Radiative forcing (Redirected from Climate forcing agents)

Earth receives a nearly constant global & annual average of about 340 Watts per square meter of incoming solar radiation.

Radiative forcing is the difference between solar irradiance (sunlight) absorbed by the Earth and energy radiated back to space.[1] It is the scientific basis for the greenhouse effect on planets, and plays an important role in computational models of Earth's energy balance and climate. Changes to Earth's radiative equilibrium that cause temperatures to rise or fall over decadal periods are called climate forcings.[2]

Positive radiative forcing means Earth receives more incoming energy from sunlight than it radiates to space. This net gain of energy will cause warming. Conversely, negative radiative forcing means that Earth loses more energy to space than it receives from the sun, which produces cooling. A planet in radiative equilibrium with its parent star and the rest of space can be characterized by net-zero radiative forcing and by a planetary equilibrium temperature.[3]

Radiative forcing on Earth is meaningfully evaluated at the tropopause and at the top of the stratosphere. It is quantified in units of watts per square meter, and often summarized as an average over the total surface area of the globe. Radiative forcing varies with solar insolation, surface albedo, and the atmospheric concentrations of radiatively active gases - commonly known as greenhouse gases - and aerosols.

Radiation balance

Atmospheric gases only absorb some wavelengths of energy but are transparent to others. The absorption patterns of water vapor (blue peaks) and carbon dioxide (pink peaks) overlap in some wavelengths. Carbon dioxide is not as strong a greenhouse gas as water vapor, but it absorbs energy in wavelengths (12-15 micrometers) that water vapor does not, partially closing the “window” through which heat radiated by the surface would normally escape to space. (Illustration NASA, Robert Rohde)[4]

Almost all of the energy that affects Earth's climate is received as radiant energy from the Sun. The planet and its atmosphere absorb and reflect some of the energy, while long-wave energy is radiated back into space. The balance between absorbed and radiated energy determines the average global temperature. Because the atmosphere absorbs some of the re-radiated long-wave energy, the planet is warmer than it would be in the absence of the atmosphere: see greenhouse effect.

The radiation balance is altered by such factors as the intensity of solar energy, reflectivity of clouds or gases, absorption by various greenhouse gases or surfaces and heat emission by various materials. Any such alteration is a radiative forcing, and changes the balance. This happens continuously as sunlight hits the surface, clouds and aerosols form, the concentrations of atmospheric gases vary and seasons alter the groundcover.

IPCC usage

Radiative forcings, IPCC 2013.

The Intergovernmental Panel on Climate Change (IPCC) AR4 report defines radiative forcings as:[5]

"Radiative forcing is a measure of the influence a factor has in altering the balance of incoming and outgoing energy in the Earth-atmosphere system and is an index of the importance of the factor as a potential climate change mechanism. In this report radiative forcing values are for changes relative to preindustrial conditions defined at 1750 and are expressed in Watts per square meter (W/m2)."

In simple terms, radiative forcing is "...the rate of energy change per unit area of the globe as measured at the top of the atmosphere."[6] In the context of climate change, the term "forcing" is restricted to changes in the radiation balance of the surface-troposphere system imposed by external factors, with no changes in stratospheric dynamics, no surface and tropospheric feedbacks in operation (i.e., no secondary effects induced because of changes in tropospheric motions or its thermodynamic state), and no dynamically induced changes in the amount and distribution of atmospheric water (vapour, liquid, and solid forms).

Basic estimates

Radiative forcing can be evaluated for its dependence on different factors which are external to the climate system.[7] Except where necessary and noted, the basic estimates which follow do not include indirect feedbacks (positive or negative) which also occur via Earth system responses. Forcing changes (ΔF) are expressed as yearly averages over the total surface of the planet. They may be significant in the context of global climate forcing for times spanning decades or longer.[8]

Forcing due to changes in solar irradiance

The intensity of solar radiation including all wavelengths is the Total Solar Irradiance (TSI) and is sometimes also referred to erroneously as the solar constant. It is equal to about 1361 W m−2 at the distance of Earth's annual-mean orbital radius of one astronomical unit and as measured at the top of the atmosphere.[9] Earth TSI varies with both solar activity and planetary orbital dynamics. Multiple satellite-based instruments including ERB, ACRIM 1-3, VIRGO, and TIM[10][11] have continuously measured TSI with improving accuracy and precision since 1978.[12]

Approximating Earth as a sphere, the cross-sectional area exposed to the Sun () is equal to one quarter the area of the planet's surface (). The globally and annually averaged amount of solar irradiance per square meter of Earth's atmospheric surface () is therefore equal to one quarter of TSI, and has a nearly constant value of .

Annual cycles

Earth follows an elliptical orbit around the Sun such that TSI received at any instance fluctuates between about 1321 W m−2 (at aphelion in early July) and 1412 W m−2 (at perihelion in early January), or thus by about +/-3.4% during each year.[13] The change in instantaneous radiative forcing has minor influences on Earth's seasonal weather patterns and its climate zones, which primarily result from the annual cycling in Earth's relative tilt direction.[14] Such repeating cycles contribute a net-zero forcing (by definition) in the context of decades-long climate changes.

Sunspot activity

Line graph showing historical sunspot number count, Maunder and Dalton minima, and the Modern Maximum
400 year sunspot history, including the Maunder Minimum

Average annual TSI varies between about 1360 W m−2 and 1362 W m−2 (+/-0.05%) over the course of a typical 11-year sunspot activity cycle.[15] Sunspot observations have been recorded since about year 1600 and show evidence of lengthier oscillations (Gleissberg cycle, Devries/Seuss cycle, etc.) which modulate the 11-year cycle (Schwabe cycle). Despite such complex behavior, the amplitude of the 11-year cycle has been the most prominent variation throughout this long-term observation record.[16]

TSI variations associated with sunspots contribute a small but non-zero net forcing in the context of decadal climate changes.[12] Some research suggests they may have partly influenced climate shifts during the Little Ice Age, along with concurrent changes in volcanic activity and deforestation.[17] Since the late 20th century, average TSI has trended slightly lower along with a downward trend in sunspot activity.[18]

Milankovitch shifts

Climate forcing caused by variations in solar irradiance have occurred during Milankovitch cycles, which span periods of about 40,000 to 100,000 years. Milankovitch cycles consist of similarly long-duration cycles in Earth's orbital eccentricity (or ellipticity), orbital obliquity, and tilt direction.[19] Among these, the 100,000 year cycle in eccentricity causes TSI to fluctuate by about +/-0.2%.[20] Currently, Earth’s eccentricity is nearing its least elliptic (most circular) causing average annual TSI to very slowly decrease.[19] Simulations also indicate that Earth's orbital dynamics will remain stable including these variations for least the next 10 million years.[21]

Sun aging

Our Sun has consumed about half its hydrogen fuel since forming approximately 4.5 billion years ago.[22] TSI will continue to slowly increase during the aging process at a rate of about 1% each 100 million years. Such rate of change is far too small to be detectable within measurements and is insignificant on human timescales.

TSI forcing summary

TSI forcing (max. 10-yr change)
Δτ ΔF (W m−2)
Annual cycle +/- 0.034 [13] 0 (net)
Sunspot activity - 5e-4 [15] - 0.1 [18][23]
Orbital shift - 4e-7 [20] - 1e-4
Sun aging + 1e-9 [22] + 2e-7

The maximum fractional variations (Δτ) in Earth's solar irradiance during the last decade are summarized in the accompanying table. Each variation previously discussed contributes a forcing of:

,

where R=0.30 is Earth's reflectivity. The radiative and climate forcings arising from changes in the Sun's insolation are expected to continue to be minor, notwithstanding some as-of-yet undiscovered solar physics.[18][24]

Forcing due to changes in albedo

A fraction of incident solar radiation is reflected by clouds & aerosols, oceans and landforms, snow & ice, vegetation, and other natural & man-made surface features. The reflected fraction is known as Earth's bond albedo (R), is evaluated at the top of the atmosphere, and has an average annual global value of about 0.30 (30%). The overall fraction of solar power absorbed by Earth is then (1-R) or 0.70 (70%).[25]

Atmospheric components contribute about three-quarters of Earth albedo, and clouds alone are responsible for half. The pronounced roles of clouds and water vapor are linked with the majority presence of liquid water covering the planet's crust. Global patterns in cloud formation and circulation are highly complex phenomena with couplings to ocean heat flows, and with jet streams assisting their rapid transport. Moreover, the albedos of Earth's northern and southern hemispheres have been observed to be essentially equal (within 0.2%). This is noteworthy since more than two-thirds of land and 85% of the human population are distributed north.[26]

Multiple satellite-based instruments including MODIS, VIIRs, and CERES have continuously monitored Earth's albedo since 1998.[27] Landsat imagery available since 1972 has also been used in some studies.[28] Measurement accuracy has improved and results have converged in recent years, enabling more confident assessment of the recent decadal forcing influence of planetary albedo.[26] Nevertheless, the existing data record is still too short to support longer-term predictions or to address other related questions.

Annual cycles

Seasonal variations in planetary albedo can be understood as a set of system feedbacks that occur largely in response to the cycle of solar forcing. Along with the atmospheric responses, most apparent to surface dwellers are the changes in vegetation, snow, and sea-ice coverage. Intra-annual variations of about +/-0.02 (+/- 7%) around Earth's mean albedo have been observed throughout the course of a year, with maxima occurring twice per year near the time of each solar equinox.[26] This repeating cycle contributes net-zero forcing in the context of decades-long climate changes.

Interannual variability

Measured global albedo anomaly from CERES (2000-2011).

Regional albedos change from year to year due to shifts arising from natural processes, human actions, and system feedbacks. For example, human acts of deforestion typically raise Earth's reflectivity while introducing water storage and irrigation to arid lands may lower it. Likewise considering feedbacks, ice loss in arctic regions decreases albedo while expanding desertification at low to middle latitudes increases it.

During years 2000-2012, no overall trend in Earth's albedo was discernible within the 0.1% standard deviation of values measured by CERES.[26] Along with the hemispherical equivalence, some researchers interpret the remarkably small interannual differences as evidence that planetary albedo may currently be constrained by the action of complex system feedbacks. Nevertheless, historical evidence also suggests that infrequent events such as major volcanic eruptions can significantly perturb the planetary albedo for several years or longer.[29]

Albedo forcing summary

Albedo forcing (max. 10-yr change)
Δα ΔF (W m−2)
Annual cycle +/- 0.07[26] 0 (net)
Interannual variation +/- 0.001[26] -/+ 0.1

The measured fractional variations (Δα) in Earth's albedo during the first decade of the 21st century are summarized in the accompanying table. Similar to TSI, the radiative forcing due to a fractional change in planetary albedo (Δα) is:

.

Satellite observations show that various Earth system feedbacks have stabilized planetary albedo despite recent natural and human-caused shifts.[27] On longer timescales, it is more uncertain whether the net forcing which results from such external changes will remain minor.

Forcing due to changes in atmospheric gas

Radiative forcing for doubling CO
2
, as calculated by radiative transfer code Modtran. Red lines are Planck curves.

For a well-mixed greenhouse gas, radiative transfer codes that examine each spectral line for atmospheric conditions can be used to calculate the forcing change ΔF as a function of a change in its concentration. These calculations may be simplified into an algebraic formulation that is specific to that gas.

Carbon dioxide

A simplified first-order approximation expression for carbon dioxide is:[30]

,

where C is the CO
2
concentration in parts per million (ppm) by volume and C0 is the reference concentration (278 ppm in year 1750}) prior to substantial anthropogenic changes.

The atmospheric burden of greenhouse gases due to human activity has grown especially rapidly during the last several decades (since about year 1950). The 50% increase (C/C0=1.5) for CO
2
realized as of year 2020 corresponds to . By comparison, a sustained 1% increase in TSI or 2% decrease in albedo might be required to induce a similar magnitude of forcing, as per these basic estimates. Assuming no change in the emissions growth path, a doubling (C/C0=2) within the next several decades would correspond to ΔF=+3.7 W m−2.

The relationship between CO
2
and radiative forcing is logarithmic at concentrations up to around eight times the current value. Increased concentrations thus have a progressively smaller warming effect.[31] However, the first-order approximation is inaccurate at higher concentrations and there is no saturation in the absorption of infrared radiation by CO
2
.[32]

Other trace gases

Somewhat different formulae apply for other trace greenhouse gases such as methane and N
2
O
(square-root dependence) or CFCs (linear), with coefficients that may be found for example in the IPCC reports.[33] A year 2016 study suggests a significant revision to the methane IPCC formula.[34] Forcings by the most influential trace gases in Earth's atmosphere are included in the section describing recent growth trends, and in the IPCC list of greenhouse gases.

Water vapor

Water vapor is Earth's primary greenhouse gas currently responsible for about half of all atmospheric gas forcing. Its overall atmospheric concentration depends almost entirely on the average planetary temperature, and has the potential to increase by as much as 7% with every degree (°C) of temperature rise (see also: Clausius–Clapeyron relation).[35] Thus over long time scales, water vapor behaves as a system feedback that amplifies the radiative forcing driven by the growth of carbon dioxide and other trace gases.[36]

Recent growth trends

Radiative forcing (warming influence) of long-lived atmospheric greenhouse gases has nearly doubled since 1979.[37]
The industrial era growth in CO2-equivalent gas concentration and AGGI since year 1750.[38]
The growth in overall gas forcing has held steady near 2% since 1979.[39]

Radiative forcing can be a useful way to compare the growing warming influence of different anthropogenic greenhouse gases over time. The table and figures below (derived by researchers at NOAA from atmospheric radiative transfer models) show changes since year 1979 in the radiative forcing of the long-lived and well-mixed greenhouse gases that have been increasing in earth's atmosphere since the industrial revolution.[39] The table includes the direct forcing contributions from carbon dioxide (CO
2
), methane (CH
4
), nitrous oxide (N
2
O
); chlorofluorocarbons (CFCs) 12 and 11; and fifteen other halogenated gases.[40] These data do not include the significant forcing contributions from shorter-lived and less-well-mixed gases or aerosols; including those indirect forcings from the decay of methane and some halogens. They also do not account for changes in land or solar activity.

Global radiative forcing (relative to 1750, in ), CO
2
-equivalent
mixing ratio, and the Annual Greenhouse Gas Index (AGGI) since 1979[39]
Year CO
2
CH
4
N
2
O
CFC-12 CFC-11 15-minor Total CO
2
-eq
ppm
AGGI
1990 = 1
AGGI
% change
1979 1.027 0.406 0.104 0.092 0.039 0.031 1.699 382 0.786
1980 1.058 0.413 0.104 0.097 0.042 0.034 1.748 385 0.808 2.8
1981 1.077 0.420 0.107 0.102 0.044 0.036 1.786 388 0.826 2.2
1982 1.089 0.426 0.111 0.108 0.046 0.038 1.818 391 0.841 1.8
1983 1.115 0.429 0.113 0.113 0.048 0.041 1.859 394 0.860 2.2
1984 1.140 0.432 0.116 0.118 0.050 0.044 1.900 397 0.878 2.2
1985 1.162 0.437 0.118 0.123 0.053 0.047 1.940 399 0.897 2.1
1986 1.184 0.442 0.122 0.129 0.056 0.049 1.982 403 0.916 2.2
1987 1.211 0.447 0.120 0.135 0.059 0.053 2.025 406 0.936 2.2
1988 1.250 0.451 0.123 0.143 0.062 0.057 2.085 410 0.964 3.0
1989 1.274 0.455 0.126 0.149 0.064 0.061 2.130 414 0.984 2.1
1990 1.293 0.459 0.129 0.154 0.065 0.065 2.165 417 1.000 1.6
1991 1.313 0.463 0.131 0.158 0.067 0.069 2.199 419 1.016 1.6
1992 1.324 0.467 0.133 0.162 0.067 0.072 2.224 421 1.027 1.1
1993 1.334 0.467 0.134 0.164 0.068 0.074 2.239 422 1.034 0.7
1994 1.356 0.470 0.134 0.166 0.068 0.075 2.269 425 1.048 1.4
1995 1.383 0.472 0.136 0.168 0.067 0.077 2.303 428 1.064 1.6
1996 1.410 0.473 0.139 0.169 0.067 0.078 2.336 430 1.079 1.5
1997 1.426 0.474 0.142 0.171 0.067 0.079 2.357 432 1.089 1.0
1998 1.465 0.478 0.145 0.172 0.067 0.080 2.404 436 1.111 2.2
1999 1.495 0.481 0.148 0.173 0.066 0.082 2.443 439 1.129 1.8
2000 1.513 0.481 0.151 0.173 0.066 0.083 2.455 441 1.139 1.1
2001 1.535 0.480 0.153 0.174 0.065 0.085 2.492 443 1.151 1.2
2002 1.564 0.481 0.156 0.174 0.065 0.087 2.525 446 1.167 1.5
2003 1.601 0.483 0.158 0.174 0.064 0.088 2.566 449 1.186 1.9
2004 1.627 0.483 0.160 0.174 0.063 0.090 2.596 452 1.199 1.4
2005 1.655 0.482 0.162 0.173 0.063 0.092 2.626 454 1.213 1.4
2006 1.685 0.482 0.165 0.173 0.062 0.095 2.661 457 1.230 1.6
2007 1.710 0.484 0.167 0.172 0.062 0.097 2.692 460 1.244 1.4
2008 1.739 0.486 0.170 0.171 0.061 0.100 2.728 463 1.260 1.7
2009 1.760 0.489 0.172 0.171 0.061 0.103 2.755 465 1.273 1.2
2010 1.791 0.491 0.174 0.170 0.060 0.106 2.792 469 1.290 1.7
2011 1.818 0.492 0.178 0.169 0.060 0.109 2.824 471 1.305 1.5
2012 1.846 0.494 0.181 0.168 0.059 0.111 2.858 474 1.320 1.5
2013 1.884 0.496 0.184 0.167 0.059 0.114 2.901 478 1.340 2.0
2014 1.909 0.499 0.187 0.166 0.058 0.116 2.935 481 1.356 1.6
2015 1.938 0.504 0.190 0.165 0.058 0.118 2.974 485 1.374 1.8
2016 1.985 0.507 0.193 0.164 0.057 0.122 3.028 490 1.399 2.5
2017 2.013 0.509 0.195 0.163 0.057 0.124 3.062 493 1.374 1.6
2018 2.044 0.512 0.199 0.162 0.057 0.127 3.101 496 1.433 1.8
2019 2.076 0.516 0.202 0.161 0.057 0.129 3.140 500 1.451 1.8

These data show that CO
2
dominates the total forcing, with methane and chlorofluorocarbons (CFC) becoming relatively smaller contributors to the total forcing over time.[39] The five major greenhouse gases account for about 96% of the direct radiative forcing by long-lived greenhouse gas increases since 1750. The remaining 4% is contributed by the 15 minor halogenated gases.

It might be observed that the total forcing for year 2016, 3.027 W m−2, together with the commonly accepted value of climate sensitivity parameter λ, 0.8 K /(W m−2), results in an increase in global temperature of 2.4 K, much greater than the observed increase, about 1.2 K.[41] Part of this difference is due to lag in the global temperature achieving steady state with the forcing. The remainder of the difference is due to negative aerosol forcing[42][circular reference], climate sensitivity being less than the commonly accepted value, or some combination thereof.[43]

The table also includes an "Annual Greenhouse Gas Index" (AGGI), which is defined as the ratio of the total direct radiative forcing due to long-lived greenhouse gases for any year for which adequate global measurements exist to that which was present in 1990.[39] 1990 was chosen because it is the baseline year for the Kyoto Protocol. This index is a measure of the inter-annual changes in conditions that affect carbon dioxide emission and uptake, methane and nitrous oxide sources and sinks, the decline in the atmospheric abundance of ozone-depleting chemicals related to the Montreal Protocol. and the increase in their substitutes (hydrogenated CFCs (HCFCs) and hydrofluorocarbons (HFC). Most of this increase is related to CO
2
. For 2013, the AGGI was 1.34 (representing an increase in total direct radiative forcing of 34% since 1990). The increase in CO
2
forcing alone since 1990 was about 46%. The decline in CFCs considerably tempered the increase in net radiative forcing.

An alternative table prepared for use in climate model intercomparisons conducted under the auspices of IPCC and including all forcings, not just those of greenhouse gases.[44]

Direct observation

Earth's global radiation balance fluctuates as the planet rotates and orbits the Sun, and as global-scale thermal anomalies arise and dissipate within the terrestrial, oceanic and atmospheric systems (e.g. ENSO).[45] Consequently, the planet's 'instantaneous radiative forcing' (IRF) is also dynamic and naturally fluctuates between states of overall warming and cooling. The combination of periodic and complex processes that give rise to these natural variations will typically revert over periods lasting as long as a few years to produce a net-zero average IRF. Such fluctuations also mask the longer-term (decade-long) forcing trends due to human activities, and thus make direct observation of such trends challenging.[46]

NASA Earth Science Division Operating Missions[47]

Earth's radiation balance has been continuously monitored by NASA's Clouds and the Earth's Radiant Energy System (CERES) instruments since year 1998.[48][49] Each scan of the globe provides an estimate of the total (all-sky) instantaneous radiation balance. This data record captures both the natural fluctuations and human influences on IRF; including changes in greenhouse gases, aerosols, land surface, etc. The record also includes the lagging radiative responses to the radiative imbalances; occurring mainly by way of Earth system feedbacks in temperature, surface albedo, atmospheric water vapor and clouds.[50][51]

Researchers have used measurements from CERES, AIRS, CloudSat and other satellite-based instruments within NASA's Earth Observing System to parse out contributions by the natural fluctuations and system feedbacks. Removing these contributions within the multi-year data record allows observation of the anthropogenic trend in top-of-atmosphere (TOA) IRF. The data analysis has also been done in a way that is computationally efficient and independent of most related modelling methods and results. Radiative forcing was thus directly observed to have risen by +0.53 W m−2 (+/-0.11 W m−2) from years 2003 to 2018. About 20% of the increase was associated with a reduction in the atmospheric aerosol burden, and most of the remaining 80% was attributed to the rising burden of greenhouse gases.[46][52][53]

A rising trend in the radiative imbalance due to increasing global CO
2
has been previously observed by ground-based instruments. For example, such measurements have been separately gathered under clear-sky conditions at two Atmospheric Radiation Measurement (ARM) sites in Oklahoma and Alaska.[54] Each direct observation found that the associated radiative (infrared) heating experienced by surface dwellers rose by +0.2 W m−2 (+/-0.07 W m−2) during the decade ending 2010.[55][56] In addition to its focus on longwave radiation and the most influential forcing gas (CO
2
) only, this result is proportionally less than the TOA forcing due to its buffering by atmospheric absorption.

Climate sensitivity

CO2, temperature, and sunspot activity since 1850.

Radiative forcing can be used to estimate a subsequent change in steady-state (often denoted "equilibrium") surface temperature (ΔTs) arising from that forcing via the equation:

where λ is commonly denoted the climate sensitivity parameter, usually with units K/(W/m2), and ΔF is the radiative forcing in W/m2.[57] A typical value of λ, 0.8 K/(W/m2), gives an increase in global temperature of about 1.6 K above the 1750 reference temperature due to the increase in CO
2
over that time (278 to 405 ppm, for a forcing of 2.0 W/m2), and predicts a further warming of 1.4 K above present temperatures if the CO
2
mixing ratio in the atmosphere were to become double its pre-industrial value; both of these calculations assume no other forcings.[58]

Historically, radiative forcing displays the best predictive capacity for specific types of forcing such as greenhouse gases.[59] It is less effective for other anthropogenic influences like soot. A new framework called ‘effective radiative forcing’ or ERF removes the effect of rapid adjustments within the atmosphere that are unrelated to longer term surface temperature responses.[59] ERF means different factors driving climate change can be placed onto a level playing field to enable comparison of their effects and a more consistent view of how global surface temperature responds to various types of human forcing.[59]

Related metrics

Other metrics can be constructed for the same purpose as radiative forcing. For example Shine et al.[60] say "... recent experiments indicate that for changes in absorbing aerosols and ozone, the predictive ability of radiative forcing is much worse ... we propose an alternative, the 'adjusted troposphere and stratosphere forcing'. We present GCM calculations showing that it is a significantly more reliable predictor of this GCM's surface temperature change than radiative forcing. It is a candidate to supplement radiative forcing as a metric for comparing different mechanisms ...". In this quote, GCM stands for "global circulation model", and the word "predictive" does not refer to the ability of GCMs to forecast climate change. Instead, it refers to the ability of the alternative tool proposed by the authors to help explain the system response.

Therefore, the concept of radiative forcing has been evolving from the initial proposal, named nowadays instantaneous radiative forcing (IRF), to other proposals that aims to relate better the radiative imbalance with global warming (global surface mean temperature). In this sense the adjusted radiative forcing, in its different calculation methodologies, estimates the imbalance once the stratosphere temperatures has been modified to achieve a radiative equilibrium in the stratosphere (in the sense of zero radiative heating rates). This new methodology is not estimating any adjustment or feedback that could be produced on the troposphere (in addition to stratospheric temperature adjustments), for that goal another definition, named effective radiative forcing has been introduced.[61] In general the ERF is the recommendation of the CMIP6 radiative forcing analysis [62] although the stratospherically adjusted methodologies are still being applied in those cases where the adjustments and feedbacks on the troposphere are considered not critical, like in the well mixed greenhouse gases and ozone.[63][64] A methodology named radiative kernel approach allows to estimate the climate feedbacks within an offline calculation based on a linear approximation [65]

See also

References

  1. ^ Shindell, Drew (2013). "Radiative Forcing in the AR5" (PDF). Retrieved 15 September 2016.
  2. ^ Rebecca, Lindsey (14 January 2009). "Climate and Earth's Energy Budget : Feature Articles". earthobservatory.nasa.gov. Retrieved 3 April 2018.
  3. ^ Lissauer, Jack Jonathan. (2013-09-16). Fundamental planetary science : physics, chemistry, and habitability. De Pater, Imke, 1952-. New York, NY, USA. p. 90. ISBN 9780521853309. OCLC 808009225.
  4. ^ "NASA: Climate Forcings and Global Warming". 14 January 2009.
  5. ^ "Climate Change 2007: Synthesis Report" (PDF). ipcc.ch. Retrieved 3 April 2018.
  6. ^ Rockström, Johan; Steffen, Will; Noone, Kevin; Persson, Asa; Chapin, F. Stuart; Lambin, Eric F.; Lenton, Timothy F.; Scheffer, M; et al. (23 September 2009). "A safe operating space for humanity". Nature. 461 (7263): 472–475. Bibcode:2009Natur.461..472R. doi:10.1038/461472a. PMID 19779433. S2CID 205049746.
  7. ^ "The Study of Earth as an Integrated System". NASA. Retrieved 2021-05-20.
  8. ^ National Research Council (2005). Radiative Forcing of Climate Change: Expanding the Concept and Addressing Uncertainties. The National Academic Press. doi:10.17226/11175.
  9. ^ Gregg Kopp and Judith L. Lean (2011-01-14). "A new, lower value of total solar irradiance: Evidence and climate significance". Geophysical Research Letters. 38 (1). doi:10.1029/2010GL045777.
  10. ^ "Solar Radiation and Climate Experiment". University of Colorado, Laboratory for Atmospheric and Space Physcis. Retrieved 2021-05-15.
  11. ^ "TSIS-1 Mission Overview". NASA. Retrieved 2021-05-20.
  12. ^ a b Gregg Kopp (2014-04-24). "Solar variability, solar forcing, and coupling mechanisms in the terrestrial atmosphere". Journal of Space Weather and Space Climate. pp. 1–9. doi:10.1051/swsc/2014012.
  13. ^ a b Sophie Lewis (2021-01-02). "Earth reaches perihelion, closer to the sun than any other day". CBS News.
  14. ^ "The Seasons, the Equinox, and the Solstices". National Weather Service. Retrieved 2021-05-20.
  15. ^ a b Claus Fröhlich & Judith Lean (2004-12-01). "Solar radiative output and its variability: evidence and mechanisms". The Astonomy and Astrophysics Review (12): 273–320. doi:10.1007/s00159-004-0024-1.
  16. ^ David H. Hathaway (2015-09-21). "The Solar Cycle" (PDF). Living Reviews in Solar Physics (12): 1–87. doi:10.1007/lrsp-2015-4. ISSN 1614-4961.
  17. ^ Lean, Judith; Rind, David (1999-01-01). "Evaluating sun–climate relationships since the Little Ice Age". Journal of Atmospheric and Solar-Terrestrial Physics. 61 (1–2): 25–36. Bibcode:1999JASTP..61...25L. doi:10.1016/S1364-6826(98)00113-8. ISSN 1364-6826.
  18. ^ a b c Gareth S. Jones, Mike Lockwood, Peter A. Stott (2012-03-16). "What influence will future solar activity changes over the 21st century have on projected global near-surface temperature changes?". Journal of Geophysical Research - Atmospheres. 117 (D5): 1–13.CS1 maint: multiple names: authors list (link)
  19. ^ a b Alan Buis (2020-02-27). "Milankovitch (Orbital) Cycles and Their Role in Earth's Climate". NASA Jet Propulsion Laboratory.
  20. ^ a b Marie-France Loutre, Didier Paillard, Françoise Vimeux, Elsa Cortijo (2004-04-30). "Does mean annual insolation have the potential to change the climate?". Earth and Planetary Science Letters. 221 (1–4): 1–14. doi:10.1016/S0012-821X(04)00108-6.CS1 maint: multiple names: authors list (link)
  21. ^ J. Laskar (1989-03-16). "A numerical experiment on the chaotic behaviour of the Solar System". Nature (338): 237–238. doi:10.1038/338237a0.
  22. ^ a b "NASA Solar System Exploration - Our Sun". NASA. Retrieved 2021-05-15.
  23. ^ "There Is No Impending 'Mini Ice Age'". NASA Global Climate Change. 2020-02-13.
  24. ^ "What Is the Sun's Role in Climate Change?". NASA. 2019-09-06.
  25. ^ Bida Jian, Jiming Li, Guoyin Wang, Yongli He, Ying Han, Min Zhang, and Jianping Huang (2018-11-01). "The Impacts of Atmospheric and Surface Parameters on Long-Term Variations in the Planetary Albedo". Journal of Climate. 31 (21): 8705–8718. doi:10.1175/JCLI-D-17-0848.1.CS1 maint: multiple names: authors list (link)
  26. ^ a b c d e f Graeme L. Stephens, Denis O'Brien, Peter J. Webster, Peter Pilewski, Seiji Kato, Jui‐lin Li (2015-01-25). "The albedo of Earth". Reviews of Geophysics. 53 (1): 141–163. doi:10.1002/2014RG000449.CS1 maint: multiple names: authors list (link)
  27. ^ a b "Measuring Earth's Albedo". NASA Earth Observatory. Retrieved 2021-05-15.
  28. ^ "Landsat Science Team's Crystal Schaaf Discusses Albedo, Its Importance, and How It Can Affect Climate". U.S. Geological Survey. 2021-01-12.
  29. ^ Robock, Alan (2000-05-01). "Volcanic eruptions and climate". Reviews of Geophysics. 38 (2): 191–219. Bibcode:2000RvGeo..38..191R. doi:10.1029/1998RG000054.
  30. ^ Myhre, G.; Highwood, E.J.; Shine, K.P.; Stordal, F. (1998). "New estimates of radiative forcing due to well mixed greenhouse gases" (PDF). Geophysical Research Letters. 25 (14): 2715–8. Bibcode:1998GeoRL..25.2715M. doi:10.1029/98GL01908.
  31. ^ Huang, Yi; Bani Shahabadi, Maziar (28 November 2014). "Why logarithmic?". J. Geophys. Res. Atmospheres. 119 (24): 13, 683–89. Bibcode:2014JGRD..11913683H. doi:10.1002/2014JD022466.
  32. ^ Zhong, Wenyi; Haigh, Joanna D. (27 March 2013). "The greenhouse effect and carbon dioxide". Weather. 68 (4): 100–5. Bibcode:2013Wthr...68..100Z. doi:10.1002/wea.2072. ISSN 1477-8696.
  33. ^ IPCC WG-1 Archived 13 December 2007 at the Wayback Machine report
  34. ^ Etminan, M.; Myhre, G.; Highwood, E. J.; Shine, K. P. (2016-12-27). "Radiative forcing of carbon dioxide, methane, and nitrous oxide: A significant revision of the methane radiative forcing". Geophysical Research Letters. 43 (24): 12, 614–12, 623. Bibcode:2016GeoRL..4312614E. doi:10.1002/2016gl071930. ISSN 0094-8276.
  35. ^ Gavin Schmidt (2010-10-01). "Taking the Measure of the Greenhouse Effect". NASA Goddard Institute for Space Studies - Science Briefs.
  36. ^ "It's Water Vapor, Not the CO2". American Chemical Society. Retrieved 2021-05-20.
  37. ^ "Climate Change Indicators: Climate Forcing". EPA.gov. United States Environmental Protection Agency. 2021. Archived from the original on May 9, 2021.
    ● EPA credits data from "NOAA's Annual Greenhouse Gas Index (An Introduction)". NOAA.gov. National Oceanographic and Atmospheric Administration (Global Monitoring Laboratory, Earth System Research Laboratories). December 2020. Archived from the original on May 13, 2021.
  38. ^ "The NOAA Annual Greenhouse Gas Index - Figure 5". NOAA. 2020.
  39. ^ a b c d e  This article incorporates public domain material from the NOAA document: "THE NOAA ANNUAL GREENHOUSE GAS INDEX (AGGI)". NOAA/ESRL. 1 August 2013. Cite journal requires |journal=
  40. ^ CFC-113, tetrachloromethane (CCl
    4
    ), 1,1,1-trichloroethane (CH
    3
    CCl
    3
    ); hydrochlorofluorocarbons (HCFCs) 22, 141b and 142b; hydrofluorocarbons (HFCs) 134a, 152a, 23, 143a, and 125; sulfur hexafluoride (SF
    6
    ), and halons 1211, 1301 and 2402)
  41. ^ Hansen, J.E.; et al. "GISS Surface Temperature Analysis: Analysis Graphs and Plots". Goddard Institute for Space Studies, National Aeronautics and Space Administration.
  42. ^ Particulates#Climate effects
  43. ^ Schwartz, Stephen E.; Charlson, Robert J.; Kahn, Ralph A.; Ogren, John A.; Rodhe, Henning (2010). "Why hasn't Earth warmed as much as expected?" (PDF). Journal of Climate (published 15 May 2010). 23 (10): 2453–64. Bibcode:2010JCli...23.2453S. doi:10.1175/2009JCLI3461.1.
  44. ^ editor., Stocker, Thomas (24 March 2014). Climate change 2013 : the physical science basis : Working Group I contribution to the Fifth assessment report of the Intergovernmental Panel on Climate Change. ISBN 978-1-107-66182-0. OCLC 1120509660.CS1 maint: extra text: authors list (link) datafile
  45. ^ Rebecca, Lindsey (14 January 2009). "Climate and Earth's Energy Budget". earthobservatory.nasa.gov. Retrieved 15 April 2021.
  46. ^ a b Kramer, R.J., H. He, B.J. Soden, L. Oreopoulos, G. Myhre, P.M. Forster, and C.J. Smith (2021-03-25). "Observational Evidence of Increasing Global Radiative Forcing". Geophysical Research Letters. 48 (7). doi:10.1029/2020GL091585.CS1 maint: multiple names: authors list (link)
  47. ^ "NASA's Earth Observing System homepage". NASA EOS Project Science Office. Retrieved 2021-04-16.
  48. ^ Loeb, N.G., S. Kato, K. Loukachine, and N. Manalo-Smith (2005-04-01). "Angular Distribution Models for Top-of-Atmosphere Radiative Flux Estimation from the Clouds and the Earth's Radiant Energy System Instrument on the Terra Satellite. Part I: Methodology". Journal of Atmospheric and Oceanic Technology. 22 (4): 338–351. Bibcode:2005JAtOT..22..338L. doi:10.1175/JTECH1712.1.CS1 maint: multiple names: authors list (link)
  49. ^ Loeb, N.G., F.G. Rose, S. Kato, D.A. Rutan, W. Su, H. Wang, D.R. Doelling, W.L. Smith, and A. Gettelman (2020-01-01). "Toward a Consistent Definition between Satellite and Model Clear-Sky Radiative Fluxes". Journal of Climate. 33 (1): 61–75. Bibcode:2020JCli...33...61L. doi:10.1175/JCLI-D-19-0381.1.CS1 maint: multiple names: authors list (link)
  50. ^ Sherwood, S.C., S. Bony, O. Boucher, C. Bretherton, P.M. Forster, J.M. Gregory, and B. Stevens (2015-02-01). "Adjustments in the Forcing-Feedback Framework for Understanding Climate Change". Bulletin of the American Meteorological Society. 96 (2): 217–228. Bibcode:2015BAMS...96..217S. doi:10.1175/BAMS-D-13-00167.1.CS1 maint: multiple names: authors list (link)
  51. ^ Wielicki, B.A., R.D. Cess, M.D. King, D.A. Randall, and E.F. Harrison (1995-11-01). "Mission to Planet Earth: Role of Clouds and Radiation in Climate". Bulletin of the American Meteorological Society. 76 (11): 2125–2154. Bibcode:1995BAMS...76.2125W. doi:10.1175/1520-0477(1995)076<2125:MTPERO>2.0.CO;2.CS1 maint: multiple names: authors list (link)
  52. ^ Sarah Hansen (12 April 2021). "UMBC's Ryan Kramer confirms human-caused climate change with direct evidence for first time". University of Maryland, Baltimore County.
  53. ^ "Direct observations confirm that humans are throwing Earth's energy budget off balance". phys.org. 26 March 2021.
  54. ^ "ARM Capabilities - Atmospheric Observatories". U.S. Department of Energy - Office of Science. Retrieved 2021-04-25.
  55. ^ Feldman, D.R., W.D. Collins, P.J. Gero, M.S. Torn, E.J. Mlawer, and T.R. Shippert (2015-02-25). "Observational determination of surface radiative forcing by CO2 from 2000 to 2010". Nature. pp. 339–343. doi:10.1038/nature14240.CS1 maint: multiple names: authors list (link)
  56. ^ Robert McSweeney (2015-02-25). "New study directly measures greenhouse effect at Earth's surface". Carbon Brief.
  57. ^ "IPCC Third Assessment Report - Climate Change 2001". Archived from the original on 30 June 2009.
  58. ^ "Atmosphere Changes". Archived from the original on 10 May 2009.
  59. ^ a b c Nauels, A.; Rosen, D.; Mauritsen, T.; Maycock, A.; McKenna, C.; Rogelj, J.; Schleussner, C.-F.; Smith, E.; Smith, C. (2019-12-02). "ZERO IN ON the remaining carbon budget and decadal warming rates. The CONSTRAIN Project Annual Report 2019". constrain-eu.org. doi:10.5518/100/20. Retrieved 2020-01-20.
  60. ^ Shine, Keith P.; Cook, Jolene; Highwood, Eleanor J.; Joshi, Manoj M. (23 October 2003). "An alternative to radiative forcing for estimating the relative importance of climate change mechanisms". Geophysical Research Letters. 30 (20): 2047. Bibcode:2003GeoRL..30.2047S. doi:10.1029/2003GL018141. S2CID 59514371.
  61. ^ Sherwood, Steven C.; Bony, Sandrine; Boucher, Olivier; Bretherton, Chris; Forster, Piers M.; Gregory, Jonathan M.; Stevens, Bjorn (2015-02-01). "Adjustments in the Forcing-Feedback Framework for Understanding Climate Change" (PDF). Bulletin of the American Meteorological Society. 96 (2): 217–228. Bibcode:2015BAMS...96..217S. doi:10.1175/bams-d-13-00167.1. ISSN 0003-0007.
  62. ^ Forster, Piers M.; Richardson, Thomas; Maycock, Amanda C.; Smith, Christopher J.; Samset, Bjorn H.; Myhre, Gunnar; Andrews, Timothy; Pincus, Robert; Schulz, Michael (2016-10-27). "Recommendations for diagnosing effective radiative forcing from climate models for CMIP6" (PDF). Journal of Geophysical Research: Atmospheres. 121 (20): 12, 460–12, 475. Bibcode:2016JGRD..12112460F. doi:10.1002/2016jd025320. ISSN 2169-897X.
  63. ^ Stevenson, D. S.; Young, P. J.; Naik, V.; Lamarque, J.-F.; Shindell, D. T.; Voulgarakis, A.; Skeie, R. B.; Dalsoren, S. B.; Myhre, G. (2013-03-15). "Tropospheric ozone changes, radiative forcing and attribution to emissions in the Atmospheric Chemistry and Climate Model Intercomparison Project (ACCMIP)" (PDF). Atmospheric Chemistry and Physics. 13 (6): 3063–3085. Bibcode:2013ACP....13.3063S. doi:10.5194/acp-13-3063-2013. ISSN 1680-7316.
  64. ^ Checa-Garcia, Ramiro; Hegglin, Michaela I.; Kinnison, Douglas; Plummer, David A.; Shine, Keith P. (2018-04-06). "Historical Tropospheric and Stratospheric Ozone Radiative Forcing Using the CMIP6 Database" (PDF). Geophysical Research Letters. 45 (7): 3264–3273. Bibcode:2018GeoRL..45.3264C. doi:10.1002/2017gl076770. ISSN 0094-8276.
  65. ^ Soden, Brian J.; Held, Isaac M.; Colman, Robert; Shell, Karen M.; Kiehl, Jeffrey T.; Shields, Christine A. (2008-07-01). "Quantifying Climate Feedbacks Using Radiative Kernels". Journal of Climate. 21 (14): 3504–3520. Bibcode:2008JCli...21.3504S. CiteSeerX 10.1.1.141.653. doi:10.1175/2007jcli2110.1. ISSN 0894-8755.

External links


This page was last updated at 2021-06-07 06:40 UTC. Update now. View original page.

All our content comes from Wikipedia and under the Creative Commons Attribution-ShareAlike License.


Top

If mathematical, chemical, physical and other formulas are not displayed correctly on this page, please useFirefox or Safari