From Wikipedia, the free encyclopedia


Map showing 5 circles. The first is between western Australia and eastern Africa. The second is between eastern Australia and western South America. The third is between Japan and western North America. Of the two in the Atlantic, one is in hemisphere.
North Atlantic
gyre
North Atlantic
gyre
North Atlantic
gyre
Indian
Ocean
gyre
North
Pacific
gyre
South
Pacific
gyre
South Atlantic
        gyre
Map showing 5 circles. The first is between western Australia and eastern Africa. The second is between eastern Australia and western South America. The third is between Japan and western North America. Of the two in the Atlantic, one is in hemisphere.
World map of the five major ocean gyres

In oceanography, a gyre ( /ˈaɪər/) is any large system of circulating ocean surface currents, particularly those involved with large wind movements. Gyres are caused by the Coriolis effect; planetary vorticity, horizontal friction and vertical friction determine the circulatory patterns from the wind stress curl. [1]

Gyre can refer to any type of vortex in an atmosphere or a sea, [2] but it is most commonly used in terrestrial oceanography to refer to the gyres that control the major ocean basin circulation.

Gyre Formation

Ocean gyres are wind-driven circulation, meaning that their locations and dynamics are controlled by the prevailing global wind patterns: easterlies at the tropics and westerlies at the midlatitudes. These wind patterns result in a wind stress curl that drives Ekman pumping in the subtropics and Ekman suction in subpolar regions [3]. Ekman pumping results in an increased sea surface height at the center of the gyre and anticyclonic geostrophic currents in subtropical gyres [3]. Ekman suction results in a depressed sea surface height and cyclonic geostrophic currents in subpolar gyres [3].

Ocean gyres are asymmetrical, with stronger flows on their western boundary and weaker flows throughout their interior. The weak flow that is typical over most of the gyre is a result of the conservation of potential vorticity. In the shallow water equations (applicable for basin-scale flow as the horizontal length scale is much greater than the vertical length scale), potential vorticity is a function of relative (local) vorticity , planetary vorticity , and the depth , and is conserved with respect to the material derivative [4]:

In the case of the subtropical ocean gyre, Ekman pumping results in water piling up in the center of the gyre, compressing water parcels. This results in a decrease in , so by the conservation of potential vorticity the numerator must also decrease [5]. It can be further simplified by realizing that, in basin-scale ocean gyres, the relative vorticity is small, meaning that local changes in vorticity cannot account for the decrease in [5]. Thus, the water parcel must change its planetary vorticity accordingly. The only way to decrease the planetary vorticity is by moving the water parcel equatorward, so throughout the majority of subtropical gyres there is a weak equatorward flow. Harald Sverdrup quantified this phenomena in his 1947 paper, "Wind Driven Currents in a Baroclinic Ocean" [6], in which the (depth-integrated) Sverdrup balance is defined as [7]:

Here, is the meridional mass transport (positive north), is the Rossby parameter, is the water density, and is the vertical Ekman velocity due to wind stress curl (positive up). It can be clearly seen in this equation that for a negative Ekman velocity (e.g., Ekman pumping in subtropical gyres), meridional mass transport (Sverdrup transport) is negative (south, equatorward) in the northern hemisphere (). Conversely, for a positive Ekman velocity (e.g., Ekman suction in subpolar gyres), Sverdrup transport is positive (north, poleward) in the northern hemisphere.

Western Intensification

As the Sverdrup balance argues, subtropical (subpolar) ocean gyres have a weak equatorward (poleward) flow over their area. However, there must be some return flow that goes against the Sverdrup transport in order to preserve mass balance [8]. In this respect, the Sverdrup solution is incomplete, as it has no mechanism in which to predict this return flow [8]. Contributions by both Henry Stommel and Walter Munk resolved this issue by showing that the return flow of gyres is done through an intensified western boundary current [9] [10]. Stommel's solution relies on a frictional bottom boundary layer which is not necessarily physical in a stratified ocean (currents do not always extend to the bottom) [5].

Munk's solution instead relies on friction between the return flow and the sidewall of the basin [5]. This allows us to consider two cases: one with the return flow on the western boundary (western boundary current) and one with the return flow on the eastern boundary (eastern boundary current). A qualitative argument for the presence of western boundary current solutions over eastern boundary current solutions can be found again through the conservation of potential vorticity. In order to move northward (an increase planetary vorticity for the case of a subtropical gyre in the northern hemisphere), there must be a source of relative vorticity to drive the northward flow. The relative vorticity in the shallow-water system is [11]:

Here is again the meridional velocity and is the zonal velocity. In the sense of a northward return flow, the zonal component is neglected and only the meridional velocity is important for relative vorticity. Thus, this solution requires that in order to increase the relative vorticity and have a valid northward return flow in the northern hemisphere subtropical gyre [5].

Due to friction at the boundary, the velocity of flow must go to zero at the sidewall before reaching some maximum northward velocity within the boundary layer and decaying to the southward Sverdrup transport solution far away from the boundary. Thus, the condition that can only be satisfied through a frictional boundary layer on the western boundary, as the eastern boundary frictional layer forces [5]. One can make similar arguments for subtropical gyres in the southern hemisphere and for subpolar gyres in either hemisphere and see that the result remains the same: the return flow of an ocean gyre is always in the form of a western boundary current.

The western boundary current must transport on the same order of water as the interior Sverdrup transport in a much smaller area. This means western boundary currents are much stronger than interior currents, a phenomena called "western intensification". A quantitative analysis of ocean gyres, western boundary currents, and the so-called Munk Layer can be found in numerous texts [12].

Gyre Distribution

Subtropical Gyres

The five major subtropical gyres are [13]:

They flow clockwise in the Northern hemisphere, and counterclockwise in the Southern hemisphere.

Subpolar gyres

Subpolar circulation in the southern hemisphere is dominated by the Antarctic Circumpolar Current, due to the lack of large landmasses breaking up the Southern Ocean. There are minor gyres in the Weddell Sea and the Ross Sea, the Weddell Gyre and Ross Gyre, which circulate in a clockwise direction. [14]

Indigenous Knowledge

Life in a Gyre

An animation of a year in organism density on Earth. The South Pacific Gyre is visibly low (purple) in organism density.

Gyres are sometimes described as "ocean deserts" or more precisely "biological deserts", a concept that uses the concept of desert in the sense of an environment lacking life and not necessarily water. Other places that are called oceanic deserts are hypoxic or anoxic waters such as dead zones. [15] [16] [17]

Climate change

Ocean circulation re-distributes the heat and water-resources, therefore determines the regional climate. For example, the western branches of the subtropical gyres flow from the lower latitudes towards higher latitudes, bringing relatively warm and moist air to the adjacent land, contributing to a mild and wet climate (e.g., East China, Japan). In contrast, the eastern boundary currents of the subtropical gyres streaming from the higher latitudes towards lower latitudes, corresponding to a relatively cold and dry climate (e.g., California).

Currently, the core of the subtropical gyres are around 30° in both Hemispheres. However, their positions were not always there. Satellite observational sea surface height and sea surface temperature data suggest that the world's major ocean gyres are slowly moving towards higher latitudes in the past few decades. Such feature show agreement with climate model prediction under anthropogenic global warming. [18] Paleo-climate reconstruction also suggest that during the past cold climate intervals, i.e., ice ages, some of the western boundary currents (western branches of the subtropical ocean gyres) are closer to the equator than their modern positions. [19] [20] These evidence implies that global warming is very likely to push the large-scale ocean gyres towards higher latitudes. [21] [22]

Pollution

Trash washed ashore in Hawaii from the Great Pacific Garbage Patch

A garbage patch is a gyre of marine debris particles caused by the effects of ocean currents and increasing plastic pollution by human populations. These human-caused collections of plastic and other debris are responsible for ecosystem and environmental problems that affect marine life, contaminate oceans with toxic chemicals, and contribute to greenhouse gas emissions. Once waterborne, marine debris becomes mobile. Flotsam can be blown by the wind, or follow the flow of ocean currents, often ending up in the middle of oceanic gyres where currents are weakest.

Within garbage patches, the waste is not compact, and although most of it is near the surface of the ocean, it can be found up to more than 30 metres (100 ft) deep in the water. [23] Patches contain plastics and debris in a range of sizes from Microplastics and small scale plastic pellet pollution, to large objects such as fishing nets and consumer goods and appliances lost from flood and shipping loss.

Garbage patches grow because of widespread loss of plastic from human trash collection systems. The United Nations Environmental Program estimated that "for every square mile of ocean" there are about "46,000 pieces of plastic". [24] The 10 largest emitters of oceanic plastic pollution worldwide are, from the most to the least, China, Indonesia, Philippines, Vietnam, Sri Lanka, Thailand, Egypt, Malaysia, Nigeria, and Bangladesh, [25] largely through the rivers Yangtze, Indus, Yellow, Hai, Nile, Ganges, Pearl, Amur, Niger, and the Mekong, and accounting for "90 percent of all the plastic that reaches the world's oceans". [26] [27] Asia was the leading source of mismanaged plastic waste, with China alone accounting for 2.4 million metric tons. [28]

The best known of these is the Great Pacific garbage patch which has the highest density of marine debris and plastic. The Pacific Garbage patch has two mass buildups: the western garbage patch and the eastern garbage patch, the former off the coast of Japan and the latter between Hawaii and California. These garbage patches contain 90 million tonnes (100 million short tons) of debris. [23] Other identified patches include the North Atlantic garbage patch between North America and Africa, the South Atlantic garbage patch located between eastern South America and the tip of Africa, the South Pacific garbage patch located west of South America, and the Indian Ocean garbage patch found east of South Africa listed in order of decreasing size. [29]

See also

References

  1. ^ Heinemann, B. and the Open University (1998) Ocean circulation, Oxford University Press: Page 98
  2. ^ Lissauer, Jack J.; de Pater, Imke (2019). Fundamental Planetary Sciences : physics, chemistry, and habitability. New York: Cambridge University Press. ISBN  978-1108411981.
  3. ^ a b c Talley, Lynne D.; Pickard, George L.; Emery, William J.; Swift, James H. (2011), "Introduction to Descriptive Physical Oceanography", Descriptive Physical Oceanography, Elsevier, pp. 142–145
  4. ^ Gill, Adrian E. (1982). Atmosphere-ocean dynamics. International geophysics series. New York: Academic Press. pp. 231–237. ISBN  978-0-12-283522-3.
  5. ^ a b c d e f Talley, Lynne D.; Pickard, George L.; Emery, William J.; Swift, James H. (2011), "Introduction to Descriptive Physical Oceanography", Descriptive Physical Oceanography, Elsevier, pp. 211–221
  6. ^ Sverdrup, Harald (1947). "Wind-Driven Currents in a Baroclinic Ocean; with Application to the Equatorial Currents of the Eastern Pacific". Proceedings of the National Academy of Sciences. 33 (11): 318–326. doi: 10.1073/pnas.33.11.318. ISSN  0027-8424. PMC  1079064. PMID  16588757.{{ cite journal}}: CS1 maint: PMC format ( link)
  7. ^ Gill, Adrian E. (1982). Atmosphere-ocean dynamics. International geophysics series. New York: Academic Press. pp. 326–328, 465–471. ISBN  978-0-12-283522-3.
  8. ^ a b Pedlosky, Joseph (1987). Geophysical fluid dynamics (2nd ed.). New York: Springer. pp. 263–271. ISBN  978-0-387-96387-7.
  9. ^ Stommel, Henry (1948). "The westward intensification of wind‐driven ocean currents". Eos, Transactions American Geophysical Union. 29 (2): 202–206. doi: 10.1029/tr029i002p00202. ISSN  0002-8606.
  10. ^ Munk, Walter H. (1950-04-01). "ON THE WIND-DRIVEN OCEAN CIRCULATION". Journal of the Atmospheric Sciences. 7 (2): 80–93. doi: 10.1175/1520-0469(1950)007<0080:OTWDOC>2.0.CO;2. ISSN  1520-0469.
  11. ^ Pedlosky, Joseph (1987). Geophysical fluid dynamics (2nd ed.). New York: Springer. pp. 58–65. ISBN  978-0-387-96387-7.
  12. ^ Pedlosky, Joseph (1987). Geophysical fluid dynamics (2nd ed.). New York: Springer. pp. 271–282. ISBN  978-0-387-96387-7.
  13. ^ US Department of Commerce, National Oceanic and Atmospheric Administration. "What is a gyre?". oceanservice.noaa.gov. Retrieved 2023-11-20.
  14. ^ The five most notable gyres Archived 2016-03-04 at the Wayback Machine PowerPoint Presentation
  15. ^ Renfrow, Stephanie (2009-02-06). "An Ocean full of Deserts". Earthdata. Retrieved 2022-11-12.
  16. ^ Reintjes, Greta; Tegetmeyer, Halina E.; Bürgisser, Miriam; Orlić, Sandi; Tews, Ivo; Zubkov, Mikhail; Voß, Daniela; Zielinski, Oliver; Quast, Christian; Glöckner, Frank Oliver; Amann, Rudolf; Ferdelman, Timothy G.; Fuchs, Bernhard M. (2022-09-26). "Microbes in the South Pacific Gyre". Max-Planck-Gesellschaft. Retrieved 2022-11-12.
  17. ^ "What is a dead zone?". NOAA's National Ocean Service. 2019-03-14. Retrieved 2022-11-12.
  18. ^ Poleward shift of the major ocean gyres detected in a warming climate. Geophysical Research Letters, 47, e2019GL085868 doi: 10.1029/2019GL085868
  19. ^ Bard, E., & Rickaby, R. E. (2009). Migration of the subtropical front as a modulator of glacial climate. Nature, 460(7253), 380.
  20. ^ Wind-driven evolution of the north pacific subpolar gyre over the last deglaciation. Geophys. Res. Lett. 47, 208–212 (2020).
  21. ^ Climate Change is Pushing Giant Ocean Currents Poleward Bob Berwyn, 26 February 2020 insideclimatenews.org, accessed 5 December 2021
  22. ^ Major Ocean Currents Drifting Poleward www.loe.org, accessed 5 December 2021
  23. ^ a b "Marine Debris in the North Pacific A Summary of Existing Information and Identification of Data Gaps" (PDF). United States Environmental Protection Agency. 24 July 2015.
  24. ^ Maser, Chris (2014). Interactions of Land, Ocean and Humans: A Global Perspective. CRC Press. pp. 147–48. ISBN  978-1482226393.
  25. ^ Jambeck, Jenna R.; Geyer, Roland; Wilcox, Chris (12 February 2015). "Plastic waste inputs from land into the ocean" (PDF). Science. 347 (6223): 769. Bibcode: 2015Sci...347..768J. doi: 10.1126/science.1260352. PMID  25678662. S2CID  206562155. Archived from the original (PDF) on 22 January 2019. Retrieved 28 August 2018.
  26. ^ Christian Schmidt; Tobias Krauth; Stephan Wagner (11 October 2017). "Export of Plastic Debris by Rivers into the Sea" (PDF). Environmental Science & Technology. 51 (21): 12246–12253. Bibcode: 2017EnST...5112246S. doi: 10.1021/acs.est.7b02368. PMID  29019247. The 10 top-ranked rivers transport 88–95% of the global load into the sea
  27. ^ Franzen, Harald (30 November 2017). "Almost all plastic in the ocean comes from just 10 rivers". Deutsche Welle. Retrieved 18 December 2018. It turns out that about 90 percent of all the plastic that reaches the world's oceans gets flushed through just 10 rivers: The Yangtze, the Indus, Yellow River, Hai River, the Nile, the Ganges, Pearl River, Amur River, the Niger, and the Mekong (in that order).
  28. ^ Robert Lee Hotz (13 February 2015). "Asia Leads World in Dumping Plastic in Seas". Wall Street Journal. Archived from the original on 23 February 2015.
  29. ^ Cózar, Andrés; Echevarría, Fidel; González-Gordillo, J. Ignacio; Irigoien, Xabier; Úbeda, Bárbara; Hernández-León, Santiago; Palma, Álvaro T.; Navarro, Sandra; García-de-Lomas, Juan; Ruiz, Andrea; Fernández-de-Puelles, María L. (2014-07-15). "Plastic debris in the open ocean". Proceedings of the National Academy of Sciences. 111 (28): 10239–10244. Bibcode: 2014PNAS..11110239C. doi: 10.1073/pnas.1314705111. ISSN  0027-8424. PMC  4104848. PMID  24982135.

External links

Category:Aerodynamics Category:Fluid dynamics Category:Oceanic gyres Category:Fisheries science