Battle of Old Fort Wayne

Cratons of South America and Africa during the Triassic Period when the two continents were joined as part of the Pangea supercontinent

A craton ( /ˈkrtɒn/, /ˈkrætɒn/, or /ˈkrtən/;[1][2][3] from Greek: κράτος kratos "strength") is an old and stable part of the continental lithosphere, which consists of Earth's two topmost layers, the crust and the uppermost mantle. Having often survived cycles of merging and rifting of continents, cratons are generally found in the interiors of tectonic plates; the exceptions occur where geologically recent rifting events have separated cratons and created passive margins along their edges. Cratons are characteristically composed of ancient crystalline basement rock, which may be covered by younger sedimentary rock. They have a thick crust and deep lithospheric roots that extend as much as several hundred kilometres into Earth's mantle.

Terminology

Geologic provinces of the world (USGS)

The term craton is used to distinguish the stable portion of the continental crust from regions that are more geologically active and unstable.[4] Cratons are composed of two layers: a continental shield, in which the basement rock crops out at the surface,[5] and a platform which overlays the shield in some areas with sedimentary rock.[6]

The word craton was first proposed by the Austrian geologist Leopold Kober in 1921 as Kratogen, referring to stable continental platforms, and orogen as a term for mountain or orogenic belts. Later Hans Stille shortened the former term to Kraton, from which craton derives.[7]

Examples

Examples of cratons are the Dharwar Craton[8] in India, North China Craton,[9] the East European Craton,[10] the Amazonian Craton in South America,[11] the Kaapvaal Craton in South Africa,[12] the North American Craton (also called the Laurentia Craton),[13] and the Gawler Craton in South Australia.[14]

Structure

Cratons have thick lithospheric roots. Mantle tomography shows that cratons are underlain by anomalously cold mantle corresponding to lithosphere more than twice the typical 100 km (60 mi) thickness of mature oceanic or non-cratonic, continental lithosphere. At that depth, craton roots extend into the asthenosphere,[15] and the low-velocity zone seen elsewhere at these depths is weak or absent beneath stable cratons.[16] Craton lithosphere is distinctly different from oceanic lithosphere because cratons have a neutral or positive buoyancy and a low intrinsic density. This low density offsets density increases from geothermal contraction and prevents the craton from sinking into the deep mantle. Cratonic lithosphere is much older than oceanic lithosphere—up to 4 billion years versus 180 million years.[17]

Rock fragments (xenoliths) carried up from the mantle by magmas containing peridotite have been delivered to the surface as inclusions in subvolcanic pipes called kimberlites. These inclusions have densities consistent with craton composition and are composed of mantle material residual from high degrees of partial melt. Peridotite is strongly influenced by the inclusion of moisture. Craton peridotite moisture content is unusually low, which leads to much greater strength. It also contains high percentages of low-weight magnesium instead of higher-weight calcium and iron.[18] Peridotites are important for understanding the deep composition and origin of cratons because peridotite nodules are pieces of mantle rock modified by partial melting. Harzburgite peridotites represent the crystalline residues after extraction of melts of compositions like basalt and komatiite.[19]

Formation

The process by which cratons were formed is called cratonization. There is much about this process that remains uncertain, with very little consensus in the scientific community.[20] However, the first cratonic landmasses likely formed during the Archean eon. This is indicated by the age of diamonds, which originate in the roots of cratons, and which are almost always over 2 billion years and often over 3 billion years in age.[17] Rock of Archean age makes up only 7% of the world's current cratons; even allowing for erosion and destruction of past formations, this suggests that only 5 to 40 percent of the present continental crust formed during the Archean.[21] Cratonization likely was completed during the Proterozoic. Subsequent growth of continents was by accretion at continental margins.[17]

Root origin

The origin of the roots of cratons is still debated.[22][23][18][20] However, the present understanding of cratonization began with the publication in 1978 of a paper by Thomas H. Jordan in Nature. Jordan proposes that cratons formed from a high degree of partial melting of the upper mantle, with 30 to 40 percent of the source rock entering the melt. Such a high degree of melting was possible because of the high mantle temperatures of the Archean. The extraction of so much magma left behind a solid peridotite residue that was enriched in lightweight magnesium and thus lower in chemical density than undepleted mantle. This lower chemical density compensated for the effects of thermal contraction as the craton and its roots cooled, so that the physical density of the cratonic roots matched that of the surrounding hotter, but more chemically dense, mantle.[24][17] In addition to cooling the craton roots and lowering their chemical density, the extraction of magma also increased the viscosity and melting temperature of the craton roots and prevented mixing with the surrounding undepleted mantle.[25] The resulting mantle roots have remained stable for billions of years.[23] Jordan suggests that depletion occurred primarily in subduction zones and secondarily as flood basalts.[26]

This model of melt extraction from the upper mantle has held up well with subsequent observations.[27] The properties of mantle xenoliths confirm that the geothermal gradient is much lower beneath continents than oceans.[28] The olivine of craton root xenoliths is extremely dry, which would give the roots a very high viscosity.[29] Rhenium–osmium dating of xenoliths indicates that the oldest melting events took place in the early to middle Archean. Significant cratonization continued into the late Archean, accompanied by voluminous mafic magmatism.[30]

However, melt extraction alone cannot explain all the properties of craton roots. Jordan notes in his paper that this mechanism could be effective for constructing craton roots only down to a depth of 200 kilometers (120 mi). The great depths of craton roots required further explanation.[26] The 30 to 40 percent partial melting of mantle rock at 4 to 10 GPa pressure produces komatiite magma and a solid residue very close in composition to Archean lithospheric mantle, but continental shields do not contain enough komatiite to match the expected depletion. Either much of the komatiite never reached the surface, or other processes aided craton root formation.[30] There are many competing hypotheses of how cratons have been formed.

Repeated continental collision model

Jordan's model suggests that further cratonization was a result of repeated continental collisions. The thickening of the crust associated with these collisions may have been balanced by craton root thickening according to the principle of isostacy.[26] Jordan likens this model to "kneading" of the cratons, allowing low density material to move up and higher density to move down, creating stable cratonic roots as deep as 400 km (250 mi).[29]

Molten plume model

A second model suggests that the surface crust was thickened by a rising plume of molten material from the deep mantle. This would have built up a thick layer of depleted mantle underneath the cratons.

Subducting ocean slab model

A third model suggests that successive slabs of subducting oceanic lithosphere became lodged beneath a proto-craton, underplating the craton with chemically depleted rock.[29][18][22]

Impact origin model

A fourth theory presented in a 2015 publication suggests that the origin of the cratons is similar to crustal plateaus observed on Venus, which may have been created by large asteroid impacts.[20] In this model, large impacts on the Earth's early lithosphere penetrated deep into the mantle and created enormous lava ponds.[20] The paper suggests these lava ponds cooled to form the craton's root.[20]

Evidence for each model

The chemistry of xenoliths[27] and seismic tomography both favor the two accretional models over the plume model.[29][31] However, other geochemical evidence favors mantle plumes.[32][33][34] Tomography shows two layers in the craton roots beneath North America. One is found at depths shallower than 150 km (93 mi) and may be Archean, while the second is found at depths from 180 to 240 km (110 to 150 mi) and may be younger. The second layer may be a less depleted thermal boundary layer that stagnated against the depleted "lid" formed by the first layer.[35] The impact origin model does not require plumes or accretion; this model is, however, not incompatible with either.[20]

All these proposed mechanisms rely on buoyant, viscous material separating from a denser residue due to mantle flow, and it is possible that more than one mechanism contributed to craton root formation.[30][20]

Erosion

The long-term erosion of cratons has been labelled the "cratonic regime". It involves processes of pediplanation and etchplanation that lead to the formation of flattish surfaces known as peneplains.[36] While the process of etchplanation is associated to humid climate and pediplanation with arid and semi-arid climate, shifting climate over geological time leads to the formation of so-called polygenetic peneplains of mixed origin. Another result of the longevity of cratons is that they may alternate between periods of high and low relative sea levels. High relative sea level leads to increased oceanicity, while the opposite leads to increased inland conditions.[36]

Many cratons have had subdued topographies since Precambrian times. For example, the Yilgarn Craton of Western Australia was flattish already by Middle Proterozoic times[36] and the Baltic Shield had been eroded into a subdued terrain already during the Late Mesoproterozoic when the rapakivi granites intruded.[37][38]

See also

References

  1. ^ "Definition of craton in North American English". Oxford Dictionaries. Archived from the original on 2015-04-02. Retrieved 2015-03-28.
  2. ^ "Definition of craton in British and Commonwealth English". Oxford Dictionaries. Archived from the original on 2015-04-02. Retrieved 2015-03-28.
  3. ^ Macquarie Dictionary (5th ed.). Sydney: Macquarie Dictionary Publishers Pty Ltd. 2009.
  4. ^ Jackson, Julia A., ed. (1997). "craton". Glossary of geology (Fourth ed.). Alexandria, Virginia: American Geological Institute. ISBN 0922152349.
  5. ^ Jackson 1997, "shield [tect]".
  6. ^ Jackson 1997, "platform [tect]".
  7. ^ Şengör, A.M.C. (2003). The Large-wavelength Deformations of the Lithosphere: Materials for a history of the evolution of though from the earliest times toi plate tectonics. Geological Society of America memoir. Vol. 196. p. 331.
  8. ^ Ratheesh-Kumar, R.T.; Windley, B.F.; Xiao, W.J.; Jia, X-L.; Mohanty, D.P.; Zeba-Nezrin, F.K. (October 2019). "Early growth of the Indian lithosphere: implications from the assembly of the Dharwar Craton and adjacent granulite blocks, southern India". Precambrian Research. 336: 105491. doi:10.1016/j.precamres.2019.105491. S2CID 210295037.
  9. ^ Kusky, T. M.; Windley, B. F.; Zhai, M.-G. (2007). "Tectonic evolution of the North China Block: from orogen to craton to orogen". Geological Society, London, Special Publications. 280 (1): 1–34. Bibcode:2007GSLSP.280....1K. doi:10.1144/sp280.1. S2CID 129902429.
  10. ^ Artemieva, Irina M (August 2003). "Lithospheric structure, composition, and thermal regime of the East European Craton: implications for the subsidence of the Russian platform" (PDF). Earth and Planetary Science Letters. 213 (3–4): 431–446. Bibcode:2003E&PSL.213..431A. doi:10.1016/S0012-821X(03)00327-3.
  11. ^ Cordani, U.G.; Teixeira, W.; D'Agrella-Filho, M.S.; Trindade, R.I. (June 2009). "The position of the Amazonian Craton in supercontinents". Gondwana Research. 15 (3–4): 396–407. Bibcode:2009GondR..15..396C. doi:10.1016/j.gr.2008.12.005.
  12. ^ Nguuri, T. K.; Gore, J.; James, D. E.; Webb, S. J.; Wright, C.; Zengeni, T. G.; Gwavava, O.; Snoke, J. A. (1 July 2001). "Crustal structure beneath southern Africa and its implications for the formation and evolution of the Kaapvaal and Zimbabwe cratons". Geophysical Research Letters. 28 (13): 2501–2504. doi:10.1029/2000GL012587. hdl:10919/24271. S2CID 15687067.
  13. ^ Hoffman, P F (May 1988). "United Plates of America, The Birth of a Craton: Early Proterozoic Assembly and Growth of Laurentia". Annual Review of Earth and Planetary Sciences. 16 (1): 543–603. Bibcode:1988AREPS..16..543H. doi:10.1146/annurev.ea.16.050188.002551.
  14. ^ Hand, M.; Reid, A.; Jagodzinski, L. (1 December 2007). "Tectonic Framework and Evolution of the Gawler Craton, Southern Australia". Economic Geology. 102 (8): 1377–1395. Bibcode:2007EcGeo.102.1377H. doi:10.2113/gsecongeo.102.8.1377.
  15. ^ Petit, Charles (18 December 2010). "Continental Hearts – Science News". Science News. 178 (13). Society for Science & the Public: 24. doi:10.1002/scin.5591781325. ISSN 0036-8423.
  16. ^ Kearey, P.; Klepeis, K.A.; Vine, F.J. (2009). Global tectonics (3rd ed.). Oxford: Wiley-Blackwell. p. 349. ISBN 9781405107778.
  17. ^ a b c d Petit 2010, p. 25.
  18. ^ a b c Petit 2010, pp. 25–26.
  19. ^ Philpotts, Anthony R.; Ague, Jay J. (2009). Principles of igneous and metamorphic petrology (2nd ed.). Cambridge, UK: Cambridge University Press. pp. 373, 602–603. ISBN 9780521880060.
  20. ^ a b c d e f g Hansen, Vicki L. (24 August 2015). "Impact origin of Archean cratons". Lithosphere. 7 (5): 563–578. Bibcode:2015Lsphe...7..563H. doi:10.1130/L371.1.
  21. ^ Stanley, Steven M. (1999). Earth System History. New York: W.H. Freeman and Company. pp. 297–302. ISBN 0-7167-2882-6.
  22. ^ a b Lee, C. (2006). Benn, K.; Mareschal, J.C.; Condie, K.C. (eds.). "Geochemical/petrologic constraints on the origin of cratonic mantle" (PDF). American Geophysical Union Geophysical Monograph. Geophysical Monograph Series. 164: 89. Bibcode:2006GMS...164...89L. doi:10.1029/164GM08. ISBN 978-0-87590-429-0. Retrieved 20 November 2021.
  23. ^ a b Miller, Meghan S.; Eaton, David W. (September 2010). "Formation of cratonic mantle keels by arc accretion: Evidence from S receiver functions: FORMATION OF CRATONIC MANTLE KEELS". Geophysical Research Letters. 37 (18): n/a. doi:10.1029/2010GL044366. S2CID 129901730.
  24. ^ Jordan, Thomas H. (August 1978). "Composition and development of the continental tectosphere". Nature. 274 (5671): 544–548. Bibcode:1978Natur.274..544J. doi:10.1038/274544a0. S2CID 4286280.
  25. ^ Jordan 1978, p. 546.
  26. ^ a b c Jordan 1978, p. 547.
  27. ^ a b Lee 2006.
  28. ^ Jordan 1978, p. 544.
  29. ^ a b c d Petit 2010, p. 26.
  30. ^ a b c Kearey, Klepeis & Vine 2009, p. 351.
  31. ^ Miller & Eaton 2010.
  32. ^ Tomlinson, Kirsty Y.; Condie, Kent C. (2001). "Archean mantle plumes: Evidence from greenstone belt geochemistry". Mantle Plumes: Their Identification Through Time. doi:10.1130/0-8137-2352-3.341. ISBN 9780813723525. Retrieved 21 November 2021.
  33. ^ Ernst, Richard E.; Buchan, Kenneth L.; Campbell, Ian H. (February 2005). "Frontiers in large igneous province research". Lithos. 79 (3–4): 271–297. Bibcode:2005Litho..79..271E. doi:10.1016/j.lithos.2004.09.004.
  34. ^ Kearey, Klepeis & Vine 2009, p. 352.
  35. ^ Yuan, Huaiyu; Romanowicz, Barbara (August 2010). "Lithospheric layering in the North American craton". Nature. 466 (7310): 1063–1068. Bibcode:2010Natur.466.1063Y. doi:10.1038/nature09332. PMID 20740006. S2CID 4380594.
  36. ^ a b c Fairbridge, Rhodes W.; Finkl Jr., Charles W. (1980). "Cratonic erosion unconformities and peneplains". The Journal of Geology. 88 (1): 69–86. Bibcode:1980JG.....88...69F. doi:10.1086/628474. S2CID 129231129.
  37. ^ Lindberg, Johan (April 4, 2016). "berggrund och ytformer". Uppslagsverket Finland (in Swedish). Archived from the original on January 6, 2018. Retrieved February 13, 2018.
  38. ^ Lundmark, Anders Mattias; Lamminen, Jarkko (2016). "The provenance and setting of the Mesoproterozoic Dala Sandstone, western Sweden, and paleogeographic implications for southwestern Fennoscandia". Precambrian Research. 275: 197–208. Bibcode:2016PreR..275..197L. doi:10.1016/j.precamres.2016.01.003.

Further reading

External links