Martin curve

Last updated
Surface concentrations of particulate organic matter (POM)
as imaged by satellite in 2011 POC.png
Surface concentrations of particulate organic matter (POM)
as imaged by satellite in 2011

The Martin curve is a power law used by oceanographers to describe the export to the ocean floor of particulate organic carbon (POC). The curve is controlled with two parameters: the reference depth in the water column, and a remineralisation parameter which is a measure of the rate at which the vertical flux of POC attenuates. [1] It is named after the American oceanographer John Martin.

Contents

The Martin Curve has been used in the study of ocean carbon cycling and has contributed to understanding the role of the ocean in regulating atmospheric CO2 levels.

Background

Flux of POC with depth Flux of POC with depth.png
Flux of POC with depth

The dynamics of the particulate organic carbon (POC) pool in the ocean are central to the marine carbon cycle. POC is the link between surface primary production, the deep ocean, and marine sediments. The rate at which POC is degraded in the dark ocean can impact atmospheric CO2 concentration. [2]

The biological carbon pump (BCP) is a crucial mechanism by which atmospheric CO2 is taken up by the ocean and transported to the ocean interior. Without the BCP, the pre-industrial atmospheric CO2 concentration (~280 ppm) would have risen to ~460 ppm. [3] At present, the particulate organic carbon (POC) flux from the surface layer of the ocean to the ocean interior has been estimated to be 4–13 Pg-C year−1. [4] To evaluate the efficiency of the BCP, it is necessary to quantify the vertical attenuation of the POC flux with depth because the deeper that POC is transported, the longer the CO2 will be isolated from the atmosphere. Thus, an increase in the efficiency of the BCP has the potential to cause an increase of ocean carbon sequestration of atmospheric CO2 that would result in a negative feedback on global warming. [5] Different researchers have investigated the vertical attenuation of the POC flux since the 1980s. [6] [7] [8] [9] [5]

In 1987, Martin et al. proposed the following power law function to describe the POC flux attenuation: [10]

                 (1)

where z is water depth (m), and Fz and F100 are the POC fluxes at depths of z metres and 100 metres respectively. Although other functions, such as an exponential curve, have also been proposed and validated, [11] [12] this power law function, commonly known as the "Martin curve", has been used very frequently in discussions of the BCP. The exponent b in this equation has been used as an index of BCP efficiency: the larger the exponent b, the higher the vertical attenuation rate of the POC flux and the lower the BCP efficiency. [13] Moreover, numerical simulations have shown that a change in the value of b would significantly change the atmospheric CO2 concentration. [14] [15] [16] [5]

Subsequently, other researchers have derived alternative remineralization profiles from assumptions about particle degradability and sinking speed. [17] [18] [12] [19] [20] [21] [22] [6] However, the Martin curve has become ubiquitous as the model that assumes slower-sinking and/or labile organic matter is preferentially depleted near the surface causing increasing sinking speed and/or remineralization timescale with depth. [10] [23]

The Martin curve can be expressed in a slightly more general way as: [24]

where fp(z) is the fraction of the flux of particulate organic matter from a productive layer near the surface [25] sinking through the depth horizon z [m], Cp [mb] is a scaling coefficient, and b is a nondimensional exponent controlling how fp decreases with depth. The equation is often normalised to a reference depth zo but this parameter can be readily absorbed into Cp. [24]

Vertical attenuation rate

The vertical attenuation rate of the POC flux is very dependent on the sinking velocity and decomposition rate of POC in the water column. Because POC is labile and has little negative buoyancy, it must be aggregated with relatively heavy materials called ballast to settle gravitationally in the ocean. Materials that may serve as ballast include biogenic opal (hereinafter "opal"), CaCO3, and aluminosilicates. In 1993, Ittekkot hypothesized that the drastic decrease from ~280 to ~200 ppm of atmospheric CO2 that occurred during the last glacial maximum was caused by an increase of the input of aeolian dust (aluminosilicate ballast) to the ocean, which strengthened the BCP. [26] In 2002, Klaas and Archer , as well as Francois et al. who compiled and analyzed global sediment trap data, suggested that CaCO3, which has the largest density among possible ballast minerals, is globally the most important and effective facilitator of vertical POC transport, because the transfer efficiency (the ratio of the POC flux in the deep sea to that at the bottom of the surface mixed layer) is higher in subtropical and tropical areas where CaCO3 is a major component of marine snow. [27] [28] [5]

Reported sinking velocities of CaCO3-rich particles are high. [29] [30] Numerical simulations that take into account these findings have indicated that future ocean acidification will reduce the efficiency of the BCP by decreasing ocean calcification. [31] In addition, the POC export ratio (the ratio of the POC flux from an upper layer (a fixed depth such as 100 metres, or the euphotic zone or mixed layer) to net primary productivity) in subtropical and tropical areas is low because high temperatures in the upper layer increase POC decomposition rates. [32] The result might be a higher transfer efficiency and a strong positive correlation between POC and CaCO3 in these low-latitude areas: labile POC, which is fresher and easier for microbes to break down, decomposes in the upper layer, and relatively refractory POC is transported to the ocean interior in low-latitude areas. [28] [33] [5]

Effective Martin curve slope for POC
(according to a global biogeochemical model called M AGO  ) Effective Martin curve slope for POC.png
Effective Martin curve slope for POC
(according to a global biogeochemical model called M AGO )

On the basis of observations that revealed a large increase of POC fluxes in high-latitude areas during diatom blooms and on the fact that diatoms are much bigger than coccolithophores, [35] [36] [37] Honda and Watanabe proposed in 2010 that opal, rather than CaCO3, is crucial as ballast for effective POC vertical transport in subarctic regions. [38] Weber et al. reported in 2016 a strong negative correlation between transfer efficiency and the picoplankton fraction of plankton as well as higher transfer efficiencies in high-latitude areas, where large phytoplankton such as diatoms predominate. [39] They also calculated that the fraction of vertically transported CO2 that has been sequestered in the ocean interior for at least 100 years is higher in high-latitude (polar and subpolar) regions than in low-latitude regions. [39] [5]

In contrast, Bach et al.conducted in 2019 a mesocosm experiment to study how the plankton community structure affected sinking velocities and reported that during more productive periods the sinking velocity of aggregated particles was not necessarily higher, because the aggregated particles produced then were very fluffy; rather, the settling velocity was higher when the phytoplankton were dominated by small cells. [40] In 2012, Henson et al. revisited the global sediment trap data and reported the POC flux is negatively correlated with the opal export flux and uncorrelated with the CaCO3 export flux. [33] [5]

Zones in the water column Pelagiczone.svg
Zones in the water column

Key factors affecting the rate of biological decomposition of sinking POC in the water column are water temperature and the dissolved oxygen (DO) concentration: the lower the water temperature and the DO concentration, the slower the biological respiration rate and, consequently, the POC flux decomposition rate. [41] [42] [43] [44] For example, in 2015 Marsay with other analysed POC flux data from neutrally buoyant sediment traps in the upper 500 m of the water column and found a significant positive correlation between the exponent b in equation (1) above and water temperature (i.e., the POC flux was attenuated more rapidly when the water was warmer). [41] In addition, Bach et al. found POC decomposition rates are high (low) when diatoms and Synechococcus (harmful algae) are the dominant phytoplankton because of increased (decreased) zooplankton abundance and the consequent increase (decrease) in grazing pressure. [40] [5]

Using radiochemical observations (234Th-based POC flux observations), Pavia et al. found in 2019 that the exponent b of the Martin curve was significantly smaller in the low-oxygen (hypoxic) eastern Pacific equatorial zone than in other areas; that is, vertical attenuation of the POC flux was smaller in the hypoxic area. [44] They pointed out that a more hypoxic ocean in the future would lead to a lower attenuation of the POC flux and therefore increased BCP efficiency and could thereby be a negative feedback on global warming. [44] McDonnell et al. reported in 2015 that vertical transport of POC is more effective in the Antarctic, where the sinking velocity is higher and the biological respiration rate is lower than in the subtropical Atlantic. [45] Henson et al. also reported in 2019 a high export ratio during the early bloom period, when primary productivity is low, and a low export ratio during the late bloom period, when primary productivity is high. [46] They attributed the low export ratio during the late bloom to grazing pressure by microzooplankton and bacteria. [46] [5]

Despite these many investigations of the BCP, the factors governing the vertical attenuation of POC flux are still under debate. Observations in subarctic regions have shown that the transfer efficiency between depths of 1000 and 2000 m is relatively low and that between the bottom of the euphotic zone and a depth of 1000 m it is relatively high. [41] Marsay et al. therefore proposed in 2015 that the Martin curve does not appropriately express the vertical attenuation of POC flux in all regions and that a different equation should instead be developed for each region. [41] Gloege et al. discussed in 2017 parameterization of the vertical attenuation of POC flux, [11] and reported that vertical attenuation of the POC flux in the twilight zone (from the base of the euphotic zone to 1000 m) can be parameterised well not only by a power law model (Martin curve) but also by an exponential model [18] and a ballast model. [17] [5]

However, the exponential model tends to underestimate the POC flux in the midnight zone (depths greater than 1000 metres). Cael and Bisson reported in 2018 that the exponential model (power law model) tends to underestimate the POC flux in the upper layer, and overestimate it in the deep layer. [12] However, the abilities of both models to describe POC fluxes were comparable statistically when they were applied to the POC flux dataset from the eastern Pacific that was used to propose the "Martin curve". [10] In a long-term study in the northeastern Pacific, Smith et al. observed in 2018 a sudden increase of the POC flux accompanied by an unusually high transfer efficiency; they have suggested that because the Martin curve cannot express such a sudden increase, it may sometimes underestimate BCP strength. [47] In addition, contrary to previous findings, some studies have reported a significantly higher transfer efficiency, especially to the deep sea, in subtropical regions than in subarctic regions. [33] [41] [42] This pattern may be attributable to small temperature and DO concentration differences in the deep sea between high-latitude and low-latitude regions, as well as to a higher sinking velocity in subtropical regions, where CaCO3 is a major component of deep-sea marine snow. Moreover, it is also possible that POC is more refractory in low-latitude areas than in high-latitude areas. [33] [41] [42] [5]

Uncertainty in the biological pump

The ocean's biological pump regulates atmospheric carbon dioxide levels and climate by transferring organic carbon produced at the surface by phytoplankton to the ocean interior via marine snow, where the organic carbon is consumed and respired by marine microorganisms. This surface to deep transport is usually described by a power law relationship of sinking particle concentration with depth. Uncertainty in biological pump strength can be related to different variable values (parametric uncertainty) or the underlying equations (structural uncertainty) that describe organic matter export. In 2021, Lauderdale evaluated structural uncertainty using an ocean biogeochemistry model by systematically substituting six alternative remineralisation profiles fit to a reference power-law curve. Structural uncertainty makes a substantial contribution, about one-third in atmospheric pCO2 terms, to the total uncertainty of the biological pump, highlighting the importance of improving biological pump characterisation from observations and its mechanistic inclusion in climate models. [24] [48]

Carbon and nutrients are consumed by phytoplankton in the surface ocean during primary production, leading to a downward flux of organic matter. This "marine snow" is transformed, respired, and degraded by heterotrophic organisms in deeper waters, ultimately releasing those constituents back into dissolved inorganic form. Oceanic overturning and turbulent mixing return resource-rich deep waters back to the sunlit surface layer, sustaining global ocean productivity. The biological pump maintains this vertical gradient in nutrients through uptake, vertical transport, and remineralisation of organic matter, storing carbon in the deep ocean that is isolated from the atmosphere on centennial and millennial timescales, lowering atmospheric CO2 levels by several hundred microatmospheres. [49] [50] The biological pump resists simple mechanistic characterisation due to the complex suite of biological, chemical, and physical processes involved, [51] so the fate of exported organic carbon is typically described using a depth-dependent profile to evaluate the degradation of sinking particulate matter. [24]

See also

Related Research Articles

<span class="mw-page-title-main">Carbon cycle</span> Natural processes of carbon exchange

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.

<span class="mw-page-title-main">Biological pump</span> Carbon capture process in oceans

The biological pump (or ocean carbon biological pump or marine biological carbon pump) is the ocean's biologically driven sequestration of carbon from the atmosphere and land runoff to the ocean interior and seafloor sediments. In other words, it is a biologically mediated process which results in the sequestering of carbon in the deep ocean away from the atmosphere and the land. The biological pump is the biological component of the "marine carbon pump" which contains both a physical and biological component. It is the part of the broader oceanic carbon cycle responsible for the cycling of organic matter formed mainly by phytoplankton during photosynthesis (soft-tissue pump), as well as the cycling of calcium carbonate (CaCO3) formed into shells by certain organisms such as plankton and mollusks (carbonate pump).

<span class="mw-page-title-main">Biogeochemical cycle</span> Chemical transfer pathway between Earths biological and non-biological parts

A biogeochemical cycle, or more generally a cycle of matter, is the movement and transformation of chemical elements and compounds between living organisms, the atmosphere, and the Earth's crust. Major biogeochemical cycles include the carbon cycle, the nitrogen cycle and the water cycle. In each cycle, the chemical element or molecule is transformed and cycled by living organisms and through various geological forms and reservoirs, including the atmosphere, the soil and the oceans. It can be thought of as the pathway by which a chemical substance cycles the biotic compartment and the abiotic compartments of Earth. The biotic compartment is the biosphere and the abiotic compartments are the atmosphere, lithosphere and hydrosphere.

The mesopelagiczone, also known as the middle pelagic or twilight zone, is the part of the pelagic zone that lies between the photic epipelagic and the aphotic bathypelagic zones. It is defined by light, and begins at the depth where only 1% of incident light reaches and ends where there is no light; the depths of this zone are between approximately 200 to 1,000 meters below the ocean surface.

The bathypelagic zone or bathyal zone is the part of the open ocean that extends from a depth of 1,000 to 4,000 m below the ocean surface. It lies between the mesopelagic above and the abyssopelagic below. The bathypelagic is also known as the midnight zone because of the lack of sunlight; this feature does not allow for photosynthesis-driven primary production, preventing growth of phytoplankton or aquatic plants. Although larger by volume than the photic zone, human knowledge of the bathypelagic zone remains limited by ability to explore the deep ocean.

<span class="mw-page-title-main">Iron cycle</span>

The iron cycle (Fe) is the biogeochemical cycle of iron through the atmosphere, hydrosphere, biosphere and lithosphere. While Fe is highly abundant in the Earth's crust, it is less common in oxygenated surface waters. Iron is a key micronutrient in primary productivity, and a limiting nutrient in the Southern ocean, eastern equatorial Pacific, and the subarctic Pacific referred to as High-Nutrient, Low-Chlorophyll (HNLC) regions of the ocean.

<span class="mw-page-title-main">Continental shelf pump</span> Transport of carbon from shallow waters

In oceanic biogeochemistry, the continental shelf pump is proposed to operate in the shallow waters of the continental shelves, acting as a mechanism to transport carbon from surface waters to the interior of the adjacent deep ocean.

<span class="mw-page-title-main">Dissolved inorganic carbon</span> Sum of inorganic carbon species in a solution

Dissolved inorganic carbon (DIC) is the sum of the aqueous species of inorganic carbon in a solution. Carbon compounds can be distinguished as either organic or inorganic, and as dissolved or particulate, depending on their composition. Organic carbon forms the backbone of key component of organic compounds such as – proteins, lipids, carbohydrates, and nucleic acids.

f-ratio (oceanography) In oceanic biogeochemistry, the fraction of total primary production fuelled by nitrate

In oceanic biogeochemistry, the f-ratio is the fraction of total primary production fuelled by nitrate. The ratio was originally defined by Richard Eppley and Bruce Peterson in one of the first papers estimating global oceanic production. This fraction was originally believed significant because it appeared to directly relate to the sinking (export) flux of organic marine snow from the surface ocean by the biological pump. However, this interpretation relied on the assumption of a strong depth-partitioning of a parallel process, nitrification, that more recent measurements has questioned.

The Global Ocean Data Analysis Project (GLODAP) is a synthesis project bringing together oceanographic data, featuring two major releases as of 2018. The central goal of GLODAP is to generate a global climatology of the World Ocean's carbon cycle for use in studies of both its natural and anthropogenically forced states. GLODAP is funded by the National Oceanic and Atmospheric Administration, the U.S. Department of Energy, and the National Science Foundation.

<span class="mw-page-title-main">Gelatinous zooplankton</span> Fragile and often translucent animals that live in the water column

Gelatinous zooplankton are fragile animals that live in the water column in the ocean. Their delicate bodies have no hard parts and are easily damaged or destroyed. Gelatinous zooplankton are often transparent. All jellyfish are gelatinous zooplankton, but not all gelatinous zooplankton are jellyfish. The most commonly encountered organisms include ctenophores, medusae, salps, and Chaetognatha in coastal waters. However, almost all marine phyla, including Annelida, Mollusca and Arthropoda, contain gelatinous species, but many of those odd species live in the open ocean and the deep sea and are less available to the casual ocean observer. Many gelatinous plankters utilize mucous structures in order to filter feed. Gelatinous zooplankton have also been called Gelata.

<span class="mw-page-title-main">Marine snow</span> Shower of organic detritus in the ocean

In the deep ocean, marine snow is a continuous shower of mostly organic detritus falling from the upper layers of the water column. It is a significant means of exporting energy from the light-rich photic zone to the aphotic zone below, which is referred to as the biological pump. Export production is the amount of organic matter produced in the ocean by primary production that is not recycled (remineralised) before it sinks into the aphotic zone. Because of the role of export production in the ocean's biological pump, it is typically measured in units of carbon. The term was coined by explorer William Beebe as observed from his bathysphere. As the origin of marine snow lies in activities within the productive photic zone, the prevalence of marine snow changes with seasonal fluctuations in photosynthetic activity and ocean currents. Marine snow can be an important food source for organisms living in the aphotic zone, particularly for organisms that live very deep in the water column.

<span class="mw-page-title-main">Oceanic carbon cycle</span> Ocean/atmosphere carbon exchange process

The oceanic carbon cycle is composed of processes that exchange carbon between various pools within the ocean as well as between the atmosphere, Earth interior, and the seafloor. The carbon cycle is a result of many interacting forces across multiple time and space scales that circulates carbon around the planet, ensuring that carbon is available globally. The Oceanic carbon cycle is a central process to the global carbon cycle and contains both inorganic carbon and organic carbon. Part of the marine carbon cycle transforms carbon between non-living and living matter.

<span class="mw-page-title-main">Particulate organic matter</span>

Particulate organic matter (POM) is a fraction of total organic matter operationally defined as that which does not pass through a filter pore size that typically ranges in size from 0.053 millimeters (53 μm) to 2 millimeters.

<span class="mw-page-title-main">Marine biogeochemical cycles</span>

Marine biogeochemical cycles are biogeochemical cycles that occur within marine environments, that is, in the saltwater of seas or oceans or the brackish water of coastal estuaries. These biogeochemical cycles are the pathways chemical substances and elements move through within the marine environment. In addition, substances and elements can be imported into or exported from the marine environment. These imports and exports can occur as exchanges with the atmosphere above, the ocean floor below, or as runoff from the land.

<span class="mw-page-title-main">Lipid pump</span>

The lipid pump sequesters carbon from the ocean's surface to deeper waters via lipids associated with overwintering vertically migratory zooplankton. Lipids are a class of hydrocarbon rich, nitrogen and phosphorus deficient compounds essential for cellular structures. This lipid carbon enters the deep ocean as carbon dioxide produced by respiration of lipid reserves and as organic matter from the mortality of zooplankton.

<span class="mw-page-title-main">Silica cycle</span> Biogeochemical cycle

The silica cycle is the biogeochemical cycle in which biogenic silica is transported between the Earth's systems. Silicon is considered a bioessential element and is one of the most abundant elements on Earth. The silica cycle has significant overlap with the carbon cycle and plays an important role in the sequestration of carbon through continental weathering, biogenic export and burial as oozes on geologic timescales.

<span class="mw-page-title-main">Particulate inorganic carbon</span>

Particulate inorganic carbon (PIC) can be contrasted with dissolved inorganic carbon (DIC), the other form of inorganic carbon found in the ocean. These distinctions are important in chemical oceanography. Particulate inorganic carbon is sometimes called suspended inorganic carbon. In operational terms, it is defined as the inorganic carbon in particulate form that is too large to pass through the filter used to separate dissolved inorganic carbon.

<span class="mw-page-title-main">Great Calcite Belt</span> High-calcite region of the Southern Ocean

The Great Calcite Belt (GCB) refers to a region of the ocean where there are high concentrations of calcite, a mineral form of calcium carbonate. The belt extends over a large area of the Southern Ocean surrounding Antarctica. The calcite in the Great Calcite Belt is formed by tiny marine organisms called coccolithophores, which build their shells out of calcium carbonate. When these organisms die, their shells sink to the bottom of the ocean, and over time, they accumulate to form a thick layer of calcite sediment.

Low-nutrient, low-chlorophyll (LNLC)regions are aquatic zones that are low in nutrients and consequently have low rate of primary production, as indicated by low chlorophyll concentrations. These regions can be described as oligotrophic, and about 75% of the world's oceans encompass LNLC regions. A majority of LNLC regions are associated with subtropical gyres but are also present in areas of the Mediterranean Sea, and some inland lakes. Physical processes limit nutrient availability in LNLC regions, which favors nutrient recycling in the photic zone and selects for smaller phytoplankton species. LNLC regions are generally not found near coasts, owing to the fact that coastal areas receive more nutrients from terrestrial sources and upwelling. In marine systems, seasonal and decadal variability of primary productivity in LNLC regions is driven in part by large-scale climatic regimes leading to important effects on the global carbon cycle and the oceanic carbon cycle.

References

  1. Olli, Kalle (2015). "Unraveling the uncertainty and error propagation in the vertical flux Martin curve". Progress in Oceanography. 135: 146–155. Bibcode:2015PrOce.135..146O. doi:10.1016/j.pocean.2015.05.016.
  2. Kharbush, J.J., Close, H.G., Van Mooy, B.A., Arnosti, C., Smittenberg, R.H., Le Moigne, F.A., Mollenhauer, G., Scholz-Böttcher, B., Obreht, I., Koch, B.P. and Becker, K. (2020) "Particulate Organic Carbon Deconstructed: Molecular and Chemical Composition of Particulate Organic Carbon in the Ocean". Frontiers in Marine Science, 7: 518. doi : 10.3389/fmars.2020.00518. CC-BY icon.svg Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  3. Villa-Alfageme, M.; De Soto, F. C.; Ceballos, E.; Giering, S. L. C.; Le Moigne, F. A. C.; Henson, S.; Mas, J. L.; Sanders, R. J. (2016). "Geographical, seasonal, and depth variation in sinking particle speeds in the North Atlantic". Geophysical Research Letters. 43 (16): 8609–8616. Bibcode:2016GeoRL..43.8609V. doi:10.1002/2016GL069233. hdl: 11441/98535 . S2CID   3728545.
  4. Lima, I. D.; Lam, P. J.; Doney, S. C. (2014). "Dynamics of particulate organic carbon flux in a global ocean model". Biogeosciences. 11 (4): 1177–1198. Bibcode:2014BGeo...11.1177L. doi: 10.5194/bg-11-1177-2014 . hdl: 1912/6674 .
  5. 1 2 3 4 5 6 7 8 9 10 11 Honda, Makio C. (2020). "Effective Vertical Transport of Particulate Organic Carbon in the Western North Pacific Subarctic Region". Frontiers in Earth Science. 8: 366. Bibcode:2020FrEaS...8..366H. doi: 10.3389/feart.2020.00366 . CC-BY icon.svg Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  6. 1 2 Suess, Erwin (1980). "Particulate organic carbon flux in the oceans—surface productivity and oxygen utilization". Nature. 288 (5788): 260–263. Bibcode:1980Natur.288..260S. doi:10.1038/288260a0. S2CID   4275275.
  7. Betzer, Peter R.; Showers, William J.; Laws, Edward A.; Winn, Christopher D.; Ditullio, Giacomo R.; Kroopnick, Peter M. (1984). "Primary productivity and particle fluxes on a transect of the equator at 153°W in the Pacific Ocean". Deep Sea Research Part A. Oceanographic Research Papers. 31 (1): 1–11. Bibcode:1984DSRA...31....1B. doi:10.1016/0198-0149(84)90068-2.
  8. Berger, W. H., Fischer, K., Lai, C., and Wu, G. (1987). "Ocean carbon flux: global maps of primary production and export production,". In: Biogeochemical Cycling and Fluxes between the Deep Euphotic Zone and Other Oceanic Realms", Vol. 3, ed. C. Agegian (Silver Spring, MD: NOAA), pages 87–30.
  9. Pace, Michael L.; Knauer, George A.; Karl, David M.; Martin, John H. (1987). "Primary production, new production and vertical flux in the eastern Pacific Ocean". Nature. 325 (6107): 803–804. Bibcode:1987Natur.325..803P. doi:10.1038/325803a0. S2CID   4353764.
  10. 1 2 3 Martin, John H.; Knauer, George A.; Karl, David M.; Broenkow, William W. (1987). "VERTEX: Carbon cycling in the northeast Pacific". Deep Sea Research Part A. Oceanographic Research Papers. 34 (2): 267–285. Bibcode:1987DSRA...34..267M. doi:10.1016/0198-0149(87)90086-0.
  11. 1 2 Gloege, Lucas; McKinley, Galen A.; Mouw, Colleen B.; Ciochetto, Audrey B. (2017). "Global evaluation of particulate organic carbon flux parameterizations and implications for atmospheric pCO2". Global Biogeochemical Cycles. 31 (7): 1192–1215. Bibcode:2017GBioC..31.1192G. doi:10.1002/2016GB005535. S2CID   133746237.
  12. 1 2 3 Cael, B. B.; Bisson, Kelsey (2018). "Particle Flux Parameterizations: Quantitative and Mechanistic Similarities and Differences". Frontiers in Marine Science. 5. doi: 10.3389/fmars.2018.00395 . hdl: 1912/23677 .
  13. Berelson, W. M. (2001). "The flux of particulate organic carbon into the ocean interior.: a comparison of four U.S. JGOFS regional studies". Oceanography, 14: 59–64.
  14. Yamanaka, Yasuhiro; Tajika, Eiichi (1996). "The role of the vertical fluxes of particulate organic matter and calcite in the oceanic carbon cycle: Studies using an ocean biogeochemical general circulation model". Global Biogeochemical Cycles. 10 (2): 361–382. Bibcode:1996GBioC..10..361Y. doi:10.1029/96gb00634.
  15. Kwon, Eun Young; Primeau, François; Sarmiento, Jorge L. (2009). "The impact of remineralization depth on the air–sea carbon balance". Nature Geoscience. 2 (9): 630–635. Bibcode:2009NatGe...2..630K. doi:10.1038/ngeo612. S2CID   59058813.
  16. Wilson, Jamie D.; Barker, Stephen; Edwards, Neil R.; Holden, Philip B.; Ridgwell, Andy (2019). "Sensitivity of atmospheric CO2 to regional variability in particulate organic matter remineralization depths". Biogeosciences. 16 (14): 2923–2936. Bibcode:2019BGeo...16.2923W. doi: 10.5194/bg-16-2923-2019 . S2CID   202987265.
  17. 1 2 Armstrong, Robert A.; Lee, Cindy; Hedges, John I.; Honjo, Susumu; Wakeham, Stuart G. (2001). "A new, mechanistic model for organic carbon fluxes in the ocean based on the quantitative association of POC with ballast minerals". Deep-Sea Research Part II: Topical Studies in Oceanography. 49 (1–3): 219–236. Bibcode:2001DSRII..49..219A. doi:10.1016/S0967-0645(01)00101-1.
  18. 1 2 Banse, Karl (1990). "New views on the degradation and disposition of organic particles as collected by sediment traps in the open sea". Deep Sea Research Part A. Oceanographic Research Papers. 37 (7): 1177–1195. Bibcode:1990DSRA...37.1177B. doi:10.1016/0198-0149(90)90058-4.
  19. Kriest, I.; Oschlies, A. (2008). "On the treatment of particulate organic matter sinking in large-scale models of marine biogeochemical cycles". Biogeosciences. 5 (1): 55–72. Bibcode:2008BGeo....5...55K. doi: 10.5194/bg-5-55-2008 . S2CID   3840184.
  20. Lutz, Michael; Dunbar, Robert; Caldeira, Ken (2002). "Regional variability in the vertical flux of particulate organic carbon in the ocean interior". Global Biogeochemical Cycles. 16 (3): 11-1–11-18. Bibcode:2002GBioC..16.1037L. doi: 10.1029/2000GB001383 .
  21. Middelburg, Jack J. (1989). "A simple rate model for organic matter decomposition in marine sediments". Geochimica et Cosmochimica Acta. 53 (7): 1577–1581. Bibcode:1989GeCoA..53.1577M. doi:10.1016/0016-7037(89)90239-1.
  22. Schwinger, Jörg; Goris, Nadine; Tjiputra, Jerry F.; Kriest, Iris; Bentsen, Mats; Bethke, Ingo; Ilicak, Mehmet; Assmann, Karen M.; Heinze, Christoph (2016). "Evaluation of NorESM-OC (Versions 1 and 1.2), the ocean carbon-cycle stand-alone configuration of the Norwegian Earth System Model (NorESM1)". Geoscientific Model Development. 9 (8): 2589–2622. Bibcode:2016GMD.....9.2589S. doi: 10.5194/gmd-9-2589-2016 . S2CID   59034696.
  23. Kriest, I.; Oschlies, A. (2011). "Numerical effects on organic-matter sedimentation and remineralization in biogeochemical ocean models". Ocean Modelling. 39 (3–4): 275–283. Bibcode:2011OcMod..39..275K. doi:10.1016/j.ocemod.2011.05.001.
  24. 1 2 3 4 Lauderdale, Jonathan Maitland; Cael, B. B. (2021). "Impact of Remineralization Profile Shape on the Air‐Sea Carbon Balance". Geophysical Research Letters. 48 (7): e2020GL091746. Bibcode:2021GeoRL..4891746L. doi:10.1029/2020GL091746. hdl: 1721.1/130486 . PMC   8243937 . PMID   34219838. CC-BY icon.svg Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  25. Buesseler, Ken O.; Boyd, Philip W.; Black, Erin E.; Siegel, David A. (2020). "Metrics that matter for assessing the ocean biological carbon pump". Proceedings of the National Academy of Sciences. 117 (18): 9679–9687. Bibcode:2020PNAS..117.9679B. doi: 10.1073/pnas.1918114117 . PMC   7211944 . PMID   32253312.
  26. Ittekkot, Venugopalan (1993). "The abiotically driven biological pump in the ocean and short-term fluctuations in atmospheric CO2 contents". Global and Planetary Change. 8 (1–2): 17–25. Bibcode:1993GPC.....8...17I. doi:10.1016/0921-8181(93)90060-2.
  27. Klaas, Christine; Archer, David E. (2002). "Association of sinking organic matter with various types of mineral ballast in the deep sea: Implications for the rain ratio". Global Biogeochemical Cycles. 16 (4): 63-1–63-14. Bibcode:2002GBioC..16.1116K. doi: 10.1029/2001GB001765 .
  28. 1 2 Francois, Roger; Honjo, Susumu; Krishfield, Richard; Manganini, Steve (2002). "Factors controlling the flux of organic carbon to the bathypelagic zone of the ocean". Global Biogeochemical Cycles. 16 (4): 34-1–34-20. Bibcode:2002GBioC..16.1087F. doi: 10.1029/2001GB001722 . S2CID   128876389.
  29. Fischer, G.; Karakaş, G. (2009). "Sinking rates and ballast composition of particles in the Atlantic Ocean: Implications for the organic carbon fluxes to the deep ocean". Biogeosciences. 6 (1): 85–102. Bibcode:2009BGeo....6...85F. doi: 10.5194/bg-6-85-2009 . S2CID   45117580.
  30. Sukigara, Chiho; Mino, Yoshihisa; Kawakami, Hajime; Honda, Makio C.; Fujiki, Tetsuichi; Matsumoto, Kazuhiko; Wakita, Masahide; Saino, Toshiro (2019). "Sinking dynamics of particulate matter in the subarctic and subtropical regions of the western North Pacific". Deep Sea Research Part I: Oceanographic Research Papers. 144: 17–27. Bibcode:2019DSRI..144...17S. doi:10.1016/j.dsr.2018.11.004. S2CID   133654627.
  31. Heinze, C. (2004). "Simulating oceanic CaCO3export production in the greenhouse". Geophysical Research Letters. 31 (16). Bibcode:2004GeoRL..3116308H. doi: 10.1029/2004GL020613 .
  32. Laws, Edward A.; Falkowski, Paul G.; Smith, Walker O.; Ducklow, Hugh; McCarthy, James J. (2000). "Temperature effects on export production in the open ocean". Global Biogeochemical Cycles. 14 (4): 1231–1246. Bibcode:2000GBioC..14.1231L. doi: 10.1029/1999gb001229 . S2CID   53998465.
  33. 1 2 3 4 Henson, Stephanie A.; Sanders, Richard; Madsen, Esben (2012). "Global patterns in efficiency of particulate organic carbon export and transfer to the deep ocean". Global Biogeochemical Cycles. 26 (1): n/a. Bibcode:2012GBioC..26.1028H. doi: 10.1029/2011GB004099 .
  34. Maerz, Joeran; Six, Katharina D.; Stemmler, Irene; Ahmerkamp, Soeren; Ilyina, Tatiana (2020). "Microstructure and composition of marine aggregates as co-determinants for vertical particulate organic carbon transfer in the global ocean". Biogeosciences. 17 (7): 1765–1803. Bibcode:2020BGeo...17.1765M. doi: 10.5194/bg-17-1765-2020 . hdl: 21.11116/0000-0004-BD3E-3 . S2CID   216206324.
  35. Buesseler, Ken O. (1998). "The decoupling of production and particulate export in the surface ocean". Global Biogeochemical Cycles. 12 (2): 297–310. Bibcode:1998GBioC..12..297B. doi:10.1029/97GB03366.
  36. Buesseler, K. O.; Lamborg, C. H.; Boyd, P. W.; Lam, P. J.; Trull, T. W.; Bidigare, R. R.; Bishop, J. K. B.; Casciotti, K. L.; Dehairs, F.; Elskens, M.; Honda, M.; Karl, D. M.; Siegel, D. A.; Silver, M. W.; Steinberg, D. K.; Valdes, J.; Van Mooy, B.; Wilson, S. (2007). "Revisiting Carbon Flux Through the Ocean's Twilight Zone". Science. 316 (5824): 567–570. Bibcode:2007Sci...316..567B. doi:10.1126/science.1137959. PMID   17463282. S2CID   8423647.
  37. Buesseler, K.O.; Trull, T.W.; Steinberg, D.K.; Silver, M.W.; Siegel, D.A.; Saitoh, S.-I.; Lamborg, C.H.; Lam, P.J.; Karl, D.M.; Jiao, N.Z.; Honda, M.C.; Elskens, M.; Dehairs, F.; Brown, S.L.; Boyd, P.W.; Bishop, J.K.B.; Bidigare, R.R. (2008). "VERTIGO (VERtical Transport in the Global Ocean): A study of particle sources and flux attenuation in the North Pacific". Deep-Sea Research Part II: Topical Studies in Oceanography. 55 (14–15): 1522–1539. Bibcode:2008DSRII..55.1522B. doi:10.1016/j.dsr2.2008.04.024.
  38. Honda, Makio C.; Watanabe, Shuichi (2010). "Importance of biogenic opal as ballast of particulate organic carbon (POC) transport and existence of mineral ballast-associated and residual POC in the Western Pacific Subarctic Gyre". Geophysical Research Letters. 37 (2): n/a. Bibcode:2010GeoRL..37.2605H. doi:10.1029/2009GL041521.
  39. 1 2 Weber, Thomas; Cram, Jacob A.; Leung, Shirley W.; Devries, Timothy; Deutsch, Curtis (2016). "Deep ocean nutrients imply large latitudinal variation in particle transfer efficiency". Proceedings of the National Academy of Sciences. 113 (31): 8606–8611. Bibcode:2016PNAS..113.8606W. doi: 10.1073/pnas.1604414113 . PMC   4978250 . PMID   27457946.
  40. 1 2 Bach, L. T.; Stange, P.; Taucher, J.; Achterberg, E. P.; Algueró‐Muñiz, M.; Horn, H.; Esposito, M.; Riebesell, U. (2019). "The Influence of Plankton Community Structure on Sinking Velocity and Remineralization Rate of Marine Aggregates". Global Biogeochemical Cycles. 33 (8): 971–994. Bibcode:2019GBioC..33..971B. doi: 10.1029/2019GB006256 . S2CID   201329875.
  41. 1 2 3 4 5 6 Marsay, Chris M.; Sanders, Richard J.; Henson, Stephanie A.; Pabortsava, Katsiaryna; Achterberg, Eric P.; Lampitt, Richard S. (2015). "Attenuation of sinking particulate organic carbon flux through the mesopelagic ocean". Proceedings of the National Academy of Sciences. 112 (4): 1089–1094. Bibcode:2015PNAS..112.1089M. doi: 10.1073/pnas.1415311112 . PMC   4313834 . PMID   25561526.
  42. 1 2 3 Devries, Tim; Weber, Thomas (2017). "The export and fate of organic matter in the ocean: New constraints from combining satellite and oceanographic tracer observations". Global Biogeochemical Cycles. 31 (3): 535–555. Bibcode:2017GBioC..31..535D. doi:10.1002/2016GB005551. S2CID   132192896.
  43. Laufkötter, C.; John, Jasmin G.; Stock, Charles A.; Dunne, John P. (2017). "Temperature and oxygen dependence of the remineralization of organic matter". Global Biogeochemical Cycles. 31 (7): 1038–1050. Bibcode:2017GBioC..31.1038L. doi:10.1002/2017GB005643. S2CID   102953187.
  44. 1 2 3 Pavia, Frank J.; Anderson, Robert F.; Lam, Phoebe J.; Cael, B. B.; Vivancos, Sebastian M.; Fleisher, Martin Q.; Lu, Yanbin; Zhang, Pu; Cheng, Hai; Edwards, R. Lawrence (2019). "Shallow particulate organic carbon regeneration in the South Pacific Ocean". Proceedings of the National Academy of Sciences. 116 (20): 9753–9758. Bibcode:2019PNAS..116.9753P. doi: 10.1073/pnas.1901863116 . PMC   6525517 . PMID   31036647.
  45. McDonnell, A. M. P.; Boyd, P. W.; Buesseler, K. O. (2015). "Effects of sinking velocities and microbial respiration rates on the attenuation of particulate carbon fluxes through the mesopelagic zone". Global Biogeochemical Cycles. 29 (2): 175–193. Bibcode:2015GBioC..29..175M. doi:10.1002/2014GB004935. hdl: 1912/7333 . S2CID   53453555.
  46. 1 2 Henson, Stephanie; Le Moigne, Fred; Giering, Sarah (2019). "Drivers of Carbon Export Efficiency in the Global Ocean". Global Biogeochemical Cycles. 33 (7): 891–903. Bibcode:2019GBioC..33..891H. doi:10.1029/2018GB006158. PMC   7006809 . PMID   32063666.
  47. Smith, Kenneth L.; Ruhl, Henry A.; Huffard, Christine L.; Messié, Monique; Kahru, Mati (2018). "Episodic organic carbon fluxes from surface ocean to abyssal depths during long-term monitoring in NE Pacific". Proceedings of the National Academy of Sciences. 115 (48): 12235–12240. Bibcode:2018PNAS..11512235S. doi: 10.1073/pnas.1814559115 . PMC   6275536 . PMID   30429327.
  48. New Research Reveals Uncertainty in How Much Carbon the Ocean Absorbs Over Time – Climate Projections Could Be Off SciTechDaily, 8 May 2021.
  49. Ito, Takamitsu; Follows, Michael J. (2005). "Preformed phosphate, soft tissue pump and atmospheric CO2". Journal of Marine Research. 63 (4): 813–839. doi:10.1357/0022240054663231.
  50. Volk, Tyler; Hoffert, Martin I. (2013). "Ocean Carbon Pumps: Analysis of Relative Strengths and Efficiencies in Ocean-Driven Atmospheric CO2 Changes". The Carbon Cycle and Atmospheric CO2 : Natural Variations Archean to Present. Geophysical Monograph Series. pp. 99–110. doi:10.1029/GM032p0099. ISBN   9781118664322.
  51. Boyd, Philip W.; Claustre, Hervé; Levy, Marina; Siegel, David A.; Weber, Thomas (2019). "Multi-faceted particle pumps drive carbon sequestration in the ocean" (PDF). Nature. 568 (7752): 327–335. Bibcode:2019Natur.568..327B. doi:10.1038/s41586-019-1098-2. PMID   30996317. S2CID   119513489.