banner



Example Of An Ocean Trench

Long and narrow depressions of the sea floor

Oceanic chaff is formed at an oceanic ridge, while the lithosphere is subducted dorsum into the asthenosphere at trenches

Oceanic trenches are prominent long, narrow topographic depressions of the ocean floor. They are typically 50 to 100 kilometers (30 to 60 mi) wide and 3 to 4 km (1.9 to 2.v mi) beneath the level of the surrounding oceanic floor, but tin be thousands of kilometers in length. There are about l,000 kilometers (31,000 mi) of oceanic trenches worldwide, mostly around the Pacific Ocean, just besides in the eastern Indian Ocean and a few other locations. The greatest ocean depth measured is in the Challenger Deep of the Mariana Trench, at a depth of xi,034 m (36,201 ft) beneath sea level.

Oceanic trenches are a feature of the Earth'southward distinctive plate tectonics. They mark the locations of convergent plate boundaries, along which lithospheric plates move towards each other at rates that vary from a few millimeters to over ten centimeters per year. Oceanic lithosphere moves into trenches at a global rate of nearly 3 kmii/yr.[1] A trench marks the position at which the flexed, subducting slab begins to descend below another lithospheric slab. Trenches are generally parallel to and near 200 km (120 mi) from a volcanic arc.

Much of the fluid trapped in sediments of the subducting slab returns to the surface at the oceanic trench, producing mud volcanoes and cold seeps. These support unique biomes based on chemotrophic microorganisms. In that location is business that plastic debris is accumulating in trenches and threatening these communities.

Geographic distribution [edit]

In that location are approximately 50,000 km (31,000 mi) of convergent plate margins worldwide. These are mostly located around the Pacific Bounding main, but are also establish in the eastern Indian Ocean, with a few shorter convergent margin segments in other parts of the Indian Ocean, in the Atlantic Body of water, and in the Mediterranean.[2] They are found on the oceanward side of island arcs and Andean-blazon orogens.[three] Globally, there are over 50 major body of water trenches covering an area of 1.9 1000000 kmii or about 0.5% of the oceans.[4]

Trenches are geomorphologically distinct from troughs. Troughs are elongated depressions of the sea floor with steep sides and flat bottoms, while trenches are characterized by a V-shaped profile.[four] Trenches that are partially infilled are sometimes described as troughs (such as the Makran Trough[five]) and sometimes trenches are completely cached and lack bathymetric expression (such as the Cascadia subduction zone,[6] which is completely filled with sediments[seven]) but the key plate tectonics structures that these stand for are those of oceanic trenches. However, many troughs represent dissimilar kinds of tectonic structures, such as the Lesser Antilles Trough, which is the forearc basin of the Bottom Antilles subduction zone;[eight] the New Caledonia trough, which is an extensional sedimentary bowl related to the Tonga-Kermadec subduction zone;[ix] and the Cayman Trough, which is a pull-apart basin inside a transform mistake zone.[10]

Trenches, along with volcanic arcs and Wadati-Benioff zones (zones of earthquakes that dip under the volcanic arc as deeply as 700 kilometers (430 mi)) are diagnostic of convergent plate boundaries and their deeper manifestations, subduction zones.[2] [iii] [11] Here two tectonic plates are globe-trotting into each other at a charge per unit of a few millimeters to over 10 centimeters (4 in) per year. At least one of the plates is oceanic lithosphere, which plunges under the other plate to be recycled in the Earth's curtain. Trenches are related to but distinguished from continental standoff zones (such equally that between India and Asia forming the Himalaya), where continental chaff enters a subduction zone. When buoyant continental crust enters a trench, subduction comes to a halt and the area becomes a zone of continental collision. Features analogous to trenches are associated with collision zones, including peripheral foreland basins, which are sediment-filled foredeeps. Examples of peripheral foreland basins include the floodplains of the Ganges River and the Tigris-Euphrates river system.[2]

History of the term "trench" [edit]

Trenches were not conspicuously defined until the late 1940s and 1950s. The bathymetry of the ocean was poorly known prior to the Challenger expedition of 1872–1876,[12] which took 492 soundings of the deep body of water.[13] At station #225, the expedition discovered Challenger Deep,[14] at present known to be the southern terminate of the Mariana Trench. The laying of transatlantic telegraph cables on the seafloor betwixt the continents during the late 19th and early 20th centuries provided further motivation for improved bathymetry.[fifteen] The term trench, in its modern sense of a prominent elongated depression of the sea bottom, was get-go used by Johnstone in his 1923 textbook An Introduction to Oceanography.[16] [ii]

During the 1920s and 1930s, Felix Andries Vening Meinesz measured gravity over trenches using a newly adult gravimeter that could measure out gravity from aboard a submarine.[11] He proposed the tectogene hypothesis to explain the belts of negative gravity anomalies that were found nigh island arcs. According to this hypothesis, the belts were zones of downwelling of light crustal rock arising from subcrustal convection currents. The tectogene hypothesis was further developed by Griggs in 1939, using an counterpart model based on a pair of rotating drums. Harry Hammond Hess substantially revised the theory based on his geological analysis.[17]

World State of war Two in the Pacific led to dandy improvements of bathymetry, particularly in the western Pacific, and the linear nature of these deeps became articulate. The rapid growth of deep ocean research efforts, specially the widespread use of echosounders in the 1950s and 1960s, confirmed the morphological utility of the term. Important trenches were identified, sampled, and mapped via sonar. The early phase of trench exploration reached its peak with the 1960 descent of the Bathyscaphe Trieste to the bottom of the Challenger Deep. Following Robert S. Dietz' and Harry Hess' promulgation of the seafloor spreading hypothesis in the early on 1960s and the plate tectonic revolution in the belatedly 1960s, the oceanic trench became an of import concept in plate tectonic theory.[11]

Morphology [edit]

The Republic of peru–Chile Trench is located just left of the sharp line between the blueish deep ocean (on the left) and the light blueish continental shelf, forth the due west declension of South America. It runs along an oceanic-continental boundary, where the oceanic Nazca Plate subducts below the continental South American Plate

Oceanic trenches are 50 to 100 kilometers (thirty to sixty mi) wide and have an asymmetric Five-shape, with the steeper slope (8 to 20 degrees) on the inner (overriding) side of the trench and the gentler gradient (around 5 degrees) on the outer (subducting) side of the trench.[eighteen] [xix] The bottom of the trench marks the purlieus between the subducting and overriding plates, known every bit the basal plate boundary shear[twenty] or the subduction décollement.[ii] The depth of the trench depends on the starting depth of the oceanic lithosphere as it begins its plunge into the trench, the bending at which the slab plunges, and the amount of sedimentation in the trench. Both starting depth and subduction angle are greater for older oceanic lithosphere, which is reflected in the deep trenches of the western Pacific. Here the bottoms of the Marianas and the Tonga-Kermadec trenches are up to 10–11 kilometers (6.two–half dozen.8 mi) below sea level. In the eastern Pacific, where the subducting oceanic lithosphere is much younger, the depth of the Republic of peru-Chile trench is around 7 to 8 kilometers (four.3 to 5.0 mi).[21]

Though narrow, oceanic trenches are remarkably long and continuous, forming the largest linear depressions on world. An individual trench tin be thousands of kilometers long.[3] Almost trenches are convex towards the subducting slab, which is attributed to the spherical geometry of the Globe.[22]

The trench asymmetry reflects the dissimilar physical mechanisms that determine the inner and outer slope angle. The outer gradient angle of the trench is determined by the angle radius of the subducting slab, as determined past its elastic thickness. Since oceanic lithosphere thickens with age, the outer slope angle is ultimately adamant past the age of the subducting slab.[23] [xx] The inner slope bending is adamant by the bending of repose of the overriding plate edge.[20] This reflects frequent earthquakes along the trench that prevent oversteepening of the inner slope.[2]

As the subducting plate approaches the trench, it bends slightly upwards before outset its plunge into the depths. As a issue, the outer trench slope is bounded by an outer trench high. This is subtle, often only tens of meters high, and is typically located a few tens of kilometers from the trench centrality. On the outer gradient itself, where the plate begins to bend down into the trench, the upper part of the subducting slab is cleaved past bending faults that give the outer trench slope a horst and graben topography. The formation of these angle faults is suppressed where oceanic ridges or large seamounts are subducting into the trench, but the angle faults cut right across smaller seamounts. Where the subducting slab is only thinly veneered with sediments, the outer slope will oftentimes evidence seafloor spreading ridges oblique to the horst and graben ridges.[20]

Sedimentation [edit]

Trench morphology is strongly modified by the amount of sedimentation in the trench. This varies from practically no sedimentation, every bit in the Tonga-Kermadec trench, to almost completely filled with sediments, equally with the southern Bottom Antilles trench or the eastern Alaskan trench. Sedimentation is largely controlled by whether the trench is near a continental sediment source.[22] The range of sedimentation is well illustrated past the Chilean trench. The north Chile portion of the trench, which lies along the Atacama Desert with its very irksome rate of weathering, is sediment-starved, with from 20 to a few hundred meters of sediments on the trench flooring. The tectonic morphology of this trench segment is fully exposed on the ocean bottom. The key Chile segment of the trench is moderately sedimented, with sediments onlapping onto pelagic sediments or sea basement of the subducting slab, just the trench morphology is still conspicuously discernible. The southern Chile segment of the trench is fully sedimented, to the point where the outer ascent and slope are no longer discernible. Other fully sedimented trenches include the Makran Trough, where sediments are up to 7.5 kilometers (4.7 mi) thick; the Cascadia subduction zone, which is completed cached by 3 to four kilometers (1.9 to two.5 mi) of sediments; and the northernmost Sumatra subduction zone, which is buried nether 6 kilometers (iii.7 mi) of sediments.[24]

Sediments are sometimes transported along the axis of an oceanic trench. The primal Chile trench experiences transport of sediments from source fans along an centric channel.[25] Like transport of sediments has been documented in the Aleutian trench.[ii]

In addition to sedimentation from rivers draining into a trench, sedimentation likewise takes place from landslides on the tectonically steepened inner gradient, often driven by megathrust earthquakes. The Reloca Slide of the central Chile trench is an example of this process.[26]

Erosive versus accretionary margins [edit]

Convergent margins are classified equally erosive or accretionary, and this has a strong influence on the morphology of the inner slope of the trench. Erosive margins, such as the northern Republic of peru-Republic of chile, Tonga-Kermadec, and Mariana trenches, represent to sediment-starved trenches.[3] The subducting slab erodes material from the lower part of the overriding slab, reducing its volume. The edge of the slab experiences subsidence and steepening, with normal faulting. The slope is underlain by relative stiff igneous and metamorphic stone, which maintains a high bending of repose.[27] Over half of all convergent margins are erosive margins.[2]

Accretionary margins, such as the southern Peru-Republic of chile, Cascadia, and Aleutians, are associated with moderately to heavily sedimented trenches. Equally the slab subducts, sediments are "bulldozed" onto the border of the overriding plate, producing an accretionary wedge or accretionary prism. This builds the overriding plate outwards. Because the sediments lack forcefulness, their bending of placidity is gentler than the rock making upwards the inner slope of erosive margin trenches. The inner slope is underlain past imbricated thrust sheets of sediments. The inner slope topography is roughened past localized mass wasting.[27] Cascadia has practically no bathymetric expression of the outer rise and trench, due to complete sediment filling, but the inner trench slope is circuitous, with many thrust ridges. These compete with canyon germination by rivers draining into the trench. Inner trench slopes of erosive margins rarely show thrust ridges.[19]

Accretionary prisms grow in two ways. The kickoff is past frontal accretion, in which sediments are scraped off the downgoing plate and emplaced at the front of the accretionary prism.[2] As the accretionary wedge grows, older sediments further from the trench go increasingly lithified, and faults and other structural features are steepened by rotation towards the trench.[28] The other mechanism for accretionary prism growth is underplating[2] (also known equally basal accretion[29]) of subducted sediments, together with some oceanic chaff, along the shallow parts of the subduction decollement. The Franciscan Group of California is interpreted as an ancient accretionary prism in which underplating is recorded every bit tectonic mélanges and duplex structures.[2]

Earthquakes [edit]

Frequent megathrust earthquakes modify the inner slope of the trench past triggering massive landslides. These exit semicircular landslide scarps with slopes of up to twenty degrees on the headwalls and sidewalls.[30]

Subduction of seamounts and aseismic ridges into the trench may increase aseismic creep and reduce the severity of earthquakes. Contrariwise, subduction of large amounts of sediments may allow ruptures along the subduction décollement to propagate for great distances to produce megathrust earthquakes.[31]

Trench rollback [edit]

Trenches seem positionally stable over time, but scientists believe that some trenches—particularly those associated with subduction zones where ii oceanic plates converge—move backward into the subducting plate.[32] [33] This is called trench rollback or hinge retreat (also hinge rollback) and is i explanation for the existence of dorsum-arc basins.

Forces perpendicular to the slab (the portion of the subducting plate within the drapery) are responsible for steepening of the slab and, ultimately, the movement of the hinge and trench at the surface.[34] These forces arise from the negative buoyancy of the slab with respect to the drapery[35] modified past the geometry of the slab itself.[36] The extension in the overriding plate, in response to the subsequent subhorizontal drape flow from the deportation of the slab, can result in formation of a back-arc basin.[37]

Processes involved [edit]

Several forces are involved in the process of slab rollback. Two forces interim against each other at the interface of the two subducting plates exert forces confronting ane another. The subducting plate exerts a bending force (FPB) that supplies pressure during subduction, while the overriding plate exerts a force against the subducting plate (FTS). The slab pull forcefulness (FSP) is caused by the negative buoyancy of the plate driving the plate to greater depths. The resisting strength from the surrounding mantle opposes the slab pull forces. Interactions with the 660-km discontinuity cause a deflection due to the buoyancy at the phase transition (F660).[36] The unique interplay of these forces is what generates slab rollback. When the deep slab department obstructs the down-going motion of the shallow slab department, slab rollback occurs. The subducting slab undergoes backward sinking due to the negative buoyancy forces causing a retrogradation of the trench hinge along the surface. Upwelling of the mantle around the slab can create favorable weather for the formation of a back-arc basin.[37]

Seismic tomography provides evidence for slab rollback. Results demonstrate high temperature anomalies within the mantle suggesting subducted material is present in the drapery.[38] Ophiolites are viewed every bit evidence for such mechanisms every bit loftier pressure level and temperature rocks are rapidly brought to the surface through the processes of slab rollback, which provides space for the exhumation of ophiolites.

Slab rollback is not ever a continuous process suggesting an episodic nature.[35] The episodic nature of the rollback is explained past a change in the density of the subducting plate, such every bit the arrival of buoyant lithosphere (a continent, arc, ridge, or plateau), a modify in the subduction dynamics, or a alter in the plate kinematics. The age of the subducting plates does non have any upshot on slab rollback.[36] Nearby continental collisions accept an effect on slab rollback. Continental collisions induce mantle flow and extrusion of drape material, which causes stretching and arc-trench rollback.[37] In the area of the Southeast Pacific, there have been several rollback events resulting in the formation of numerous back-arc basins.[35]

Mantle interactions [edit]

Interactions with the mantle discontinuities play a significant part in slab rollback. Stagnation at the 660-km discontinuity causes retrograde slab motion due to the suction forces acting at the surface.[36] Slab rollback induces mantle return flow, which causes extension from the shear stresses at the base of operations of the overriding plate. Equally slab rollback velocities increase, round mantle flow velocities also increment, accelerating extension rates.[34] Extension rates are altered when the slab interacts with the discontinuities within the mantle at 410 km and 660 km depth. Slabs can either penetrate directly into the lower mantle, or can be retarded due to the phase transition at 660 km depth creating a difference in buoyancy. An increment in retrograde trench migration (slab rollback) (2–4 cm/yr) is a result of flattened slabs at the 660-km aperture where the slab does not penetrate into the lower mantle.[39] This is the instance for the Japan, Java and Izu–Bonin trenches. These flattened slabs are merely temporarily arrested in the transition zone. The subsequent displacement into the lower pall is acquired past slab pull forces, or the destabilization of the slab from warming and broadening due to thermal diffusion. Slabs that penetrate direct into the lower curtain upshot in slower slab rollback rates (~1–three cm/twelvemonth) such as the Mariana arc, Tonga arcs.[39]

Hydrothermal activity and associated biomes [edit]

As sediments are subducted at the bottom of trenches, much of their fluid content is expelled and moves back along the subduction décollement to emerge on the inner slope equally mud volcanoes and cold seeps. Methyl hydride clathrates and gas hydrates also accumulate in the inner gradient, and there is concern that their breakdown could contribute to global warming.[2]

The fluids released at mud volcanoes and common cold seeps are rich in methane and hydrogen sulfide, providing chemic free energy for chemotrophic microorganisms that form the base of a unique trench biome. Cold seep communities have been identified in the inner trench slopes of the western Pacific (especially Japan[40]), Southward America, Barbados, the Mediterranean, Makran, and the Sunda trench. These are found at depths as bully as half dozen,000 meters (20,000 ft).[two] The genome of the extremophile Deinococcus from Challenger Deep has sequenced for its ecological insights and potential industrial uses.[41]

Because trenches are the lowest points in the bounding main floor, there is concern that plastic debris may accumulate in trenches and endanger the fragile trench biomes.[42]

Deepest oceanic trenches [edit]

Recent measurements, where the salinity and temperature of the water was measured throughout the swoop, accept uncertainties of about 15 1000 (49 ft).[43] Older measurements may exist off by hundreds of meters.

Trench Ocean Lowest Indicate Maximum Depth Source
Mariana Trench Pacific Ocean Challenger Deep 10,920 m (35,830 ft) [43]
Tonga Trench Pacific Bounding main Horizon Deep 10,820 k (35,500 ft) [43]
Philippine Trench Pacific Ocean Emden Deep 10,540 1000 (34,580 ft) [44]
Kuril–Kamchatka Trench Pacific Bounding main 10,542 chiliad (34,587 ft) [44]
Kermadec Trench Pacific Sea 10,047 m (32,963 ft) [44]
Izu–Bonin Trench (Izu–Ogasawara Trench) Pacific Ocean ix,810 m (32,190 ft) [44]
New Britain Trench Pacific Ocean (Solomon Ocean) Planet Deep nine,140 m (29,990 ft) [45]
Puerto Rico Trench Atlantic Ocean Brownson Deep 8,380 thou (27,490 ft) [43]
S Sandwich Trench Atlantic Bounding main Shooting star Deep 8,265 m (27,116 ft) [43]
Republic of peru–Chile Trench or Atacama Trench Pacific Sea Richards Deep viii,055 m (26,427 ft) [44]
Japan Trench Pacific Sea 8,412 grand (27,598 ft) [44]

Notable oceanic trenches [edit]

Trench Location
Aleutian Trench South of the Aleutian Islands, west of Alaska
Bougainville Trench South of New Guinea
Cayman Trench Western Caribbean
Cedros Trench (inactive) Pacific coast of Baja California
Hikurangi Trench East of New Zealand
Hjort Trench Southwest of New Zealand
Izu–Ogasawara Trench Nigh Izu and Bonin islands
Japan Trench East of Japan
Kermadec Trench * Northeast of New Zealand
Kuril–Kamchatka Trench * Well-nigh Kuril islands
Manila Trench W of Luzon, Philippines
Mariana Trench * Western Pacific Ocean; due east of Mariana Islands
Middle America Trench Eastern Pacific Ocean; off coast of Mexico, Guatemala, Republic of el salvador, Nicaragua, Republic of costa rica
New Hebrides Trench West of Vanuatu (New Hebrides Islands).
Peru–Chile Trench Eastern Pacific Body of water; off declension of Peru & Chile
Philippine Trench * East of the Philippines
Puerto Rico Trench Purlieus of Caribbean and Atlantic sea
Puysegur trench Southwest of New Zealand
Ryukyu Trench Eastern edge of Japan's Ryukyu Islands
South Sandwich Trench East of the Southward Sandwich Islands
Sunda Trench Curves from due south of Java to w of Sumatra and the Andaman and Nicobar Islands
Tonga Trench * Near Tonga
Yap Trench Western Pacific Body of water; between Palau Islands and Mariana Trench

(*) The five deepest trenches in the world

Ancient oceanic trenches [edit]

Trench Location
Intermontane Trench Western North America; betwixt the Intermontane Islands and North America
Insular Trench Western North America; between the Insular Islands and the Intermontane Islands
Farallon Trench Western Due north America
Tethyan Trench South of Turkey, Iran, Tibet and Southeast Asia

See too [edit]

  • Listing of landforms
  • Listing of submarine topographical features
  • Mid-ocean ridge
  • Physical oceanography
  • Band of Fire

References [edit]

  1. ^ Rowley 2002.
  2. ^ a b c d eastward f g h i j k l m Stern 2005.
  3. ^ a b c d Kearey, Klepeis & Vine 2009, p. 250.
  4. ^ a b Harris et al. 2014.
  5. ^ Dastanpour 1996.
  6. ^ Thomas, Burbidge & Cummins 2007.
  7. ^ Goldfinger et al. 2012.
  8. ^ Westbrook, Mascle & Biju-Duval 1984.
  9. ^ Hackney, Sutherland & Collot 2012.
  10. ^ Einsele 2000.
  11. ^ a b c Geersen, Voelker & Behrmann 2018.
  12. ^ Eiseley 1946.
  13. ^ Weyl 1969, p. 49.
  14. ^ Thomson & Murray 1895.
  15. ^ McConnell 1990.
  16. ^ Johnstone 1923.
  17. ^ Allwrardt 1993.
  18. ^ Kearey, Klepeis & Vine 2009, pp. 250–251.
  19. ^ a b Geersen, Voelker & Behrmann 2018, p. 420.
  20. ^ a b c d Geersen, Voelker & Behrmann 2018, pp. 411–412.
  21. ^ Kearey, Klepeis & Vine 2009, pp. =250-251.
  22. ^ a b Kearey, Klepeis & Vine 2009, p. 251.
  23. ^ Bodine & Watts 1979.
  24. ^ Geersen, Voelker & Behrmann 2018, pp. 412–416.
  25. ^ Völker et al. 2013.
  26. ^ Völker et al. 2009.
  27. ^ a b Geersen, Voelker & Behrmann 2018, p. 416.
  28. ^ Kearey, Klepeis & Vine 2009, pp. 264–266.
  29. ^ Bangs et al. 2020.
  30. ^ Völker et al. 2014.
  31. ^ Geersen, Voelker & Behrmann 2018, p. 421.
  32. ^ Dvorkin et al. 1993.
  33. ^ Garfunkel, Anderson & Schubert 1986.
  34. ^ a b Schellart & Moresi 2013.
  35. ^ a b c Schellart, Lister & Toy 2006.
  36. ^ a b c d Nakakuki & Mura 2013.
  37. ^ a b c Flower & Dilek 2003.
  38. ^ Hall & Spakman 2002.
  39. ^ a b Christensen 1996.
  40. ^ Fujikura et al. 2010.
  41. ^ Zhang et al. 2021.
  42. ^ Peng et al. 2020.
  43. ^ a b c d eastward Amos 2021.
  44. ^ a b c d e f Jamieson et al.
  45. ^ Gallo et al. 2015.

Bibliography [edit]

  • Allwrardt, Allan O. (1993). "Evolution of the tectogene concept, 1930-1965" (PDF). Proceedings of the Fifth International Congress on the History of Oceanography . Retrieved 29 September 2021.
  • Amos, Jonathan (xi May 2021). "Oceans' extreme depths measured in precise detail". News. BBC. Retrieved 2 Oct 2021.
  • Bangs, Due north. L.; Morgan, J. K.; Tréhu, A. Thousand.; Contreras‐Reyes, E.; Arnulf, A. F.; Han, South.; Olsen, One thousand. Yard.; Zhang, Eastward. (November 2020). "Basal Accretion Along the S Central Chilean Margin and Its Relationship to Keen Earthquakes". Periodical of Geophysical Research: Solid Earth. 125 (11). Bibcode:2020JGRB..12519861B. doi:ten.1029/2020JB019861. S2CID 225154312.
  • Bodine, J.H.; Watts, A.B> (1979). "On lithospheric flexure seaward of the Bonin and Mariana trenches". Earth and Planetary Science Letters. 43 (1): 132-148. Bibcode:1979E&PSL..43..132B. doi:10.1016/0012-821X(79)90162-6.
  • Christensen, UR (1996). "The Influence of Trench Migration on Slab Penetration into the Lower Mantle". Globe and Planetary Science Letters. 140 (1–4): 27–39. Bibcode:1996E&PSL.140...27C. doi:10.1016/0012-821x(96)00023-4.
  • Dastanpour, Mohammad (March 1996). "The Devonian System in Islamic republic of iran: a review". Geological Magazine. 133 (two): 159–170. Bibcode:1996GeoM..133..159D. doi:x.1017/S0016756800008670. S2CID 129199671.
  • Dvorkin, Jack; Nur, Amos; Mavko, Gary; Ben-Avraham, Zvi (1993). "Narrow subducting slabs and the origin of backarc basins". Tectonophysics. 227 (1–4): 63–79. Bibcode:1993Tectp.227...63D. doi:10.1016/0040-1951(93)90087-Z.
  • Einsele, Gerhard (2000). Sedimentary Basins: Development, Facies, and Sediment Budget (2nd ed.). Springer. p. 630. ISBN978-iii-540-66193-1.
  • Eiseley, Loren (1946). "The Corking Deeps". The Immense Journey (1959 ed.). The states: Vintage Books. p. 38-41. ISBN0394701577.
  • Ellouz-Zimmermann, N.; Deville, E.; Müller, C.; Lallemant, S.; Subhani, A. B.; Tabreez, A. R. (2007). "Bear on of Sedimentation on Convergent Margin Tectonics: Case of the Makran Accretionary Prism (Pakistan)". Thrust Belts and Foreland Basins. Frontiers in Earth Sciences: 327–350. doi:10.1007/978-iii-540-69426-7_17. ISBN978-3-540-69425-0.
  • Fujikura, Yard.; Lindsay, D.; Kitazato, H.; Nishida, Due south.; Shirayama, Y. (2010). "Marine Biodiversity in Japanese Waters". PLoS 1. five (eight): e11836. Bibcode:2010PLoSO...511836F. doi:10.1371/journal.pone.0011836. PMC2914005. PMID 20689840.
  • "Deep-sea trench". McGraw-Colina Encyclopedia of Science & Applied science (8th ed.). 1997.
  • Flower, MFJ; Dilek, Y (2003). "Arc–trench Rollback and Forearc Accretion: ane. A Collision–Induced Pall Menses Model for Tethyan Ophiolites". Pub. Geol. Soc. Lond. 218 (i): 21–41. Bibcode:2003GSLSP.218...21F. doi:ten.1144/gsl.sp.2003.218.01.03. S2CID 128899276.
  • Fisher, R. L. & Hess, H. H. & M. N. Hill (Editor) (1963). "Trenches". The Sea v. 3 The Earth Beneath the Sea. New York: Wiley-Interscience. pp. 411–436. {{cite news}}: CS1 maint: uses authors parameter (link)
  • Gallo, Northward.D.; Cameron, J; Hardy, K.; Fryer, P.; Bartlett, D.H.; Levin, L.A. (2015). "Submersible- and lander-observed community patterns in the Mariana and New Britain trenches: Influence of productivity and depth on epibenthic and scavenging communities". Deep Sea Research Part I: Oceanographic Research Papers. 99: 119–133. Bibcode:2015DSRI...99..119G. doi:10.1016/j.dsr.2014.12.012.
  • Garfunkel, Z; Anderson, C. A.; Schubert, 1000 (10 June 1986). "Mantle circulation and the lateral migration of subducted slabs". Periodical of Geophysical Research: Solid Earth. 91 (B7): 7205–7223. Bibcode:1986JGR....91.7205G. doi:x.1029/JB091iB07p07205.
  • Geersen, Jacob; Voelker, David; Behrmann, Jan H. (2018). "Oceanic Trenches". Submarine Geomorphology. Springer Geology: 409–424. doi:10.1007/978-3-319-57852-1_21. ISBN978-3-319-57851-4.
  • Goldfinger, Chris; Nelson, C. Hans; Morey, Ann E.; Johnson, Joel E.; Patton, Jason R.; Karabanov, Eugene B.; Gutierrez-Pastor, Julia; Eriksson, Andrew T.; Gracia, Eulalia; Dunhill, Gita; Enkin, Randolph J.; Dallimore, Audrey; Vallier, Tracy (2012). Kayen, Robert (ed.). "Turbidite consequence history—Methods and implications for Holocene paleoseismicity of the Cascadia subduction zone". U.Southward. Geological Survey Professional Newspaper. Professional Paper. 1661-E. doi:10.3133/pp1661F.
  • Hackney, Ron; Sutherland, Rupert; Collot, Julien (June 2012). "Rifting and subduction initiation history of the New Caledonia Trough, southwest Pacific, constrained by procedure-oriented gravity models: Gravity modelling of the New Caledonia Trough". Geophysical Periodical International. 189 (3): 1293–1305. doi:ten.1111/j.1365-246X.2012.05441.x.
  • Hall, R; Spakman, W (2002). "Subducted Slabs Beneath the Eastern Indonesia–Tonga Region: Insights from Tomography". Globe and Planetary Scientific discipline Messages. 201 (2): 321–336. Bibcode:2002E&PSL.201..321H. CiteSeerX10.i.1.511.9094. doi:10.1016/s0012-821x(02)00705-7. S2CID 129884170.
  • Hamilton, W. B. (1988). "Plate tectonics and island arcs". Geological Society of America Bulletin. Vol. 100, no. 10. pp. 1503–1527.
  • Harris, P.T.; MacMillan-Lawler, M.; Rupp, J.; Baker, E.K. (2014). "Geomorphology of the oceans". Marine Geology. 352: iv–24. Bibcode:2014MGeol.352....4H. doi:10.1016/j.margeo.2014.01.011.
  • Hawkins, J. Due west.; Bloomer, S. H.; Evans, C. A.; Melchior, J. T. (1984). "Evolution of Intra-Oceanic Arc-Trench Systems". Tectonophysics. 102 (1–4): 175–205. Bibcode:1984Tectp.102..175H. doi:10.1016/0040-1951(84)90013-i.
  • Jamieson, A.J.; Fujii, T.; Mayor, D.J.; Solan`, M.; Priede, I.K. (2010). "Hadal trenches: the ecology of the deepest places on Earth". Trends in Ecology & Evolution. 25 (3): 190–197. doi:10.1016/j.tree.2009.09.009. PMID 19846236.
  • Jarrard, R. D. (1986). "Relations amid subduction parameters". Reviews of Geophysics. 24 (ii): 217–284. Bibcode:1986RvGeo..24..217J. doi:10.1029/RG024i002p00217.
  • Johnstone, James (1923). An Introduction to Oceanography, With Special Reference to Geography and Geophysics. ISBN978-1340399580.
  • Kearey, P.; Klepeis, K.A.; Vine, F.J. (2009). Global tectonics (3rd ed.). Oxford: Wiley-Blackwell. pp. 184–188. ISBN9781405107778.
  • Ladd, J.Due west. & Holcombe, T. Fifty. & Westbrook, M. K. & Edgar, North. T. & Dengo, G. (Editor) & Case, J. (Editor) (1990). "Caribbean area Marine Geology: Active margins of the plate boundary". The Geology of North America, Vol. H, The Caribbean Region. Geological Order of America. pp. 261–290. {{cite news}}: CS1 maint: uses authors parameter (link)
  • Lemenkova, Paulina (2021). "Topography of the Aleutian Trench southward-east off Bowers Ridge, Bering Sea, in the context of the geological development of North Pacific Bounding main". Baltica. 34 (i): 27–46. doi:10.5200/baltica.2021.1.iii. S2CID 247031368. SSRN 3854076. Retrieved 30 September 2021.
  • McConnell, A. (1990). "The art of submarine cable- laying: its contribution to physical oceanography". Deutsche hydrographische Zeitschrift, Erganzungs-heft, (B). 22: 467–473.
  • Nakakuki, T; Mura, E (2013). "Dynamics of Slab Rollback and Induced Back-Arc Basin Formation". World and Planetary Scientific discipline Letters. 361 (B11): 287–297. Bibcode:2013E&PSL.361..287N. doi:10.1016/j.epsl.2012.10.031.
  • Peng, Guyu; Bellerby, Richard; Zhang, Feng; Lord's day, Xuerong; Li, Daoji (January 2020). "The ocean'due south ultimate trashcan: Hadal trenches as major depositories for plastic pollution". Water Inquiry. 168: 115121. doi:10.1016/j.watres.2019.115121. PMID 31605833. S2CID 204122125.
  • Rowley, David B. (2002). "Rate of plate cosmos and destruction: 180 Ma to present". Geological Lodge of America Bulletin. 114 (8): 927–933. Bibcode:2002GSAB..114..927R. doi:10.1130/0016-7606(2002)114<0927:ROPCAD>2.0.CO;2.
  • Schellart, WP; Lister, GS (2004). "Orogenic Curvature: Paleomagnetic and Structural Analyses". Geological Society of America: 237–254.
  • Schellart, WP; Lister, GS; Toy, VG (2006). "A Late Cretaceous and Cenozoic Reconstruction of the Southwest Pacific Region: Tectonics Controlled by Subduction and Slab Rollback Processes". Earth-Scientific discipline Reviews. 76 (3–4): 191–233. Bibcode:2006ESRv...76..191S. doi:10.1016/j.earscirev.2006.01.002.
  • Schellart, WP; Moresi, 50 (2013). "A New Driving Mechanism for Backarc Extension and Backarc Shortening Through Slab Sinking Induced Toroidal and Poloidal Mantle Flow: Results from dynamic subduction models with an overriding plate". Periodical of Geophysical Enquiry. 118 (6): 3221–3248. Bibcode:2013JGRB..118.3221S. doi:10.1002/jgrb.50173.
  • Scholl, D. W.; Scholl, D (1993). "The render of sialic material to the drape indicated by terrigeneous material subducted at convergent margins". Tectonophysics. 219 (ane–3): 163–175. Bibcode:1993Tectp.219..163V. doi:10.1016/0040-1951(93)90294-T.
  • Sibuet, M.; Olu, 1000. (1998). "Biogeography, biodiversity and fluid dependence of deep-ocean cold-seep communities at active and passive margins". Abyssal Research. II (45): 517–567. Bibcode:1998DSRII..45..517S. doi:x.1016/S0967-0645(97)00074-X.
  • Smith, Westward. H. F.; Sandwell, D. T. (1997). "Global bounding main floor topography from satellite altimetry and transport depth soundings". Science. 277 (5334): 1956–1962. doi:x.1126/scientific discipline.277.5334.1956.
  • Stern, R. J. (2002). "Subduction Zones". Reviews of Geophysics. forty (4): 1012–1049. Bibcode:2002RvGeo..40.1012S. doi:10.1029/2001RG000108. S2CID 247695067.
  • Stern, R.J. (2005). "TECTONICS | Body of water Trenches". Encyclopedia of Geology: 428–437. doi:ten.1016/B0-12-369396-9/00141-6. ISBN9780123693969.
  • Thomas, C.; Burbidge, D.; Cummins, P. (2007). A preliminary study into the tsunami run a risk faced past southwest Pacific nations. Risk and Impact Analysis Group, Geoscience Australia. Retrieved 26 September 2021.
  • Thomson, C.Westward.; Murray, J. (1895). "Report on the scientific results of the voyage of H.One thousand.S. Challenger during the years of 1872–76 (folio 877)". 19thcenturyscience.org. Archived from the original on 17 Apr 2012. Retrieved 26 March 2012.
  • Völker, David; Geersen, Jacob; Contreras-Reyes, Eduardo; Sellanes, Javier; Pantoja, Silvio; Rabbel, Wolfgang; Thorwart, Martin; Reichert, Christian; Block, Martin; Weinrebe, Wilhelm Reimer (October 2014). "Morphology and geology of the continental shelf and upper slope of southern Primal Chile (33°South–43°S)" (PDF). International Journal of World Sciences. 103 (seven): 1765–1787. Bibcode:2014IJEaS.103.1765V. doi:ten.1007/s00531-012-0795-y. S2CID 129460412.
  • Völker, D.; Weinrebe, W.; Behrmann, J. H.; Bialas, J.; Klaeschen, D. (2009). "Mass wasting at the base of the south primal Chilean continental margin: The Reloca Slide". Advances in Geosciences. 22: 155–167. Bibcode:2009AdG....22..155V. doi:10.5194/adgeo-22-155-2009.
  • Völker, David; Geersen, Jacob; Contreras-Reyes, Eduardo; Reichert, Christian (2013). "Sedimentary fill up of the Republic of chile Trench (32–46°S): Volumetric distribution and causal factors". Journal of the Geological Society. 170 (5): 723–736. Bibcode:2013JGSoc.170..723V. doi:10.1144/jgs2012-119. S2CID 128432525.
  • Watts, A.B. (2001). Isostasy and Flexure of the Lithosphere. Cambridge University Press. 458p.
  • Weyl, Peter K. (1969). Oceanography: an introduction to the marine environment. New York: Wiley. ISBN978-0471937449. </ref>
  • Westbrook, G.M.; Mascle, A.; Biju-Duval, B. (1984). "Geophysics and the structure of the Lesser Antilles forearc" (PDF). Initial Reports of the Deep Sea Drilling Projection. 78: 23–38. Retrieved 26 September 2021.
  • Wright, D. J.; Bloomer, S. H.; MacLeod, C. J.; Taylor, B.; Goodlife, A. M. (2000). "Bathymetry of the Tonga Trench and Forearc: a map series". Marine Geophysical Researches. 21 (489–511): 2000. Bibcode:2000MarGR..21..489W. doi:10.1023/A:1026514914220. S2CID 6072675.
  • Zhang, Ru-Yi; Huang, Ying; Qin, Wen-Jing; Quan, Zhe-Xue (June 2021). "The complete genome of extracellular protease-producing Deinococcus sp. D7000 isolated from the hadal region of Mariana Trench Challenger Deep". Marine Genomics. 57: 100832. doi:x.1016/j.margen.2020.100832. PMID 33867118. S2CID 229392459.

External links [edit]

  • "HADEX: Enquiry project to explore ocean trenches". Woods Pigsty Oceanographic Institution.
  • "Body of water Trenches". Woods Hole Oceanographic Institution.

Example Of An Ocean Trench,

Source: https://en.wikipedia.org/wiki/Oceanic_trench

Posted by: lipseyforged.blogspot.com

0 Response to "Example Of An Ocean Trench"

Post a Comment

Iklan Atas Artikel

Iklan Tengah Artikel 1

Iklan Tengah Artikel 2

Iklan Bawah Artikel