Carbon cycle

Source: Wikipedia, the free encyclopedia.

Fast carbon cycle showing the movement of carbon between land, atmosphere, and oceans in billions of tons (gigatons) per year. Yellow numbers are natural fluxes, red are human contributions, and white are stored carbon. The effects of the slow (or deep) carbon cycle, such as volcanic and tectonic activity are not included.[1]

The carbon cycle is that part of the biogeochemical cycle by which carbon is exchanged among the biosphere, pedosphere, geosphere, hydrosphere, and atmosphere of Earth. Other major biogeochemical cycles include the nitrogen cycle and the water cycle. Carbon is the main component of biological compounds as well as a major component of many minerals such as limestone. The carbon cycle comprises a sequence of events that are key to making Earth capable of sustaining life. It describes the movement of carbon as it is recycled and reused throughout the biosphere, as well as long-term processes of carbon sequestration (storage) to and release from carbon sinks.

To describe the dynamics of the carbon cycle, a distinction can be made between the fast and slow carbon cycle. The fast carbon cycle is also referred to as the biological carbon cycle. Fast carbon cycles can complete within years, moving substances from atmosphere to biosphere, then back to the atmosphere. Slow or geological cycles (also called deep carbon cycle) can take millions of years to complete, moving substances through the Earth's crust between rocks, soil, ocean and atmosphere.[2]

Humans have disturbed the carbon cycle for many centuries. They have done so by

global warming.[3] The increased carbon dioxide has also caused a reduction in the ocean's pH value and is fundamentally altering marine chemistry.[4]

Main compartments

The carbon cycle was first described by

carbon pools) interconnected by pathways of exchange:[6]
: 5–6 

The carbon exchanges between reservoirs occur as the result of various chemical, physical, geological, and biological processes. The ocean contains the largest active pool of carbon near the surface of the Earth.[7] The natural flows of carbon between the atmosphere, ocean, terrestrial ecosystems, and sediments are fairly balanced; so carbon levels would be roughly stable without human influence.[8][9]

Atmosphere

Computer model showing a year in the life of atmospheric carbon dioxide and how it travels around the globe [10]

Carbon in the Earth's atmosphere exists in two main forms: carbon dioxide and methane. Both of these gases absorb and retain heat in the atmosphere and are partially responsible for the greenhouse effect.[7] Methane produces a larger greenhouse effect per volume as compared to carbon dioxide, but it exists in much lower concentrations and is more short-lived than carbon dioxide. Thus, carbon dioxide contributes more to the global greenhouse effect than methane.[11]

Carbon dioxide is removed from the atmosphere primarily through photosynthesis and enters the terrestrial and oceanic biospheres. Carbon dioxide also dissolves directly from the atmosphere into bodies of water (ocean, lakes, etc.), as well as dissolving in precipitation as raindrops fall through the atmosphere. When dissolved in water, carbon dioxide reacts with water molecules and forms carbonic acid, which contributes to ocean acidity. It can then be absorbed by rocks through weathering. It also can acidify other surfaces it touches or be washed into the ocean.[12]

CO2 concentrations over the last 800,000 years as measured from ice cores (blue/green) and directly (black)

Human activities over the past two centuries have increased the amount of carbon in the atmosphere by nearly 50% as of year 2020, mainly in the form of carbon dioxide, both by modifying ecosystems' ability to extract carbon dioxide from the atmosphere and by emitting it directly, e.g., by burning fossil fuels and manufacturing concrete.[3][7]

In the far future (2 to 3 billion years), the rate at which carbon dioxide is absorbed into the soil via the carbonate–silicate cycle will likely increase due to expected changes in the sun as it ages. The expected increased luminosity of the Sun will likely speed up the rate of surface weathering.[13] This will eventually cause most of the carbon dioxide in the atmosphere to be squelched into the Earth's crust as carbonate.[14][15][16] Once the concentration of carbon dioxide in the atmosphere falls below approximately 50 parts per million (tolerances vary among species), C3 photosynthesis will no longer be possible.[15] This has been predicted to occur 600 million years from the present, though models vary.[17]

Once the oceans on the Earth evaporate in about 1.1 billion years from now,[13] plate tectonics will very likely stop due to the lack of water to lubricate them. The lack of volcanoes pumping out carbon dioxide will cause the carbon cycle to end between 1 billion and 2 billion years into the future.[18]

Terrestrial biosphere

Amount of carbon stored in Earth's various terrestrial ecosystems, in gigatonnes.[19]

The terrestrial biosphere includes the organic carbon in all land-living organisms, both alive and dead, as well as carbon stored in

heterotrophs
receive carbon by consuming other organisms.

Because carbon uptake in the terrestrial biosphere is dependent on biotic factors, it follows a diurnal and seasonal cycle. In CO2 measurements, this feature is apparent in the

hemisphere
because this hemisphere has more land mass than the southern hemisphere and thus more room for ecosystems to absorb and emit carbon.

A portable soil respiration system measuring soil CO2 flux.

Carbon leaves the terrestrial biosphere in several ways and on different time scales. The combustion or respiration of organic carbon releases it rapidly into the atmosphere. It can also be exported into the ocean through rivers or remain sequestered in soils in the form of inert carbon.[23] Carbon stored in soil can remain there for up to thousands of years before being washed into rivers by erosion or released into the atmosphere through soil respiration. Between 1989 and 2008 soil respiration increased by about 0.1% per year.[24] In 2008, the global total of CO2 released by soil respiration was roughly 98 billion tonnes[citation needed], about 3 times more carbon than humans are now putting into the atmosphere each year by burning fossil fuel (this does not represent a net transfer of carbon from soil to atmosphere, as the respiration is largely offset by inputs to soil carbon).[citation needed] There are a few plausible explanations for this trend, but the most likely explanation is that increasing temperatures have increased rates of decomposition of soil organic matter, which has increased the flow of CO2. The length of carbon sequestering in soil is dependent on local climatic conditions and thus changes in the course of climate change.[25]

Size of major carbon pools on the Earth (year 2000 estimates)[7]
Pool Quantity
(gigatons)
Atmosphere 720
Ocean (total) 38,400
Total inorganic 37,400
Total organic 1,000
Surface layer 670
Deep layer 36,730
Lithosphere
Sedimentary carbonates > 60,000,000
Kerogens 15,000,000
Terrestrial biosphere (total) 2,000
Living biomass 600 – 1,000
Dead biomass 1,200
Aquatic biosphere 1 – 2
Fossil fuels (total) 4,130
Coal 3,510
Oil 230
Gas 140
Other (peat) 250

Ocean

The ocean can be conceptually divided into a surface layer within which water makes frequent (daily to annual) contact with the atmosphere, and a deep layer below the typical mixed layer depth of a few hundred meters or less, within which the time between consecutive contacts may be centuries. The dissolved inorganic carbon (DIC) in the surface layer is exchanged rapidly with the atmosphere, maintaining equilibrium. Partly because its concentration of DIC is about 15% higher[26] but mainly due to its larger volume, the deep ocean contains far more carbon—it is the largest pool of actively cycled carbon in the world, containing 50 times more than the atmosphere[7]—but the timescale to reach equilibrium with the atmosphere is hundreds of years: the exchange of carbon between the two layers, driven by thermohaline circulation, is slow.[7]

Carbon enters the ocean mainly through the dissolution of atmospheric carbon dioxide, a small fraction of which is converted into carbonate. It can also enter the ocean through rivers as dissolved organic carbon. It is converted by organisms into organic carbon through photosynthesis and can either be exchanged throughout the food chain or precipitated into the oceans' deeper, more carbon-rich layers as dead soft tissue or in shells as calcium carbonate. It circulates in this layer for long periods of time before either being deposited as sediment or, eventually, returned to the surface waters through thermohaline circulation.[8]

Oceans are basic (with a current

carbon sequestering. The projected rate of pH reduction could slow the biological precipitation of calcium carbonates, thus decreasing the ocean's capacity to absorb CO2.[27][28]

Geosphere

Diagram showing relative sizes (in gigatonnes) of the main storage pools of carbon on Earth. Cumulative changes (thru year 2014) from land use and emissions of fossil carbon are included for comparison.[19]

The geologic component of the carbon cycle operates slowly in comparison to the other parts of the global carbon cycle. It is one of the most important determinants of the amount of carbon in the atmosphere, and thus of global temperatures.[29]

Most of the Earth's carbon is stored inertly in the Earth's lithosphere.[7] Much of the carbon stored in the Earth's mantle was stored there when the Earth formed.[30] Some of it was deposited in the form of organic carbon from the biosphere.[31] Of the carbon stored in the geosphere, about 80% is limestone and its derivatives, which form from the sedimentation of calcium carbonate stored in the shells of marine organisms. The remaining 20% is stored as kerogens formed through the sedimentation and burial of terrestrial organisms under high heat and pressure. Organic carbon stored in the geosphere can remain there for millions of years.[29]

Carbon can leave the geosphere in several ways. Carbon dioxide is released during the

fossil fuels
. After extraction, fossil fuels are burned to release energy and emit the carbon they store into the atmosphere.

Types of dynamic

The slow (or deep) carbon cycle operates through rocks
The fast carbon cycle operates through the biosphere, see diagram at start of article ↑

There is a fast and a slow carbon cycle. The fast cycle operates in the biosphere and the slow cycle operates in rocks. The fast or biological cycle can complete within years, moving carbon from atmosphere to biosphere, then back to the atmosphere. The slow or geological cycle may extend deep into the mantle and can take millions of years to complete, moving carbon through the Earth's crust between rocks, soil, ocean and atmosphere.[2]

The fast carbon cycle involves relatively short-term

biogeochemical processes between the environment and living organisms in the biosphere (see diagram at start of article). It includes movements of carbon between the atmosphere and terrestrial and marine ecosystems, as well as soils and seafloor sediments. The fast cycle includes annual cycles involving photosynthesis and decadal cycles involving vegetative growth and decomposition. The reactions of the fast carbon cycle to human activities will determine many of the more immediate impacts of climate change.[32][33][34][35][36]

The slow (or deep) carbon cycle involves medium to long-term

Mountain building processes result in the return of this geologic carbon to the Earth's surface. There the rocks are weathered and carbon is returned to the atmosphere by degassing and to the ocean by rivers. Other geologic carbon returns to the ocean through the hydrothermal emission of calcium ions. In a given year between 10 and 100 million tonnes of carbon moves around this slow cycle. This includes volcanoes returning geologic carbon directly to the atmosphere in the form of carbon dioxide. However, this is less than one percent of the carbon dioxide put into the atmosphere by burning fossil fuels.[2][32][37]

Sub-processes within fast carbon cycle

Terrestrial carbon in the water cycle

Where terrestrial carbon goes when water flows [38]

The movement of terrestrial carbon in the water cycle is shown in the diagram on the right and explained below: [38]

  1. Atmospheric particles act as cloud condensation nuclei, promoting cloud formation.[39][40]
  2. Raindrops absorb
    inorganic carbon through particle scavenging and adsorption of organic vapors while falling toward Earth.[41][42]
  3. Burning and volcanic eruptions produce highly condensed polycyclic aromatic molecules (i.e. black carbon) that is returned to the atmosphere along with greenhouse gases such as CO2.[43][44]
  4. Terrestrial plants fix atmospheric CO2 through photosynthesis, returning a fraction back to the atmosphere through respiration.[45] Lignin and celluloses represent as much as 80% of the organic carbon in forests and 60% in pastures.[46][47]
  5. Litterfall and root organic carbon mix with sedimentary material to form organic soils where plant-derived and petrogenic organic carbon is both stored and transformed by microbial and fungal activity.[48][49][50]
  6. Water absorbs plant and settled aerosol-derived
    overland flow occurs when soils are completely saturated,[54] or rainfall occurs more rapidly than saturation into soils.[55]
  7. Organic carbon derived from the terrestrial biosphere and in situ
    photo-oxidation), resulting in a flux of CO2 from rivers to the atmosphere that are the same order of magnitude as the amount of carbon sequestered annually by the terrestrial biosphere.[56][57][58] Terrestrially-derived macromolecules such as lignin [59] and black carbon[60] are decomposed into smaller components and monomers, ultimately being converted to CO2, metabolic intermediates, or biomass
    .
  8. Lakes, reservoirs, and
    heterotrophy in the water column, resulting in a net flux of CO2 to the atmosphere that is roughly one order of magnitude less than rivers.[61][58] Methane production is also typically high in the anoxic sediments of floodplains, lakes, and reservoirs.[62]
  9. Primary production is typically enhanced in
    estuarine waters are a source of CO2 to the atmosphere, globally.[65]
  10. Coastal marshes both store and export blue carbon.[66][67][68] Marshes and wetlands are suggested to have an equivalent flux of CO2 to the atmosphere as rivers, globally.[69]
  11. open ocean typically absorb CO2 from the atmosphere.[65]
  12. The marine biological pump sequesters a small but significant fraction of the absorbed CO2 as organic carbon in marine sediments (see below).[70][38]

Terrestrial runoff to the ocean

How carbon moves from inland waters to the ocean
Carbon dioxide exchange, photosynthetic production and respiration of terrestrial vegetation, rock weathering, and sedimentation occur in terrestrial ecosystems. Carbon transports to the ocean through the land-river-estuary continuum in the form of organic carbon and inorganic carbon. Carbon exchange at the air-water interface, transportation, transformation and sedimentation occur in oceanic ecosystems.[71]

Terrestrial and marine ecosystems are chiefly connected through

dioxygen (O2) pools.[71]

Riverine transport, being the main connective channel of these pools, will act to transport

Gg C y−1.[78][71]

Biological pump in the ocean

Flow of carbon through the open ocean

The ocean

ppm higher than the present day.[80][81][82]

Most carbon incorporated in organic and inorganic biological matter is formed at the sea surface where it can then start sinking to the ocean floor. The deep ocean gets most of its nutrients from the higher water column when they sink down in the form of marine snow. This is made up of dead or dying animals and microbes, fecal matter, sand and other inorganic material.[83]

The biological pump is responsible for transforming

advected and mixed into the deep sea. DOM and aggregates exported into the deep water are consumed and respired, thus returning organic carbon into the enormous deep ocean reservoir of DIC.[84]

A single phytoplankton cell has a sinking rate around one metre per day. Given that the average depth of the ocean is about four kilometres, it can take over ten years for these cells to reach the ocean floor. However, through processes such as coagulation and expulsion in predator fecal pellets, these cells form aggregates. These aggregates have sinking rates orders of magnitude greater than individual cells and complete their journey to the deep in a matter of days.[85]

About 1% of the particles leaving the surface ocean reach the seabed and are consumed, respired, or buried in the sediments. The net effect of these processes is to remove carbon in organic form from the surface and return it to DIC at greater depths, maintaining a surface-to-deep ocean gradient of DIC.

subducted into the earth's mantle and stored for millions of years as part of the slow carbon cycle (see next section).[84]

Sub-processes within slow carbon cycle

Movement of oceanic plates—which carry carbon compounds—through the mantle

Slow or deep carbon cycling is an important process, though it is not as well-understood as the relatively fast carbon movement through the atmosphere, terrestrial biosphere, ocean, and geosphere.[86] The deep carbon cycle is intimately connected to the movement of carbon in the Earth's surface and atmosphere. If the process did not exist, carbon would remain in the atmosphere, where it would accumulate to extremely high levels over long periods of time.[87] Therefore, by allowing carbon to return to the Earth, the deep carbon cycle plays a critical role in maintaining the terrestrial conditions necessary for life to exist.

Furthermore, the process is also significant simply due to the massive quantities of carbon it transports through the planet. In fact, studying the composition of basaltic

core extend from 660 to 2,891 km and 2,891 to 6,371  km deep into the Earth respectively. Accordingly, not much is conclusively known regarding the role of carbon in the deep Earth. Nonetheless, several pieces of evidence—many of which come from laboratory simulations of deep Earth conditions—have indicated mechanisms for the element's movement down into the lower mantle, as well as the forms that carbon takes at the extreme temperatures and pressures of said layer. Furthermore, techniques like seismology
have led to a greater understanding of the potential presence of carbon in the Earth's core.

Carbon in the lower mantle

Carbon outgassing through various processes [89]

Carbon principally enters the mantle in the form of

lower mantle. The study analyzed rare, super-deep diamonds at a site in Juina, Brazil, determining that the bulk composition of some of the diamonds' inclusions matched the expected result of basalt melting and crystallisation under lower mantle temperatures and pressures.[90] Thus, the investigation's findings indicate that pieces of basaltic oceanic lithosphere act as the principle transport mechanism for carbon to Earth's deep interior. These subducted carbonates can interact with lower mantle silicates, eventually forming super-deep diamonds like the one found.[91]

However, carbonates descending to the lower mantle encounter other fates in addition to forming diamonds. In 2011, carbonates were subjected to an environment similar to that of 1800 km deep into the Earth, well within the lower mantle. Doing so resulted in the formations of

reduction as they descend into the mantle before being stabilised at depth by low oxygen fugacity environments. Magnesium, iron, and other metallic compounds act as buffers throughout the process.[94]
The presence of reduced, elemental forms of carbon like graphite would indicate that carbon compounds are reduced as they descend into the mantle.

Carbon is tetrahedrally bonded to oxygen

Polymorphism alters carbonate compounds' stability at different depths within the Earth. To illustrate, laboratory simulations and density functional theory calculations suggest that tetrahedrally coordinated carbonates are most stable at depths approaching the core–mantle boundary.[95][92] A 2015 study indicates that the lower mantle's high pressure causes carbon bonds to transition from sp2 to sp3 hybridised orbitals, resulting in carbon tetrahedrally bonding to oxygen.[96] CO3 trigonal groups cannot form polymerisable networks, while tetrahedral CO4 can, signifying an increase in carbon's coordination number, and therefore drastic changes in carbonate compounds' properties in the lower mantle. As an example, preliminary theoretical studies suggest that high pressure causes carbonate melt viscosity to increase; the melts' lower mobility as a result of its increased viscosity causes large deposits of carbon deep into the mantle.[97]

Accordingly, carbon can remain in the lower mantle for long periods of time, but large concentrations of carbon frequently find their way back to the lithosphere. This process, called carbon outgassing, is the result of carbonated mantle undergoing decompression melting, as well as mantle plumes carrying carbon compounds up towards the crust.[98] Carbon is oxidised upon its ascent towards volcanic hotspots, where it is then released as CO2. This occurs so that the carbon atom matches the oxidation state of the basalts erupting in such areas.[99]

Knowledge about carbon in the core can be gained by analysing shear wave velocities

Carbon in the core

Although the presence of carbon in the Earth's core is well-constrained, recent studies suggest large inventories of carbon could be stored in this region.[

Shear (S) waves moving through the inner core travel at about fifty percent of the velocity expected for most iron-rich alloys.[100] Because the core's composition is believed to be an alloy of crystalline iron and a small amount of nickel, this seismic anomaly indicates the presence of light elements, including carbon, in the core. In fact, studies using diamond anvil cells to replicate the conditions in the Earth's core indicate that iron carbide (Fe7C3) matches the inner core's wave speed and density. Therefore, the iron carbide model could serve as an evidence that the core holds as much as 67% of the Earth's carbon.[101] Furthermore, another study found that in the pressure and temperature condition of the Earth's inner core, carbon dissolved in iron and formed a stable phase with the same Fe7C3 composition—albeit with a different structure from the one previously mentioned.[102]
In summary, although the amount of carbon potentially stored in the Earth's core is not known, recent studies indicate that the presence of iron carbides can explain some of the geophysical observations.

Human influence on fast carbon cycle

Carbon dioxide emissions and partitioning
Emissions of CO2 have been caused by different sources ramping up one after the other (Global Carbon Project)
Partitioning of CO2 emissions show that most emissions are being absorbed by carbon sinks, including plant growth, soil uptake, and ocean uptake (Global Carbon Project)
WWII, human activity has substantially disturbed the global carbon cycle by redistributing massive amounts of carbon from the geosphere.[1] Humans have also continued to shift the natural component functions of the terrestrial biosphere with changes to vegetation and other land use.[7] Man-made (synthetic) carbon compounds have been designed and mass-manufactured that will persist for decades to millennia in air, water, and sediments as pollutants.[103][104] Climate change is amplifying and forcing further indirect human changes to the carbon cycle as a consequence of various positive and negative feedbacks.[25]

Climate change

Climate–carbon cycle feedbacks and state variables
as represented in a stylised model
Carbon stored on land in vegetation and soils is aggregated into a single stock ct. Ocean mixed layer carbon, cm, is the only explicitly modelled ocean stock of carbon; though to estimate carbon cycle feedbacks the total ocean carbon is also calculated.[105]

Current trends in climate change lead to higher ocean temperatures and

acidity, thus modifying marine ecosystems.[106] Also, acid rain and polluted runoff from agriculture and industry change the ocean's chemical composition. Such changes can have dramatic effects on highly sensitive ecosystems such as coral reefs,[107]
thus limiting the ocean's ability to absorb carbon from the atmosphere on a regional scale and reducing oceanic biodiversity globally.

The exchanges of carbon between the atmosphere and other components of the Earth system, collectively known as the carbon cycle, currently constitute important negative (dampening) feedbacks on the effect of anthropogenic carbon emissions on climate change. Carbon sinks in the land and the ocean each currently take up about one-quarter of anthropogenic carbon emissions each year.[108][105]

These feedbacks are expected to weaken in the future, amplifying the effect of anthropogenic carbon emissions on climate change.[109] The degree to which they will weaken, however, is highly uncertain, with Earth system models predicting a wide range of land and ocean carbon uptakes even under identical atmospheric concentration or emission scenarios.[110][105][111] Arctic methane emissions indirectly caused by anthropogenic global warming also affect the carbon cycle and contribute to further warming.

Fossil carbon extraction and burning

Detail of anthropogenic carbon flows, showing cumulative mass in gigatons during years 1850–2018 (left) and the annual mass average during 2009–2018 (right).[112]

The largest and one of the fastest growing human impacts on the carbon cycle and biosphere is the extraction and burning of

clinker production.[113] Clinker is an industrial precursor of cement
.

As of 2020[update], about 450 gigatons of fossil carbon have been extracted in total; an amount approaching the carbon contained in all of Earth's living terrestrial biomass.[112] Recent rates of global emissions directly into the atmosphere have exceeded the uptake by vegetation and the oceans.[114][115][116][117] These sinks have been expected and observed to remove about half of the added atmospheric carbon within about a century.[112][118][119] Nevertheless, sinks like the ocean have evolving saturation properties, and a substantial fraction (20–35%, based on coupled models) of the added carbon is projected remain in the atmosphere for centuries to millennia.[120][121]

Halocarbons

Halocarbons are less prolific compounds developed for diverse uses throughout industry; for example as

perfluorocarbon gases in the atmosphere is responsible for about 10% of the total direct radiative forcing from all long-lived greenhouse gases (year 2019); which includes forcing from the much larger concentrations of carbon dioxide and methane.[122] Chlorofluorocarbons also cause stratospheric ozone depletion. International efforts are ongoing under the Montreal Protocol and Kyoto Protocol to control rapid growth in the industrial manufacturing and use of these environmentally potent gases. For some applications more benign alternatives such as hydrofluoroolefins have been developed and are being gradually introduced.[123]

Land use changes

Since the invention of agriculture, humans have directly and gradually influenced the carbon cycle over century-long timescales by modifying the mixture of vegetation in the terrestrial biosphere.[118] Over the past several centuries, direct and indirect human-caused land use and land cover change (LUCC) has led to the loss of biodiversity, which lowers ecosystems' resilience to environmental stresses and decreases their ability to remove carbon from the atmosphere. More directly, it often leads to the release of carbon from terrestrial ecosystems into the atmosphere.

Deforestation for agricultural purposes removes forests, which hold large amounts of carbon, and replaces them, generally with agricultural or urban areas. Both of these replacement land cover types store comparatively small amounts of carbon so that the net result of the transition is that more carbon stays in the atmosphere. However, the effects on the atmosphere and overall carbon cycle can be intentionally and/or naturally reversed with reforestation.

See also

References

  1. ^ a b c Riebeek, Holli (16 June 2011). "The Carbon Cycle". Earth Observatory. NASA. Archived from the original on 5 March 2016. Retrieved 5 April 2018.
  2. ^ .
  3. ^
    NOAA
    Global Monitoring Laboratory/Earth System Research Laboratories. Retrieved 30 October 2020.
  4. ^ "What is Ocean Acidification?". National Ocean Service, National Oceanic and Atmospheric Administration. Retrieved 30 October 2020.
  5. .
  6. .
  7. ^ .
  8. ^ .
  9. ^ "An Introduction to the Global Carbon Cycle" (PDF). University of New Hampshire. 2009. Archived (PDF) from the original on 8 October 2016. Retrieved 6 February 2016.
  10. ^ A Year In The Life Of Earth's CO2 Archived 26 December 2021 at the Wayback Machine NASA: Goddard Space Flight Center, 17 November 2014.
  11. ^ Forster, P.; Ramawamy, V.; Artaxo, P.; Berntsen, T.; Betts, R.; Fahey, D.W.; Haywood, J.; Lean, J.; Lowe, D.C.; Myhre, G.; Nganga, J.; Prinn, R.; Raga, G.; Schulz, M.; Van Dorland, R. (2007). "Changes in atmospheric constituents and in radiative forcing". Climate Change 2007: The Physical Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change.
  12. ^ "Many Planets, One Earth // Section 4: Carbon Cycling and Earth's Climate". Many Planets, One Earth. 4. Archived from the original on 17 April 2012. Retrieved 24 June 2012.
  13. ^
    S2CID 73722450
    .
  14. .
  15. ^ ].
  16. .
  17. .
  18. .
  19. ^ .
  20. ^ Rice, Charles W. (January 2002). "Storing carbon in soil: Why and how?". Geotimes. 47 (1): 14–17. Archived from the original on 5 April 2018. Retrieved 5 April 2018.
  21. .
  22. .
  23. .
  24. .
  25. ^ .
  26. ^ Sarmiento, J.L.; Gruber, N. (2006). Ocean Biogeochemical Dynamics. Princeton University Press, Princeton, New Jersey, US.
  27. PMID 10102806
    .
  28. .
  29. ^ a b "The Slow Carbon Cycle". NASA. 16 June 2011. Archived from the original on 16 June 2012. Retrieved 24 June 2012.
  30. ^ a b The Carbon Cycle and Earth's Climate Archived 23 June 2003 at the Wayback Machine Information sheet for Columbia University Summer Session 2012 Earth and Environmental Sciences Introduction to Earth Sciences I
  31. ^ Berner, Robert A. (November 1999). "A New Look at the Long-term Carbon Cycle" (PDF). GSA Today. 9 (11): 1–6. Archived (PDF) from the original on 13 February 2019.
  32. ^
    S2CID 210305910
    .
  33. ^ NASA Earth Observatory (16 June 2011). "The Fast Carbon Cycle". Archive. Public Domain This article incorporates text from this source, which is in the public domain.
  34. PMID 11904360
    .
  35. doi:10.3390/sci1010017. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License "Archived copy". Archived from the original on 16 October 2017. Retrieved 5 July 2020.{{cite web}}: CS1 maint: archived copy as title (link) CS1 maint: bot: original URL status unknown (link
    ).
  36. .
  37. ^ NASA Earth Observatory (16 June 2011). "The Slow Carbon Cycle". Archive. Public Domain This article incorporates text from this source, which is in the public domain.
  38. ^
    doi:10.3389/fmars.2017.00007. Modified material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License Archived 16 October 2017 at the Wayback Machine
    .
  39. .
  40. .
  41. .
  42. .
  43. .
  44. .
  45. .
  46. .
  47. .
  48. .
  49. .
  50. .
  51. .
  52. .
  53. .
  54. ^ Linsley, Ray K. (1975). "Solutions Manual to Accompany Hydrology for Engineers".
  55. .
  56. .
  57. .
  58. ^ .
  59. .
  60. .
  61. .
  62. .
  63. .
  64. .
  65. ^ .
  66. .
  67. .
  68. .
  69. .
  70. .
  71. ^
    S2CID 251168582. Modified material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License
    .
  72. .
  73. .
  74. .
  75. .
  76. ^ .
  77. .
  78. .
  79. ^ Sigman DM & GH Haug. 2006. The biological pump in the past. In: Treatise on Geochemistry; vol. 6, (ed.). Pergamon Press, pp. 491–528
  80. .
  81. .
  82. .
  83. .
  84. ^
    doi:10.5670/oceanog.2001.06. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License Archived 16 October 2017 at the Wayback Machine
    .
  85. ^ De La Rocha C.L. (2006) "The Biological Pump". In: Treatise on Geochemistry; vol. 6, Pergamon Press, pp. 83–111.
  86. S2CID 204027259
    .
  87. ^ "The Deep Carbon Cycle and our Habitable Planet | Deep Carbon Observatory". deepcarbon.net. Archived from the original on 27 July 2020. Retrieved 19 February 2019.
  88. .
  89. ^ Dasgupta, Rajdeep (10 December 2011). "From Magma Ocean to Crustal Recycling: Earth's Deep Carbon Cycle". Archived from the original on 24 April 2016. Retrieved 9 March 2019.
  90. ^ "Carbon cycle reaches Earth's lower mantle: Evidence of carbon cycle found in 'superdeep' diamonds From Brazil". ScienceDaily. Retrieved 6 February 2019.
  91. S2CID 129243867
    .
  92. ^ .
  93. (PDF) from the original on 18 July 2021.
  94. .
  95. . Retrieved 7 February 2019.
  96. .
  97. .
  98. .
  99. .
  100. ^ "Does Earth's Core Host a Deep Carbon Reservoir? | Deep Carbon Observatory". deepcarbon.net. Archived from the original on 27 July 2020. Retrieved 9 March 2019.
  101. PMID 25453077
    .
  102. .
  103. ^ "Overview of greenhouse gases". U.S. Environmental Protection Agency. 23 December 2015. Retrieved 2 November 2020.
  104. ^ "The known unknowns of plastic pollution". The Economist. 3 March 2018. Retrieved 17 June 2018.
  105. ^
    hdl:1885/163968. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License Archived 16 October 2017 at the Wayback Machine
    .
  106. .
  107. (PDF) from the original on 26 April 2019.
  108. .
  109. .
  110. .
  111. ^ Hausfather, Zeke; Betts, Richard (14 April 2020). "Analysis: How 'carbon-cycle feedbacks' could make global warming worse". Carbon Brief. Archived from the original on 16 April 2020. Retrieved 4 January 2022.
  112. ^
    doi:10.5194/essd-11-1783-2019. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License Archived 16 October 2017 at the Wayback Machine
    .
  113. ^ IPCC (2007) 7.4.5 Minerals Archived 25 May 2016 at the Wayback Machine in Climate Change 2007: Working Group III: Mitigation of Climate Change,
  114. ^ Buis, Alan; Ramsayer, Kate; Rasmussen, Carol (12 November 2015). "A Breathing Planet, Off Balance". NASA. Archived from the original on 14 November 2015. Retrieved 13 November 2015.
  115. ^ "Audio (66:01) - NASA News Conference - Carbon & Climate Telecon". NASA. 12 November 2015. Archived from the original on 17 November 2015. Retrieved 12 November 2015.
  116. ^ St. Fleur, Nicholas (10 November 2015). "Atmospheric Greenhouse Gas Levels Hit Record, Report Says". The New York Times. Archived from the original on 11 November 2015. Retrieved 11 November 2015.
  117. AP News. Archived
    from the original on 17 November 2015. Retrieved 11 November 2015.
  118. ^ .
  119. ^ "Figure 8.SM.4" (PDF). Intergovernmental Panel on Climate Change Fifth Assessment Report. p. 8SM-16. Archived (PDF) from the original on 13 March 2019.
  120. .
  121. .
  122. NOAA
    Global Monitoring Laboratory/Earth System Research Laboratories.
  123. ^ Sciance, Fred (29 October 2013). "The Transition from HFC- 134a to a Low -GWP Refrigerant in Mobile Air Conditioners HFO -1234yf" (PDF). General Motors Public Policy Center. Archived (PDF) from the original on 15 October 2015. Retrieved 1 August 2018.

External links