Andean orogeny

Last updated

Simplified sketch of the present-situation along most of the Andes Oceanic-continental convergence Fig21oceancont.gif
Simplified sketch of the present-situation along most of the Andes

The Andean orogeny (Spanish : Orogenia andina) is an ongoing process of orogeny that began in the Early Jurassic and is responsible for the rise of the Andes mountains. The orogeny is driven by a reactivation of a long-lived subduction system along the western margin of South America. On a continental scale the Cretaceous (90 Ma) and Oligocene (30 Ma) were periods of re-arrangements in the orogeny. The details of the orogeny vary depending on the segment and the geological period considered.

Contents

Overview

Subduction orogeny has been occurring in what is now western South America since the break-up of the supercontinent Rodinia in the Neoproterozoic. [1] The Paleozoic Pampean, Famatinian and Gondwanan orogenies are the immediate precursors to the later Andean orogeny. [2] The first phases of Andean orogeny in the Jurassic and Early Cretaceous were characterized by extensional tectonics, rifting, the development of back-arc basins and the emplacement of large batholiths. [1] [3] This development is presumed to have been linked to the subduction of cold oceanic lithosphere. [3] During the mid to Late Cretaceous (ca. 90 million years ago) the Andean orogeny changed significantly in character. [1] [3] Warmer and younger oceanic lithosphere is believed to have started to be subducted beneath South America around this time. Such kind of subduction is held responsible not only for the intense contractional deformation that different lithologies were subject to, but also the uplift and erosion known to have occurred from the Late Cretaceous onward. [3] Plate tectonic reorganization since the mid-Cretaceous might also have been linked to the opening of the South Atlantic Ocean. [1] Another change related to mid-Cretaceous plate tectonic changes was the change of subduction direction of the oceanic lithosphere that went from having south-east motion to having a north-east motion at about 90 million years ago. [4] While subduction direction changed it remained oblique (and not perpendicular) to the coast of South America, and the direction change affected several subduction zone-parallel faults including Atacama, Domeyko and Liquiñe-Ofqui. [3] [4]

Paleogeography of the Late Cretaceous South America. Areas subject to the Andean orogeny are shown in light grey while the stable cratons are shown as grey squares. The sedimentary formations of Los Alamitos and La Colonia that formed in the Late Cretaceous are indicated. Paleogeography Gondwana - Late Cretaceous-Early Paleogene - around 85-63 Ma.jpg
Paleogeography of the Late Cretaceous South America. Areas subject to the Andean orogeny are shown in light grey while the stable cratons are shown as grey squares. The sedimentary formations of Los Alamitos and La Colonia that formed in the Late Cretaceous are indicated.

Low angle subduction or flat-slab subduction has been common during the Andean orogeny leading to crustal shortening and deformation and the suppression of arc volcanism. Flat-slab subduction has occurred at different times in various part of the Andes, with northern Colombia (6–10° N), Ecuador (0–2° S), northern Peru (3–13° S) and north-central Chile (24–30° S) experiencing these conditions at present. [1]

The tectonic growth of the Andes and the regional climate have evolved simultaneously and have influenced each other. [5] The topographic barrier formed by the Andes stopped the income of humid air into the present Atacama desert. This aridity, in turn, changed the normal superficial redistribution of mass via erosion and river transport, modifying the later tectonic deformation. [5]

In the Oligocene the Farallon Plate broke up, forming the modern Cocos and Nazca plates ushering a series of changes in the Andean orogeny. The new Nazca Plate was then directed into an orthogonal subduction with South America causing ever-since uplift in the Andes, but causing most impact in the Miocene. While the various segments of the Andes have their own uplift histories, as a whole the Andes have risen significantly in last 30 million years (Oligocene–present). [6]

Orogeny by segment

Colombia, Ecuador and Venezuela (12° N–3° S)

Map of a north-south sea-parallel pattern of rock ages in western Colombia. This pattern is a result of the Andean orogeny. Geologische kaart van Valle del Cauca.png
Map of a north-south sea-parallel pattern of rock ages in western Colombia. This pattern is a result of the Andean orogeny.

Tectonic blocks of continental crust that had separated from northwestern South America in the Jurassic re-joined the continent in the Late Cretaceous by colliding obliquely with it. [6] This episode of accretion occurred in a complex sequence. The accretion of the island arcs against northwestern South America in the Early Cretaceous led to the development of a magmatic arc caused by subduction. The Romeral Fault in Colombia forms the suture between the accreted terranes and the rest of South America. Around the Cretaceous–Paleogene boundary (ca. 65 million years ago) the oceanic plateau of the Caribbean large igneous province collided with South America. The subduction of the lithosphere as the oceanic plateau approached South America led to the formation of a magmatic arc now preserved in the Cordillera Real of Ecuador and the Cordillera Central of Colombia. In the Miocene an island arc and terrane (Chocó terrane) collided against northwestern South America. This terrane forms parts of what is now Chocó Department and Western Panamá. [1]

The Caribbean Plate collided with South America in the Early Cenozoic but shifted then its movement eastward. [6] [7] Dextral fault movement between the South American and Caribbean plate started 17–15 million years ago. This movement was canalized along a series of strike-slip faults, but these faults alone do not account for all deformation. [8] The northern part of the Dolores-Guayaquil Megashear forms part of the dextral fault systems while in the south the megashear runs along the suture between the accreted tectonic blocks and the rest of South America. [9]

Northern Peru (3–13° S)

The seaward tilting of the sedimentary strata of Salto del Fraile Formation in Peru was caused by the Andean orogeny. Salto del Fraile Lima Peru.jpg
The seaward tilting of the sedimentary strata of Salto del Fraile Formation in Peru was caused by the Andean orogeny.

Long before the Andean orogeny the northern half of Peru was subject of the accretion of terranes in the Neoproterozoic and Paleozoic. [10] Andean orogenic deformation in northern Peru can be traced to the Albian (Early Cretaceous). [11] This first phase of deformation —the Mochica Phase [upper-alpha 1] — is evidenced in the folding of Casma Group sediments near the coast. [10]

Sedimentary basins in western Peru changed from marine to continental conditions in the Late Cretaceous as a consequence of a generalized vertical uplift. The uplift in northern Peru is thought to be associated with the contemporary accretion of the Piñón terrane in Ecuador. This stage of orogeny is called the Peruvian Phase. [10] Besides coastal Peru the Peruvian Phase affected or caused crustal shortening along the Cordillera Oriental and the tectonic inversion of Santiago Basin in the Sub-Andean zone. The bulk of the Sub-Andean zone was however unaffected by the Peruvian Phase. [12]

After a period without much tectonic activity in the Early Eocene the Incaic Phase of orogeny occurred in the Mid and Late Eocene. [11] [12] No other tectonic event in the western Peruvian Andes compare with the Incaic Phase in magnitude. [11] [12] Horizontal shortening during the Incaic Phase resulted in the formation of the Marañón fold and thrust belt. [11] An unconformity cutting across the Marañón fold and thrust belt show the Incaic Phase ended no later than 33 million years ago in the earliest Oligocene. [10]

Topographic map of the Andes by the NASA. The southern and northern ends of the Andes are not shown. The Bolivian Orocline is visible as a bend in the coastline and the Andes lower half of the map. Nasa anden.jpg
Topographic map of the Andes by the NASA. The southern and northern ends of the Andes are not shown. The Bolivian Orocline is visible as a bend in the coastline and the Andes lower half of the map.

In the period after the Eocene the Northern Peruvian Andes were subject to the Quechua Phase of orogeny. The Quechua Phase is divided into the sub-phases Quechua 1, Quechua 2 and Quechua 3. [upper-alpha 2] The Quechua 1 Phase lasted from 17 to 15 million years ago and included a reactivation of Inca Phase structures in the Cordillera Occidental. [upper-alpha 3] 9–8 million years ago, in the Quechua 2 Phase, the older parts of the Andes in northern Peru were thrusted to the northeast. [10] Most of the Sub-Andean zone of northern Peru deformed 7–5 million years ago (Late Miocene) during the Quechua 3 Phase. [10] [12] The Sub-Andean stacked in a thrust belt. [10]

The Miocene rise of the Andes in Peru and Ecuador led to increased orographic precipitation along its eastern parts and to the birth of the modern Amazon River. One hypothesis links these two changes by assuming that increased precipitation led to increased erosion and this erosion led to filling the Andean foreland basins beyond their capacity and that it would have been the basin over-sedimentation rather than the rise of the Andes that made drainage basins flow to the east. [12] Previously the interior of northern South America drained to the Pacific.

Bolivian Orocline (13–26° S)

Early Andean subduction in the Jurassic formed a volcanic arc in northern Chile known as La Negra Arc. [upper-alpha 4] The remnants of this arc are now exposed in the Chilean Coast Range. Several plutons were emplaced in the Chilean Coast Range in the Jurassic and Early Cretaceous including the Vicuña Mackenna Batholith. [14] Further east at similar latitudes, in Argentina and Bolivia, the Salta rift system developed during the Late Jurassic and the Early Cretaceous. [15] Salar de Atacama Basin, which is thought to be the western arm of the rift system, [16] accumulated during the Late Cretaceous and Early Paleogene a >6,000 m thick pile of sediments now known as the Purilactis Group. [17]

Pisco Basin, around latitude 14° S, was subject to a marine transgression in the Oligocene and Early Miocene epochs (25–16 Ma [18] ). [19] In contrast Moquegua Basin to the southeast and the coast to south of Pisco Basin saw no transgression during this time but a steadily rise of the land. [19]

From the Late Miocene onward the region that would become the Altiplano rose from low elevations to more than 3,000 m.a.s.l. It is estimated that the region rose 2000 to 3000 meters in the last ten million years. [20] Together with this uplift several valleys incised in the western flank of the Altiplano. In the Miocene the Atacama Fault moved, uplifting the Chilean Coast Range and creating sedimentary basins east of it. [21] At the same time the Andes around the Altiplano region broadened to exceed any other Andean segment in width. [6] Possibly about 1000 km of lithosphere has been lost due to lithospheric shortening. [22] During subduction the western end of the forearc region [upper-alpha 5] flexured downward forming a giant monocline. [23] [24] Somewhat to the south, tectonic inversion belonging during the "Incaic Phase" (Eocene?) have tilted the strata of Purilactis Group and in some localities also thrust younger strata on top of it. [25]

The Altiplano and its largest lake as seen from Ancohuma. The uplift of the Altiplano plateau is one of the most striking features of the Andean orogeny. Lake-Titicaca-from-Ancohuma.JPG
The Altiplano and its largest lake as seen from Ancohuma. The uplift of the Altiplano plateau is one of the most striking features of the Andean orogeny.

The region east of the Altiplano is characterized by deformation and tectonics along a complex fold and thrust belt. [23] Over-all the region surrounding the Altiplano and Puna plateaux has been horizontally shortened since the Eocene. [26] In southern Bolivia lithospheric shortening has made the Andean foreland basin to move eastward relative to the continent at an average rate of ca. 12–20 mm per year during most of the Cenozoic. [22] [upper-alpha 6] Along the Argentine Northwest the Andean uplift has caused Andean foreland basins to separate into several minor isolated intermontane sedimentary basins. [27] Towards the east the piling up of crust in Bolivia and the Argentine Norwest caused a north-south forebulge known as Asunción arch to develop in Paraguay. [28]

The uplift of the Altiplano is thought to be indebted to a combination of horizontal shortening of the crust and to increased temperatures in the mantle (thermal thinning). [1] [23] The bend in the Andes and the west coast of South America known as the Bolivian Orocline was enhanced by Cenozoic horizontal shortening but existed already independently of it. [23]

Meso-scale tectonic processes aside, the particular characteristics of the Bolivian Orocline–Altiplano region have been attributed to a variety of deeper causes. These causes include a local steepening of the subduction angle of Nazca Plate, increased crustal shortening and plate convergence between the Nazca and South American plates, an acceleration in the westward drift of the South American Plate, and a rise in the shear stress between the Nazca and South American plates. This increase in shear stress could in turn be related to the scarcity of sediments in the Atacama trench which is caused by the arid conditions along Atacama Desert. [6] Capitanio et al. attributes the rise of Altiplano and the bending of the Bolivian Orocline to the varying ages of the subducted Nazca Plate with the older parts of the plate subducting at the centre of the orocline. [29] As Andrés Tassara puts it the rigidity of the Bolivian Orocline crust is derivative of the thermal conditions. The crust of the western region (forearc) of the orocline has been cold and rigid, resisting and damming up the westward flow of warmer and weaker ductile crustal material from beneath the Altiplano. [24]

The Cenozoic orogeny at the Bolivian orocline has produced a significative anatexis of crustal rocks including metasediments and gneisses resulting in the formation of peraluminous magmas. These characteristics imply that the Cenozoic tectonics and magmatism in parts of Bolivian Andes is similar to that seen in collisional orogens. The peralumineous magmatism in Cordillera Oriental is the cause of the world-class mineralizations of the Bolivian tin belt. [30]

Tilted strata of the Yacoraite Formation at Serrania de Hornocal in northernmost Argentina. The Andean orogeny caused the tilting of these originally horizontal strata. Serrania de Hornocal up close near Humahuaca.jpg
Tilted strata of the Yacoraite Formation at Serranía de Hornocal in northernmost Argentina. The Andean orogeny caused the tilting of these originally horizontal strata.

The rise of the Altiplano is thought by scientist Adrian Hartley to have enhanced an already prevailing aridity or semi-aridity in Atacama Desert by casting a rain shadow over the region. [31]

Central Chile and Western Argentina (26–39° S)

At the latitudes between 17 and 39° S the Late Cretaceous and Cenozoic development of the Andean orogeny is characterized by an eastward migration of the magmatic belt and the development of several foreland basins. [3] The eastward migration of the arc is thought to be caused by subduction erosion. [32]

At the latitudes of 32–36° S —that is Central Chile and most of Mendoza Province— the Andean orogeny proper began in the Late Cretaceous when backarc basins were inverted. Immediately east of the early Andes foreland basins developed and their flexural subsidence caused the ingression of waters from the Atlantic all the way to the front of the orogen in the Maastrichtian. [33] The Andes at the latitudes of 32–36° S experienced a sequence of uplift in the Cenozoic that started in the west and spread to the east. Beginning about 20 million years ago in the Miocene the Principal Cordillera (east of Santiago) began an uplift that lasted until about 8 million years ago. [33] From the Eocene to the early Miocene, sediments [upper-alpha 7] accumulated in the Abanico Extensional Basin, a north-south elongated basin in Chile that spanned from 29° to 38° S. Tectonic inversion from 21 to 16 million years ago made the basin to collapse and the sediments to be incorporated to the Andean cordillera. [34] Lavas and volcanic material that are now part of Farellones Formation accumulated while the basin was being inverted and uplifted. [35] The Miocene continental divide was about 20 km to the west of the modern water divide that makes up the Argentina–Chile border. [35] Subsequent river incision shifted the divide to the east leaving old flattish surfaces hanging. [35] Compression and uplift in this part of the Andes has continued into the present. [35] The Principal Cordillera had risen to heights that allowed for the development of valley glaciers about 1 million years ago. [35]

Before the Miocene uplift of the Principal Cordillera was over, the Frontal Cordillera to the east started a period of uplift that lasted from 12 to 5 million years ago. Further east the Precordillera was uplifted in the last 10 million years and the Sierras Pampeanas has experienced a similar uplift in the last 5 million years. The more eastern part of the Andes at these latitudes had their geometry controlled by ancient faults dating to the San Rafael orogeny of the Paleozoic. [33] The Sierras de Córdoba (part of the Sierras Pampeanas) where the effects of the ancient Pampean orogeny can be observed, owes it modern uplift and relief to the Andean orogeny in the late Cenozoic. [36] [37] Similarly the San Rafael Block east of the Andes and south of Sierras Pampeanas was raised in the Miocene during the Andean orogeny. [38] In broad terms the most active phase of orogeny in area of southern Mendoza Province and northern Neuquén Province (34–38° S) happened in the Late Miocene while arc volcanism occurred east of the Andes. [38]

At more southern latitudes (36–39° S) various Jurassic and Cretaceous marine transgressions from the Pacific are recorded in the sediments of Neuquén Basin. [upper-alpha 8] In the Late Cretaceous conditions changed. A marine regression occurred and the fold and thrust belts of Malargüe (36°00 S), Chos Malal (37° S) and Agrio (38° S) started to develop in the Andes and did so in until Eocene times. This meant an advance of the orogenic deformation since the Late Cretaceous that caused the western part of Neuquén Basin to stack in the Malargüe and Agrio fold and thrust belts. [39] [38] In the Oligocene the western part of the fold and thrust belt was subject to a short period of extensional tectonics whose structures were inverted in the Miocene. [39] [upper-alpha 9] After a period of quiescence the Agrio fold and thrust belt resumed limited activity in the Eocene and then again in the Late Miocene. [38]

In the south of Mendoza Province the Guañacos fold and thrust belt (36.5° S) appeared and grew in the Pliocene and Pleistocene consuming the western fringes of the Neuquén Basin. [39] [38]

Northern Patagonian Andes (39–48° S)

Southern Patagonian Andes (48–55° S)

Syncline next to Nordenskjold Lake in Torres del Paine National Park. The syncline formed during the Andean orogeny. Sinclino Nordenskiold.jpg
Syncline next to Nordenskjöld Lake in Torres del Paine National Park. The syncline formed during the Andean orogeny.

The early development of the Andean orogeny in southernmost South America affected also the Antarctic Peninsula. [42] In southern Patagonia at the onset of the Andean orogeny in the Jurassic, extensional tectonics created the Rocas Verdes Basin, a back-arc basin whose southeastern extension survives as the Weddell Sea in Antarctica. [42] [43] In the Late Cretaceous the tectonic regime of Rocas Verdes Basin changed leading to its transformation into a compressional foreland basin –the Magallanes Basin– in the Cenozoic. This change was associated with an eastward move of the basin depocenter and the obduction of ophiolites. [42] [43] The closure of Rocas Verdes Basin in the Cretaceous is linked to the high-grade metamorphism of Cordillera Darwin Metamorphic Complex in southern Tierra del Fuego. [44]

As the Andean orogeny went on, South America drifted away from Antarctica during the Cenozoic leading first to the formation of an isthmus and then to the opening of the Drake Passage 45 million years ago. The separation from Antarctica changed the tectonics of the Fuegian Andes into a transpressive regime with transform faults. [42] [upper-alpha 10]

About 15 million years ago in the Miocene the Chile Ridge begun to subduct beneath the southern tip of Patagonia (55° S). The point of subduction, the triple junction has gradually moved to the north and lies at present at 47° S. The subduction of the ridge has created a northward moving "window" or gap in the asthenosphere beneath South America. [45]


Notes

  1. The Mochica Phase and the other phases in Peru were named by Gustav Steinmann (1856–1929) who established the first chronology of structural events in central Peru. [10]
  2. The validity in of this subdivision to describe the latest Andean orogeny in Peru has been questioned considering that deformation could have been continuous and migrating along the Andes. [12]
  3. The Quechua 1 Phase did also affect southern Peru and the Cordillera Oriental of Ecuador. [10]
  4. A series of iron ore deposits in the northern Chilean Coast Range known as the Chilean Iron Belt are related to the magmatism of La Negra Arc. [13]
  5. Northern Chile and the westernmost fringes of Bolivia.
  6. At least during the last 55 millions years.
  7. These sediments are grouped in the Abanico and Farellones Formation. [34]
  8. These marine sediments belong to Cuyo Group, Tordillo Formation, Auquilco Formation and Vaca Muerta Formation. [39]
  9. This inversion is thought to have led to the closure of Cura-Mallín Basin as evidenced by structural studies of Loncopué Trough. [40] However, evidence for Oligocene extension and rifting in the south-central Andes has been questioned. [41]
  10. Currently these faults have been carved into glacial valleys. [42]

Related Research Articles

<span class="mw-page-title-main">Andes</span> Mountain range in South America

The Andes, Andes Mountains or Andean Mountain Range are the longest continental mountain range in the world, forming a continuous highland along the western edge of South America. The range is 8,900 km (5,530 mi) long, 200 to 700 km wide, and has an average height of about 4,000 m (13,123 ft). The Andes extend from north to south through seven South American countries: Venezuela, Colombia, Ecuador, Peru, Bolivia, Chile, and Argentina.

<span class="mw-page-title-main">Laramide orogeny</span> Period of mountain building in North America

The Laramide orogeny was a time period of mountain building in western North America, which started in the Late Cretaceous, 70 to 80 million years ago, and ended 35 to 55 million years ago. The exact duration and ages of beginning and end of the orogeny are in dispute. The Laramide orogeny occurred in a series of pulses, with quiescent phases intervening. The major feature that was created by this orogeny was deep-seated, thick-skinned deformation, with evidence of this orogeny found from Canada to northern Mexico, with the easternmost extent of the mountain-building represented by the Black Hills of South Dakota. The phenomenon is named for the Laramie Mountains of eastern Wyoming. The Laramide orogeny is sometimes confused with the Sevier orogeny, which partially overlapped in time and space.

<span class="mw-page-title-main">Nazca Plate</span> Oceanic tectonic plate in the eastern Pacific Ocean basin

The Nazca Plate or Nasca Plate, named after the Nazca region of southern Peru, is an oceanic tectonic plate in the eastern Pacific Ocean basin off the west coast of South America. The ongoing subduction, along the Peru–Chile Trench, of the Nazca Plate under the South American Plate is largely responsible for the Andean orogeny. The Nazca Plate is bounded on the west by the Pacific Plate and to the south by the Antarctic Plate through the East Pacific Rise and the Chile Rise respectively. The movement of the Nazca Plate over several hotspots has created some volcanic islands as well as east-west running seamount chains that subduct under South America. Nazca is a relatively young plate both in terms of the age of its rocks and its existence as an independent plate having been formed from the break-up of the Farallon Plate about 23 million years ago. The oldest rocks of the plate are about 50 million years old.

<span class="mw-page-title-main">Salar de Atacama</span>

Salar de Atacama is the largest salt flat in Chile. It is located 55 km (34 mi) south of San Pedro de Atacama, is surrounded by mountains, and has no drainage outlets. In the east it is enclosed by the main chain of the Andes, while to the west lies a secondary mountain range of the Andes called Cordillera de Domeyko. Large volcanoes dominate the landscape, including the Licancabur, Acamarachi, Aguas Calientes and the Láscar. The last is one of the most active volcanoes in Chile. All of them are located along the eastern side of the Salar de Atacama, forming a generally north-south trending line of volcanoes that separate it from smaller endorheic basins.

<span class="mw-page-title-main">Andean Volcanic Belt</span> Volcanic belt in South America

The Andean Volcanic Belt is a major volcanic belt along the Andean cordillera in Argentina, Bolivia, Chile, Colombia, Ecuador, and Peru. It is formed as a result of subduction of the Nazca Plate and Antarctic Plate underneath the South American Plate. The belt is subdivided into four main volcanic zones which are separated by volcanic gaps. The volcanoes of the belt are diverse in terms of activity style, products, and morphology. While some differences can be explained by which volcanic zone a volcano belongs to, there are significant differences within volcanic zones and even between neighboring volcanoes. Despite being a type location for calc-alkalic and subduction volcanism, the Andean Volcanic Belt has a broad range of volcano-tectonic settings, as it has rift systems and extensional zones, transpressional faults, subduction of mid-ocean ridges and seamount chains as well as a large range of crustal thicknesses and magma ascent paths and different amounts of crustal assimilations.

<span class="mw-page-title-main">Geology of Chile</span>

The geology of Chile is a characterized by processes linked to subduction, such as volcanism, earthquakes, and orogeny. The building blocks of Chile's geology were assembled during the Paleozoic Era when Chile was the southwestern margin of the supercontinent Gondwana. In the Jurassic, Gondwana began to split, and the ongoing period of crustal deformation and mountain building known as the Andean orogeny began. In the Late Cenozoic, Chile definitely separated from Antarctica, and the Andes experienced a significant rise accompanied by a cooling climate and the onset of glaciations.

<span class="mw-page-title-main">Gondwana</span> Neoproterozoic to Cretaceous landmass

Gondwana was a large landmass, often referred to as a supercontinent, that formed during the late Neoproterozoic and began to break up during the Jurassic period. The final stages of break-up, involving the separation of Antarctica from South America and Australia, occurred during the Paleogene. Gondwana was not considered a supercontinent by the earliest definition, since the landmasses of Baltica, Laurentia, and Siberia were separated from it. To differentiate it from the Indian region of the same name, it is also commonly called Gondwanaland.

<span class="mw-page-title-main">Geology of Bolivia</span>

The geology of Bolivia comprises a variety of different lithologies as well as tectonic and sedimentary environments. On a synoptic scale, geological units coincide with topographical units. The country is divided into a mountainous western area affected by the subduction processes in the Pacific and an eastern lowlands of stable platforms and shields. The Bolivian Andes is divided into three main ranges; these are from west to east: the Cordillera Occidental that makes up the border to Chile and host several active volcanoes and geothermal areas, Cordillera Central once extensively mined for silver and tin and the relatively low Cordillera Oriental that rather than being a range by its own is the eastern continuation of the Central Cordillera as a fold and thrust belt. Between the Occidental and Central Cordillera the approximately 3,750-meter-high Altiplano high plateau extends. This basin hosts several freshwater lakes, including Lake Titicaca as well as salt-covered dry lakes that bring testimony of past climate changes and lake cycles. The eastern lowlands and sub-Andean zone in Santa Cruz, Chuquisaca, and Tarija Departments was once an old Paleozoic sedimentary basin that hosts valuable hydrocarbon reserves. Further east close to the border with Brazil lies the Guaporé Shield, made up of stable Precambrian crystalline rock.

<span class="mw-page-title-main">Kutai Basin</span>

The Kutai sedimentary basin extends from the central highlands of Borneo, across the eastern coast of the island and into the Makassar Strait. With an area of 60,000 km2, and depths up to 15 km, the Kutai is the largest and deepest Tertiary age basin in Indonesia. Plate tectonic evolution in the Indonesian region of SE Asia has produced a diverse array of basins in the Cenozoic. The Kutai is an extensional basin in a general foreland setting. Its geologic evolution begins in the mid Eocene and involves phases of extension and rifting, thermal sag, and isostatic subsidence. Rapid, high volume, sedimentation related to uplift and inversion began in the Early Miocene. The different stages of Kutai basin evolution can be roughly correlated to regional and local tectonic events. It is also likely that regional climate, namely the onset of the equatorial ever wet monsoon in early Miocene, has affected the geologic evolution of Borneo and the Kutai basin through the present day. Basin fill is ongoing in the lower Kutai basin, as the modern Mahakam River delta progrades east across the continental shelf of Borneo.

<span class="mw-page-title-main">Famatinian orogeny</span> Paleozoic geological event in South America

The Famatinian orogeny is an orogeny that predates the rise of the Andes and that took place in what is now western South America during the Paleozoic, leading to the formation of the Famatinian orogen also known as the Famatinian belt. The Famatinian orogeny lasted from the Late Cambrian to at least the Late Devonian and possibly the Early Carboniferous, with orogenic activity peaking about 490 to 460 million years ago. The orogeny involved metamorphism and deformation in the crust and the eruption and intrusion of magma along a Famatinian magmatic arc that formed a chain of volcanoes. The igneous rocks of the Famatinian magmatic arc are of calc-alkaline character and include gabbros, tonalites, granodiorites and trondhjemites. The youngest igneous rocks of the arc are granites.

Abanico Formation is a 3 kilometres (9,800 ft) thick sedimentary formation exposed in the Andes of Central Chile. The formation has been deposited in a timespan from the Eocene to the Miocene. Abanico Formation's contact with the overlying Miocene Farellones Formation has been the subject of differing interpretations since the 1960s. A small part of the formation crops out in the Mendoza Province of western Argentina.

Lauca is a 5,140 metres (16,860 ft) high andesitic stratovolcano in the Central Volcanic Zone of the Andes on the Altiplano in northern Chile. Administratively it is located in Putre, Arica y Parinacota Region. The volcano was active during the Late Miocene from 10.5 million years ago onwards. A major ignimbrite collapsed the volcano in the Late Pliocene.

<span class="mw-page-title-main">Flat slab subduction</span> Subduction characterized by a low subduction angle

Flat slab subduction is characterized by a low subduction angle beyond the seismogenic layer and a resumption of normal subduction far from the trench. A slab refers to the subducting lower plate. Although, some would characterize flat slab subduction as any shallowly dipping lower plate as in western Mexico. Flat slab subduction is associated with the pinching out of the asthenosphere, an inland migration of arc magmatism, and an eventual cessation of arc magmatism. The coupling of the flat slab to the upper plate is thought to change the style of deformation occurring on the upper plate's surface and form basement-cored uplifts like the Rocky Mountains. The flat slab also may hydrate the lower continental lithosphere and be involved in the formation of economically important ore deposits. During the subduction, a flat slab itself may be deformed, or buckling, causing sedimentary hiatus in marine sediments on the slab. The failure of a flat slab is associated with ignimbritic volcanism and the reverse migration of arc volcanism. Multiple working hypotheses about the cause of flat slabs are subduction of thick, buoyant oceanic crust (15–20 km) and trench rollback accompanying a rapidly overriding upper plate and enhanced trench suction. The west coast of South America has two of the largest flat slab subduction zones. Flat slab subduction is occurring at 10% of subduction zones.

Tectonic erosion or subduction erosion is the loss of crust from an overriding tectonic plate due to subduction. Two types of tectonic erosion exist: frontal erosion at the outer margin of a plate and basal erosion at the base of the plate's crust. Basal erosion causes a thinning of the overriding plate. When frontal tectonic erosion consumes a crustal block at the outer margin it may induce a domino effect on upper crustal tectonics causing the remaining blocks to fault and tilt to fill the “gap” left by the consumed block. Subduction erosion is believed to be enhanced by high convergence rates and low sediment supply to the trench.

<span class="mw-page-title-main">Tectonic evolution of Patagonia</span>

Patagonia comprises the southernmost region of South America, portions of which lie on either side of the Argentina-Chile border. It has traditionally been described as the region south of the Rio Colorado, although the physiographic border has more recently been moved southward to the Huincul fault. The region's geologic border to the north is composed of the Rio de la Plata craton and several accreted terranes comprising the La Pampa province. The underlying basement rocks of the Patagonian region can be subdivided into two large massifs: the North Patagonian Massif and the Deseado Massif. These massifs are surrounded by sedimentary basins formed in the Mesozoic that underwent subsequent deformation during the Andean orogeny. Patagonia is known for its vast earthquakes and the damage they cause.

The Bolivian tin belt is a mineral-rich region in the Cordillera Oriental of Bolivia. Being a metallogenetic province the Bolivian tin belt is rich in tin, tungsten, silver and base metals. The Bolivian tin belt follows the same bend as the Bolivian orocline. The mineralizations of the belt were formed episodically beginning in the Triassic and with the youngest known mineralizations dating to the Miocene.

The geology of Ecuador includes ancient Precambrian basement rock and a complex tectonic assembly of new sections of crust from formerly separate landmasses, often uplifted as the Andes or transformed into basins.

The geology of Argentina includes ancient Precambrian basement rock affected by the Grenville orogeny, sediment filled basins from the Mesozoic and Cenozoic as well as newly uplifted areas in the Andes.

<span class="mw-page-title-main">Geology of Peru</span>

The geology of Peru includes ancient Proterozoic rocks, Paleozoic and Mesozoic volcanic and sedimentary rocks, and numerous basins and the Andes Mountains formed in the Cenozoic.

Purilactis Group is a heterogeneous group of volcanic, volcano-sedimentary and formations of Cretaceous to Eocene age in Salar de Atacama basin, northern Chile. The group has a stratigraphic thickness of more than 6000 m. The group overlies basement rocks of Late Paleozoic age. The north-south El Bordo Escarpment of Cordillera Domeyko contain the main outcrops of the group. The group has been difficult to date in detail since it hosts few fossils and dateable minerals. The sediments of the group deposited when volcanism in the area was mainly occurring to the west of it, rather than to east as in the present-day. In geological terms this qualifies the basin as a back-arc basin.

References

  1. 1 2 3 4 5 6 7 Ramos, Víctor A. (2009). "Anatomy and global context of the Andes: Main geologic features and the Andean orogenic cycle". Backbone of the Americas: Shallow Subduction, Plateau Uplift, and Ridge and Terrane Collision. Geological Society of America Memoirs. Vol. 204. pp. 31–65. doi:10.1130/2009.1204(02). ISBN   9780813712048 . Retrieved December 15, 2015.
  2. Charrier et al. 2006, pp. 113–114.
  3. 1 2 3 4 5 6 Charrier et al. 2006, pp. 45–46.
  4. 1 2 Hoffmann-Rothe, Arne; Kukowski, Nina; Dresen, Georg; Echtler, Helmut; Oncken, Onno; Klotz, Jürgen; Scheuber, Ekkehard; Kellner, Antje (2006). "Oblique Convergence along the Chilean Margin: Partitioning, Margin-Parallel Faulting and Force Interaction at the Plate Interface". In Oncken, Onno; Chong, Guillermo; Franz, Gerhard; Giese, Peter; Götze, Hans-Jürgen; Ramos, Víctor A.; Strecker, Manfred R.; Wigger, Peter (eds.). The Andes: Active Subduction Orogeny . pp.  125–146. ISBN   978-3-540-24329-8.
  5. 1 2 Garcia-Castellanos, D (2007). "The role of climate in high plateau formation. Insights from numerical experiments". Earth Planet. Sci. Lett. 257 (3–4): 372–390. Bibcode:2007E&PSL.257..372G. doi:10.1016/j.epsl.2007.02.039. hdl: 10261/67302 .
  6. 1 2 3 4 5 Orme, Antony R. (2007). "The Tectonic Framework of South America". In Veblen, Thomas T.; Young, Kenneth R.; Orme, Anthony R. (eds.). Physical Geography of South America . Oxford University Press. pp.  12–17.
  7. Kerr, Andrew C.; Tarney, John (2005). "Tectonic evolution of the Caribbean and northwestern South America: The case for accretion of two Late Cretaceous oceanic plateaus". Geology . 33 (4): 269–272. Bibcode:2005Geo....33..269K. doi:10.1130/g21109.1.
  8. Audemard M., Franck A.; Singer P., André; Soulas, Jean-Pierre (2006). "Quaternary faults and stress regime of Venezuela" (PDF). Revista de la Asociación Geológica Argentina . 61 (4): 480–491. Retrieved November 24, 2015.
  9. Frutos, J. (1990). "The Andes Cordillera: A Synthesis of the Geologic Evolution". In Fontboté, L.; Amstutz, G.C.; Cardozo, M.; Cedillo, E.; Frustos, J. (eds.). Stratabound Ore Deposits in the Andes. Springer-Verlag. pp. 12–15.
  10. 1 2 3 4 5 6 7 8 9 Pfiffner, Adrian O.; Gonzalez, Laura (2013). "Mesozoic–Cenozoic Evolution of the Western Margin of South America: Case Study of the Peruvian Andes". Geosciences. 3 (2): 262–310. Bibcode:2013Geosc...3..262P. doi: 10.3390/geosciences3020262 .
  11. 1 2 3 4 Mégard, F. (1984). "The Andean orogenic period and its major structures in central and northern Peru". Journal of the Geological Society, London. 141 (5): 893–900. Bibcode:1984JGSoc.141..893M. doi:10.1144/gsjgs.141.5.0893. S2CID   128738174 . Retrieved December 26, 2015.
  12. 1 2 3 4 5 6 Mora, Andres; Baby, Patrice; Roddaz, Martin; Parra, Mauricio; Brusset, Stéphane; Hermoza, Wilber; Espurt, Nicolas (2010). "Tectonic history of the Andes and sub-Andean zones: implications for the development of the Amazon drainage basin". In Hoorn, C.; Wesselingh, F.P. (eds.). Amazonia, Landscape and Species Evolution: A Look into the Past . pp.  38–60.
  13. Tornos, Fernando; Hanchar, John M.; Munizaga, Rodrigo; Velasco, Francisco; Galindo, Carmen (2020). "The role of the subducting slab and melt crystallization in the formation of magnetite-(apatite) systems, Coastal Cordillera of Chile". Mineralium Deposita . 56 (2): 253–278. doi:10.1007/s00126-020-00959-9. ISSN   0026-4598. S2CID   212629723.
  14. Charrier et al. 2006, pp. 47–48.
  15. Salfity, J.A.; Marquillas, R.A. (1994). "Tectonic and Sedimentary Evolution of the Cretaceous-Eocene Salta Group Basin, Argentina". In Salfity, J.A. (ed.). Cretaceous Tectonics of the Andes. pp. 266–315.
  16. Reutter, Klaus-J.; Charrier, Reynaldo; Götze, Hans-J.; Schurr, Bernd; Wigger, Peter; Scheuber, Ekkehard; Giese, Peter; Reuther, Claus-Dieter; Schmidt, Sabine; Rietbrock, Andreas; Chong, Guillermo; Belmonte-Pool, Arturo (2006). "The Salar de Atacama Basin: a Subsiding Block within the Western Edge of the Altiplano-Puna Plateau". In Oncken, Onno; Chong, Guillermo; Franz, Gerhard; Giese, Peter; Götze, Hans-Jürgen; Ramos, Víctor A.; Strecker, Manfred R.; Wigger, Peter (eds.). The Andes: Active Subduction Orogeny. pp. 303–325. ISBN   978-3-540-24329-8.
  17. Mpodozis, Constantino; Arriagada, César; Roperch, Pierrick (October 6, 1999). Cretaceous to Paleogene geology of the Salar de Atacama basin, northern Chile: A reappraisal of the Purilactis Group stratigraphy. Fourth ISAG, Goettingen. Goettingen, Germany.
  18. Devries, T.J. (1998). "Oligocene deposition and Cenozoic sequence boundaries in the Pisco Basin (Peru)". Journal of South American Earth Sciences . 11 (3): 217–231. Bibcode:1998JSAES..11..217D. doi:10.1016/S0895-9811(98)00014-5.
  19. 1 2 Macharé, José; Devries, Thomas; Barron, John; Fourtanier, Élisabeth (1988). "Oligo-Miocene transgression along the Pacifie margin of South America: new paleontological and geological evidence from the Pisco basin (Peru)" (PDF). Geódynamique. 3 (1–2): 25–37.
  20. Charrier et al. 2006, pp. 100–101.
  21. Charrier et al. 2006, p. 97.
  22. 1 2 DeCelles, Peter G.; Horton, Brian K. (2003). "Early to middle Tertiary foreland basin development and the history of Andean crustal shortening in Bolivia". Geological Society of America Bulletin. 115 (1): 58–77. Bibcode:2003GSAB..115...58D. doi:10.1130/0016-7606(2003)115<0058:etmtfb>2.0.co;2.
  23. 1 2 3 4 Isacks, Bryan L. (1988). "Uplift of the Central Andean Plateau and Bending of the Bolivian Orocline". Journal of Geophysical Research . 93 (B4): 3211–3231. Bibcode:1988JGR....93.3211I. doi:10.1029/jb093ib04p03211.
  24. 1 2 Tassara, Andrés (2005). "Interaction between the Nazca and South American plates and formation of the Altiplano-Puna Plateau: Review of a flexural analysis along the Andean margin (15°-34°S)". Tectonophysics . 399 (1–4): 39–57. Bibcode:2005Tectp.399...39T. doi:10.1016/j.tecto.2004.12.014.
  25. Charrier, Reynaldo; Reutter, Klaus-J. (1990). "The Purilactis Group of Northern Chile: Boundary Between Arc and Backarc from Late Cretaceous to Eocene". In Reutter, Klaus-Joachim; Scheuber, Ekkehard; Wigger, Peter J. (eds.). Tectonics of the Southern Central Andes. Springer, Berlin, Heidelberg. pp. 189–202. doi:10.1007/978-3-642-77353-2. ISBN   978-3-642-77353-2.
  26. Hongn, F.; del Papa, C.; Powell, J.; Petrinovic, I.; Mon, R.; Deraco, V. (2007). "Middle Eocene deformation and sedimentation in the Puna–Eastern Cordillera transition (23°–26°S): Control by preexisting heterogeneities on the pattern of initial Andean shortening". Geology. 35 (3): 271–274. Bibcode:2007Geo....35..271H. doi:10.1130/G23189A.1. hdl: 11336/55884 .
  27. Pingel, Heiko; Strecker, Manfred R.; Alonso, Ricardo N.; Schmitt, Axel K. (2012). "Neotectonic basin and landscape evolution in the Eastern Cordillera of NW Argentina, Humahuaca Basin (~24°S)". Basin Research. 25 (5): 554–573. Bibcode:2013BasR...25..554P. doi:10.1111/bre.12016. S2CID   111384903 . Retrieved December 26, 2015.
  28. Milani, José; Zalán, Pedro Victor (1999). "An outline of the geology and petroleum systems of the Paleozoic interior basins of South America". Episodes . 22 (3): 199–205. doi: 10.18814/epiiugs/1999/v22i3/007 .
  29. Capitanio, F.A.; Faccenna, C.; Zlotnik, S.; Stegman, D.R. (2011). "Subduction dynamics and the origin of Andean orogeny and the Bolivian orocline". Nature. 480 (7375): 83–86. Bibcode:2011Natur.480...83C. doi:10.1038/nature10596. hdl: 2117/16106 . PMID   22113613. S2CID   205226860.
  30. Mlynarczyk, Michael S.J.; Williams-Jones, Anthony E. (2005). "The role of collisional tectonics in the metallogeny of the Central Andean tin belt". Earth and Planetary Science Letters. 240 (3–4): 656–667. Bibcode:2005E&PSL.240..656M. doi:10.1016/j.epsl.2005.09.047.
  31. Hartley, Adrian J. (2003). "Andean uplift and climate change". Journal of the Geological Society, London. 160 (1): 7–10. Bibcode:2003JGSoc.160....7H. doi:10.1144/0016-764902-083. S2CID   128703154.
  32. Charrier et al. 2006, p. 21.
  33. 1 2 3 Giambiagi, Laura; Mescua, José; Bechis, Florencia; Hoke, Gregory; Suriano, Julieta; Spagnotto, Silvana; Moreiras, Stella Maris; Lossada, Ana; Mazzitelli, Manuela; Toural Dapoza, Rafael; Folguera, Alicia; Mardonez, Diego; Pagano, Diego Sebastián (2016). "Cenozoic Orogenic Evolution of the Southern Central Andes (32–36°S)". In Folguera, Andrés; Naipauer, Maximiliano; Sagripanti, Lucía; Ghiglione, Matías C.; Orts, Darío L.; Giambiagi, Laura (eds.). Growth of the Southern Andes. Springer. pp. 63–98. ISBN   978-3-319-23060-3.
  34. 1 2 Charrier et al. 2006, pp. 93–94.
  35. 1 2 3 4 5 Charrier, Reynaldo; Iturrizaga, Lafasam; Charretier, Sebastién; Regard, Vincent (2019). "Geomorphologic and Glacial Evolution of the Cachapoal and southern Maipo catchments in the Andean Principal Cordillera, Central Chile (34°-35º S)". Andean Geology . 46 (2): 240–278. doi: 10.5027/andgeoV46n2-3108 . Retrieved June 9, 2019.
  36. Rapela, C.W.; Pankhurst, R.J; Casquet, C.; Baldo, E.; Saavedra, J.; Galindo, C.; Fanning, C.M. (1998). "The Pampean Orogeny of the southern proto-Andes: Cambrian continental collision in the Sierras de Córdoba" (PDF). In Pankhurst, R.J; Rapela, C.W. (eds.). The Proto-Andean Margin of Gondwana. Geological Society, London, Special Publications. Vol. 142. pp. 181–217. doi:10.1144/GSL.SP.1998.142.01.10. S2CID   128814617 . Retrieved December 7, 2015.
  37. Ramos, Victor A.; Cristallini, E.O.; Pérez, Daniel J. (2002). "The Pampean flat-slab of the Central Andes". Journal of South American Earth Sciences . 15 (1): 59–78. Bibcode:2002JSAES..15...59R. doi:10.1016/S0895-9811(02)00006-8.
  38. 1 2 3 4 5 Ramos, Víctor A.; Mahlburg Kay, Suzanne (2006). "Overview of the tectonic evolution of the southern Central Andes of Mendoza and Neuquén (35°–39°S latitude)". In Mahlburg Kay, Suzanne; Ramos, Víctor A. (eds.). Evolution of an Andean Margin: A Tectonic and Magmatic View from the Andes to the Neuquén Basin (35–39°S lat) . pp.  1–17.
  39. 1 2 3 4 Rojas Vera, Emilio Agustín; Orts, Darío L.; Folguera, Andrés; Zamora Valcarce, Gonzalo; Bottesi, Germán; Fennell, Lucas; Chiachiarelli, Francisco; Ramos, Víctor A. (2016). "The Transitional Zone Between the Southern Central and Northern Patagonian Andes (36–39°S)". In Folguera, Andrés; Naipauer, Maximiliano; Sagripanti, Lucía; Ghiglione, Matías C.; Orts, Darío L.; Giambiagi, Laura (eds.). Growth of the Southern Andes. Springer. pp. 99–114. ISBN   978-3-319-23060-3.
  40. Rojas Vera, Emilio A.; Folguera, Andrés; Zamora Valcarce, Gonzalo; Gímenez, Mario; Martínez, Patricia; Ruíz, Francisco; Bottesi, Germán; Ramos, Víctor A. (2011). "La fosa de Loncopué en el piedemonte de la cordillera neuquina.". Relatorio del XVIII Congreso Geológico Argentino. XVIII Congreso Geológico Argentino (in Spanish). Neuquén. pp. 375–383.
  41. Cobbold, Peter R.; Rossello, Eduardo A.; Marques, Fernando O. (2008). "Where is the evidence for Oligocene rifting in the Andes? Is it in the Loncopué Basin of Argentina?". Extended abstracts. 7th International Symposium on Andean Geodynamics. Nice. pp. 148–151.
  42. 1 2 3 4 5 Ghiglione, Matías C. (2016). "Orogenic Growth of the Fuegian Andes (52–56°) and Their Relation to Tectonics of the Scotia Arc". In Folguera, Andrés; Naipauer, Maximiliano; Sagripanti, Lucía; Ghiglione, Matías C.; Orts, Darío L.; Giambiagi, Laura (eds.). Growth of the Southern Andes. Springer. pp. 241–267. ISBN   978-3-319-23060-3.
  43. 1 2 Wilson, T.J. (1991). "Transition from back-arc to foreland basin development in the southernmost Andes: Stratigraphic record from the Ultima Esperanza District, Chile". Geological Society of America Bulletin. 103 (1): 98–111. Bibcode:1991GSAB..103...98W. doi:10.1130/0016-7606(1991)103<0098:tfbatf>2.3.co;2.
  44. Hervé, F.; Fanning, C.M.; Pankhurst, R.J.; Mpodozis, C.; Klepeis, K.; Calderón, M.; Thomson, S.N. (2010). "Detrital zircon SHRIMP U–Pb age study of the Cordillera Darwin Metamorphic Complex of Tierra del Fuego: sedimentary sources and implications for the evolution of the Pacific margin of Gondwana" (PDF). Journal of the Geological Society, London. 167 (3): 555–568. Bibcode:2010JGSoc.167..555H. doi:10.1144/0016-76492009-124. S2CID   129413187.
  45. Charrier et al. 2006, p. 112.

Further reading