Phytoplankton

Source: Wikipedia, the free encyclopedia.

Phytoplankton (/ˌftˈplæŋktən/) are the autotrophic (self-feeding) components of the plankton community and a key part of ocean and freshwater ecosystems. The name comes from the Greek words φυτόν (phyton), meaning 'plant', and (planktos), meaning 'wanderer' or 'drifter'.[1][2][3]

Phytoplankton obtain their energy through

euphotic zone
) of oceans and lakes. In comparison with terrestrial plants, phytoplankton are distributed over a larger surface area, are exposed to less seasonal variation and have markedly faster turnover rates than trees (days versus decades). As a result, phytoplankton respond rapidly on a global scale to climate variations.

Phytoplankton form the base of marine and freshwater food webs and are key players in the global carbon cycle. They account for about half of global photosynthetic activity and at least half of the oxygen production, despite amounting to only about 1% of the global plant biomass. Phytoplankton are very diverse, varying from photosynthesizing bacteria to plant-like algae to armour-plated coccolithophores. Important groups of phytoplankton include the diatoms, cyanobacteria and dinoflagellates, although many other groups are represented.[2]

Most phytoplankton are too small to be individually seen with the unaided eye. However, when present in high enough numbers, some varieties may be noticeable as colored patches on the water surface due to the presence of chlorophyll within their cells and accessory pigments (such as phycobiliproteins or xanthophylls) in some species.

Types

Some types of phytoplankton (not to scale)
Left to right: cyanobacteria, diatom, dinoflagellate, green algae and coccolithophore

Phytoplankton are

photosynthesizing microscopic protists and bacteria that inhabit the upper sunlit layer of marine and fresh water bodies of water on Earth. Paralleling plants on land, phytoplankton undertake primary production in water,[2] creating organic compounds from carbon dioxide dissolved in the water. Phytoplankton form the base of — and sustain — the aquatic food web,[4] and are crucial players in the Earth's carbon cycle.[5]

"Marine photosynthesis is dominated by microalgae, which together with cyanobacteria, are collectively called phytoplankton."[6] Phytoplankton are extremely diverse, varying from photosynthesizing bacteria (cyanobacteria), to plant-like diatoms, to armour-plated coccolithophores.[7][2]

Diatoms are one of the most common types
of phytoplankton
Coccolithophores
are armour-plated
The dinoflagellate Dinophysis acuta

Ecology

Global distribution of ocean phytoplankton – NASA
This visualization shows a model simulation of the dominant phytoplankton types averaged over the period 1994–1998. * Red = diatoms (big phytoplankton, which need silica) * Yellow = flagellates (other big phytoplankton) * Green = prochlorococcus (small phytoplankton that cannot use nitrate) * Cyan = synechococcus (other small phytoplankton) Opacity indicates concentration of the carbon biomass. In particular, the role of the swirls and filaments (mesoscale features) appear important in maintaining high biodiversity in the ocean.[5][8]

Phytoplankton obtain

is a notable exception).

While almost all phytoplankton

Noctiluca and Dinophysis, that obtain organic carbon by ingesting other organisms or detrital
material.

Phytoplankton live in the

viral lysis. Although some phytoplankton cells, such as dinoflagellates, are able to migrate vertically, they are still incapable of actively moving against currents, so they slowly sink and ultimately fertilize the seafloor with dead cells and detritus.[16]

Cycling of marine phytoplankton [16]

Phytoplankton are crucially dependent on a number of

ferrous sulfate) to the oceans to promote phytoplankton growth and draw atmospheric CO2 into the ocean. Controversy about manipulating the ecosystem and the efficiency of iron fertilization has slowed such experiments.[18]

Phytoplankton depend on B vitamins for survival. Areas in the ocean have been identified as having a major lack of some B Vitamins, and correspondingly, phytoplankton.[19]

The effects of

anthropogenic warming on the global population of phytoplankton is an area of active research. Changes in the vertical stratification of the water column, the rate of temperature-dependent biological reactions, and the atmospheric supply of nutrients are expected to have important effects on future phytoplankton productivity.[20][21]

The effects of anthropogenic ocean acidification on phytoplankton growth and community structure has also received considerable attention. The cells of coccolithophore phytoplankton are typically covered in a calcium carbonate shell called a coccosphere that is sensitive to ocean acidification. Because of their short generation times, evidence suggests some phytoplankton can adapt to changes in pH induced by increased carbon dioxide on rapid time-scales (months to years).[22][23]

Phytoplankton serve as the base of the aquatic food web, providing an essential ecological function for all aquatic life. Under future conditions of anthropogenic warming and ocean acidification, changes in phytoplankton mortality due to changes in rates of zooplankton grazing may be significant.[24] One of the many food chains in the ocean – remarkable due to the small number of links – is that of phytoplankton sustaining krill (a crustacean similar to a tiny shrimp), which in turn sustain baleen whales.

The El Niño-Southern Oscillation (ENSO) cycles in the Equatorial Pacific area can affect phytoplankton.[25] Biochemical and physical changes during ENSO cycles modify the phytoplankton community structure.[25] Also, changes in the structure of the phytoplankton, such as a significant reduction in biomass and phytoplankton density, particularly during El Nino phases can occur.[26] The sensitivity of phytoplankton to environmental changes is why they are often used as indicators of estuarine and coastal ecological condition and health.[27] To study these events satellite ocean color observations are used to observe these changes. Satellite images help to have a better view of their global distribution.[25]

Diversity

eddies. Phytoplankton concentrates along the boundaries of the eddies, tracing the motion of the water.
Algal bloom off south west England

The term phytoplankton encompasses all photoautotrophic microorganisms in aquatic

In terms of numbers, the most important groups of phytoplankton include the

atmosphere. DMS is oxidized to form sulfate which, in areas where ambient aerosol particle concentrations are low, can contribute to the population of cloud condensation nuclei, mostly leading to increased cloud cover and cloud albedo according to the so-called CLAW hypothesis.[30][31] Different types of phytoplankton support different trophic levels within varying ecosystems. In oligotrophic oceanic regions such as the Sargasso Sea or the South Pacific Gyre, phytoplankton is dominated by the small sized cells, called picoplankton and nanoplankton (also referred to as picoflagellates and nanoflagellates), mostly composed of cyanobacteria (Prochlorococcus, Synechococcus) and picoeucaryotes such as Micromonas. Within more productive ecosystems, dominated by upwelling or high terrestrial inputs, larger dinoflagellates are the more dominant phytoplankton and reflect a larger portion of the biomass.[32]

Growth strategies

In the early twentieth century, Alfred C. Redfield found the similarity of the phytoplankton's elemental composition to the major dissolved nutrients in the deep ocean.[33] Redfield proposed that the ratio of carbon to nitrogen to phosphorus (106:16:1) in the ocean was controlled by the phytoplankton's requirements, as phytoplankton subsequently release nitrogen and phosphorus as they are remineralized. This so-called "Redfield ratio" in describing stoichiometry of phytoplankton and seawater has become a fundamental principle to understand marine ecology, biogeochemistry and phytoplankton evolution.[34] However, the Redfield ratio is not a universal value and it may diverge due to the changes in exogenous nutrient delivery[35] and microbial metabolisms in the ocean, such as nitrogen fixation, denitrification and anammox.

The dynamic stoichiometry shown in unicellular algae reflects their capability to store nutrients in an internal pool, shift between enzymes with various nutrient requirements and alter osmolyte composition.[36][37] Different cellular components have their own unique stoichiometry characteristics,[34] for instance, resource (light or nutrients) acquisition machinery such as proteins and chlorophyll contain a high concentration of nitrogen but low in phosphorus. Meanwhile, growth machinery such as ribosomal RNA contains high nitrogen and phosphorus concentrations.

Based on allocation of resources, phytoplankton is classified into three different growth strategies, namely survivalist, bloomer[38] and generalist. Survivalist phytoplankton has a high ratio of N:P (>30) and contains an abundance of resource-acquisition machinery to sustain growth under scarce resources. Bloomer phytoplankton has a low N:P ratio (<10), contains a high proportion of growth machinery, and is adapted to exponential growth. Generalist phytoplankton has similar N:P to the Redfield ratio and contain relatively equal resource-acquisition and growth machinery.

Factors affecting abundance

The

atmospheric aerosols, clouds, and climate (NAAMES stands for the North Atlantic Aerosols and Marine Ecosystems Study). The study focused on the sub-arctic region of the North Atlantic Ocean, which is the site of one of Earth's largest recurring phytoplankton blooms. The long history of research in this location, as well as relative ease of accessibility, made the North Atlantic an ideal location to test prevailing scientific hypotheses[39] in an effort to better understand the role of phytoplankton aerosol emissions on Earth's energy budget.[40]

NAAMES was designed to target specific phases of the annual phytoplankton cycle: minimum, climax and the intermediary decreasing and increasing biomass, in order to resolve debates on the timing of bloom formations and the patterns driving annual bloom re-creation.[40] The NAAMES project also investigated the quantity, size, and composition of aerosols generated by primary production in order to understand how phytoplankton bloom cycles affect cloud formations and climate.[41]

Competing hypothesis of plankton variability[39]
Figure adapted from Behrenfeld & Boss 2014.[42]
Courtesy of NAAMES, Langley Research Center, NASA[43]
World concentrations of surface ocean chlorophyll as viewed by satellite during the northern spring, averaged from 1998 to 2004. Chlorophyll is a marker for the distribution and abundance of phytoplankton.
Global patterns of monthly phytoplankton species richness and species turnover
(A) Annual mean of monthly species richness and (B) month-to-month species turnover projected by SDMs. Latitudinal gradients of (C) richness and (D) turnover. Colored lines (regressions with local polynomial fitting) indicate the means per degree latitude from three different SDM algorithms used (red shading denotes ±1 SD from 1000 Monte Carlo runs that used varying predictors for GAM). Poleward of the thin horizontal lines shown in (C) and (D), the model results cover only <12 or <9 months, respectively.[44]

Factors affecting productivity

Environmental factors that affect phytoplankton productivity [45][46]

Phytoplankton are the key mediators of the

marine food chains.[48] Climate change may greatly restructure phytoplankton communities leading to cascading consequences for marine food webs, thereby altering the amount of carbon transported to the ocean interior.[49][45]

The diagram on the right gives an overview of the various environmental factors that together affect

oceanic stratification, circulation and changes in cloud cover and sea ice, resulting in an increased light supply to the ocean surface. Also, reduced nutrient supply is predicted to co-occur with ocean acidification and warming, due to increased stratification of the water column and reduced mixing of nutrients from the deep water to the surface.[51][45]

Role of phytoplankton

Role of phytoplankton on various compartments of the marine environment [52]

In the diagram on the right, the compartments influenced by phytoplankton include the atmospheric gas composition, inorganic nutrients, and trace element fluxes as well as the transfer and cycling of organic matter via biological processes. The photosynthetically fixed carbon is rapidly recycled and reused in the surface ocean, while a certain fraction of this biomass is exported as sinking particles to the deep ocean, where it is subject to ongoing transformation processes, e.g., remineralization.[52]

Phytoplankton contribute to not only a basic pelagic marine food web but also to the microbial loop. Phytoplankton are the base of the marine food web and because they do not rely on other organisms for food, they make up the first trophic level. Organisms such as zooplankton feed of these phytoplankton which are fed on by other organisms and so forth until the fourth trophic level is reached with apex predators. Approximately 90% of total carbon is lost between trophic levels due to respiration, detritus, and dissolved organic matter. This makes the remineralization process and nutrient cycling performed by phytoplankton and bacteria important in maintaining efficiency.[53]

Phytoplankton blooms in which a species increases rapidly under conditions favorable to growth can produce harmful algal blooms (HABs).

Aquaculture

Phytoplankton are a key food item in both

molluscs, including pearl oysters and giant clams. A 2018 study estimated the nutritional value of natural phytoplankton in terms of carbohydrate, protein and lipid across the world ocean using ocean-colour data from satellites,[55] and found the calorific value of phytoplankton to vary considerably across different oceanic regions and between different time of the year.[55][56]

The production of phytoplankton under artificial conditions is itself a form of aquaculture. Phytoplankton is cultured for a variety of purposes, including foodstock for other aquacultured organisms,

colour temperature of illumination should be approximately 6,500 K, but values from 4,000 K to upwards of 20,000 K have been used successfully. The duration of light exposure should be approximately 16 hours daily; this is the most efficient artificial day length.[54]

Anthropogenic changes

Plot demonstrating increases in phytoplankton species richness with increased temperature

Marine phytoplankton perform half of the global photosynthetic CO2 fixation (net global primary production of ~50 Pg C per year) and half of the oxygen production despite amounting to only ~1% of global plant biomass.[57] In comparison with terrestrial plants, marine phytoplankton are distributed over a larger surface area, are exposed to less seasonal variation and have markedly faster turnover rates than trees (days versus decades).[57] Therefore, phytoplankton respond rapidly on a global scale to climate variations. These characteristics are important when one is evaluating the contributions of phytoplankton to carbon fixation and forecasting how this production may change in response to perturbations. Predicting the effects of climate change on primary productivity is complicated by phytoplankton bloom cycles that are affected by both bottom-up control (for example, availability of essential nutrients and vertical mixing) and top-down control (for example, grazing and viruses).[58][57][59][60][61][62] Increases in solar radiation, temperature and freshwater inputs to surface waters strengthen ocean stratification and consequently reduce transport of nutrients from deep water to surface waters, which reduces primary productivity.[57][62][63] Conversely, rising CO2 levels can increase phytoplankton primary production, but only when nutrients are not limiting.[64][65][66][24]

Some studies indicate that overall global oceanic phytoplankton density has decreased in the past century,[67] but these conclusions have been questioned because of the limited availability of long-term phytoplankton data, methodological differences in data generation and the large annual and decadal variability in phytoplankton production.[68][69][70][71] Moreover, other studies suggest a global increase in oceanic phytoplankton production[72] and changes in specific regions or specific phytoplankton groups.[73][74] The global Sea Ice Index is declining,[75] leading to higher light penetration and potentially more primary production;[76] however, there are conflicting predictions for the effects of variable mixing patterns and changes in nutrient supply and for productivity trends in polar zones.[62][24]

The effect of human-caused climate change on phytoplankton biodiversity is not well understood. Should greenhouse gas emissions continue rising to high levels by 2100, some phytoplankton models predict an increase in species richness, or the number of different species within a given area. This increase in plankton diversity is traced to warming ocean temperatures. In addition to species richness changes, the locations where phytoplankton are distributed are expected to shift towards the Earth's poles. Such movement may disrupt ecosystems, because phytoplankton are consumed by zooplankton, which in turn sustain fisheries. This shift in phytoplankton location may also diminish the ability of phytoplankton to store carbon that was emitted by human activities. Human (anthropogenic) changes to phytoplankton impact both natural and economic processes.[77]

Image Gallery

See also

References

  1. ]
  2. ^ .
  3. .
  4. ^ Ghosal; Rogers; Wray, S.; M.; A. "The Effects of Turbulence on Phytoplankton". Aerospace Technology Enterprise. NTRS. Retrieved 16 June 2011.{{cite web}}: CS1 maint: multiple names: authors list (link)
  5. ^ a b Modeled Phytoplankton Communities in the Global Ocean NASA Hyperwall, 30 September 2015. Public Domain This article incorporates text from this source, which is in the public domain.
  6. PMID 18983264
    .
  7. ^ Lindsey, R.; Scott, M.; Simmon, R. (16 July 2010). "What are Phytoplankton?". NASA Earth Observatory.
  8. ^ "MIT Darwin Project". Massachusetts Institute of Technology.
  9. S2CID 38043167
    .
  10. ^ "NASA Satellite Detects Red Glow to Map Global Ocean Plant Health" Archived 10 April 2021 at the Wayback Machine NASA, 28 May 2009.
  11. ^ "Satellite Sees Ocean Plants Increase, Coasts Greening". NASA. 2 March 2005. Archived from the original on 29 October 2011. Retrieved 9 June 2014.
  12. PMID 26927496
    .
  13. .
  14. .
  15. .
  16. ^
    doi:10.1007/978-3-319-93284-2_5. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License
    .
  17. ^ Richtel, M. (1 May 2007). "Recruiting Plankton to Fight Global Warming". The New York Times.
  18. JSTOR 4018225
    .
  19. ^ Sañudo-Wilhelmy, Sergio (23 June 2012). "Existence of vitamin 'deserts' in the ocean confirmed". ScienceDaily.
  20. .
  21. .
  22. .
  23. .
  24. ^
    PMID 31213707. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License
    .
  25. ^ .
  26. .
  27. .
  28. .
  29. .
  30. .
  31. .
  32. .
  33. .
  34. ^ .
  35. .
  36. ]
  37. .
  38. .
  39. ^ .
  40. ^ .
  41. .
  42. .
  43. ^ NAAMES: Science - Objectives Langley Research Center, NASA, Updated: 6 June 2020. Retrieved: 15 June 2020.
  44. doi:10.1126/sciadv.aau6253. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License
    .
  45. ^ .
  46. doi:10.3390/su10030869. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License
    .
  47. .
  48. .
  49. .
  50. .
  51. .
  52. ^
    S2CID 210308256. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License
    .
  53. .
  54. ^ a b c d McVey, James P., Nai-Hsien Chao, and Cheng-Sheng Lee. CRC Handbook of Mariculture Vol. 1 : Crustacean Aquaculture. New York: CRC Press LLC, 1993.[page needed]
  55. ^
    PMID 29434313
    .
  56. ^ "Nutrition study reveals instability in world's most important fishing regions".
  57. ^ .
  58. .
  59. .
  60. S2CID 2476669. Archived from the original
    (PDF) on 2 November 2020. Retrieved 29 October 2020.
  61. .
  62. ^ .
  63. .
  64. .
  65. .
  66. .
  67. .
  68. .
  69. .
  70. .
  71. .
  72. .
  73. .
  74. .
  75. ^ Sea Ice Index National Snow and Ice Data Center. Accessed 30 October 2020.
  76. S2CID 31230080
    .
  77. .

Further reading

External links