by Josef Stefan in 1879 and theoretically derived by Ludwig Boltzmann in 1884 using thermodynamic arguments. Wien's displacement law, derived from Planck's formula, specifies that the wavelength \lambda_{\max} at which the spectral radiance peaks is inversely proportional to the temperature: \lambda_{\max} T = b where b \approx 2.898 \times 10^{-3} m K is Wien's displacement constant. This relation shifts the peak emission to shorter wavelengths for hotter blackbodies, providing a practical tool for inferring temperatures from spectra. These principles of blackbody radiation form the foundation for the concept of effective temperature, which applies similar radiative equilibrium ideas to real objects.

Definition and Formula

The effective temperature of an object, such as a star or planet, is defined as the temperature of a blackbody that would emit the same total energy flux as the actual object across its entire spectrum. This concept arises from blackbody radiation principles, where the effective temperature represents an equivalent thermal emission temperature without assuming the object is a perfect blackbody. The formula for effective temperature T_\text{eff} is derived from the Stefan-Boltzmann law applied to the object's total luminosity L and radius R. The luminosity is given by L = 4\pi R^2 \sigma T_\text{eff}^4, where \sigma is the Stefan-Boltzmann constant ($5.67 \times 10^{-8} W m^{-2} K^{-4}). Solving for T_\text{eff} yields: T_\text{eff} = \left( \frac{L}{4\pi R^2 \sigma} \right)^{1/4}. This equation equates the object's observed total radiated power to that of a blackbody sphere of the same size. Unlike the local thermodynamic (physical) temperature, which measures the kinetic energy of particles in a specific region, effective temperature is a spectral average that characterizes the overall energy output, independent of the object's internal temperature gradients or non-blackbody emission properties. For real objects, this leads to differences from actual surface temperatures due to deviations from ideal blackbody behavior. Effective temperature is expressed in Kelvin (K). For example, the Sun's effective temperature is approximately 5772 K. The derivation assumes isotropic emission (uniform radiation in all directions) and spherical symmetry of the object.

Stellar Applications

Calculation for Stars

The effective temperature of a star is calculated using the Stefan-Boltzmann law, which relates the star's total luminosity L to its radius R and effective temperature T_\mathrm{eff} via the formula T_\mathrm{eff} = \left( \frac{L}{4 \pi R^2 \sigma} \right)^{1/4}, where \sigma = 5.670374419 \times 10^{-8} W m^{-2} K^{-4} is the Stefan-Boltzmann constant. This yields the temperature at which a blackbody of radius R would radiate the observed luminosity L. Luminosity L is derived from the bolometric flux measured at Earth, integrated over all wavelengths, multiplied by the square of the distance d to the star, such that L = 4 \pi d^2 F_\mathrm{bol}. The stellar radius R is obtained either from direct measurements of the angular diameter \theta (typically via optical interferometry) combined with distance, giving R = (\theta / 2) d where \theta is in radians, or from theoretical stellar evolution models calibrated to observed properties. For the Sun, using the nominal values L_\odot = 3.828 \times 10^{26} W and R_\odot = 6.957 \times 10^8 m, the effective temperature is calculated as follows: T_\mathrm{eff,\odot} = \left( \frac{3.828 \times 10^{26}}{4 \pi (6.957 \times 10^8)^2 (5.670374419 \times 10^{-8})} \right)^{1/4} \approx 5772~\mathrm{K}. These nominal solar values are defined by the for consistent astrophysical computations. The calculation varies significantly across star types due to differences in radius. Main-sequence stars, fusing hydrogen in compact cores, have relatively small radii (e.g., comparable to or smaller than the Sun's for solar-mass stars), resulting in higher T_\mathrm{eff} and hotter, bluer appearances for a given luminosity. In contrast, red giants and supergiants, which have expanded envelopes from post-main-sequence evolution, possess radii hundreds to thousands of times larger than main-sequence counterparts, leading to much lower T_\mathrm{eff} (often below 4000 K) and cooler, redder colors despite similar or higher luminosities. This approach assumes an idealized blackbody with uniform , but it overlooks atmospheric effects such as , where the stellar disk appears brighter at the center and dimmer at the edges due to temperature gradients in the outer layers; this can introduce uncertainties of a few percent in estimates from diameters, indirectly affecting T_\mathrm{eff}. In the Hertzsprung-Russell diagram, T_\mathrm{eff} forms the horizontal axis (decreasing from left to right), plotted against luminosity on the vertical axis to classify stars by evolutionary stage, with main-sequence stars forming a diagonal band and giants branching upward at cooler temperatures.

Observational Determination

The effective temperature of a star can be determined observationally through several empirical methods that leverage photometric, spectroscopic, and interferometric data, often calibrated against theoretical blackbody approximations or detailed atmospheric models. These techniques provide independent estimates that can be compared to the fundamental relation T_{\text{eff}} = \left( \frac{L}{4\pi R^2 \sigma} \right)^{1/4}, where luminosity L and radius R are derived from other observations. One common photometric approach is the color index method, which correlates broadband color indices like the —measuring the difference in magnitude between blue (B) and visual (V) filters—with effective temperature. For main-sequence stars spanning F0 to K5 spectral types (approximately 4000–7000 ), calibrations relate dereddened (B-V)_0 to T_{\text{eff}} using polynomial fits that account for metallicity [Fe/H] and surface gravity log g, such as T_{\text{eff}} = c_0 + c_1 (B-V)_0 + c_2 \log T_{\text{eff}} + \cdots + h_1 \log g (B-V)_0, where coefficients are derived from infrared flux temperatures and empirical color data. This relies on the fact that cooler stars appear redder due to the peak of their blackbody spectrum shifting to longer wavelengths, with typical residuals of about 62 for well-calibrated samples. It is widely applied to large stellar catalogs for initial temperature estimates, though it assumes minimal interstellar reddening and standard atmospheric opacity. Spectroscopic methods estimate T_{\text{eff}} by analyzing the strengths, widths, and profiles of absorption lines in a star's spectrum, which depend on temperature through excitation and ionization balances. The Saha ionization equation, which relates the ratio of ionized to neutral atom abundances to temperature and electron pressure, is used to infer T_{\text{eff}} from the relative strengths of lines from different ionization states (e.g., neutral vs. singly ionized iron lines in solar-type stars). For instance, in fitting observed spectra to synthetic ones, line profile shapes and equivalent widths are minimized against models assuming local thermodynamic equilibrium, yielding T_{\text{eff}} values like 4300 K for Arcturus (α Boo) with adjustments for excitation equilibrium. These techniques are particularly effective for high-resolution spectra (R > 20,000) and provide constraints on metallicity and gravity simultaneously. Interferometry offers a direct geometric measurement by resolving a star's angular diameter θ using long-baseline optical or near-infrared interferometers, such as those at the CHARA array. Combined with the star's parallax-derived distance d to obtain physical radius R = (θ d)/2 and bolometric flux F_bol (integrated over the spectrum), T_{\text{eff}} is calculated via T_{\text{eff}} = \left( \frac{4 F_{\text{bol}}}{ \theta^2 \sigma} \right)^{1/4}, where σ is the Stefan-Boltzmann constant; limb-darkening corrections are applied to θ for accuracy. This method has achieved precisions of 1% in T_{\text{eff}} (about 50–70 K) for nearby bright stars, such as 9700 ± 400 K for β UMa, and is essential for calibrating indirect methods since it bypasses model assumptions. It is limited to angular diameters larger than ~0.5 mas, favoring giants and supergiants within 100 pc. For higher precision, especially in crowded fields or for fainter , observed spectra are fitted to grids of synthetic spectra generated from stellar model atmospheres, such as those computed with the code under assumptions of , , and statistical equilibrium. These models span parameters like T_{\text{eff}} from 2300–8000 K, log g from 0–6, and [Fe/H] from -4 to +1, allowing least-squares minimization of residuals across wavelength ranges (e.g., 500–55,000 ) to derive T_{\text{eff}} by matching continuum shapes and line profiles; for example, fitting VLT/ spectra of globular cluster yields T_{\text{eff}} with ~100 K uncertainty. Such fittings incorporate non-local thermodynamic equilibrium effects for hot and are validated against , providing robust estimates when combined with photometric data. Uncertainties in observationally determined T_{\text{eff}} vary by method and stellar type but are typically 50–100 K (1–2%) for nearby FGK stars using modern data. Photometric color indices introduce errors of ~100–200 K from calibration scatter and reddening, while spectroscopic fittings suffer ~140 K from line selection and non-LTE effects; interferometric and model-based methods achieve <1% precision for fundamentals but are sample-limited. For benchmark stars like those in the Gaia FGK catalog, direct angular diameter measurements contribute <0.5% uncertainty, though systematic discrepancies between methods (e.g., 100–300 K offsets for metal-poor dwarfs) highlight the need for multi-technique validation.

Planetary Applications

Blackbody Temperature for Planets

The effective temperature of a is the uniform temperature it would achieve in blackbody radiative equilibrium, where the stellar energy it absorbs balances the it emits. This concept models as passive absorbers and re-emitters of stellar radiation, without self-luminosity. The energy balance for a planet states that the average absorbed stellar equals the emitted thermal radiation . The incident stellar S (the adjusted for orbital distance) is intercepted over the planet's cross-sectional area but distributed over its full spherical surface, yielding an average incoming of S/4. Accounting for reflection, the absorbed is S(1 - A)/4, where A is the Bond albedo representing the fraction of total incident energy reflected across all wavelengths and angles. This absorbed equals the blackbody emission \epsilon \sigma T_{\text{eff}}^4, where \epsilon is the emissivity (typically \approx 1 for planets approximating blackbodies) and \sigma = 5.67 \times 10^{-8} W m^{-2} K^{-4} is the Stefan-Boltzmann constant. Solving for the effective temperature gives: T_{\text{eff}} = \left[ \frac{S (1 - A)}{4 \epsilon \sigma} \right]^{1/4} This derivation assumes a blackbody spectrum for emission and rapid redistribution of heat to maintain uniform temperature, either through fast rotation or efficient conduction in the surface or atmosphere; it also excludes internal heat sources like radioactive decay or tidal heating, focusing purely on stellar input. The Bond albedo A is critical for accurate energy balance, as it integrates reflectivity over the full spectrum and phase angles, unlike the geometric albedo, which measures visible-light brightness at full phase relative to a Lambertian disk and overestimates reflection for energy calculations in some cases. Applying this to Solar System examples, Venus—with S \approx 2614 W m^{-2} at 0.72 AU and Bond albedo A \approx 0.77—yields T_{\text{eff}} \approx 232 K. Mars, at 1.52 AU with S \approx 590 W m^{-2} and A \approx 0.25, has T_{\text{eff}} \approx 210 K. These values illustrate the scaling with distance and reflectivity under idealized conditions.

Relation to Surface Temperature

The surface temperature of a planet refers to the local thermodynamic temperature measured at the ground level or the base of the atmosphere, representing the actual physical heat content available for processes like , , and . In contrast, the effective temperature serves as a baseline blackbody approximation of the planet's overall radiative balance, but real surface temperatures often deviate significantly due to various dynamical and compositional influences. Several key factors contribute to the discrepancies between effective and surface temperatures on planets. Atmospheric composition, particularly the presence of greenhouse gases such as carbon dioxide and water vapor, alters radiative transfer by absorbing and re-emitting infrared radiation, thereby modifying the energy distribution. Heat transport mechanisms, including atmospheric circulation and oceanic currents, redistribute absorbed solar energy from illuminated regions to shadowed areas, smoothing out extremes. Day-night cycles induce thermal contrasts, with rapid heating on the dayside and cooling on the nightside, especially on worlds with thin or no atmospheres like Mercury. Obliquity, or axial tilt, influences seasonal variations in insolation, leading to latitudinal temperature gradients that can amplify or mitigate global averages. A prominent mechanism driving surface temperatures above the effective temperature is the greenhouse effect, where atmospheric gases trap outgoing longwave radiation, warming the lower layers through downward infrared emission. This process elevates planetary surface temperatures beyond the effective value; for instance, on Earth, it raises the mean surface temperature by approximately 33 K compared to the no-atmosphere baseline. Similar effects occur on Venus, where dense CO₂ creates extreme warming, and on Titan, where methane contributes to a modest enhancement. Equilibrium temperature calculations can vary depending on whether they represent the —the location of maximum insolation—or the global average over the planetary disk. The subsolar equilibrium temperature assumes instantaneous local balance and is typically higher than the disk-averaged effective temperature, which accounts for the full illuminated and shadowed hemispheres; for example, on airless bodies, the subsolar value can exceed the average by factors related to rate and . Rapid rotators like approach a more uniform global average, while slow rotators like the exhibit stark subsolar highs. Observationally, the effective temperature and related surface inferences are derived from infrared emission spectra, which reveal the planet's total outgoing flux and spectral signatures of atmospheric layers. Space-based telescopes measure broadband thermal emission to compute the effective temperature via the Stefan-Boltzmann law applied to the integrated radiance, while detailed spectroscopy identifies temperature profiles and greenhouse influences through molecular absorption bands. This approach has been used for solar system planets like Mars and exoplanets, providing constraints on surface conditions without direct in-situ measurements.

Earth's Effective Temperature

The effective temperature of Earth, representing the temperature at which the planet radiates energy to space as a blackbody in equilibrium with incoming solar radiation, is approximately 255 K (-18°C). This value is derived from the solar constant of 1366 W/m², which measures the average incoming solar flux at Earth's distance from the Sun, and the Bond albedo of 0.3, indicating that 30% of incident sunlight is reflected back to space primarily by clouds, ice, and aerosols. The absorbed solar energy, after averaging over the planet's spherical geometry, balances the outgoing longwave radiation emitted at this effective temperature. In comparison, Earth's observed global mean surface is 288 (15°C), resulting in a 33 warming attributable to the natural . This enhancement arises from the and re-emission of by atmospheric constituents, particularly , , , and , which trap that would otherwise to . Early conceptual estimates of Earth's thermal balance date to 1824, when calculated that solar heating alone would yield a much colder average than observed, proposing that the atmosphere acts as an insulating layer to retain . Modern determinations rely on observations from the Clouds and the Earth's () instruments aboard missions such as and Aqua, which measure top-of-atmosphere reflected shortwave and emitted fluxes to quantify the global energy and confirm the 255 effective through the Stefan-Boltzmann applied to outgoing . As a baseline for climate dynamics, Earth's effective temperature underpins global warming models by establishing the pre-industrial energy equilibrium; perturbations like anthropogenic radiative forcing from greenhouse gas increases disrupt this balance, leading to net heat accumulation and surface warming. For instance, positive radiative forcing enhances the greenhouse effect, raising the effective temperature and amplifying climate feedbacks such as water vapor changes. CERES observations reveal seasonal variations in effective temperature driven by Earth's 23.5° axial tilt, with the summer hemisphere exhibiting higher values due to increased insolation and reduced cloud cover, while the winter hemisphere cools. Latitudinally, effective temperature gradients span from about 210 K over polar regions to 280 K near the equator, reflecting stronger solar absorption and emission at low latitudes compared to high latitudes where ice and persistent clouds elevate albedo.

References

  1. [1]
  2. [2]
    Glossary term: Effective Temperature
    The effective temperature of a star is the temperature of a theoretical perfect emitter (a "blackbody") that has the same surface area and the same total light ...
  3. [3]
    4.3 blackbody radiation - NASA
    An ideal blackbody absorbs all radiant energy that strikes it and re-emits all of the absorbed energy. Blackbody radiation emission depends on the temperature ...Introduction · The Spectrum of Blackbody... · Blackbody Emissions and...<|control11|><|separator|>
  4. [4]
    Black Body Radiation - Galileo
    Planck's essential assumption in deriving his formula was that the oscillators only exchange energy with the radiation in quanta hf. Einstein made clear that ...How is Radiation Absorbed? · Understanding the Black Body...
  5. [5]
    [PDF] Lecture Notes: Blackbody Radiation and Compton Scattering
    In December 1900, Max Planck introduced a formula to describe the specific pattern of light emitted in the form of blackbody radiation. Recall that in the late ...Missing: law | Show results with:law
  6. [6]
    Black-Body Radiation - AstroBaki - CASPER
    Sep 3, 2021 · The frequency content of blackbody radiation is given by the Planck Function: B ν = 2 h ν 3 c 2 ( e h ν k T − 1 ) {\displaystyle B_{\nu }={2h\ ...
  7. [7]
  8. [8]
    Stefan-Boltzmann Law - HyperPhysics
    The thermal energy radiated by a blackbody radiator per second per unit area is proportional to the fourth power of the absolute temperature.
  9. [9]
    Stefan-Boltzmann Law - JILA
    The Stefan-Boltzmann Law. The Stefan-Boltzmann Law was first discovered experimentally in 1879 by Stefan, then derived theoretically in 1884 by his student ...
  10. [10]
    Blackbody Radiation - HyperPhysics
    Wien's Displacement Law. For a blackbody radiator, the temperature can be found from the wavelength at which the radiation curve peaks.
  11. [11]
    Effective Temperature | COSMOS
    ### Summary of Effective Temperature
  12. [12]
    Effective Temperature - an overview | ScienceDirect Topics
    Effective temperature refers to the temperature of a star's surface that is determined based on its spectral energy distribution.
  13. [13]
    Solar Temperature - an overview | ScienceDirect Topics
    Solar temperature refers to the effective temperature of the Sun's surface, which can be approximated as 5772 K, characterized by the peak of its emitted ...
  14. [14]
    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
  15. [15]
    Effective Temperature | COSMOS
    The surface temperature, calculated by assuming a perfect blackbody radiating the same amount of energy per unit area as the star, is known as the effective ...
  16. [16]
    Stellar Diameters - CHARA Array
    The angular diameters range between 0.2 mas to 21 mas, with an average value of 2.5 mas. The parallaxes range between 0.3 mas to 797 mas (corresponding to ...
  17. [17]
    Direct measurement of stellar angular diameters by the VERITAS ...
    Apr 12, 2019 · Direct measurement of stellar angular diameters is difficult: at interstellar distances stars are generally too small to resolve by any ...
  18. [18]
    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 · IAU 2015 RESOLUTION B3 · THE RATIONALE AND...
  19. [19]
    Evolution from the Main Sequence to Red Giants | Astronomy
    The fusion of hydrogen to form helium changes the interior composition of a star, which in turn results in changes in its temperature, luminosity, and radius.
  20. [20]
    I. Stellar structures in the red giant branch phase
    1 where it can be seen that differences in effective temperature appear on the main sequence and further increase during the red-giant-branch phase. To better ...
  21. [21]
    The effect of stellar limb darkening values on the accuracy of the ...
    This means that the apparent stellar disc cannot be characterized by only one effective surface temperature value, and that is why the surface brightness ...
  22. [22]
    Limb Darkening - an overview | ScienceDirect Topics
    This effect, called limb darkening, is produced by the temperature gradient of the outer stellar atmosphere compared with the photospheres. The effect of limb ...
  23. [23]
    Lecture 10: Synthesis: The Hertzsprung-Russell Diagram
    May 3, 2008 · The Hertzsprung-Russell (H-R) Diagram is a plot of stellar luminosity vs. effective temperature, created by Hertzsprung and Russell.
  24. [24]
    Estimating stellar parameters from spectra - Astronomy & Astrophysics
    Figure 10 shows the results for model 125 (Teff = 4440 K, log g = 1.80 dex, [Fe/H] = 0.00 dex), which fit the data poorly according to the least squares ...
  25. [25]
    [PDF] Abundance determination in stellar atmospheres
    Effective temperature. 2. Surface gravity. 3. « Metallicity ». The general ... assumption): Saha's equation with. •. Enough atoms or ions excited in the ...
  26. [26]
    Towards stellar effective temperatures and diameters at 1 per cent ...
    While interferometric measurements of stellar angular diameters are the most direct method to gauge these, they are still limited to relatively nearby and ...
  27. [27]
    Angular Diameter of β UMa: Stellar Intensity Interferometry
    The VERITAS-SII limb-darkened angular diameter yields an effective temperature of 9700 ± 200 ± 200 K, consistent with ultraviolet spectrophotometry, and an ...
  28. [28]
    [1203.1941] A New Extensive Library of Synthetic Stellar Spectra ...
    Mar 8, 2012 · Abstract:We present a new library of synthetic spectra based on the stellar atmosphere code PHOENIX. It covers the wavelength range from ...
  29. [29]
    [PDF] Gaia FGK benchmark stars: Effective temperatures and surface ...
    The measurements of bolometric flux seem to be accurate to 5% or better, which contributes about 1% or less to the uncertainties in effective temperature. The ...
  30. [30]
    How do we calculate the Earth's effective temperature?
    Oct 10, 2001 · Earth's effective temperature is calculated using the Stefan-Boltzmann Law, equating absorbed energy (239.7 W/m2) to emitted radiation, and ...
  31. [31]
    Strange New Worlds - UMD Geology
    Incorporating albedo into estimates of effective temperature, we get: Venus: 227 K; Earth: 255 K; Mars: 217 K. Makes sense. Mars is farther from the sun than ...
  32. [32]
    Atmosphere/mantle coupling and feedbacks on Venus - AGU Journals
    Apr 29, 2014 · The average surface temperature of Venus is around 740 K due to the ... Effective temperature, Teff, 232 K. Reference oxygen escape rate ...
  33. [33]
    Transient reducing greenhouse warming on early Mars - Wordsworth
    Jan 5, 2017 · ... Mars would have had an equilibrium temperature of only 210 K [Wordsworth, 2016]. Carbon dioxide provides some greenhouse warming but not ...
  34. [34]
    1.1 The Earth's mean temperature | OpenLearn - The Open University
    The two main factors that influence the mean temperature of the Earth and its neighbouring planets are distance from the Sun and the nature and extent of their ...
  35. [35]
    Solar System Temperatures - NASA Science
    Feb 15, 2022 · Venus is the hottest at 867°F, Earth is 59°F, Mars is -85°F, and Pluto is -375°F. Temperatures generally decrease with distance from the Sun.Missing: blackbody | Show results with:blackbody
  36. [36]
    Planetary Temperature Factors
    the average distance of a planet from the Sun (the semi-major axis of its orbit) is overwhelmingly the most important factor in ...<|separator|>
  37. [37]
    The Climate Model: Behind The Scenes - SpaceEngine
    Nov 1, 2023 · The temperatures of the SpaceEngine climate model are based on energy transport calculations and account for planetary albedo, presence of an atmosphere, ...
  38. [38]
    Attribution of the present‐day total greenhouse effect - AGU Journals
    Oct 16, 2010 · The actual mean surface temperature is larger (by around 33°C, assuming a constant planetary albedo) due to the absorption and emission of long- ...
  39. [39]
    [PDF] 19740012369.pdf
    ... subsolar temperature is 116°K. The corresponding disk-averaged infrared brightness temperature would be about. 0.92 times this, or 107°K. These temperatures ...
  40. [40]
    The inference of temperature from the infrared spectra of planets.
    The paper is concerned primarily with the determination of the temperature of planetary atmospheres from ground-based observations of the vibration-rotation ...Missing: observational | Show results with:observational
  41. [41]
    Lacis 2018: Explaining climate - NASA GISS
    Nov 28, 2022 · However, the global-mean surface temperature of the Earth is known to be about 288 K, which implies that the Planck emission from the ground ...
  42. [42]
    Planetary Energy Balance – Introduction to Global Change
    Sep 20, 2010 · If we do this calculation for Earth, using a solar flux of 1367 Wm-2 and an albedo of 0.3, we calculate a planetary temperature of 255 K. This ...<|separator|>
  43. [43]
    [PDF] Chapter 2 The global energy balance
    The global average mean surface temperature of the earth is 288K (Table. 2.1). Above we deduced that the emission temperature of the Earth is 255K, considerably ...
  44. [44]
    Evidence - NASA Science
    Oct 23, 2024 · In 1824, Joseph Fourier calculated that an Earth-sized planet, at our distance from the Sun, ought to be much colder. He suggested something ...
  45. [45]
    CERES – Clouds and the Earth's Radiant Energy System
    The CERES instruments provide direct measurements of reflected solar radiation and emission of thermal infrared radiation to space across all wavelengths ...Data · CERES Operations · CERES Acronyms · Satellite MissionsMissing: effective | Show results with:effective
  46. [46]
    CERES Radiation Balance - NASA Scientific Visualization Studio
    Apr 16, 2021 · The CERES instrument aboard many Earth-orbiting satellites records the flow of reflected Solar radiation and reprocessed longwave radiation in ...
  47. [47]
    Chapter 7: The Earth's Energy Budget, Climate Feedbacks, and ...
    Effective radiative forcing (ERF) quantifies the energy gained or lost by the Earth system following an imposed perturbation (for instance in GHGs, aerosols or ...
  48. [48]
    Effective radiative forcing and adjustments in CMIP6 models - ACP
    Aug 17, 2020 · We evaluate effective radiative forcing and adjustments in 17 contemporary climate models that are participating in the Coupled Model Intercomparison Project ( ...
  49. [49]
    The Annual Cycle of Earth Radiation Budget from Clouds and the ...
    The seasonal cycle of the Earth radiation budget is investigated by use of data from the Clouds and the Earth's Radiant Energy System (CERES).Missing: effective | Show results with:effective
  50. [50]
    CERES EBAF: Clouds and Earth's Radiant Energy Systems (CERES ...
    EBAF is used for climate model evaluation, estimating the Earth's global mean energy budget, and to infer meridional heat transport.