Proterozoic

From Wikipedia, the free encyclopedia
Proterozoic
2500 – 538.8 ± 0.2 Ma
From left to right: Four main Proterozoic events: Great Oxidation Event and subsequent Huronian glaciation; First eukaryotes, like red algae; Snowball Earth in Cryogenian period; Ediacaran biota[1]
Chronology
Etymology
Name formalityFormal
Usage information
Celestial bodyEarth
Regional usageGlobal (ICS)
Time scale(s) usedICS Time Scale
Definition
Chronological unitEon
Stratigraphic unitEonothem
Time span formalityFormal
Lower boundary definitionDefined Chronometrically
Lower GSSA ratified1991[2]
Upper boundary definitionAppearance of the Ichnofossil Treptichnus pedum
Upper boundary GSSPFortune Head section, Newfoundland, Canada
47°04′34″N 55°49′52″W / 47.0762°N 55.8310°W / 47.0762; -55.8310
Upper GSSP ratified1992

The Proterozoic (IPA: /ˌprtərəˈzɪk, ˌprɒt-, -ər-, -trə-, -tr-/ PROH-tər-ə-ZOH-ik, PROT-, -⁠ər-oh-, -⁠trə-, -⁠troh-)[3][4][5] is the third of the four geologic eons of Earth's history, spanning the time interval from 2500 to 538.8 Mya,[6] the longest eon of the Earth's geologic time scale. It is preceded by the Archean and followed by the Phanerozoic, and is the most recent part of the Precambrian "supereon".

The Proterozoic is subdivided into three geologic eras (from oldest to youngest): the Paleoproterozoic, Mesoproterozoic and Neoproterozoic.[7] It covers the time from the appearance of free oxygen in Earth's atmosphere to just before the proliferation of complex life on the Earth during the Cambrian Explosion. The name Proterozoic combines two words of Greek origin: protero- meaning "former, earlier", and -zoic, meaning "of life".[8]

Well-identified events of this eon were the transition to an oxygenated atmosphere during the Paleoproterozoic; the evolution of eukaryotes via symbiogenesis; several global glaciations, which produced the 300 million years-long Huronian glaciation (during the Siderian and Rhyacian periods of the Paleoproterozoic) and the hypothesized Snowball Earth (during the Cryogenian period in the late Neoproterozoic); and the Ediacaran period (635 to 538.8 Ma), which is characterized by the evolution of abundant soft-bodied multicellular organisms such as sponges, algae, cnidarians, bilaterians and the sessile Ediacaran biota (some of which had evolved sexual reproduction) and provides the first obvious fossil evidence of life on Earth.

The Proterozoic record[edit]

The geologic record of the Proterozoic Eon is more complete than that for the preceding Archean Eon. In contrast to the deep-water deposits of the Archean, the Proterozoic features many strata that were laid down in extensive shallow epicontinental seas; furthermore, many of those rocks are less metamorphosed than Archean rocks, and many are unaltered.[9]: 315  Studies of these rocks have shown that the eon continued the massive continental accretion that had begun late in the Archean Eon. The Proterozoic Eon also featured the first definitive supercontinent cycles and wholly modern[clarify] mountain building activity (orogeny).[9]: 315–18, 329–32 

There is evidence that the first known glaciations occurred during the Proterozoic. The first began shortly after the beginning of the Proterozoic Eon, and evidence of at least four during the Neoproterozoic Era at the end of the Proterozoic Eon, possibly climaxing with the hypothesized Snowball Earth of the Sturtian and Marinoan glaciations.[9]: 320–1, 325 

The accumulation of oxygen[edit]

One of the most important events of the Proterozoic was the accumulation of oxygen in the Earth's atmosphere. Though oxygen is believed to have been released by photosynthesis as far back as the Archean Eon, it could not build up to any significant degree until mineral sinks of unoxidized sulfur and iron had been exhausted. Until roughly 2.3 billion years ago, oxygen was probably only 1% to 2% of its current level.[9]: 323  The banded iron formations, which provide most of the world's iron ore, are one mark of that mineral sink process. Their accumulation ceased after 1.9 billion years ago, after the iron in the oceans had all been oxidized.[9]: 324 

Red beds, which are colored by hematite, indicate an increase in atmospheric oxygen 2 billion years ago. Such massive iron oxide formations are not found in older rocks.[9]: 324  The oxygen buildup was probably due to two factors: exhaustion of the chemical sinks, and an increase in carbon sequestration, which sequestered organic compounds that would have otherwise been oxidized by the atmosphere.[9]: 325 

A second surge in oxygen concentrations, known as the Neoproterozoic Oxygenation Event,[10] occurred during the Middle and Late Neoproterozoic[11] and drove the rapid evolution of multicellular life towards the end of the era.[12][13]

Subduction processes[edit]

The Proterozoic Eon was a very tectonically active period in the Earth's history.

The late Archean Eon to Early Proterozoic Eon corresponds to a period of increasing crustal recycling, suggesting subduction. Evidence for this increased subduction activity comes from the abundance of old granites originating mostly after 2.6 Ga.[14]

The occurrence of eclogite (a type of metamorphic rock created by high pressure, > 1 GPa), is explained using a model that incorporates subduction. The lack of eclogites that date to the Archean Eon suggests that conditions at that time did not favor the formation of high grade metamorphism and therefore did not achieve the same levels of subduction as was occurring in the Proterozoic Eon.[15]

As a result of remelting of basaltic oceanic crust due to subduction, the cores of the first continents grew large enough to withstand the crustal recycling processes.

The long-term tectonic stability of those cratons is why we find continental crust ranging up to a few billion years in age.[16] It is believed that 43% of modern continental crust was formed in the Proterozoic, 39% formed in the Archean, and only 18% in the Phanerozoic.[14] Studies by Condie (2000)[17] and Rino et al. (2004)[18] suggest that crust production happened episodically. By isotopically calculating the ages of Proterozoic granitoids it was determined that there were several episodes of rapid increase in continental crust production. The reason for these pulses is unknown, but they seemed to have decreased in magnitude after every period.[14]

Tectonic history (supercontinents)[edit]

Evidence of collision and rifting between continents raises the question as to what exactly were the movements of the Archean cratons composing Proterozoic continents. Paleomagnetic and geochronological dating mechanisms have allowed the deciphering of Precambrian Supereon tectonics. It is known that tectonic processes of the Proterozoic Eon resemble greatly the evidence of tectonic activity, such as orogenic belts or ophiolite complexes, we see today. Hence, most geologists would conclude that the Earth was active at that time. It is also commonly accepted that during the Precambrian, the Earth went through several supercontinent breakup and rebuilding cycles (Wilson cycle).[14]

In the late Proterozoic (most recent), the dominant supercontinent was Rodinia (~1000–750 Ma). It consisted of a series of continents attached to a central craton that forms the core of the North American Continent called Laurentia. An example of an orogeny (mountain building processes) associated with the construction of Rodinia is the Grenville orogeny located in Eastern North America. Rodinia formed after the breakup of the supercontinent Columbia and prior to the assemblage of the supercontinent Gondwana (~500 Ma).[19] The defining orogenic event associated with the formation of Gondwana was the collision of Africa, South America, Antarctica and Australia forming the Pan-African orogeny.[20]

Columbia was dominant in the early-mid Proterozoic and not much is known about continental assemblages before then. There are a few plausible models that explain tectonics of the early Earth prior to the formation of Columbia, but the current most plausible hypothesis is that prior to Columbia, there were only a few independent cratons scattered around the Earth (not necessarily a supercontinent, like Rodinia or Columbia).[14]

Life[edit]

The emergence of advanced single-celled eukaryotes began after the Great Oxidation Event.[21] This may have been due to an increase in the oxidized nitrates that eukaryotes use, as opposed to cyanobacteria.[9]: 325  It was also during the Proterozoic that the first symbiotic relationships between mitochondria (found in nearly all eukaryotes) and chloroplasts (found in plants and some protists only) and their hosts evolved.[9]: 321–2 

By the late Palaeoproterozoic, eukaryotic organisms had become moderately biodiverse.[22] The blossoming of eukaryotes such as acritarchs did not preclude the expansion of cyanobacteria; in fact, stromatolites reached their greatest abundance and diversity during the Proterozoic, peaking roughly 1200 million years ago.[9]: 321–3 

The earliest fossils possessing features typical of fungi date to the Paleoproterozoic Era, some 2,400 million years ago; these multicellular benthic organisms had filamentous structures capable of anastomosis.[23]

Classically, the boundary between the Proterozoic and the Phanerozoic eons was set at the base of the Cambrian Period when the first fossils of animals, including trilobites and archeocyathids, as well as the animal-like Caveasphaera, appeared. In the second half of the 20th century, a number of fossil forms have been found in Proterozoic rocks, particularly in ones from the Ediacaran, proving that multicellular life had already become widespread tens of millions of years before the Cambrian Explosion in what is known as the Avalon Explosion.[24] Nonetheless, the upper boundary of the Proterozoic has remained fixed at the base of the Cambrian, which is currently placed at 538.8 Ma.

See also[edit]

References[edit]

  1. ^ Smithsonian National Museum flickr.
  2. ^ Plumb, K. A. (June 1, 1991). "New Precambrian time scale". Episodes. 14 (2): 139–140. doi:10.18814/epiiugs/1991/v14i2/005.
  3. ^ "Proterozoic – definition of Proterozoic in English from the Oxford dictionary". OxfordDictionaries.com. Archived from the original on July 24, 2012. Retrieved 2016-01-20.
  4. ^ "Proterozoic". Merriam-Webster.com Dictionary.
  5. ^ "Proterozoic". Dictionary.com Unabridged (Online). n.d.
  6. ^ "Stratigraphic Chart 2022" (PDF). International Stratigraphic Commission. February 2022. Retrieved 22 April 2022.
  7. ^ Speer, Brian. "The Proterozoic Eon". University of California Museum of Paleontology.
  8. ^ "Proterozoic, adj. and n." OED Online. Oxford University Press. June 2021. Retrieved 25 June 2021.
  9. ^ a b c d e f g h i j Stanley, Steven M. (1999). Earth System History. New York: W.H. Freeman and Company. ISBN 978-0-7167-2882-5.
  10. ^ Shields-Zhou, Graham A.; Och, Lawrence M. (March 2011). "The case for a Neoproterozoic Oxygenation Event: Geochemical evidence and biological consequences" (PDF). GSA Today. 21 (3): 4–11. doi:10.1130/GSATG102A.1. Retrieved 10 November 2022.
  11. ^ Och, Lawrence M.; Shields-Zhou, Graham A. (January 2012). "The Neoproterozoic oxygenation event: Environmental perturbations and biogeochemical cycling". Earth-Science Reviews. 110 (1–4): 26–57. doi:10.1016/j.earscirev.2011.09.004. Retrieved 10 November 2022.
  12. ^ Canfield, Donald Eugene; Poulton, Simon W.; Narbonne, Guy M. (5 January 2007). "Late-Neoproterozoic Deep-Ocean Oxygenation and the Rise of Animal Life". Science. 315 (5808): 92–95. doi:10.1126/science.1135013. PMID 17158290. S2CID 24761414. Retrieved 10 November 2022.
  13. ^ Fan, Haifeng; Zhu, Xiangkun; Wen, Hanjie; Yan, Bin; Li, Jin; Feng, Lianjun (1 September 2014). "Oxygenation of Ediacaran Ocean recorded by iron isotopes". Geochimica et Cosmochimica Acta. 140: 80–94. doi:10.1016/j.gca.2014.05.029. Retrieved 10 November 2022.
  14. ^ a b c d e Kearey, P.; Klepeis, K.; Vine, F. (2008). Precambrian Tectonics and the Supercontinent Cycle. Global Tectonics (Third ed.). pp. 361–377.
  15. ^ Bird, P. (2003). "An updated digital model of plate boundaries". Geochemistry, Geophysics, Geosystems. 4 (3): 1027. Bibcode:2003GGG.....4.1027B. doi:10.1029/2001GC000252.
  16. ^ Mengel, F. (1998). Proterozoic History. Earth System: History and Variablility. Vol. 2.
  17. ^ Condie, K. (2000). Episodic continental growth models: afterthoughts and extensions. Tectonophysics, 322(1), 153–162. doi:10.1016/S0040-1951(00)00061-5
  18. ^ Rino, Shuji; Komiya, Tsuyoshi; Windley, Brian F.; Katayama, Ikuo; Motoki, Akihisa; Hirata, Takafumi (August 2004). "Major episodic increases of continental crustal growth determined from zircon ages of river sands; implications for mantle overturns in the Early Precambrian". Physics of the Earth and Planetary Interiors. 146 (1–2): 369–394. Bibcode:2004PEPI..146..369R. doi:10.1016/j.pepi.2003.09.024. S2CID 140166194.
  19. ^ Condie, K. C.; O'Neill, C. (2011). "The Archean-Proterozoic boundary: 500 my of tectonic transition in Earth history". American Journal of Science. 310 (9): 775–790. Bibcode:2010AmJS..310..775C. doi:10.2475/09.2010.01. S2CID 128469935.
  20. ^ Huntly, C. (2002). The Mozambique Belt, Eastern Africa: Tectonic evolution of the Mozambique Ocean and Gondwana amalgamation. The Geological Society of America.
  21. ^ Fakhraee, Mojtaba; Tarhan, Lidya G.; Reinhard, Christopher T.; Crowe, Sean A.; Lyons, Timothy W.; Planavsky, Noah J. (May 2023). "Earth's surface oxygenation and the rise of eukaryotic life: Relationships to the Lomagundi positive carbon isotope excursion revisited". Earth-Science Reviews. 240: 104398. doi:10.1016/j.earscirev.2023.104398. S2CID 257761993.
  22. ^ Miao, Lanyun; Moczydłowska, Małgorzata; Zhu, Shixing; Zhu, Maoyan (February 2019). "New record of organic-walled, morphologically distinct microfossils from the late Paleoproterozoic Changcheng Group in the Yanshan Range, North China". Precambrian Research. 321: 172–198. doi:10.1016/j.precamres.2018.11.019. S2CID 134362289. Retrieved 29 December 2022.
  23. ^ Bengtson, Stefan; Rasmussen, Birger; Ivarsson, Magnus; Muhling, Janet; Broman, Curt; Marone, Federica; Stampanoni, Marco; Bekker, Andrey (2017-04-24). "Fungus-like mycelial fossils in 2.4-billion-year-old vesicular basalt". Nature Ecology & Evolution. 1 (6): 141. doi:10.1038/s41559-017-0141. hdl:20.500.11937/67718. ISSN 2397-334X. PMID 28812648. S2CID 25586788.
  24. ^ Xiao, Shuhai; Laflamme, Marc (January 2009). "On the eve of animal radiation: phylogeny, ecology and evolution of the Ediacara biota". Trends in Ecology & Evolution. 24 (1): 31–40. doi:10.1016/j.tree.2008.07.015. PMID 18952316. Retrieved 10 November 2022.

External links[edit]