Schwarzschild's equation for radiative transfer

Last updated

In the study of heat transfer, Schwarzschild's equation [1] [2] [3] is used to calculate radiative transfer (energy transfer via electromagnetic radiation) through a medium in local thermodynamic equilibrium that both absorbs and emits radiation.

Contents

The incremental change in spectral intensity, [4] (dIλ, [W/sr/m2/μm]) at a given wavelength as radiation travels an incremental distance (ds) through a non-scattering medium is given by:

where

This equation and various equivalent expressions are known as Schwarzschild's equation. The second term describes absorption of radiation by the molecules in a short segment of the radiation's path (ds) and the first term describes emission by those same molecules. In a non-homogeneous medium, these parameters can vary with altitude and location along the path, formally making these terms n(s), σλ(s), T(s), and Iλ(s). Additional terms are added when scattering is important. Integrating the change in spectral intensity [W/sr/m2/μm] over all relevant wavelengths gives the change in intensity [W/sr/m2]. Integrating over a hemisphere then affords the flux perpendicular to a plane (F, [W/m2]).

Schwarzschild's equation contains the fundamental physics needed to understand and quantify how increasing greenhouse gases (GHGs) in the atmosphere reduce the flux of thermal infrared radiation to space. If no other fluxes change, the law of conservation of energy demands that the Earth warm (from one steady state to another) until balance is restored between inward and outward fluxes. Schwarzschild's equation alone says nothing about how much warming would be required to restore balance. When meteorologists and climate scientists refer to "radiative transfer calculations" or "radiative transfer equations" (RTE), the phenomena of emission and absorption are handled by numerical integration of Schwarzschild's equation over a path through the atmosphere. Weather forecasting models and climate models use versions of Schwarzschild's equation optimized to minimize computation time. Online programs are available that perform computations using Schwarzschild's equation.

History

The Schwarzschild equation first appeared in Karl Schwarzschild's 1906 paper “Ueber das Gleichgewicht der Sonnenatmosphäre” (On the equilibrium of the solar atmosphere). [5]

Background

Radiative transfer refers to energy transfer through an atmosphere or other medium by means of electromagnetic waves or (equivalently) photons.  The simplest form of radiative transfer involves a collinear beam of radiation traveling through a sample to a detector.  That flux can be reduced by absorption, scattering or reflection, resulting in energy transmission over a path of less than 100%.  The concept of radiative transfer extends beyond simple laboratory phenomena to include thermal emission of radiation by the medium - which can result in more photons arriving at the end of a path than entering it.  It also deals with radiation arriving at a detector from a large source - such as the surface of the Earth or the sky.  Since emission can occur in all directions, atmospheric radiative transfer (like Planck's Law) requires units involving a solid angle, such as W/sr/m2.  

At the most fundamental level, the absorption and emission of radiation are controlled by the Einstein coefficients for absorption, emission and stimulated emission of a photon (B12, A21 and B21) and the density of molecules in the ground and excited states (n1 and n2). However, in the simplest physical situation – blackbody radiation – radiation and the medium through which it is passing are in thermodynamic equilibrium, and the rate of absorption and emission are equal. The spectral intensity [W/sr/m2/μm] and intensity [W/sr/m2] of blackbody radiation are given by the Planck function Bλ(T) and the Stefan–Boltzmann law. These expressions are independent of Einstein coefficients. Absorption and emission often reach equilibrium inside dense, non-transparent materials, so such materials often emit thermal infrared of nearly blackbody intensity. Some of that radiation is internally reflected or scattered at a surface, producing emissivity less than 1. The same phenomena makes the absorptivity of incoming radiation less than 1 and equal to emissivity (Kirchhoff's law).

When radiation has not passed far enough through a homogeneous medium for emission and absorption to reach thermodynamic equilibrium or when the medium changes with distance, Planck's Law and the Stefan-Boltzmann equation do not apply. This is often the case when dealing with atmospheres. If a medium is in Local Thermodynamic Equilibrium (LTE), then Schwarzschild's equation can be used to calculate how radiation changes as it travels through the medium. A medium is in LTE when the fraction of molecules in an excited state is determined by the Boltzmann distribution. LTE exists when collisional excitation and collisional relaxation of any excited state occur much faster than absorption and emission. [6] (LTE does not require the rates of absorption and emission to be equal.) The vibrational and rotational excited states of greenhouse gases that emit thermal infrared radiation are in LTE up to about 60 km. [7] Radiative transfer calculations show negligible change (0.2%) due to absorption and emission above about 50 km. Schwarzschild's equation therefore is appropriate for most problems involving thermal infrared in the Earth's atmosphere. The absorption cross-sections (σλ) used in Schwarzschild's equation arise from Einstein coefficients and processes that broaden absorption lines. In practice, these quantities have been measured in the laboratory; not derived from theory.

When radiation is scattered (the phenomena that makes the sky appear blue) or when the fraction of molecules in an excited state is not determined by the Boltzmann distribution (and LTE doesn't exist), more complicated equations are required. For example, scattering from clear skies reflects about 32 W/m2 (about 13%) of incoming solar radiation back to space. [8] Visible light is also reflected and scattered by aerosol particles and water droplets (clouds). Neither of these phenomena have a significant impact on the flux of thermal infrared through clear skies.

Schwarzschild's equation can not be used without first specifying the temperature, pressure, and composition of the medium through which radiation is traveling. When these parameters are first measured with a radiosonde, the observed spectrum of the downward flux of thermal infrared (DLR) agrees closely with calculations and varies dramatically with location. [9] [10] Where dI is negative, absorption is greater than emission, and net effect is to locally warm the atmosphere. Where dI is positive, the net effect is "radiative cooling". By repeated approximation, Schwarzschild's equation can be used to calculate the equilibrium temperature change caused by an increase in GHGs, but only in the upper atmosphere where heat transport by convection is unimportant.

Derivation

Schwarzschild's equation can be derived from Kirchhoff's law of thermal radiation, which states that absorptivity must equal emissivity at a given wavelength. (Like Schwarzschild's equation, Kirchhoff's law only applies to media in LTE.) Given a thin slab of atmosphere of incremental thickness ds, by definition its absorptivity is where I is the incident radiation and dIa is radiation absorbed by the slab. According to Beer's Law:

Also by definition, emissivity is equal to where dIe is the radiation emitted by the slab and Bλ(T) is the maximum radiation any object in LTE can emit. Setting absorptivity equal to emissivity affords:

The total change in radiation, dI, passing through the slab is given by:

Schwarzschild's equation has also been derived from Einstein coefficients by assuming a Maxwell–Boltzmann distribution of energy between a ground and excited state (LTE). [11] The oscillator strength for any transition between ground and excited state depends on these coefficients. The absorption cross-section (σλ) is empirically determined from this oscillator strength and the broadening of the absorption/emission line by collisions, the Doppler effect and the uncertainty principle.

Equivalent equations

Schwarzschild's equation has been expressed in different forms and symbols by different authors. The quantity λ is known as the absorption coefficient (βa), a measure of attenuation with units of [cm−1]. The absorption coefficient is fundamentally the product of a quantity of absorbers per unit volume, [cm−3], times an efficiency of absorption (area/absorber, [cm2]). Several sources [2] [12] [3] replace λ with kλr, where kλ is the absorption coefficient per unit density and r is the density of the gas. The absorption coefficient for spectral flux (a beam of radiation with a single wavelength, [W/m2/μm]) differs from the absorption coefficient for spectral intensity [W/sr/m2/μm] used in Schwarzschild's equation.

Integration of an absorption coefficient over a path from s1 and s2 affords the optical thickness (τ) of that path, a dimensionless quantity that is used in some variants of the Schwarzschild equation. When emission is ignored, the incoming radiation is reduced by a factor for 1/e when transmitted over a path with an optical thickness of 1.

When expressed in terms of optical thickness, Schwarzschild's equation becomes: [1] [13]

After integrating between a sensor located at τ = 0 and an arbitrary starting point in the medium, τ', the spectral intensity of the radiation reaching the sensor, Iλ(0), is:

where I(τ') is the spectral intensity of the radiation at the beginning of the path, is the transmittance along the path, and the final term is the sum of all of the emission along the path attenuated by absorption along the path yet to be traveled. [1]

Relationship to Planck's and Beer's laws

Both Beer's Law and Planck's Law can be derived from Schwarzschild's equation. [14] In a sense, they are corollaries of Schwarzschild's equation.

When the spectral intensity of radiation is not changing as it passes through a medium, dIλ = 0. In that situation, Schwarzschild's equation simplifies to Planck's law:

When Iλ > Bλ(T), dI is negative and when Iλ < Bλ(T), dI is positive. As a consequence, the intensity of radiation traveling through any medium is always approaching the blackbody intensity given by Planck's law and the local temperature. The rate of approach depends on the density of absorbing/emitting molecules (n) and their absorption cross-section (σλ).

When the intensity of the incoming radiation, Iλ, is much greater than the intensity of blackbody radiation, Bλ(T), the emission term can be neglected. This is usually the case when working with a laboratory spectrophotometer, where the sample is near 300 K and the light source is a filament at several thousand K.

If the medium is homogeneous, λ doesn't vary with location. Integration over a path of length s affords the form of Beer's Law used most often in the laboratory experiments:

Greenhouse effect

Schwarzschild's equation provides a simple explanation for the existence of the greenhouse effect and demonstrates that it requires a non-zero lapse rate. [15] Rising air in the atmosphere expands and cools as the pressure on it falls, producing a negative temperature gradient in the Earth's troposphere. When radiation travels upward through falling temperature, the incoming radiation, I, (emitted by the warmer surface or by GHGs at lower altitudes) is more intense than that emitted locally by Bλ(T). [Bλ(T) − I] is generally less than zero throughout the troposphere, and the intensity of outward radiation decreases as it travels upward. According to Schwarzschild's equation, the rate of fall in outward intensity is proportional to the density of GHGs (n) in the atmosphere and their absorption cross-sections (σλ). Any anthropogenic increase in GHGs will slow down the rate of radiative cooling to space, i.e. produce a radiative forcing until a saturation point is reached.

At steady state, incoming and outgoing radiation at the top of the atmosphere (TOA) must be equal. When the presence of GHGs in the atmosphere causes outward radiation to decrease with altitude, then the surface must be warmer than it would be without GHGs - assuming nothing else changed. Some scientists quantify the greenhouse effect as the 150 W/m2 difference between the average outward flux of thermal IR from the surface (390 W/m2) and the average outward flux at the TOA.

If the Earth had an isothermal atmosphere, Schwarzschild's equation predicts that there would be no greenhouse effect or no enhancement of the greenhouse effect by rising GHGs. In fact, the troposphere over the Antarctic plateau is nearly isothermal. Both observations and calculations show a slight "negative greenhouse effect" – more radiation emitted from the TOA than the surface. [16] [17] Although records are limited, the central Antarctic Plateau has seen little or no warming. [18]

Saturation

In the absence of thermal emission, wavelengths that are strongly absorbed by GHGs can be significantly attenuated within 10 m in the lower atmosphere. Those same wavelengths, however, are the ones where emission is also strongest. In an extreme case, roughly 90% of 667.5 cm−1 photons are absorbed within 1 meter by 400 ppm of CO2 at surface density, [19] but they are replaced by emission of an equal number of 667.5 cm−1 photons. The radiation field thereby maintains the blackbody intensity appropriate for the local temperature. At equilibrium, Iλ = Bλ(T) and therefore dIλ = 0 even when the density of the GHG (n) increases.

This has led some to falsely believe that Schwarzschild's equation predicts no radiative forcing at wavelengths where absorption is "saturated". However, such reasoning reflects what some refer to as the surface budget fallacy. This fallacy involves reaching erroneous conclusions by focusing on energy exchange near the planetary surface rather than at the top of the atmosphere (TOA). At wavelengths where absorption is saturated, increasing the concentration of a greenhouse gas does not change thermal radiation levels at low altitudes, but there are still important differences at high altitudes where the air is thinner. [20] :413

As density decreases with altitude, even the strongest absorption bands eventually become semi-transparent. Once that happens, radiation can travel far enough that the local emission, Bλ(T), can differ from the absorption of incoming Iλ. The altitude where the transition to semi-transparency occurs is referred to as the "effective emission altitude" or "effective radiating level." Thermal radiation from this altitude is able to escape to space. Consequently, the temperature at this level sets the intensity of outgoing longwave radiation. This altitude varies depending on the particular wavelength involved. [21] [20] :413

Increasing concentration increases the "effective emission altitude" at which emitted thermal radiation is able to escape to space. The lapse rate (change in temperature with altitude) at the effective radiating level determines how a change in concentration will affect outgoing emissions to space. For most wavelengths, this level is in the troposphere, where temperatures decrease with increasing altitude. This means that increasing concentrations of greenhouse gas lead to decreasing emissions to space (a positive incremental greenhouse effect), creating an energy imbalance that makes the planet warmer than it would be otherwise. Thus, the presence or absence of absorption saturation at low altitudes does not necessarily indicate that absence of radiative forcing in response to increased concentrations. [20] :413

The radiative forcing from doubling carbon dioxide occurs mostly on the flanks of the strongest absorption band. [22]

Temperature rises with altitude in the lower stratosphere, and increasing CO2 there increases radiative cooling to space and is predicted by some to cause cooling above 14–20 km. [21]

Application to climate science

Schwarzschild's equation is used to calculate the outward radiative flux from the Earth (measured in W/m2 perpendicular to the surface) at any altitude, especially the "top of the atmosphere" or TOA. This flux originates at the surface (I0) for clear skies or cloud tops. dI increments are calculated for layers thin enough to be effectively homogeneous in composition and flux (I). These increments are numerically integrated from the surface to the TOA to give the flux of thermal infrared to space, commonly referred to as outgoing long-wavelength radiation (OLR). OLR is the only mechanism by which the Earth gets rid of the heat delivered continuously by the sun. The net downward radiative flux of thermal IR (DLR) produced by emission from GHGs in the atmosphere is obtained by integrating dI from the TOA (where I0 is zero) to the surface. DLR adds to the energy from the sun. Emission from each layer adds equally to the upward and downward fluxes. In contrast, different amounts of radiation are absorbed, because the upward flux entering any layer is usually greater than the downward flux.

In "line-by-line" methods, the change in spectral intensity (dIλ, W/sr/m2/μm) is numerically integrated using a wavelength increment small enough (less than 1 nm) to accurately describe the shape of each absorption line. The HITRAN database contains the parameters needed to describe 7.4 million absorption lines for 47 GHGs and 120 isotopologues. A variety of programs or radiative transfer codes can be used to process this data, including an online facility, SpectralCalc. [23] To reduce the computational demand, weather forecast and climate models use broad-band methods that handle many lines as a single "band". [24] MODTRAN [25] is a broad-band method available online with a simple interface that anyone can use.

To convert intensity [W/sr/m2] to flux [W/m2], calculations usually invoke the "two-stream" and "plane parallel" approximations. [13] [26] The radiative flux is decomposed into three components, upward (+z), downward (-z), and parallel to the surface. This third component contributes nothing to heating or cooling the planet. , where is the zenith angle (away from vertical). Then the upward and downward intensities are integrated over a forward hemisphere, a process that can be simplified by using a "diffusivity factor" or "average effective zenith angle" of 53°. Alternatively, one can integrate over all possible paths from the entire surface to a sensor positioned a specified height above surface for OLR, or over all possible paths from the TOA to a sensor on the surface for DLR.

Related Research Articles

The Beer-Lambert law is commonly applied to chemical analysis measurements to determine the concentration of chemical species that absorb light. It is often referred to as Beer's law. In physics, the Bouguer–Lambert law is an empirical law which relates the extinction or attenuation of light to the properties of the material through which the light is travelling. It had its first use in astronomical extinction. The fundamental law of extinction is sometimes called the Beer-Bouguer-Lambert law or the Bouguer-Beer-Lambert law or merely the extinction law. The extinction law is also used in understanding attenuation in physical optics, for photons, neutrons, or rarefied gases. In mathematical physics, this law arises as a solution of the BGK equation.

<span class="mw-page-title-main">Greenhouse effect</span> Atmospheric phenomenon causing planetary warming

The greenhouse effect occurs when greenhouse gases in a planet's atmosphere insulate the planet from losing heat to space, raising its surface temperature. Surface heating can happen from an internal heat source as in the case of Jupiter, or from its host star as in the case of the Earth. In the case of Earth, the Sun emits shortwave radiation (sunlight) that passes through greenhouse gases to heat the Earth's surface. In response, the Earth's surface emits longwave radiation (heat) that is mostly absorbed by greenhouse gases. That heat absorption reduces the rate at which the Earth can cool off in response to being warmed by the Sun. Adding to greenhouse gases further reduces the rate a planet emits radiation to space, raising its average surface temperature.

<span class="mw-page-title-main">Optical depth</span> Physics concept

In physics, optical depth or optical thickness is the natural logarithm of the ratio of incident to transmitted radiant power through a material. Thus, the larger the optical depth, the smaller the amount of transmitted radiant power through the material. Spectral optical depth or spectral optical thickness is the natural logarithm of the ratio of incident to transmitted spectral radiant power through a material. Optical depth is dimensionless, and in particular is not a length, though it is a monotonically increasing function of optical path length, and approaches zero as the path length approaches zero. The use of the term "optical density" for optical depth is discouraged.

<span class="mw-page-title-main">Stefan–Boltzmann law</span> Physical law on the emissive power of black body

The Stefan–Boltzmann law, also known as Stefan's law, describes the intensity of the thermal radiation emitted by matter in terms of that matter's temperature. It is named for Josef Stefan, who empirically derived the relationship, and Ludwig Boltzmann who derived the law theoretically.

<span class="mw-page-title-main">Thermal radiation</span> Electromagnetic radiation generated by the thermal motion of particles

Thermal radiation is electromagnetic radiation emitted by the thermal motion of particles in matter. Thermal radiation transmits as an electromagnetic wave through both matter and vacuum. When matter absorbs thermal radiation its temperature will tend to rise. All matter with a temperature greater than absolute zero emits thermal radiation. The emission of energy arises from a combination of electronic, molecular, and lattice oscillations in a material. Kinetic energy is converted to electromagnetism due to charge-acceleration or dipole oscillation. At room temperature, most of the emission is in the infrared (IR) spectrum. Thermal radiation is one of the fundamental mechanisms of heat transfer, along with conduction and convection.

<span class="mw-page-title-main">Planck's law</span> Spectral density of light emitted by a black body

In physics, Planck's law describes the spectral density of electromagnetic radiation emitted by a black body in thermal equilibrium at a given temperature T, when there is no net flow of matter or energy between the body and its environment.

In radiometry, radiance is the radiant flux emitted, reflected, transmitted or received by a given surface, per unit solid angle per unit projected area. Radiance is used to characterize diffuse emission and reflection of electromagnetic radiation, and to quantify emission of neutrinos and other particles. The SI unit of radiance is the watt per steradian per square metre. It is a directional quantity: the radiance of a surface depends on the direction from which it is being observed.

<span class="mw-page-title-main">Kirchhoff's law of thermal radiation</span> Law of wavelength-specific emission and absorption

In heat transfer, Kirchhoff's law of thermal radiation refers to wavelength-specific radiative emission and absorption by a material body in thermodynamic equilibrium, including radiative exchange equilibrium. It is a special case of Onsager reciprocal relations as a consequence of the time reversibility of microscopic dynamics, also known as microscopic reversibility.

<span class="mw-page-title-main">Black-body radiation</span> Thermal electromagnetic radiation

Black-body radiation is the thermal electromagnetic radiation within, or surrounding, a body in thermodynamic equilibrium with its environment, emitted by a black body. It has a specific, continuous spectrum of wavelengths, inversely related to intensity, that depend only on the body's temperature, which is assumed, for the sake of calculations and theory, to be uniform and constant.

<span class="mw-page-title-main">Emissivity</span> Capacity of an object to radiate electromagnetic energy

The emissivity of the surface of a material is its effectiveness in emitting energy as thermal radiation. Thermal radiation is electromagnetic radiation that most commonly includes both visible radiation (light) and infrared radiation, which is not visible to human eyes. A portion of the thermal radiation from very hot objects is easily visible to the eye.

The effective temperature of a body such as a star or planet is the temperature of a black body that would emit the same total amount of electromagnetic radiation. Effective temperature is often used as an estimate of a body's surface temperature when the body's emissivity curve is not known.

Radiative transfer is the physical phenomenon of energy transfer in the form of electromagnetic radiation. The propagation of radiation through a medium is affected by absorption, emission, and scattering processes. The equation of radiative transfer describes these interactions mathematically. Equations of radiative transfer have application in a wide variety of subjects including optics, astrophysics, atmospheric science, and remote sensing. Analytic solutions to the radiative transfer equation (RTE) exist for simple cases but for more realistic media, with complex multiple scattering effects, numerical methods are required. The present article is largely focused on the condition of radiative equilibrium.

The source function is a characteristic of a stellar atmosphere, and in the case of no scattering of photons, describes the ratio of the emission coefficient to the absorption coefficient. It is a measure of how photons in a light beam are removed and replaced by new photons by the material it passes through. Its units in the cgs-system are erg s−1 cm−2 sr−1 Hz−1 and in SI are W m−2 sr−1 Hz−1. The source function can be written

In astronomy, the spectral index of a source is a measure of the dependence of radiative flux density on frequency. Given frequency in Hz and radiative flux density in Jy, the spectral index is given implicitly by

<span class="mw-page-title-main">Outgoing longwave radiation</span> Energy transfer mechanism which enables planetary cooling

In climate science, longwave radiation (LWR) is electromagnetic thermal radiation emitted by Earth's surface, atmosphere, and clouds. It may also be referred to as terrestrial radiation. This radiation is in the infrared portion of the spectrum, but is distinct from the shortwave (SW) near-infrared radiation found in sunlight.

In radiometry, radiosity is the radiant flux leaving a surface per unit area, and spectral radiosity is the radiosity of a surface per unit frequency or wavelength, depending on whether the spectrum is taken as a function of frequency or of wavelength. The SI unit of radiosity is the watt per square metre, while that of spectral radiosity in frequency is the watt per square metre per hertz (W·m−2·Hz−1) and that of spectral radiosity in wavelength is the watt per square metre per metre (W·m−3)—commonly the watt per square metre per nanometre. The CGS unit erg per square centimeter per second is often used in astronomy. Radiosity is often called intensity in branches of physics other than radiometry, but in radiometry this usage leads to confusion with radiant intensity.

The linear attenuation coefficient, attenuation coefficient, or narrow-beam attenuation coefficient characterizes how easily a volume of material can be penetrated by a beam of light, sound, particles, or other energy or matter. A coefficient value that is large represents a beam becoming 'attenuated' as it passes through a given medium, while a small value represents that the medium had little effect on loss. The (derived) SI unit of attenuation coefficient is the reciprocal metre (m−1). Extinction coefficient is another term for this quantity, often used in meteorology and climatology. Most commonly, the quantity measures the exponential decay of intensity, that is, the value of downward e-folding distance of the original intensity as the energy of the intensity passes through a unit thickness of material, so that an attenuation coefficient of 1 m−1 means that after passing through 1 metre, the radiation will be reduced by a factor of e, and for material with a coefficient of 2 m−1, it will be reduced twice by e, or e2. Other measures may use a different factor than e, such as the decadic attenuation coefficient below. The broad-beam attenuation coefficient counts forward-scattered radiation as transmitted rather than attenuated, and is more applicable to radiation shielding. The mass attenuation coefficient is the attenuation coefficient normalized by the density of the material.

<span class="mw-page-title-main">Idealized greenhouse model</span> Mathematical estimate of planetary temperatures

The temperatures of a planet's surface and atmosphere are governed by a delicate balancing of their energy flows. The idealized greenhouse model is based on the fact that certain gases in the Earth's atmosphere, including carbon dioxide and water vapour, are transparent to the high-frequency solar radiation, but are much more opaque to the lower frequency infrared radiation leaving Earth's surface. Thus heat is easily let in, but is partially trapped by these gases as it tries to leave. Rather than get hotter and hotter, Kirchhoff's law of thermal radiation says that the gases of the atmosphere also have to re-emit the infrared energy that they absorb, and they do so, also at long infrared wavelengths, both upwards into space as well as downwards back towards the Earth's surface. In the long-term, the planet's thermal inertia is surmounted and a new thermal equilibrium is reached when all energy arriving on the planet is leaving again at the same rate. In this steady-state model, the greenhouse gases cause the surface of the planet to be warmer than it would be without them, in order for a balanced amount of heat energy to finally be radiated out into space from the top of the atmosphere.

The skin temperature of an atmosphere is the temperature of a hypothetical thin layer high in the atmosphere that is transparent to incident solar radiation and partially absorbing of infrared radiation from the planet. It provides an approximation for the temperature of the tropopause on terrestrial planets with greenhouse gases present in their atmospheres.

There is a strong scientific consensus that greenhouse effect due to carbon dioxide is a main driver of climate change. Following is an illustrative model meant for a pedagogical purpose, showing the main physical determinants of the effect.

References

  1. 1 2 3 Petty, Grant (2006). A First Course in Radiation Physics (2nd ed.). Madison, WI: Sundog Publishing. pp. 204–210. ISBN   978-0-9729033-1-8. Those without access to a text discussing Schwarzschild's equation can find a discussion abstracted from these pages at: https://scienceofdoom.com/2011/02/07/understanding-atmospheric-radiation-and-the-“greenhouse”-effect-–-part-six-the-equations/.
  2. 1 2 Wallace and Hobbs (2006). Atmospheric Science (2nd ed.). Amsterdam: Elsevier. pp.  134–136. ISBN   978-0-12-732951-2.
  3. 1 2 Barrett & Bellamy. "Schwarzschild's Equation" . Retrieved 25 September 2018.
  4. For readability and clarity, radiation is described herein using the intuitive terms used by meteorologists: intensity [W/m2/sr] and flux [W/m2] followed by units. For example, see: Wallace and Hobbs (2006). Atmospheric science : an introductory survey (2nd ed.). Elsevier Academic Press. pp.  114–117. ISBN   978-0-12-732951-2.
  5. Mobley, Curtis. "Notes on The Evolution of Radiative Transfer Theory" (PDF). Ocean Optics. Retrieved 19 October 2023.
  6. Petty, Grant (2006). A First Course in Radiation Physics (2nd ed.). Madison, WI: Sundog Publishing. pp. 126–127. ISBN   978-0-9729033-1-8.
  7. Wallace and Hobbs (2006). Atmospheric Science (2nd ed.). p.  60. ISBN   978-0-12-732951-2.
  8. Stephens, Graeme L.; O'Brien, Denis; Webster, Peter J.; Pilewski, Peter; Kato, Seiji; Li, Jui-lin (March 2015). "The albedo of Earth" (PDF). Reviews of Geophysics. 53 (1): 141–163. Bibcode:2015RvGeo..53..141S. doi: 10.1002/2014RG000449 .
  9. "Theory and Experiment – Atmospheric Radiation". The Science of Doom. 1 November 2010. Provides a variety of comparisons between observed and calculated spectra and fluxes.
  10. Grant, Petty (2006). A first course in atmospheric radiation (2nd ed.). Sundog Pub. pp. 219–228. ISBN   978-0-9729033-1-8.
  11. Harde, Hermann (2013). "Radiation and Heat Transfer in the Atmosphere: A Comprehensive Approach on a Molecular Basis". International Journal of Atmospheric Sciences. 2013: 1–26. doi: 10.1155/2013/503727 . ISSN   2314-4122.
  12. Taylor, F. W. (2005). Elementary climate physics. Oxford University Press. pp. 91–93. ISBN   978-0-19-856733-2.
  13. 1 2 Pierrehumbert, R. T. (2010). Principles of planetary climate . Cambridge University Press. pp.  188–204. ISBN   978-0-521-86556-2.
  14. Pierrehumbert, R. T. (2010-12-02). Principles of planetary climate . Cambridge University Press. p.  194. ISBN   978-0-521-86556-2.
  15. The requirement for a non-zero lapse rate is discussed on page 202 of Pierrehumbert.
  16. Schmithüsen, Holger; Notholt, Justus; König-Langlo, Gert; Lemke, Peter; Jung, Thomas (16 December 2015). "How increasing CO2 leads to an increased negative greenhouse effect in Antarctica". Geophysical Research Letters. 42 (23): 10, 422–10, 428. Bibcode:2015GeoRL..4210422S. doi: 10.1002/2015GL066749 .
  17. Sejas, Sergio A.; Taylor, Patrick C.; Cai, Ming (11 July 2018). "Unmasking the negative greenhouse effect over the Antarctic Plateau". npj Climate and Atmospheric Science. 1 (1): 17. doi: 10.1038/s41612-018-0031-y . PMC   7580794 . PMID   33102742.
  18. See the first paragraph of Schmithüsen (2015).
  19. Petty, Grant (2006). A First Course in Radiation Physics (2nd ed.). p. 276.
  20. 1 2 3 Pierrehumbert, Raymond T. (2010). Principles of Planetary Climate. Cambridge University Press. ISBN   978-0-521-86556-2.
  21. 1 2 Hegerl; et al. (2007). IPCC AR4 WG1 Chapter 9. Figure 9.1c (PDF). p. 675. Retrieved 25 September 2018.
  22. "Radiative forcing for doubling CO2". 2018-09-06.{{cite web}}: Check |url= value (help)
  23. "Spectral Calculator-Hi-resolution gas spectra". spectralcalc.com.
  24. Petty, Grant (2006). A first course in atmospheric radiation (2nd ed.). Sundog Pub. pp. 286–305. ISBN   978-0-9729033-1-8.
  25. "MODTRAN Infrared Light in the Atmosphere".
  26. Petty, Grant (2006). A first course in atmospheric radiation (2nd ed.). Sundog Pub. pp. 392–400. ISBN   978-0-9729033-1-8.