Subduction

From Wikipedia, the free encyclopedia

This is an old revision of this page, as edited by 2601:541:4580:8500:582c:f3cc:239b:47e0 (talk) at 18:29, 17 June 2021 (change link). The present address (URL) is a permanent link to this revision, which may differ significantly from the current revision.

Diagram of the geological process of subduction

Subduction is a geological process in which the oceanic lithosphere is recycled into the Earth's mantle at convergent boundaries. Where the oceanic lithosphere of a tectonic plate converges with the less dense lithosphere of a second plate, the heavier plate dives beneath the second plate and sinks into the mantle. A region where this process occurs is known as a subduction zone, and its surface expression is known as an arc-trench complex. The process of subduction has created most of the Earth's continental crust.[1] Rates of subduction are typically measured in centimeters per year, with the average rate of convergence being approximately two to eight centimeters per year along most plate boundaries.[2]

Subduction is possible because the cold oceanic lithosphere is slightly more dense than the underlying asthenosphere, the hot, ductile layer in the upper mantle underlying the cold, rigid lithosphere. Once initiated, stable subduction is driven mostly by the negative buoyancy of the dense subducting lithosphere. The slab sinks into the mantle largely under its weight.[3]

Earthquakes are common along the subduction zone, and fluids released by the subducting plate trigger volcanism in the overriding plate. If the subducting plate sinks at a shallow angle, the overriding plate develops a belt of deformation characterized by crustal thickening, mountain building, and metamorphism. Subduction at a steeper angle is characterized by the formation of back-arc basins.[4]

Subduction and plate tectonics

Oceanic plates are subducted creating oceanic trenches.

According to the theory of plate tectonics, the Earth's lithosphere, its rigid outer shell, is broken into sixteen larger tectonic plates and several smaller plates. These are in slow motion, due to convection in the underlying ductile mantle. This process of convection allows heat generated by radioactive decay to escape from the Earth's interior.[5]

The lithosphere consists of the outermost light crust plus the uppermost rigid portion of the mantle. Oceanic lithosphere ranges in thickness from just a few km for young lithosphere created at mid-ocean ridges to around 100 km (62 mi) for the oldest oceanic lithosphere.[6] Continental lithosphere is up to 200 km (120 mi) thick.[7] The lithosphere is relatively cold and rigid compared with the underlying asthenosphere, and so tectonic plates move as solid bodies atop the asthenosphere. Individual plates often include both regions of the oceanic lithosphere and continental lithosphere.

Subduction zones are where the cold oceanic lithosphere sinks back into the mantle and is recycled.[4][8] They are found at convergent plate boundaries, where the oceanic lithosphere of one plate converges with the less dense lithosphere of another plate. The heavier oceanic lithosphere is overridden by the leading edge of the other plate.[6] The overridden plate (the slab) sinks at an angle of approximately twenty-five to seventy-five degrees to Earth's surface.[9] This sinking is driven by the temperature difference between the slab and the surrounding asthenosphere, as the colder oceanic lithosphere has, on average, a greater density.[6] Sediments and some trapped water are carried downwards by the slab and recycled into the deep mantle.[10]

Earth is so far the only planet where subduction is known to occur, and subduction zones are its most important tectonic feature. Subduction is the driving force behind plate tectonics, and without it, plate tectonics could not occur.[11] Oceanic subduction zones are located along 55,000 km (34,000 mi) of convergent plate margins,[12] almost equal to the cumulative 60,000 km (37,000 mi) of mid-ocean ridges.[13]

Structure of subduction zones

Arc-trench complex

The surface expression of subduction zones are arc-trench complexes. On the ocean side of the complex, where the subducting plate first approaches the subduction zone, there is often an outer trench high or outer trench swell. Here the plate shallows slightly before plunging downwards, as a consequence of the rigidity of the plate.[14] The point where the slab begins to plunge downwards is marked by an oceanic trench. Oceanic trenches are the deepest parts of the ocean floor.

Beyond the trench is the forearc portion of the overriding plate. Depending on sedimentation rates, the forearc may include an accretionary wedge of sediments scraped off the subducting slab and accreted to the overriding plate. However, not all arc-trench complexes have an accretionary wedge. Accretionary arcs have a well-developed forearc basin behind the accretionary wedge, while the forearc basin is poorly developed in non-accretionary arcs.[15]

Beyond the forearc basin, volcanoes are found in long chains called volcanic arcs. The subducting basalt and sediment are normally rich in hydrous minerals and clays. Additionally, large quantities of water are introduced into cracks and fractures created as the subducting slab bends downward.[16] During the transition from basalt to eclogite, these hydrous materials break down, producing copious quantities of water, which at such great pressure and temperature exists as a supercritical fluid.[17] The supercritical water, which is hot and more buoyant than the surrounding rock, rises into the overlying mantle, where it lowers the melting temperature of the mantle rock, generating magma via flux melting.[18] The magmas, in turn, rise as diapirs because they are less dense than the rocks of the mantle.[19] The mantle-derived magmas (which are initially basaltic in composition) can ultimately reach the Earth's surface, resulting in volcanic eruptions. The chemical composition of the erupting lava depends upon the degree to which the mantle-derived basalt interacts with (melts) Earth's crust or undergoes fractional crystallization. Arc volcanoes tend to produce dangerous eruptions because they are rich in water (from the slab and sediments) and tend to be extremely explosive.[20] Krakatoa, Nevado del Ruiz, and Mount Vesuvius are all examples of arc volcanoes. Arcs are also associated with most ore deposits.[19]

Beyond the volcanic arc is a back-arc region whose character depends strongly on the angle of subduction of the subducting slab. Where this angle is shallow, the subducting slab drags the overlying continental crust, producing a zone of compression in which there may be extensive folding and thrust faulting. If the angle of subduction is deep, the crust will be put in tension instead, often producing a back-arc basin.[21]

Deep structure

The arc-trench complex is the surface expression of a much deeper structure. Though not directly accessible, the deeper portions can be studied using geophysics and geochemistry. Subduction zones are defined by an inclined zone of earthquakes, the Wadati–Benioff zone, that dips away from the trench and extends down to the 660-kilometer discontinuity. Subduction zone earthquakes occur at greater depths (up to 600 km (370 mi)) than elsewhere on Earth (typically less than 20 km (12 mi) depth); such deep earthquakes may be driven by deep phase transformations, thermal runaway, or dehydration embrittlement.[22][23] Seismic tomography shows that some slabs can penetrate the lower mantle[24][25] and sink clear to the core–mantle boundary.[26] Here the residue of the slabs may eventually heat enough to rise back to the surface as mantle plumes.[27][28]

Subduction angle

Subduction typically occurs at a moderately steep angle right at the point of the convergent plate boundary. However, anomalous shallower angles of subduction are known to exist as well as some that are extremely steep.[29]

  • Flat-slab subduction (subducting angle less than 30°) occurs when the slab subducts nearly horizontally. The relatively flat slab can extend for hundreds of kilometers. That is abnormal, as the dense slab typically sinks at a much steeper angle. Because subduction of slabs to depth is necessary to drive subduction zone volcanism, flat-slab subduction can be invoked to explain volcanic gaps.

Flat-slab subduction is ongoing beneath part of the Andes, causing segmentation of the Andean Volcanic Belt into four zones. The flat-slab subduction in northern Peru and the Norte Chico region of Chile is believed to be the result of the subduction of two buoyant aseismic ridges, the Nazca Ridge and the Juan Fernández Ridge, respectively. Around Taitao Peninsula flat-slab subduction is attributed to the subduction of the Chile Rise, a spreading ridge.[30][31]

The Laramide Orogeny in the Rocky Mountains of the United States is attributed to flat-slab subduction.[32] During this orogeny, a broad volcanic gap appeared at the southwestern margin of North America, and deformation occurred much farther inland; it was during this time that the basement-cored mountain ranges of Colorado, Utah, Wyoming, South Dakota, and New Mexico came into being. The most massive subduction zone earthquakes, so-called "megaquakes", have been found to occur in flat-slab subduction zones.[33]

  • Steep-angle subduction (subducting angle greater than 70°) occurs in subduction zones where Earth's oceanic crust and lithosphere are old and thick and have, therefore, lost buoyancy. The steepest dipping subduction zone lies in the Mariana Trench, which is also where the oceanic crust, of Jurassic age, is the oldest on Earth exempting ophiolites. Steep-angle subduction is, in contrast to flat-slab subduction, associated with back-arc extension[34] of crust, creating volcanic arcs and pulling fragments of continental crust away from continents to leave behind a marginal sea.

Life cycle of subduction zones

Initiation of subduction

Although stable subduction is fairly well understood, the process by which subduction is initiated remains a matter of discussion and continuing study. Subduction can begin spontaneously if the denser oceanic lithosphere can founder and sink beneath the adjacent oceanic or continental lithosphere through vertical forcing only; alternatively, existing plate motions can induce new subduction zones by horizontally forcing the oceanic lithosphere to rupture and sink into the asthenosphere.[35][36] Both models can eventually yield self-sustaining subduction zones, as the oceanic crust is metamorphosed at great depth and becomes denser than the surrounding mantle rocks. The compilation of subduction zone initiation events back to 100 Ma suggests horizontally-forced subduction zone initiation for most modern subduction zones,[36] which is supported by results from numerical models[37][38] and geologic studies.[39][40] Some analogue modeling shows, however, the possibility of spontaneous subduction from inherent density differences between two plates at specific locations like passive margins.[41][42] There is evidence this has taken place in the Izu-Bonin-Mariana subduction system.[43][44] Earlier in Earth's history, subduction is likely to have initiated without horizontal forcing due to the lack of relative plate motion, though an unorthodox proposal by A. Yin suggests that meteorite impacts may have contributed to subduction initiation on early Earth.[45]

End of subduction

Subduction can continue as long as the oceanic lithosphere moves into the subduction zone. However, the arrival of buoyant crust at a subduction zone can result in its failure, by disrupting downwelling. The arrival of continental crust results in a collision or terrane accretion that disrupts subduction.[46] Continental crust can subduct to depths of 100 km (62 mi) or more but then resurfaces.[47][28] Sections of crustal or intraoceanic arc crust greater than 15 km (9.3 mi) in thickness or oceanic plateau greater than 30 km (19 mi) in thickness can disrupt subduction. However, island arcs subducted end-on may cause only local disruption, while an arc arriving parallel to the zone can shut it down.[46] This has happened with the Ontong Java Plateau and the Vitiaz Trench.[48]

Effects

Metamorphism

Subduction zones host a unique variety of rock types created by the high-pressure, low-temperature conditions a subducting slab encounters during its descent.[49] The metamorphic conditions the slab passes through in this process creates and destroys water bearing (hydrous) mineral phases, releasing water into the mantle. This water lowers the melting point of mantle rock, initiating melting.[50] Understanding the timing and conditions in which these dehydration reactions occur, is key to interpreting mantle melting, volcanic arc magmatism, and the formation of continental crust.[51]

A metamorphic facies is characterized by a stable mineral assemblage specific to a pressure-temperature range and specific starting material. Subduction zone metamorphism is characterized by a low temperature, high-ultrahigh pressure metamorphic path through the zeolite, prehnite-pumpellyite, blueschist, and eclogite facies stability zones of subducted oceanic crust.[52] Zeolite and prehnite-pumpellyite facies assemblages may or may not be present, thus the onset of metamorphism may only be marked by blueschist facies conditions.[53] Subducting slabs are composed of basaltic crust topped with pelagic sediments;[54] however, the pelagic sediments may be accreted onto the forearc-hanging wall and not subducted.[55] Most metamorphic phase transitions that occur within the subducting slab are prompted by the dehydration of hydrous mineral phases. The breakdown of hydrous mineral phases typically occurs at depths greater than 10 km.[56] Each of these metamorphic facies is marked by the presence of a specific stable mineral assemblage, recording the metamorphic conditions undergone but the subducting slab. Transitions between facies causes hydrous minerals to dehydrate at certain pressure-temperature conditions and can therefore be tracked to melting events in the mantle beneath a volcanic arc.

Volcanic activity

Volcanoes that occur above subduction zones, such as Mount St. Helens, Mount Etna, and Mount Fuji, lie approximately one hundred kilometers from the trench in arcuate chains called volcanic arcs. Two kinds of arcs are generally observed on Earth: island arcs that form on the oceanic lithosphere (for example, the Mariana and the Tonga island arcs), and continental arcs such as the Cascade Volcanic Arc, that form along the coast of continents. Island arcs (intraoceanic or primitive arcs) are produced by the subduction of oceanic lithosphere beneath another oceanic lithosphere (ocean-ocean subduction) while continental arcs (Andean arcs) form during the subduction of oceanic lithosphere beneath a continental lithosphere (ocean-continent subduction).[57] An example of a volcanic arc having both island and continental arc sections is found behind the Aleutian Trench subduction zone in Alaska.[58]

The arc magmatism occurs one hundred to two hundred kilometers from the trench and approximately one hundred kilometers above the subducting slab. This depth of arc magma generation is the consequence of the interaction between hydrous fluids, released from the subducting slab, and the arc mantle wedge that is hot enough to melt with the addition of water.[59] It has also been suggested that the mixing of fluids from a subducted tectonic plate and melted sediment is already occurring at the top of the slab before any mixing with the mantle takes place.[60]

Arcs produce about 10% of the total volume of magma produced each year on Earth (approximately 0.75 cubic kilometers), much less than the volume produced at mid-ocean ridges,[61] but they have formed most continental crust.[4] Arc volcanism has the greatest impact on humans because many arc volcanoes lie above sea level and erupt violently. Aerosols injected into the stratosphere during violent eruptions can cause rapid cooling of Earth's climate and affect air travel.[59]

Earthquakes and tsunamis

Global map of subduction zones, with subducted slabs contoured by depth

The strains caused by plate convergence in subduction zones cause at least three types of earthquakes. These are deep earthquakes, megathrust earthquakes, and outer rise earthquakes.

Anomalously deep events are a characteristic of subduction zones, which produce the deepest quakes on the planet. Earthquakes are generally restricted to the shallow, brittle parts of the crust, generally at depths of less than twenty kilometers. However, in subduction zones, quakes occur at depths as great as 700 km (430 mi). These quakes define inclined zones of seismicity known as Wadati–Benioff zones which trace the descending slab.[62]

Nine of the ten largest earthquakes of the last 100 years were subduction zone megathrust earthquakes, which included the 1960 Great Chilean earthquake, which, at M 9.5, was the largest earthquake ever recorded; the 2004 Indian Ocean earthquake and tsunami; and the 2011 Tōhoku earthquake and tsunami. The subduction of cold oceanic crust into the mantle depresses the local geothermal gradient and causes a larger portion of Earth to deform in a more brittle fashion than it would in a normal geothermal gradient setting. Because earthquakes can occur only when a rock is deforming in a brittle fashion, subduction zones can cause large earthquakes. If such a quake causes rapid deformation of the sea floor, there is potential for tsunamis, such as the earthquake caused by subduction of the Indo-Australian Plate under the Euro-Asian Plate on December 26, 2004, that devastated the areas around the Indian Ocean. Small tremors which cause small, nondamaging tsunamis, also occur frequently.[62]

A study published in 2016 suggested a new parameter to determine a subduction zone's ability to generate mega-earthquakes.[63] By examining subduction zone geometry and comparing the degree of curvature of the subducting plates in great historical earthquakes such as the 2004 Sumatra-Andaman and the 2011 Tōhoku earthquake, it was determined that the magnitude of earthquakes in subduction zones is inversely proportional to the degree of the fault's curvature, meaning that "the flatter the contact between the two plates, the more likely it is that mega-earthquakes will occur."[64]

Outer rise earthquakes occur when normal faults oceanward of the subduction zone are activated by flexure of the plate as it bends into the subduction zone.[65] The 2009 Samoa earthquake is an example of this type of event. Displacement of the sea floor caused by this event generated a six-meter tsunami in nearby Samoa.

Seismic tomography has helped detect subducted lithosphere, slabs, deep in the mantle where there are no earthquakes. About one hundred slabs have been described in terms of depth and their timing and location of subduction.[66] The great seismic discontinuities in the mantle, at 410 km (250 mi) depth and 670 km (420 mi), are disrupted by the descent of cold slabs in deep subduction zones. Some subducted slabs seem to have difficulty penetrating the major discontinuity that marks the boundary between the upper mantle and lower mantle at a depth of about 670 kilometers. Other subducted oceanic plates have sunk to the core–mantle boundary at 2890 km depth. Generally, slabs decelerate during their descent into the mantle, from typically several cm/yr (up to ~10 cm/yr in some cases) at the subduction zone and in the uppermost mantle, to ~1 cm/yr in the lower mantle.[66] This leads to either folding or stacking of slabs at those depths, visible as thickened slabs in Seismic tomography. Below ~1700 km, there might be a limited acceleration of slabs due to lower viscosity as a result of inferred mineral phase changes until they approach and finally stall at the core–mantle boundary.[66] Here the slabs are heated up by the ambient heat and are not detected anymore ~300 Myr after subduction.[66]

Orogeny

Orogeny is the process of mountain building. Subducting plates can lead to orogeny by bringing oceanic islands, oceanic plateaus, and sediments to convergent margins. The material often does not subduct with the rest of the plate but instead is accreted (scraped off) to the continent, resulting in exotic terranes. The collision of this oceanic material causes crustal thickening and mountain-building. The accreted material is often referred to as an accretionary wedge or prism. These accretionary wedges can be identified by ophiolites (uplifted ocean crust consisting of sediments, pillow basalts, sheeted dykes, gabbro, and peridotite).[67]

Subduction may also cause orogeny without bringing in oceanic material that collides with the overriding continent. When the subducting plate subducts at a shallow angle underneath a continent (something called "flat-slab subduction"), the subducting plate may have enough traction on the bottom of the continental plate to cause the upper plate to contract to lead to folding, faulting, crustal thickening, and mountain building. Flat-slab subduction causes mountain building and volcanism moving into the continent, away from the trench, and has been described in North America (i.e. Laramide orogeny), South America, and East Asia.[66]

The processes described above allow subduction to continue while mountain building happens progressively, which is in contrast to continent-continent collision orogeny, which often leads to the termination of subduction.

Beginnings of subduction on Earth

Modern-style subduction is characterized by low geothermal gradients and the associated formation of high-pressure low-temperature rocks such as eclogite and blueschist.[68][69] Likewise, rock assemblages called ophiolites, associated with modern-style subduction, also indicate such conditions.[68] Eclogite xenoliths found in the North China Craton provide evidence that modern-style subduction occurred at least as early as 1.8 Ga ago in the Paleoproterozoic Era.[68] Nevertheless, the eclogite itself was produced by oceanic subduction during the assembly of supercontinents at about 1.9–2.0 Ga.

Blueschist is a rock typical for present-day subduction settings. The absence of blueschist older than Neoproterozoic reflects more magnesium-rich compositions of Earth's oceanic crust during that period.[70] These more magnesium-rich rocks metamorphose into greenschist at conditions when modern oceanic crust rocks metamorphose into blueschist.[70] The ancient magnesium-rich rocks mean that Earth's mantle was once hotter, but not that subduction conditions were hotter. Previously, the lack of pre-Neoproterozoic blueschist was thought to indicate a different type of subduction.[70] Both lines of evidence refute previous conceptions of modern-style subduction having been initiated in the Neoproterozoic Era 1.0 Ga ago.[68][70]

History of investigation

Harry Hammond Hess, who during World War II served in the United States Navy Reserve and became fascinated in the ocean floor, studied the Mid-Atlantic Ridge and proposed that hot molten rock was added to the crust at the ridge and expanded the seafloor outward. This theory was to become known as seafloor spreading. Since the Earth's circumference has not changed over geologic time, Hess concluded that older seafloor has to be consumed somewhere else, and suggested that this process takes place at oceanic trenches, where the crust would be melted and recycled in the Earth's mantle.[71]

In 1964, George Plafker researched the Good Friday earthquake in Alaska. He concluded that the cause of the earthquake was a megathrust reaction in the Aleutian Trench, a result of the Alaskan continental crust overlapping the Pacific oceanic crust. This meant that the Pacific crust was being forced downward, or subducted, beneath the Alaskan crust. The concept of subduction would play a role in the development of the plate tectonics theory.[72]

First geologic attestations of the "subduct" words date to 1970,[73] In ordinary English to subduct, or to subduce (from Latin subducere, “to lead away”)[74] are transitive verbs requiring a subject to perform an action on an object not itself, here the lower plate, which has then been subducted (“removed”). The geological term is "consumed," which happens the geological moment the lower plate slips under, even though it may persist for some time until its remelting and dissipation. In this conceptual model, plate is continually being used up.[75] The identity of the subject, the consumer, or agent of consumption, is left unstated. Some sources accept this subject-object construct.

Geology makes to subduct into an intransitive verb and a reflexive verb. The lower plate itself is the subject. It subducts, in the sense of retreat, or removes itself, and while doing so, is the "subducting plate." Moreover, the word slab is specifically attached to the "subducting plate," even though in English the upper plate is just as much of a slab.[76] The upper plate is left hanging, so to speak. To express it geology must switch to a different verb, typically to override. The upper plate, the subject, performs the action of overriding the object, the lower plate, which is overridden.[77]

Importance

Subduction zones are important for several reasons:

  • Subduction zone physics: Sinking of the oceanic lithosphere (sediments, crust, mantle), by the contrast of density between the cold and old lithosphere and the hot asthenospheric mantle wedge, is the strongest force (but not the only one) needed to drive plate motion and is the dominant mode of mantle convection.
  • Subduction zone chemistry: The subducted sediments and crust dehydrate and release water-rich (aqueous) fluids into the overlying mantle, causing mantle melting and fractionation of elements between the surface and deep mantle reservoirs, producing island arcs and continental crust. Hot fluids in subduction zones also alter the mineral compositions of the subducting sediments and potentially the habitability of the sediments for microorganisms.[78]
  • Subduction zones drag down subducted oceanic sediments, oceanic crust, and mantle lithosphere that interact with the hot asthenospheric mantle from the over-riding plate to produce calc-alkaline series melts, ore deposits, and continental crust.
  • Subduction zones pose significant threats to lives, property, economic vitality, cultural and natural resources, and quality of life. The tremendous magnitudes of earthquakes or volcanic eruptions can also have knock-on effects with global impact.[79]

Subduction zones have also been considered as possible disposal sites for nuclear waste in which the action of subduction itself would carry the material into the planetary mantle, safely away from any possible influence on humanity or the surface environment. However, that method of disposal is currently banned by international agreement.[80][81][82][83] Furthermore, plate subduction zones are associated with very large megathrust earthquakes, making the effects of using any specific site for disposal unpredictable and possibly adverse to the safety of long-term disposal.[81]

See also

References

  1. ^ Stern, Robert J. (2002), "Subduction zones", Reviews of Geophysics, 40 (4): 1012, Bibcode:2002RvGeo..40.1012S, doi:10.1029/2001RG000108
  2. ^ Defant, M. J. (1998). Voyage of Discovery: From the Big Bang to the Ice Age. Mancorp. p. 325. ISBN 978-0-931541-61-2.
  3. ^ Stern 2002, p. 3.
  4. ^ a b c Stern 2002.
  5. ^ Schmincke, Hans-Ulrich (2003). Volcanism. Berlin: Springer. pp. 13–20. ISBN 9783540436508.
  6. ^ a b c Stern 2002, p. 5.
  7. ^ Rudnick, Roberta L.; McDonough, William F.; O'Connell, Richard J. (April 1998). "Thermal structure, thickness and composition of continental lithosphere". Chemical Geology. 145 (3–4): 395–411. Bibcode:1998ChGeo.145..395R. doi:10.1016/S0009-2541(97)00151-4.
  8. ^ Zheng, YF; Chen, YX (2016). "Continental versus oceanic subduction zones". National Science Review. 3 (4): 495–519. doi:10.1093/nsr/nww049.
  9. ^ Tovish, Aaron; Schubert, Gerald; Luyendyk, Bruce P. (10 December 1978). "Mantle flow pressure and the angle of subduction: Non-Newtonian corner flows". Journal of Geophysical Research: Solid Earth. 83 (B12): 5892–5898. Bibcode:1978JGR....83.5892T. doi:10.1029/JB083iB12p05892.
  10. ^ Stern 2002, p. 15.
  11. ^ Stern 2002, pp. 1–4.
  12. ^ Lallemand, S (1999). La Subduction Oceanique (in French). Newark, New Jersey: Gordon and Breach.
  13. ^ Stern 2002, p. 4.
  14. ^ Whitman, Dean (May 1999). "The Isostatic Residual Gravity Anomaly of the Central Andes, 12° to 29° S: A Guide to Interpreting Crustal Structure and Deeper Lithospheric Processes". International Geology Review. 41 (5): 457–475. Bibcode:1999IGRv...41..457W. doi:10.1080/00206819909465152.
  15. ^ Stern 2002, pp. 25–26.
  16. ^ Fujie, Gou; et al. (2013). "Systematic changes in the incoming plate structure at the Kuril trench". Geophysical Research Letters. 40 (1): 88–93. Bibcode:2013GeoRL..40...88F. doi:10.1029/2012GL054340.
  17. ^ Stern 2002, pp. 6–10.
  18. ^ Schmincke 2003, pp. 18, 113–126.
  19. ^ a b Stern 2002, pp. 19–22.
  20. ^ Stern 2002, p. 27-28.
  21. ^ Stern 2002, p. 31.
  22. ^ Frolich, C. (1989). "The Nature of Deep Focus Earthquakes". Annual Review of Earth and Planetary Sciences. 17: 227–254. Bibcode:1989AREPS..17..227F. doi:10.1146/annurev.ea.17.050189.001303.
  23. ^ Hacker, B.; et al. (2003). "Subduction factory 2. Are intermediate-depth earthquakes in subducting slabs linked to metamorphic dehydration reactions?" (PDF). Journal of Geophysical Research. 108 (B1): 2030. Bibcode:2003JGRB..108.2030H. doi:10.1029/2001JB001129.
  24. ^ Domeier, Mathew; Doubrovine, Pavel V.; Torsvik, Trond H.; Spakman, Wim; Bull, Abigail L. (28 May 2016). "Global correlation of lower mantle structure and past subduction". Geophysical Research Letters. 43 (10): 4945–4953. doi:10.1002/2016GL068827. PMC 6686211. PMID 31413424.
  25. ^ Faccenna, Claudio; Oncken, Onno; Holt, Adam F.; Becker, Thorsten W. (2017). "Initiation of the Andean orogeny by lower mantle subduction". Earth and Planetary Science Letters. 463: 189–201. doi:10.1016/j.epsl.2017.01.041.
  26. ^ Hutko, Alexander R.; Lay, Thorne; Garnero, Edward J.; Revenaugh, Justin (2006). "Seismic detection of folded, subducted lithosphere at the core–mantle boundary". Nature. 441 (7091): 333–336. doi:10.1038/nature04757. PMID 16710418. S2CID 4408681.
  27. ^ Li, Mingming; McNamara, Allen K. (2013). "The difficulty for subducted oceanic crust to accumulate at the Earth's core-mantle boundary". Journal of Geophysical Research: Solid Earth. 118 (4): 1807–1816. doi:10.1002/jgrb.50156.
  28. ^ a b Stern 2002, p. 1.
  29. ^ Zheng, YF; Chen, RX; Xu, Z; Zhang, SB (2016). "The transport of water in subduction zones". Science China Earth Sciences. 59 (4): 651–682. Bibcode:2016ScChD..59..651Z. doi:10.1007/s11430-015-5258-4. S2CID 130912355.
  30. ^ Sillitoe, Richard H. (August 1974). "Tectonic segmentation of the Andes: implications for magmatism and metallogeny". Nature. 250 (5467): 542–545. Bibcode:1974Natur.250..542S. doi:10.1038/250542a0. S2CID 4173349.
  31. ^ Jordan, Teresa E.; Isacks, Bryan L.; Allmendinger, Richard W.; Brewer, Jon A.; Ramos, Victor A.; Ando, Clifford J. (1 March 1983). "Andean tectonics related to geometry of subducted Nazca plate". GSA Bulletin. 94 (3): 341–361. Bibcode:1983GSAB...94..341J. doi:10.1130/0016-7606(1983)94<341:ATRTGO>2.0.CO;2.
  32. ^ W. P. Schellart; D. R. Stegman; R. J. Farrington; J. Freeman & L. Moresi (16 July 2010). "Cenozoic Tectonics of Western North America Controlled by Evolving Width of Farallon Slab". Science. 329 (5989): 316–319. Bibcode:2010Sci...329..316S. doi:10.1126/science.1190366. PMID 20647465. S2CID 12044269.
  33. ^ Bletery, Quentin; Thomas, Amanda M.; Rempel, Alan W.; Karlstrom, Leif; Sladen, Anthony; De Barros, Louis (2016-11-24). "Fault curvature may control where big quakes occur, Eurekalert 24-NOV-2016". Science. 354 (6315): 1027–1031. Bibcode:2016Sci...354.1027B. doi:10.1126/science.aag0482. PMID 27885027. Retrieved 2018-06-05.
  34. ^ Lallemand, Serge; Heuret, Arnauld; Boutelier, David (8 September 2005). "On the relationships between slab dip, back-arc stress, upper plate absolute motion, and crustal nature in subduction zones" (PDF). Geochemistry, Geophysics, Geosystems. 6 (9): Q09006. Bibcode:2005GGG.....609006L. doi:10.1029/2005GC000917.
  35. ^ Stern, R.J. (2004). "Subduction initiation: spontaneous and induced". Earth and Planetary Science Letters. 226 (3–4): 275–292. Bibcode:2004E&PSL.226..275S. doi:10.1016/j.epsl.2004.08.007.
  36. ^ a b Crameri, Fabio; Magni, Valentina; Domeier, Mathew; Shephard, Grace E.; Chotalia, Kiran; Cooper, George; Eakin, Caroline M.; Grima, Antoniette Greta; Gürer, Derya; Király, Ágnes; Mulyukova, Elvira (2020-07-27). "A transdisciplinary and community-driven database to unravel subduction zone initiation". Nature Communications. 11 (1): 3750. Bibcode:2020NatCo..11.3750C. doi:10.1038/s41467-020-17522-9. ISSN 2041-1723. PMC 7385650. PMID 32719322.
  37. ^ Hall, C.E.; et al. (2003). "Catastrophic initiation of subduction following forced convergence across fracture zones". Earth and Planetary Science Letters. 212 (1–2): 15–30. Bibcode:2003E&PSL.212...15H. doi:10.1016/S0012-821X(03)00242-5.
  38. ^ Gurnis, M.; et al. (2004). "Evolving force balance during incipient subduction". Geochemistry, Geophysics, Geosystems. 5 (7): Q07001. Bibcode:2004GGG.....5.7001G. doi:10.1029/2003GC000681.
  39. ^ Keenan, Timothy E.; Encarnación, John; Buchwaldt, Robert; Fernandez, Dan; Mattinson, James; Rasoazanamparany, Christine; Luetkemeyer, P. Benjamin (2016). "Rapid conversion of an oceanic spreading center to a subduction zone inferred from high-precision geochronology". PNAS. 113 (47): E7359–E7366. Bibcode:2016PNAS..113E7359K. doi:10.1073/pnas.1609999113. PMC 5127376. PMID 27821756.
  40. ^ House, M. A.; Gurnis, M.; Kamp, P. J. J.; Sutherland, R. (September 2002). "Uplift in the Fiordland Region, New Zealand: Implications for Incipient Subduction" (PDF). Science. 297 (5589): 2038–2041. Bibcode:2002Sci...297.2038H. doi:10.1126/science.1075328. PMID 12242439. S2CID 31707224.
  41. ^ Mart, Y., Aharonov, E., Mulugeta, G., Ryan, W.B.F., Tentler, T., Goren, L. (2005). "Analog modeling of the initiation of subduction". Geophys. J. Int. 160 (3): 1081–1091. Bibcode:2005GeoJI.160.1081M. doi:10.1111/j.1365-246X.2005.02544.x.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  42. ^ Goren, L.; E. Aharonov; G. Mulugeta; H. A. Koyi; Y. Mart (2008). "Ductile Deformation of Passive Margins: A New Mechanism for Subduction Initiation". J. Geophys. Res. 113: B08411. Bibcode:2008JGRB..11308411G. doi:10.1029/2005JB004179. S2CID 130779676.
  43. ^ Stern, R.J.; Bloomer, S.H. (1992). "Subduction zone infancy: examples from the Eocene Izu-Bonin-Mariana and Jurassic California arcs". Geological Society of America Bulletin. 104 (12): 1621–1636. Bibcode:1992GSAB..104.1621S. doi:10.1130/0016-7606(1992)104<1621:SZIEFT>2.3.CO;2.
  44. ^ Arculus, R.J.; et al. (2015). "A record of spontaneous subduction initiation in the Izu–Bonin–Mariana arc" (PDF). Nature Geoscience. 8 (9): 728–733. Bibcode:2015NatGe...8..728A. doi:10.1038/ngeo2515.
  45. ^ Yin, A. (2012). "An episodic slab-rollback model for the origin of the Tharsis rise on Mars: Implications for initiation of local plate subduction and final unification of a kinematically linked global plate-tectonic network on Earth". Lithosphere. 4 (6): 553–593. Bibcode:2012Lsphe...4..553Y. doi:10.1130/L195.1.
  46. ^ a b Stern 2002, pp. 6–7.
  47. ^ Ernst, W. G. (June 1999). "Metamorphism, partial preservation, and exhumation of ultrahigh‐pressure belts". Island Arc. 8 (2): 125–153. doi:10.1046/j.1440-1738.1999.00227.x.
  48. ^ Cooper, P. A.; Taylor, B. (1985). "Polarity reversal in the Solomon Islands arc" (PDF). Nature. 314 (6010): 428–430. Bibcode:1985Natur.314..428C. doi:10.1038/314428a0. S2CID 4341305. Retrieved 4 December 2020.
  49. ^ Zheng, Y.-F., Chen, Y.-X., 2016. Continental versus oceanic subduction zones. National Science Review 3, 495-519.
  50. ^ "How Volcanoes work – Subduction Zone Volcanism". San Diego State University Department of Geological Science.
  51. ^ Mibe, Kenji; et al. (2011). "Slab melting versus slab dehydration in subduction zones". Proceedings of the National Academy of Sciences. 108 (20): 8177–8182. doi:10.1073/pnas.1010968108. PMC 3100975. PMID 21536910.
  52. ^ Zheng, Y.-F., Chen, R.-X., 2017. Regional metamorphism at extreme conditions: Implications for orogeny at convergent plate margins. Journal of Asian Earth Sciences 145, 46-73.
  53. ^ Winter, John D. (2010). Principles of Igneous and Metamorphic Petrology. Prentice Hall. pp. 541–548. ISBN 978-0-321-59257-6.
  54. ^ Reynolds, Stephen (2012-01-09). Exploring Geology. McGraw-Hill. p. 124. ISBN 978-0073524122.
  55. ^ Bebout, Grey E. (May 31, 2007). "Metamorphic Chemical Geodynamics of Subduction". Earth and Planetary Science Letters. 260: 375. Bibcode:2007E&PSL.260..373B. doi:10.1016/j.epsl.2007.05.050.
  56. ^ Peacock, Simon M. (1 January 2004). "Thermal Structure and Metamorphic Evolution of Subducting Slabs". In Eiler, John (ed.). Inside the subduction factory. Geophysical Monograph Series. Vol. 138. American Geophysical Union. pp. 12–15. ISBN 9781118668573.
  57. ^ Stern 2002, pp. 24–25.
  58. ^ Carver, Gary; Plafker, George (19 March 2013). "Paleoseismicity and Neotectonics of the Aleutian Subduction Zone-An Overview". Geophysical Monograph Series: 43–63. doi:10.1029/179GM03. ISBN 9781118666395.
  59. ^ a b Stern 2002, pp. 27–31.
  60. ^ "Volcanic arcs form by deep melting of rock mixtures: Study changes our understanding of processes inside subduction zones". ScienceDaily. Retrieved 2017-06-21.
  61. ^ Fisher, Richard V.; Schmincke, H.-U. (1984). Pyroclastic rocks. Berlin: Springer-Verlag. p. 5. ISBN 3540127569.
  62. ^ a b Stern 2002, pp. 17–18.
  63. ^ Bletery, Quentin; Thomas, Amanda M.; Rempel, Alan W.; Karlstrom, Leif; Sladen, Anthony; Barros, Louis De (2016-11-25). "Mega-earthquakes rupture flat megathrusts". Science. 354 (6315): 1027–1031. Bibcode:2016Sci...354.1027B. doi:10.1126/science.aag0482. ISSN 0036-8075. PMID 27885027.
  64. ^ "Subduction zone geometry: Mega-earthquake risk indicator". ScienceDaily. Retrieved 2017-06-21.
  65. ^ Garcia-Castellanos, D.; M. Torné; M. Fernàndez (2000). "Slab pull effects from a flexural analysis of the Tonga and Kermadec Trenches (Pacific Plate)". Geophys. J. Int. 141 (2): 479–485. Bibcode:2000GeoJI.141..479G. doi:10.1046/j.1365-246x.2000.00096.x.
  66. ^ a b c d e "Atlas of the Underworld | Van der Meer, D.G., van Hinsbergen, D.J.J., and Spakman, W., 2017, Atlas of the Underworld: slab remnants in the mantle, their sinking history, and a new outlook on lower mantle viscosity, Tectonophysics". www.atlas-of-the-underworld.org. Retrieved 2017-12-02.
  67. ^ Matthews, John A., ed. (2014). Encyclopedia of Environmental Change. Vol. 1. Los Angeles: SAGE Reference.
  68. ^ a b c d Xu, Cheng; Kynický, Jindřich; Song, Wenlei; Tao, Renbiao; Lü, Zeng; Li, Yunxiu; Yang, Yueheng; Miroslav, Pohanka; Galiova, Michaela V.; Zhang, Lifei; Fei, Yingwei (2018). "Cold deep subduction recorded by remnants of a Paleoproterozoic carbonated slab". Nature Communications. 9 (1): 2790. Bibcode:2018NatCo...9.2790X. doi:10.1038/s41467-018-05140-5. PMC 6050299. PMID 30018373.
  69. ^ Stern, Robert J. (2005). "Evidence from ophiolites, blueschists, and ultrahigh-pressure metamorphic terranes that the modern episode of subduction tectonics began in Neoproterozoic time". Geology. 33 (7): 557–560. Bibcode:2005Geo....33..557S. doi:10.1130/G21365.1. S2CID 907243.
  70. ^ a b c d Palin, Richard M.; White, Richard W. (2016). "Emergence of blueschists on Earth linked to secular changes in oceanic crust composition". Nature Geoscience. 9 (1): 60. Bibcode:2016NatGe...9...60P. doi:10.1038/ngeo2605.
  71. ^ Wilson, J. Tuzo (December 1968). "A Revolution in Earth Science". Geotimes. 13 (10). Washington DC: 10–16.
  72. ^ Geological Society of America (July 6, 2017). "Geological Society of America honors Excellence in Geoscience for 2017" (Press release). Eurekalert!.
  73. ^ "subduction". Online Etymology Dictionary. Retrieved 31 December 2020.
  74. ^ John Ogilvie; Charles Annandale (1883). "Subduce, Subduct". Imperial Dictionary of the English Language. Vol. Volume IV Scream-Zythus (New Edition Carefully Reviewed and Greatly Augmented ed.). London: Blackie & Son. {{cite encyclopedia}}: |volume= has extra text (help)
  75. ^ "What is a tectonic plate?". U.S. Geological Survey (USGS). 1999.
  76. ^ "Subduction Zone". Database of Individual Seismogenic Sources (DISS). Istituto Nazionale di Geofisica e Vulcanologia (INGV). Retrieved 4 January 2021.
  77. ^ Schultz, C. (2015). "Overriding plate's properties affect subduction". Eos. 96. doi:10.1029/2015EO026911.
  78. ^ Tsang, Man-Yin; Bowden, Stephen A.; Wang, Zhibin; Mohammed, Abdalla; Tonai, Satoshi; Muirhead, David; Yang, Kiho; Yamamoto, Yuzuru; Kamiya, Nana; Okutsu, Natsumi; Hirose, Takehiro (2020-02-01). "Hot fluids, burial metamorphism and thermal histories in the underthrust sediments at IODP 370 site C0023, Nankai Accretionary Complex". Marine and Petroleum Geology. 112: 104080. doi:10.1016/j.marpetgeo.2019.104080. ISSN 0264-8172.
  79. ^ "USGS publishes a new blueprint that can help make subduction zone areas more resilient". www.usgs.gov. Retrieved 2017-06-21.
  80. ^ Hafemeister, David W. (2007). Physics of societal issues: calculations on national security, environment, and energy. Berlin: Springer Science & Business Media. p. 187. ISBN 978-0-387-95560-5.
  81. ^ a b Kingsley, Marvin G.; Rogers, Kenneth H. (2007). Calculated risks: highly radioactive waste and homeland security. Aldershot, Hants, England: Ashgate. pp. 75–76. ISBN 978-0-7546-7133-6.
  82. ^ "Dumping and Loss overview". Oceans in the Nuclear Age. Archived from the original on June 5, 2011. Retrieved 18 September 2010.
  83. ^ "Storage and Disposal Options. World Nuclear Organization (date unknown)". Archived from the original on July 19, 2011. Retrieved February 8, 2012.

Additional reading

External links