Cloud feedback

Source: Wikipedia, the free encyclopedia.
Attribution of individual atmospheric component contributions to the greenhouse effect, separated into feedback and forcing categories (NASA)

Cloud feedback is a type of

climate change feedback that has been difficult to quantify in contemporary climate models. It can affect the magnitude of internally generated climate variability[1][2] or they can affect the magnitude of climate change resulting from external radiative forcings.[3] Cloud representations vary among global climate models, and small changes in cloud cover have a large impact on the climate.[4][5]

Global warming is expected to change the distribution and type of clouds.

optical thickness actually increases with increasing temperature.[10]
Whether the net effect is warming or cooling depends on details such as the type and altitude of the cloud; details that are difficult to represent in climate models.

The closely related effective climate sensitivity has increased substantially in the latest generation of global climate models. Differences in the physical representation of clouds in models drive this enhanced sensitivity relative to the previous generation of models.[11][12][13]

Mechanisms

Examples of some effects of global warming that can amplify (positive feedbacks) or reduce (negative feedbacks) global warming[14][15] Observations and modeling studies indicate that there is a net positive feedback to Earth's current global warming.[16]

Global warming is expected to change the distribution and type of clouds. Seen from below, clouds emit infrared radiation back to the surface, and so exert a warming effect; seen from above, clouds reflect sunlight and emit infrared radiation to space, and so exert a cooling effect. Whether the net effect is warming or cooling depends on details such as the type and altitude of the cloud. Low clouds are brighter and optically thicker, while high clouds are optically thin (transparent) in the visible and trap IR. Reduction of low clouds tends to increase incoming solar radiation and therefore have a positive feedback, while a reduction in high clouds (since they mostly just trap IR) would result in a negative feedback. These details were poorly observed before the advent of satellite data and are difficult to represent in climate models.[17] Global climate models were showing a near-zero to moderately strong positive net cloud feedback, but the effective climate sensitivity has increased substantially in the latest generation of global climate models. Differences in the physical representation of clouds in models drive this enhanced climate sensitivity relative to the previous generation of models.[18][19][20]

A 2019 simulation predicts that if greenhouse gases reach three times the current level of atmospheric carbon dioxide that stratocumulus clouds could abruptly disperse, contributing to additional global warming.[21][22]

Aerosols

The extent to which physical factors in the atmosphere or on land affect climate change, including the cooling provided by sulfate aerosols and the dimming they cause. The large error bar shows that there are still substantial unresolved uncertainties.

Atmospheric aerosols—fine partices suspended in the air—affect cloud formation and properties, which also alters their impact on climate. While some aerosols, such as black carbon particles, make the clouds darker and thus contribute to warming,[23] by far the strongest effect is from sulfates, which increase the number of cloud droplets, making the clouds more reflective, and helping them cool the climate more. That is known as a direct aerosol effect; however, aerosols also have an indirect effect on liquid water path, and determing it involves computationally heavy continuous calculations of evaporation and condensation within clouds. Climate models generally assume that aerosols increase liquid water path, which makes the clouds even more reflective.[24] However, satellite observations taken in 2010s suggested that aerosols decreased liquid water path instead, and in 2018, this was reproduced in a model which integrated more complex cloud microphysics.[25] Yet, 2019 research found that earlier satellite observations were biased by failing to account for the thickest, most water-heavy clouds naturally raining more and shedding more particulates: very strong aerosol cooling was seen when comparing clouds of the same thickness.[26]

Moreover, large-scale observations can be confounded by changes in other atmospheric factors, like humidity: i.e. it was found that while post-1980 improvements in air quality would have reduced the number of clouds over the

AMOC slowdown.[27] Similarly, while the initial research looking at sulfates from the 2014–2015 eruption of Bárðarbunga found that they caused no change in liquid water path,[28] it was later suggested that this finding was confounded by counteracting changes in humidity.[27]

Visible ship tracks in the Northern Pacific, on 4 March 2009

To avoid confounders, many observations of aerosol effects focus on ship tracks, but post-2020 research found that visible ship tracks are a poor proxy for other clouds, and estimates derived from them overestimate aerosol cooling by as much as 200%.[29] At the same time, other research found that the majority of ship tracks are "invisible" to satellites, meaning that the earlier research had underestimated aerosol cooling by overlooking them.[30] Finally, 2023 research indicates that all climate models have underestimated sulfur emissions from volcanoes which occur in the background, outside of major eruptions, and so had consequently overestimated the cooling provided by anthropogenic aerosols, especially in the Arctic climate.[31]

Early 2010s estimates of past and future anthropogenic global sulfur dioxide emissions, including the Representative Concentration Pathways. While no climate change scenario may reach Maximum Feasible Reductions (MFRs), all assume steep declines from today's levels. By 2019, sulfate emission reductions were confirmed to proceed at a very fast rate.[32]

Estimates of how much aerosols affect cloud cooling are very important, because the amount of sulfate aerosols in the air had undergone dramatic changes in the recent decades. First, it had increased greatly from 1950s to 1980s, largely due to the widespread burning of sulfur-heavy coal, which caused an observable reduction in visible sunlight that had been described as global dimming.[33][34] Then, it started to decline substantially from the 1990s onwards and is expected to continue to decline in the future, due to the measures to combat acid rain and other impacts of air pollution.[35] Consequently, the aerosols provided a considerable cooling effect which counteracted or "masked" some of the greenhouse effect from human emissions, and this effect had been declining as well, which contributed to acceleration of climate change.[36] Climate models do account for the presence of aerosols and their recent and future decline in their projections, and typically estimate that the cooling they provide in 2020s is similar to the warming from human-added atmospheric methane, meaning that simultaneous reductions in both would effectively cancel each other out.[37] However, the existing uncertainty about aerosol-cloud interactions likewise introduces uncertainty into models, particularly when concerning predictions of changes in weather events over the regions with a poorer historical record of atmospheric observations.[38][34][39][40]

Role as contributor to climate sensitivity

Clouds, such as those in this image (viewed from space), snow, and ice have the biggest influence on how reflective Earth is. When any of these factors change, Earth’s albedo can change.

Changes in cloud cover is one of several contributors to climate change and climate sensitivity.

Radiative forcing is one component of climate change. The radiative forcing caused by a doubling of atmospheric CO2 levels (from the pre-industrial 280 ppm) is approximately 3.7

global warming. That figure is straightforward to calculate by using the Stefan–Boltzmann law[41][42] and is undisputed.[43]

A further contribution arises from
water vapour feedback, ice–albedo feedback, cloud feedback, and lapse rate feedback.[43] Suppressing feedbacks tend to counteract warming by increasing the rate at which energy is radiated to space from a warmer planet. Exacerbating feedbacks increase warming; for example, higher temperatures can cause ice to melt, which reduces the ice area and the amount of sunlight the ice reflects, which in turn results in less heat energy being radiated back into space. Climate sensitivity depends on the balance between those feedbacks.[42]

Current understanding in climate models

When the IPCC began to produce its

Equilibrium Climate Sensitivity changed from 3.2 °C to 3.7 °C and the estimates for the Transient climate response from 1.8 °C, to 2.0 °C. That is probably because of better understanding of the role of clouds and aerosols.[46]

In preparation for the 2021 IPCC Sixth Assessment Report, a new generation of climate models have been developed by scientific groups around the world.[47][48] The average estimated climate sensitivity has increased in Coupled Model Intercomparison Project Phase 6 (CMIP6) compared to the previous generation, with values spanning 1.8 to 5.6 °C (3.2 to 10.1 °F) across 27 global climate models and exceeding 4.5 °C (8.1 °F) in 10 of them.[49][50] The cause of the increased equilibrium climate sensitivity (ECS) lies mainly in improved modelling of clouds. Temperature rises are now believed to cause sharper decreases in the number of low clouds, and fewer low clouds means more sunlight is absorbed by the planet and less reflected to space.[49][51][52] Models with the highest ECS values, however, are not consistent with observed warming.[53]

A 2019 simulation predicts that if greenhouse gases reach three times the current level of atmospheric carbon dioxide that stratocumulus clouds could abruptly disperse, contributing to additional global warming.[54]

Relationship with other feedbacks

In addition to how clouds themselves will respond to increased temperatures, other feedbacks affect clouds properties and formation. The amount and vertical distribution of water vapor is closely linked to the formation of clouds. Ice crystals have been shown to largely influence the amount of water vapor.[55] Water vapor in the subtropical upper troposphere has been linked to the convection of water vapor and ice. Changes in subtropical humidity could provide a negative feedback that decreases the amount of water vapor which in turn would act to mediate global climate transitions.[56]

Changes in cloud cover are closely coupled with other feedback, including the

polar regions is expected in increase the amount of low-level clouds, whose stratification prevents the convection of moisture to upper levels. This feedback would partially cancel the increased surface warming due to the cloudiness. This negative feedback has less effect than the positive feedback. The upper atmosphere more than cancels negative feedback that causes cooling, and therefore the increase of CO2 is actually exacerbating the positive feedback as more CO2 enters the system.[58]

See also

References

  1. (PDF) from the original on 2018-07-19.
  2. .
  3. .
  4. (PDF) on 2018-07-22. Retrieved 2017-10-27.
  5. ^ Stocker, T.F.; et al. (2001). "Physical climate processes and feedbacks". In J.T. Houghton; et al. (eds.). Climate Change 2001: The Scientific Basis, Contributions of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, U.K.: Cambridge University Press.
  6. ^ GREENFIELDBOYCE, Nell (July 11, 2016). "Climate Change May Already Be Shifting Clouds Toward The Poles". NPR. Retrieved July 11, 2016.
  7. ^ "How climate change is altering Earth's cloud cover". The Christian Science Monitor. July 12, 2016. Retrieved July 14, 2016.
  8. .
  9. .
  10. .
  11. .
  12. ^ Watts, Jonathan (2020-06-13). "Climate worst-case scenarios may not go far enough, cloud data shows". the Guardian. Retrieved 2020-06-19.
  13. PMID 32457461
    .
  14. ^ "The Study of Earth as an Integrated System". nasa.gov. NASA. 2016. Archived from the original on November 2, 2016.
  15. ^ Fig. TS.17, Technical Summary, Sixth Assessment Report (AR6), Working Group I, IPCC, 2021, p. 96. Archived from the original on July 21, 2022.
  16. ^ Stocker, Thomas F.; Dahe, Qin; Plattner, Gian-Kaksper (2013). IPCC AR5 WG1. Technical Summary (PDF). Archived (PDF) from the original on 16 July 2023. See esp. TFE.6: Climate Sensitivity and Feedbacks at p. 82.
  17. . Interestingly, the true feedback is consistently weaker than the constant relative humidity value, implying a small but robust reduction in relative humidity in all models on average clouds appear to provide a positive feedback in all models
  18. .
  19. . Retrieved 2020-06-19.
  20. .
  21. .
  22. .
  23. .
  24. .
  25. .
  26. .
  27. ^ .
  28. .
  29. .
  30. .
  31. .
  32. .
  33. ^ "Aerosol pollution has caused decades of global dimming". American Geophysical Union. 18 February 2021. Archived from the original on 27 March 2023. Retrieved 18 December 2023.
  34. ^ .
  35. ^ "Air Emissions Trends – Continued Progress Through 2005". U.S. Environmental Protection Agency. 8 July 2014. Archived from the original on 2007-03-17. Retrieved 2007-03-17.
  36. .
  37. ^ Zeke Hausfather (29 April 2021). "Explainer: Will global warming 'stop' as soon as net-zero emissions are reached?". Carbon Brief. Retrieved 2023-03-23.
  38. .
  39. .
  40. .
  41. ^ The calculation is as follows. In equilibrium, the energy of incoming and outgoing radiation have to balance. The outgoing radiation is given by the Stefan–Boltzmann law: . When incoming radiation increases, the outgoing radiation and therefore temperature must increase as well. The temperature rise directly caused by the additional radiative forcing, because of the doubling of CO2 is then given by
    .
    Given an
    Stefan–Boltzmann constant
    of 5.67 W/m2 K−4 and around 4 W/m2, the equation gives a climate sensitivity of a world without feedback of approximately 1 K.
  42. ^
    S2CID 66109238
    .
  43. ^ a b Rahmstorf S (2008). "Anthropogenic Climate Change: Revisiting the Facts" (PDF). In Zedillo E (ed.). Global Warming: Looking Beyond Kyoto. Brookings Institution Press. pp. 34–53. Archived (PDF) from the original on 14 July 2019. Retrieved 14 August 2008.
  44. PMID 31776487
    .
  45. .
  46. ^ "Increased warming in latest generation of climate models likely caused by clouds: New representations of clouds are making models more sensitive to carbon dioxide". Science Daily. 24 June 2020. Archived from the original on 26 June 2020. Retrieved 26 June 2020.
  47. ISSN 1758-6798
    .
  48. ^ "New climate models suggest Paris goals may be out of reach". France 24. 2020-01-14. Archived from the original on 14 January 2020. Retrieved 2020-01-18.
  49. ^
    ISSN 1944-8007
    .
  50. ^ "International analysis narrows range of climate's sensitivity to CO2". UNSW Newsroom. 2020-07-23. Archived from the original on 23 July 2020. Retrieved 2020-07-23.
  51. PMID 32457461
    .
  52. from the original on 19 June 2020. Retrieved 2020-06-19.
  53. ^ Bender M (2020-02-07). "Climate Change Predictions Have Suddenly Gone Catastrophic. This Is Why". Vice. Archived from the original on 10 February 2020. Retrieved 2020-02-09.
  54. S2CID 134307699
    .
  55. .
  56. (PDF) from the original on 2003-06-08.
  57. .
  58. .