You are on page 1of 34

1

Ramos, Victor A., 2018, Tectonic evolution of the central Andes: From terrane accre-
tion to crustal delamination, in G. Zamora, K. R. McClay, and V. A. Ramos, eds.,
Petroleum basins and hydrocarbon potential of the Andes of Peru and Bolivia:
AAPG Memoir 117, p. 1–34.

Tectonic Evolution of the Central Andes:


From Terrane Accretion to Crustal
Delamination
Victor A. Ramos
Instituto de Estudios Andinos Don Pablo Groeber, Facultad de Ciencias Exactas y Naturales (FCEN),
Universidad de Buenos Aires— Consejo Nacional de Investigaciones Científicas y Tecnológicas (CONICET)
Ciudad Universitaria – Pabellón 2, 1428, Buenos Aires, Argentina
Email: andes@gl.fcen.uba.ar

ABSTRACT
The analysis of the pre-Andean history of the Central Andes shows a complex tectonic evo-
lution. The basement of the Andean continental margin was formed by the accretion of
­Precambrian blocks during the formation of Rodinia in late Mesoproterozoic times. There
are two magmatic arcs of Grenvillian age, one developed on the margin of the craton, known
as the Sunsas belt, and another on the accreted terranes. The suture between these blocks with
the Amazonian craton has been continuously reactivated by tectonic and magmatic processes.
The terranes of Paracas and Arequipa, both of Grenvillian age, have a contrasting Paleozoic
evolution. The Arequipa terrane amalgamated to the craton by the end of the Mesoproterozoic,
and during the Paleozoic its suture acted as a crustal weakness zone. This zone concentrated
the extension and the formation of a large platform in the retro-arc basin, where the Eopaleo-
zoic sediments accumulated. The Famatinian magmatic arc of Ordovician age 14752460 Ma2
is preserved in this segment along the continental margin. The Eopaleozoic extension that
affected the Paracas terrane reopened the old suture and formed oceanic crust between Am-
azonia and Paracas. The subduction of this oceanic crust developed a magmatic arc over the
cratonic margin, which is preserved in the Eastern Cordillera of Peru as orthogneisses asso-
ciated with metamorphic rocks of Famatinian age. There are ophiolitic assemblages, paired
metamorphic belts, and intense deformation associated with the Paracas collision (∼460 Ma)
against the Amazonian craton. In northern Eastern Cordillera of Peru the late Paleozoic oro-
gen has within-plate granitic belts and was far away from the active margin. The orogen was
deformed and uplifted in two phases (336–285 Ma and 280–235 Ma) known as the early and
late Gondwanide orogenies. They are preserved as medium grade metamorphic belts devel-
oped along the Paracas segment. Further south along the Arequipa segment in southern Peru
and Bolivia, the late Paleozoic–Triassic rocks are represented by granites and acidic volcanic
rocks, which are not metamorphosed and are associated with sedimentary rocks. Relics of a

Copyright ©2018 by The American Association of Petroleum Geologists.


DOI:10.1306/13622115M1172855

1
Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf
by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 1 11/05/18 12:18 PM


2 Ramos

magmatic arc are exposed as tonalites and metamorphic rocks 1∼260 Ma2 along the northern
continental margin of Peru and in the near offshore platform. The extensional regime that
dominated most of the Mesozoic developed rift basins in the hanging-wall of the terrane su-
tures, which controlled the structural highs and basin margins. The Peruvian Late Cretaceous
orogeny produced the emplacement of the Coastal batholith, the beginning of deformation
along the coast, and the first foreland basins. The giant Ayabacas submarine syn-tectonic col-
lapse is also controlled by previous sutures. The Cenozoic Andean evolution was dominated
by a wave of shallowing of the subducted slab, the migration of the magmatism to the fore-
land, the steepening of the oceanic plate, and the consequent “inner arc” magmatism. The
“inner arc” plutonic and volcanic rocks are the expression of deep crustal melts, associated
with crustal delamination and lithospheric mantle removal. The flattening of the oceanic slab
is related to ablative subduction and shortening in the Altiplano and Eastern Cordillera. The
steepening is associated with rapid removal of mantle lithosphere and crustal delamination,
expressed at surface by the “inner arc” magmatism. The suture crustal weakness zones be-
tween different terranes partially controlled the location of the delaminated blocks and the
“inner arc” magmatism. Both processes triggered the lower crust ductile shortening and sub-
sequent upper crustal brittle development of the sub-Andean fold-and-thrust belt.

INTRODUCTION understanding of new plays that may have potentially


developed below productive Mesozoic sequences.
The understanding of the geology of the Central Therefore, the objective of this chapter is to present a
Andes of Peru and Bolivia has increased notably succinct appraisal of these new data, highlighting those
in the last decade from numerous investigations, that may have a direct relation with the evolution of
both from academia and from industry. This rapid the adjacent sub-Andean basins. Special attention will
advance of knowledge with specific works in dif- be paid to the terrane boundaries and sutures, because
ferent disciplines makes it difficult to update the these inherited crustal weakness zones have an import-
understanding and correlation of these contribu- ant role in the control of the extensional basins during
tions for the explorationists, whose specific objec- the early Mesozoic and the subsequent Andean defor-
tives are concentrated, most of the time, in the mation. Many processes such as lower crustal delam-
evaluation of sedimentary basins, generally located ination and lithospheric mantle removal are strongly
in the sub-Andean foreland. influenced by this inherited architecture, focusing up-
However, the many processes that have been iden- lift and deformation in the upper crust. I am aware
tified in this sector of the Central Andes are important that, because of these ambitious objectives, we will not
to understand the evolution of these sub-Andean ba- be able to treat them with the same detail, but we will
sins. For example, displacements of the volcanic front try to draw attention to those that are considered the
of the Andes, changes in the composition of magmatic most important for the purposes of this volume.
arcs, and the rapid uplift of mountain ranges have
been associated with crustal erosion by subduction,
changes in the Benioff zone geometry including steep- THE PRE-ANDEAN BASEMENT
ening of the subducted oceanic slab, with important
volcanic flare-ups on the surface. It is now known There is abundant evidence that recognizes the exis-
that part of these processes is related to crustal de- tence of a block of Precambrian basement attached to
lamination and thermal and mechanical removal of the Amazonian craton. The early seismological studies
the lithosphere, which produce important changes in of Dorbath et al. (1993) together with the gravimetric,
the dynamics of the fold-and-thrust belts of the sub-­ geochemical, and isotopic analyses of Mamani (2006)
Andean region. and Mamani et al. (2008, 2010) have outlined the limits
Besides these processes, the accretion and collision of a cratonic block, known as the Arequipa Massif. The
of terranes with their respective paired metamorphic boundary with Amazonia was also confirmed by the
belts, ophiolitic rocks and deformation, have been rec- occurrence of potassic and ultrapotassic mafic rocks of
ognized particularly in the Paleozoic. Our knowledge Cenozoic age, which delineate two deep lithospheric
of the geographical boundaries of these terranes and mantle blocks that have undergone different depletion
their deformation and uplift histories is critical in the and enrichment events (Carlier et al., 2005; Ramos,

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 2 11/05/18 12:18 PM


Tectonic Evolution of the Central Andes: From Terrane Accretion to Crustal Delamination  3

2008; Carlotto et al., 2009a). Further north, the pres- The present knowledge of the basement in Peru
ence of a cratonic block underlying the Western Cor- was set by the pioneering work of Dalmayrac et al.
dillera and adjacent offshore was demonstrated by the (1977, 1980; Figure 1a). These studies recognized the
studies of Romero Fernández et al. (2011) and Romero existence of an old early Proterozoic block named
Fernández et al. (2013), who confirmed a Precambrian as the Arequipa Massif, as well as the correlation
basement underlying the Paracas terrane. of the metamorphic rocks of the Eastern Cordillera
with similar rocks of Sierras Pampeanas (Figure 1a).
The existence of a ~2000 Ma-old granulitic base-
The Grenvillian-Age Basement ment along the coast of southern Peru was a tectonic
enigma that led to different hypotheses (see, e.g.,
Several attempts have been made to understand the Coira et al., 1982; Dalziel and Forsythe, 1985; R
­ amos,
regional tectonic evolution of the Precambrian and 1988, 2008; Bahlburg and Hervé, 1997; Wörner
Paleozoic basement rocks and the processes involved et al., 2000; Loewy et al., 2003, 2004). There is pres-
in the Andes of Peru and Bolivia. A milestone and ently some consensus that the Arequipa Massif, as
classic contribution to this comprehension of the part of the Arequipa terrane, was first accreted in
Andean geology of Peru was carried out by Mégard the late Mesoproterozoic during the amalgamation
(1978, 1979) and other ORSTOM’s researchers in the of the Rodinia supercontinent (Loewy et al., 2004;
1980s such as Dalmayrac, Laubacher, Marocco, as well ­R eimann et al., 2010) and later reactivated during
as those carried out in the 1990s in Bolivia, by Herail, the Ordovician but without involvement of oceanic
Sempere, Martinez, and several others. crust (Ramos, 2008).

Figure 1. A. First correlation between metamorphic belts of the Cordillera de Marañón in Peru and Sierras Pampeanas in
Argentina. Note the Grenville age of the Río Picharí outcrop and the outline of the Arequipa Massif (based on Dalmayrac
et al., 1977, 1980); B. Present knowledge of the metamorphic belt and the extension of the Paracas-Arequipa-Antofalla
basement and its off-shore continuation; the Grenvillian ages are based on Wörner et al. (2000); Loewy et al. (2004);
Miškovic’ et al. (2009a) and Romero et al. (2013). Ages in brackets are protolith ages. The proposed sutures and their
ages are modified from Ramos (2008, 2009, 2018). Note that the metamorphic belt is interrupted between Cordillera de
Vilcabamba and Sierras Pampeanas by the coeval sedimentary rocks of southern Peru and Bolivia.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 3 11/05/18 12:18 PM


4 Ramos

The first occurrence of Grenville-age basement in The Brasiliano–Pampean Basement


the Central Andes was reported by Dalmayrac (1978)
and Dalmayrac et al. (1980) along the Río Picharí valley There are no clear indications of late Neoproterozoic oro-
(Figure 1b). These outcrops have been later extended to genic rocks in the basement of the Central Andes between
other exposures by Miškovic’ et al. (2009a), who dated foli- 6 and 14°S latitude (Cardona et al., 2005a), in spite of the
ated late Mesoproterozoic rocks such as the Satipo tonalite late Neoproterozoic bulk zircon ages obtained by Dal-
by U-Pb in zircons at 985 { 14 Ma, the Querobamba mayrac et al. (1980). However, some authors interpreted
granite at 1123 { 13 Ma, and the Mariposa alkali feldspar the presence of a Neoproterozoic belt that would extend
granite at 1071 { 23 Ma. Most of these intrusives have in- below the younger Phanerozoic sequences or sub-Andean
herited zircon ages clustering at 1100 Ma, interpreted as cover (Chew et al., 2008; Escayola et al., 2011; Reimann et
the crystallization age, which allowed their assignment al., 2015), as a northern continuation of the Puncoviscana
to the Sunsas magmatic arc (Cordani et al., 2000). These belt. Although in the Laurentian margin in the vicinity
rocks represent a dismembered orogenic belt of Grenville of the southern Appalachians, the early formation of an
age as paraautochthonous domains that have remained ocean at 700 Ma, subsequently subducted during the
in proximity to the margin of Amazonia according to Car- approach of peri-Gondwanan terranes has been inferred
dona et al. (2010) and Chew et al. (2011, 2016). (Chew et al., 2008), no evidence of an active margin has
The main outcrops of granulitic rocks of late Meso- been recorded along the Andes at these latitudes.
proterozoic Grenville age were described and dated in There is consensus that during the late Neopro-
southern Peru by Wasteneys et al. (1995) and Loewy et terozoic the Central Andean margin was dominated
al. (2004), who recognized a late Mesoproterozoic met- by the opening of the proto-Iapetus Ocean (Miškovic’
amorphic event, superimposed on a Paleoproterozoic et al., 2009a), with some minor evidence of magmatic
basement of 2000 Ma. On Isla Hormigas de Afuera, activity related to the rift phase (Baldo et al., 2006;
along the offshore Paracas High in the continental ­Ramos, 2009). The active margin was developed in the
platform to the west of Lima (Figure 1b), high-grade Puncoviscana belt and their associated calc-alkaline
metamorphic rocks have been subsequently identified batholith, south of 22°S latitude (Ramos, 2008).
and dated by Romero et al. (2013), with similar ages of
crystallization and metamorphism.
One of the few exposures of Grenville-age rocks in The Famatinian Orogeny
Bolivia is at Cerro Uyarani (Figure 1b), in the northern
Altiplano near the boundary with Chile, which was dis- The Famatinian orogenic cycle was defined in north-
covered by Troëng et al. (1994). This hill of medium to western Argentina by Aceñolaza and Toselli (1976) as
high-grade metamorphic basement stands barely 500 m the orogeny that occurred after the Pampean (or late
(1640 ft) above the surface of the piedmont of the Western Brasiliano) orogeny during the late Neoproterozoic
Cordillera and is partially obliterated by the products of and persisted up to Middle-Late Devonian times.
extensive stratovolcanoes and Oligocene–Miocene syn- Subsequent studies have demonstrated that a major
orogenic deposits. These authors recognized other smaller orogeny ended at Middle Ordovician times, and as a
exposures further south in the vicinity of the town of Wila consequence, the Famatinian orogeny was restricted
Kkollu. The Uyarani metamorphic rocks have been dated from the base of the Cambrian up to the Middle Ordo-
by Wörner et al. (2000) at 1157 { 57 Ma with inherited vician in the Sierras Pampeanas of west central Argen-
zircons yield a 2024 { 133 Ma age. Different C ­ enozoic tina (Haller and Ramos, 1984; Astini and Benedetto,
volcanic rocks in the Altiplano have also recorded these 1993; Astini et al., 1995; Pankhurst et al., 1996, 1998).
Paleoproterozoic ages in several localities, as shown by The occurrence of metamorphic rocks associated
McLeod et al. (2013). These studies extended the base- with orthogneisses with ages contemporaneous with
ment from the eastern margin of the central region of the those of Sierras Pampeanas led Chew et al. (2007) to
Altiplano to the Huarina belt along the western margin identify the Famatinian orogeny in the Eastern Cordil-
of the Eastern Cordillera (Jiménez and López-Velásquez, lera of northern Peru. Rocks of similar ages are also
2008; Jiménez et al., 2009; Ramos and Jiménez, 2014). exposed further north in Venezuela and Colombia and
In summary the Mesoproterozoic metamorphism of were described by several authors in the Mérida and
Paracas and Arequipa terranes in Peru and Bolivia was Chibcha terranes (Bellizzia and Pimentel, 1994; Re-
associated with the formation of Rodinia, when these strepo-Pace et al., 1997; Alemán and Ramos, 2000). In
terranes collided with the Amazonian craton (Carlotto recent years further geochronology has enhanced that
et al., 2008, 2009a; Ramos, 2008, 2009; Miškovic’ et al., correlation, allowing the Famatinian orogeny to be ex-
2009a). The basement units of both terranes typically tended along the entire proto-Pacific margin of South
have an older Paleoproterozoic protolith. America, north of Patagonia (Ramos, 2018).

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 4 11/05/18 12:18 PM


Tectonic Evolution of the Central Andes: From Terrane Accretion to Crustal Delamination  5

Figure 2. Reconstruction of the Famatinian Orogen in the Central Andes. Ages of the Ordovician rocks in the
Marañón Domain based on Cardona et al. (2005a, 2007, 2009), Cardona (2006), and Chew et al. (2007,
2016); in the Vilcabamba Domain on Reitsma (2012) and Spikings et al. (2016); in the Paracas terrane on
Romero et al. (2013); in the Arequipa terrane on Wörner et al. (2000) and Loewy et al. (2004). Proposed
sutures and terrane boundaries modified from Ramos (2008, 2009, and 2018).

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 5 11/05/18 12:18 PM


6 Ramos

The Famatinian orogeny in the Central Andes of 2.422.6 kbar and 3002330°C 15722626°F2 estimated
Peru and Bolivia was controlled by the old Grenvillian for a phyllite-greenschist assemblage. According to
sutures, in a similar way to the northern Andes, where Willner et al. (2014) these contrasting metamorphic
Precambrian blocks were detached with the associated conditions defined a paired metamorphic belt, which
formation of oceanic crust and re-amalgamated to is characteristic of a magmatic arc environment. The
Gondwana, as proposed by Restrepo-Pace et al. (1997). age of the peak metamorphism is 465 { 24 Ma dated
These processes can be recognized north of the Aban- by a Sm-Nd mineral-whole rock isochron.
cay deflection in Peru, as described by Ramos (2008, The occurrence of ultramafic rocks in the depart-
2010c). However, the old suture between Arequipa ment of Huanuco was known from the pioneering re-
and the Amazonian craton did not separate to form connaissance of Antonio Raimondi in the 19th century,
oceanic crust (Figure 2) but instead controlled the sub- as testified by the talc schists and peridotites that are
sidence of a large retro-arc basin from Ordovician to exhibited in the Museum of Natural Sciences of the
Devonian times (Sempere, 1995; Suárez Soruco, 1999). National Major University of San Marcos (see locali-
Several distinct domains have been recognized in ties in Raimondi, 1929). The extent of these ultrabasic
the Famatinian orogen based on different rock types outcrops in Huanuco, Huancapallac, and Chinchao is
and inferred tectonic settings in this sector of the Cen- known from surveys of Aumaitre et al. (1977). These
tral Andes (see Figure 2). rocks currently known as the Tapo Ultramafic Com-
plex were interpreted as relics of oceanic crust, which
The Olmos-Loja Domain  This domain as defined were subducted and exhumed along a collisional su-
by Carlotto et al. (2009a) encompasses the region ture (Castroviejo et al., 2009a, b, 2010a, b; Fanlo et al.,
adjacent to the Huancabamba deflection along the 2009). These mafic-ultramafic complexes described for
border with Ecuador. Although most authors recog- 200 km (124 mi) from Huanuco to Tapo correspond
nize that the rotation associated with the deflection is to the suture between Paracas and Amazonia (Ra-
Late Cretaceous or younger (Mitouard et al., 1990), it mos, 2018). These rocks are associated with chromite
seems that an old structural Paleozoic terrane bound- ores, which extend further north and south and are
ary (Ramos, 2009) enhanced this rotation. The bound- of regional importance (Grandin and Zegarra, 1979).
aries between juxtaposed basement blocks together According to V. Benavides Cáceres (personal com-
with the later collision of Laurentia in the Late Pa- munication, 2017), these metamorphic and associated
leozoic are the candidates for controlling the present mafic rocks, together with serpentines, are exposed
deflection. south of the Abancay deflection associated with the
This domain is characterized by Ordovician sed- “Phyllite formation” of Steinmann (1929, p. 11). This
imentary and low-grade metamorphic rocks, which unit has been identified above 4000 m (13,123 ft) in
contrast with the high- to medium-grade metamorphic the Cordillera de Vilcabamba by Egeler and de Booy
rocks of the Marañón Massif further south (Reyes and (1961), who found low-grade metamorphic rocks bear-
Caldas, 1987). The Ordovician phyllites and schists of ing early Ordovician graptolites and the “Phyllite for-
the Olmos Complex and the metasedimentary rocks mation.” The ultramafic rocks are also exposed north
of the Salas Group continue to the north in Ecuador, of the Cordillera de Marañón; they have been de-
where similar rocks have been described in the vicini- scribed in Ñaupe in the Olmos Massif as serpentinized
ties of Loja (Aspden et al., 1995). peridotites (Grandin and Zegarra, 1979) and continue
even further north in the locality of Zumbia in the
The Marañón Domain  Chew et al. (2007) and Car- Cordillera Real of Ecuador according to Benavides
dona et al. (2009) dated Famatinian rocks in the Cáceres (personal communication, 2017). Based on
northern segment of the Eastern Cordillera, north this evidence, it is possible to recognize a discontin-
of the Abancay deflection. These authors described uous belt over 1200 km (746 mi) of these ultramafic
magmatic arc rocks (now orthogneisses) with ages rocks along the Eastern Cordillera of Peru and south-
similar to the Famatina type locality in the Sierras ern part of the Real Cordillera of Ecuador.
Pampeanas. U-Pb ages in zircons in these gneisses Whole rock geochemical data from siliciclastic de-
range from 475 to 460 Ma (Figure 2). tritus in the Paleozoic cover show a high Cr content,
Willner et al. (2014) identified high-­p ressure which may indicate that ophiolitic rocks were exposed
conditions west of the magmatic arc of to erosion in the source areas (Reimann et al., 2015).
11213 kbar>5002540°C 193221004°F2 during maxi- Structural studies of this ophiolitic belt in Huanuco
mum burial derived from a garnet amphibolite in the and Tarma, west of the Marañón Massif, demonstrated
Tapo Ultramafic Massif. The metamorphic rocks in the the development of phyllonites and mylonites along
western Marañón Complex have low PT conditions at the contacts, ruling out a previous interpretation of

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 6 11/05/18 12:18 PM


Tectonic Evolution of the Central Andes: From Terrane Accretion to Crustal Delamination  7

these rocks as peridotite igneous i­ ntrusives (­ Rodrigues (1978) disregarded the previously postulated Pre-
et al., 2010a). The ductile deformation D3 that affects cambrian ages of the green-schists exposed near
the peridotites and serpentinites in Huanuco has a Cusco (Heim, 1948a; Egeler and de Booy, 1961) and
southwest vergence (Rodrigues et al., 2010b), coher- recognized that these rocks were part of an Eoher-
ent with an east-dipping polarity of the previous cynian metamorphic belt. This assumption has been
subduction. confirmed in the Machu Picchu inlier of this region
These data led Ramos (2008, 2009) to define the by the granitic intrusions reported by Reitsma (2012)
Paracas terrane, a Grenvillian-age basement micro- and Spikings et al. (2016), who identified Famatinian
continent, which collided against the proto-margin of ages between 479.9 { 2.3 and 472.3 { 4.8 Ma in these
Gondwana in Ordovician times (see details in Ramos, rocks. South of the Vilcabamba domain, early Paleo-
2018). Based on these data, it is likely that the Paracas zoic sedimentary rocks are dominant (Figure 2).
microcontinent collided against the proto-margin with
a subduction zone dipping beneath the Marañón Mas- The Altiplano Domain  Clastic sedimentary se-
sif (Ramos, 2008, 2009, 2010a, c; Carlotto et al., 2011; quences of Ordovician age widely exposed in the Al-
Willner et al., 2014; Chew et al., 2016, among others). tiplano domain of southern Peru and Bolivia, south
The Paracas terrane has been interpreted as a paraau- of Cusco, have been described by Reimann et al.
tochthonous block based on the similar characteristics (2010, 2015). Igneous rocks are almost absent along
described by Romero et al. (2013), between the base- this segment, with some minor volcano-sedimentary
ments of Paracas and Arequipa. Both blocks share in rocks described by Bahlburg et al. (2006) in the Ol-
common a Grenville-age metamorphism, early Pro- lantaytambo and Umachiri Formations, which are
terozoic protoliths, and evidence of an overimposed exposed in the Machu Picchu inlier and on the Al-
Famatinian arc. tiplano, respectively. Detrital zircon U-Pb ages and
It is worth mentioning here that the studies of graptolite faunas indicate that the Ollantaytambo
Chew et al. (2007) and Cardona et al. (2009) recognize Formation is Ordovician in age (Bahlburg et al., 2011;
metamorphism around 480–470 Ma, with similar ages Reitsma, 2012).
as the main metamorphic peak described by Willner et These clastic deposits have been interpreted as a
al. (2014). However, Chew et al. (2016) identified a late continuous Ordovician retro-arc platform developed
Famatinian event in the Early Silurian, represented by inland from the magmatic arc, which is along the coast
high-grade ~435 Ma orthogneiss in the Marañón Com- of southern Peru and northern Chile (Ramos, 2008).
plex. These authors correlate this event with Ocloyic Further south, these rocks gradually pass again to
magmatism and deformation described by Bahlburg low-grade metamorphic facies, which are in continu-
et al. (2016) in the eastern Puna Eruptive Belt of north- ity with the higher grade metamorphic rocks of Sierras
ern Argentina (Faja Eruptiva de la Puna of Figure 2). Pampeanas.
This belt was formed in a back-arc setting of the main
Famatinian arc along the hanging-wall of the Grenville Summary of the Famatinian Orogeny  In summary
suture with volcanic rocks of 476 Ma interbedded with it is probable that the old Precambrian terrane that
graptolite-bearing sediments. This deformed sequence was accreted to the Amazonian Craton as part of Lau-
has been affected by 444 Ma old intrusives that Bahl- rentia during the Grenville orogeny was detached or
burg et al. (2016) relate to the Ocloyic orogenesis. This partially detached prior to the Famatinian Orogeny.
last event is part of a series of episodes recognized The northern block, known as the Paracas terrane, de-
in the Terra Australis orogen (Cawood, 2005; Ramos, tached sufficiently to form oceanic crust, which was
2009; Chew et al., 2016). later subducted, to form the Famatinian arc along
the Cordillera de Marañón (Chew et al., 2007, 2016;
The Vilcabamba Domain  The metamorphic rocks of Ramos, 2008). The southern block, the Arequipa ter-
the Cordillera de Vilcabamba south of the Cordillera rane, did not form oceanic crust, but the extension
de Marañón, in the vicinities of the Abancay deflec- produced a large retro-arc basin during Ordovician
tion (Figure 2), were described by Marocco (1978). to Devonian times (Sempere, 1995; Bahlburg et al.,
This author was the first to document the gradual 2006) coeval with the development of the magmatic
transition from the “Phyllite formation,” the fossil- arc along the coast on the Arequipa Massif. The
iferous Ordovician rocks, slightly metamorphic (of- boundary between both basement terranes coincides
ten with some schistosity or strain-slip cleavage), to with a major crustal boundary defined by Marocco
higher grade metamorphic rocks without evidence of (1978, p. 5) as the “Abancay deflection” (Figure 2).
a major discontinuity along a traverse across this cor- The change in metamorphic grade along strike of the
dillera. Based on these field observations, ­Marocco early Paleozoic rocks is related to the collision of the

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 7 11/05/18 12:18 PM


8 Ramos

­ araautochthonous and allochthonous blocks of Para-


p Litherland et al., 1994; Sánchez et al., 2006). These
cas (northern Peru) and Cuyania (northern Argentina gneissic rocks and foliated granitoids are also exposed
and Chile), respectively. The segments that recorded along the Huancabamba deflection in the Amotape–
collisions expose middle to deeper crustal levels and Tahuin domain. The rocks in this region are associated
are characterized by metamorphic rocks (Otamendi with the Río Piedras amphibolite, a mafic complex
et al., 2008). of Triassic age, interpreted as possible oceanic rocks
The northern sub-Andean foreland records a hiatus (Feininger, 1978; Litherland et al., 1994; Ramos, 2009).
with no sedimentation or erosion during Late Ordo- The Amotape–Tahuin basement has been consid-
vician–Silurian times, and only the Contaya Forma- ered by Feininger (1987) and Mourier et al. (1988) as
tion of possible Early Ordovician age is preserved in an allochthonous terrane, mainly based on the com-
the Huallaga and Marañón Basins of northern Peru parison with the Paleozoic sequences exposed in the
(Baby et al., 2018). Molasse deposits in these basins are Olmos Massif of southern Ecuador and the paleo-
represented by the Cabanillas Formation of Devonian magnetic data obtained from these areas. Mourier et
age (Calderon et al., 2017) and may represent the sy- al. (1988) proposed that this block was accreted by a
norogenic deposits of what Chew et al. (2016) identi- right-lateral strike-slip during Cretaceous times.
fied as the late Famatinian orogeny. Further south in Based on the U/Pb sensitive high-resolution ion
the Ucayali and Madre de Dios Basins, Alemán et al. microprobe (SHRIMP) and Ar-Ar dating of the met-
(2003) recognized an unconformity between the Caba- amorphic event that affected the Illescas Massif (Car-
nillas and Contaya Formations as related to the Oclo- dona et al., 2005b, 2008), Ramos (2009), following
yic deformation, which to the south is considered as these authors, interpreted that the collision took place
part of the late Famatinian orogeny. during the Permian. This last author proposed that the
The Bolivian sub-Andean region also records the Tahuín terrane of Feininger (1987) collided against the
Famatinian orogeny through thick sequences of Late Gondwana margin during the Alleghenide orogeny,
Ordovician–Devonian synorogenic deposits (Sempere, possibly as a part of Laurentia, and it was left on the
1995; Jiménez et al., 2009) in the Tarija Basin. Gondwana margin after being detached from Lauren-
As pointed out by Bahlburg and Hervé (1997) and tia. Some other authors, although correlated this de-
Chew et al. (2016), after the Famatinian deformation formation with the Alleghenide collision, interpreted
there is a magmatic and metamorphic gap between that the Tahuín terrane as an autochthonous part of
420 and 350 Ma (Late Silurian–Early Carboniferous). Gondwana (Bellido et al., 2009; Timoteo et al., 2012).
However, contractional deformation persisted during The new studies by Witt et al. (2017) in the Amotape
the Devonian with the development of foreland ba- region showed that the maximum age of the me-
sins in northwestern Argentina and southern Bolivia taquartzite of Illescas as well as the low-grade schists
(Starck, 1995). The tectonic control changed in the of Quebrada Seca are from locally derived igneous
Early Carboniferous when flexural loading subsidence source ranging from 331 to 329 Ma and from 321 to 319
of the foreland basin gave place to viscoelastic relax- Ma, respectively (Figure 3). These data also show that
ation of bending stresses (Tankard, 1986). This author the abundant occurrence of Pampean zircons in Illes-
also described subsequent upwarping along the ba- cas and Quebrada Seca (560–521 Ma and 575–525 Ma,
sin margin and subsidence in the basin axis during a respectively) indicate an autochthonous Gondwanan
period of tectonic quiescence. These viscoelastic ba- origin for the Amotape basement (Witt et al., 2017).
sins lose the previous asymmetry of a foreland basin These authors related the first phase of metamor-
by regional isostatic adjustments of the lithosphere to phism affecting the protolith at ~310 Ma, to terrane
thrust-belt loading. accretion of subduction dynamics following Cardona
et al. (2008). This Late Carboniferous episode could be
related to one of the early phases of the Alleghenide
The Alleghenide Orogeny orogen described by Hatcher (2010).
Five exploratory wells were drilled by oil companies
The northwestern sector of Peru together with the along the continental platform between Chiclayo and
adjacent ranges of Ecuador, in the Cerro de Amotapes Talara in recent years (Figure 3). Two of these wells are
(Figure 3), have been the focus of different tectonic located in the Outer Shelf High and intersected meta-
interpretations. Gneisses and granitoids of Late Paleo- morphic rocks interpreted as possible Precambrian in age
zoic to Triassic age such as the Tres Lagunas and (Romero Fernández et al., 2011). However, subsequent
Jubones foliated granitoids are well known along Ar-Ar dating showed Late Paleozoic ages in the San Mi-
the eastern margin of the Cordillera Real of Ecuador guel x-1 well (Timoteo et al., 2012; Romero et al., 2013).
(Aspden and Litherland, 1992; Aspden et al., 1992; The metadiorites intersected by the well were Ar-Ar

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 8 11/05/18 12:18 PM


Tectonic Evolution of the Central Andes: From Terrane Accretion to Crustal Delamination  9

Figure 3. The Amotape–Tahuin domain in northwestern Peru and Ecuador (based on Cardona et al., 2005b,
2008; Sánchez et al., 2006; Bellido et al., 2009; Romero Fernández et al., 2011; Timoteo et al., 2012; Romero
et al., 2013; Witt et al., 2017). Note the offshore extent of late Paleozoic rocks south of Illescas. (331–329 Ma):
Ages between brackets are maximum ages from igneous zircons in metamorphic rocks; (221 Ma): Ages in
italic are nondeformed granites of Triassic age.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 9 11/05/18 12:18 PM


10 Ramos

dated by these authors and an age of 259.5 { 12.3 Ma of Carboniferous to Early Permian between 336 and
was obtained. The age is similar to the metamorphic 285 Ma was characterized by a transition from mildly
257 { 8 Ma age obtained by Cardona et al. (2008) in the peraluminous and calc-alkaline diorites, ­granodiorites,
tonalitic gneisses of the Illescas Massif. Both ages show and monzogranites to a main Gondwanide phase of
an important metamorphic episode at about 260 Ma, juvenile granitoids lacking a subduction-­related signa-
subsequent to the emplacement of tonalities and diorites ture between 280 and 235 Ma (Miškovic’ et al., 2009a).
of a possible Late Paleozoic magmatic arc, which could According to these authors, the highly peraluminous
be related to late episodes of the Alleghenide orogeny. leucogranites of the Late Triassic (230–205 Ma) show
The new data of Witt et al. (2017), together with the termination of arc magmatism by detachment of
those provided by Romero et al. (2013), indicate the the subducting oceanic slab (Figure 4).
existence of subduction in the Western Cordillera The Gondwanide magmatism started in Carbonif-
along the continental margin in this sector of Western erous times in the northern Eastern Cordillera, shifted
Gondwana for the Late Paleozoic. However, the new to the south in the Permian being dominant in its cen-
data question the allochthonous nature of the Tahuín tral part, and, during the Triassic, ended in the south-
terrane for this time period. Later tectonic reactiva- ern segment of Eastern Cordillera of Peru, passing to
tions during the Mesozoic produced the rotation de- the Cordillera Real of Bolivia (Ávila Salinas, 1991).
tected by paleomagnetism as identified by Mourier The early Gondwanide deformation at about ~305 Ma
et al. (1988). (Chew et al., 2016) produced the angular unconfor-
The orthogneisses and foliated granitoids of several mity of Pennsylvanian age between Ambo and Tarma
sectors of the Northern Andes yielded Triassic Ar-Ar Formations in the Huallaga Basin and between Caba-
ages between 237 and 211 Ma (Vinasco, 2004; Sánchez nillas and Tarma Formations in the Marañón Basin
et al., 2006; Spikings et al., 2015), which were inter- (Calderon et al., 2017).
preted as crustal melts associated with the collapse of These magmatic phases can be related to the late
the Alleghenide orogen and subsequent breakup of Gondwanide metamorphism characterized by the
Pangea (Alemán and Ramos, 2000; Cediel et al., 2003; high-grade paragneisses of the Huaytapallana Com-
Martin-Gombojav and Winkler, 2008; Ramos, 2009). plex, constrained to ~260 Ma by U–Pb and Th–Pb
An alternative interpretation was advanced by Riel monazite age data (Chew et al., 2016). We agree with
et al. (2013, 2015), who proposed a high-temperature these authors that this orogenic phase cannot be re-
low-pressure metamorphism in the forearc zone for lated neither to flat-slab subduction along the margin
the El Oro metamorphic complex of the Tahuin block nor to terrane accretion because there is no evidence of
at 229.3 { 2.1 Ma. This extension was followed by un- collision of microcontinents during the Late Paleozoic
derplating of a high-pressure low-temperature unit at in this segment of the Central Andes (Ramos, 2009;
225.8 { 1.8 Ma during a contractional episode. Chew et al., 2016).
Subsequent extension during Triassic times indi- The mildly peraluminous nature of the initial Car-
cates the beginning of Pangea breakup after the colli- boniferous granitoids is similar to other intraplate
sion of Laurentia with Gondwana. magmatic events of this age developed along the pro-
to-margin of Gondwana. Carboniferous granites have
been described as weakly peraluminous in northern
The Gondwanide Orogeny and central Argentina (see Grosse et al., 2009). The
emplacement of the Carboniferous batholiths of Peru
After the pioneering studies of Steinmann (1929) and inboard of the continental margin can be explained
Steinmann et al. (1930), who described Paleozoic met- in a similar fashion to the Sierras Pampeanas, as a
amorphic rocks in the Eastern Cordillera of Peru, sev- response to the collapse of the previous orogenic de-
eral authors such as Mégard (1978) and Dalmayrac formation with extension of the crust and mantle up-
et al. (1980), among others, recognized an orogenic welling (Grosse et al., 2009). This extension shifted
episode characterized by igneous intrusions, defor- to the south in the Eastern Cordillera of Peru, east of
mation, and metamorphism as part of the Hercynian the previous sutures. The intrusion of the Permian
deformation. These rocks have been recently dated granitoids is characterized by the lack of a subduc-
and correlated with the Gondwanide orogeny, widely tion-related signature. The emplacement of these Late
recorded in southern Argentina and Chile (Miškovic’ et Paleozoic granites, not directly related to subduction
al., 2009a; Chew et al., 2016). These authors recognized in the asthenospheric wedge, is controlled by localized
discrete episodes of Gondwanide magmatism with extension along the hanging-wall of the Grenvillian–
variable intrusive facies and changes in styles of sub- Famatinian sutures. This reactivation of the suture
duction. An early Gondwanide phase from the middle explains, as proposed by Carlotto et al. (2011), the

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 10 11/05/18 12:18 PM


Tectonic Evolution of the Central Andes: From Terrane Accretion to Crustal Delamination  11

Figure 4. Late Paleozoic igneous rocks and representative ages of the Triassic phase after Jiménez and López-Velásquez
(2008), Jiménez et al. (2009), Chew et al. (2016), and Spikings et al. (2016). The outline of the Arequipa terrane is based
on Carlier et al. (2005) and Mamani et al. (2008, 2010); other boundaries are based on Ramos (2008, 2009) and Carlotto
et al. (2009a).

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 11 11/05/18 12:18 PM


12 Ramos

emplacement along this zone of structural weakness Medium-grade metamorphic rocks of this age are pre-
of the Pataz batholith, among others (Figure 4). served in the Cordillera Real of Ecuador and in the
This late Paleozoic “Hercynian” magmatism was in- Central ­Cordillera of ­Colombia. This clearly indicates
terpreted by Carlier et al. (1982) as connected to exten- that the d ­ eformation, uplift, and main peak of meta-
sional episodes affecting the continental crust. These morphism are associated with the collision of Lauren-
igneous rocks are now associated with an “inner arc,” tia and ­Gondwana in the Early Permian times.
as defined by Clark et al. (1984, p. 270), which was the This intense compressive deformation that occurred
locus of widespread crustal anatexis. Its tectonic set- at ∼260 Ma, the main Gondwanide orogeny at these
ting will be discussed in the analysis of the Oligocene latitudes, uplifted the Marañón Massif, exposed the
granites based on Ramos and Jiménez (2014). Paleozoic metamorphic rocks, and developed a Late
The extension that controlled the emplacement of Permian fold-and-thrust belt in the sub-Andean fore-
these Late Paleozoic granitoids in the Eastern Cor- land of the Huallaga and Marañón Basins ­(Calderon
dillera not only shifted from north to south but also et al., 2017). This deformation phase was so import-
decreased in intensity in that direction before ending ant in the northern sub-Andean basins that part of
nearby the Abancay deflection (Figure 4). The suture the underlying Paleozoic sequences was eroded away
associated with these intrusions corresponds to the according to Baby et al. (2018). This Late Permian up-
contact between the Paracas terrane and the Amazo- lift is also detected in the Ucayali and Madre de Dios
nian craton. Basins (Alemán et al., 2003). The angular unconfor-
It is interesting to note that the main phase of defor- mity between the synorogenic molasse deposits of the
mation identified in the northern Eastern Cordillera, Tarma and Copacabana Formations (Late Carbonifer-
as late Gondwanide at 260 Ma by Chew et al. (2016), ous–Early Permian) and the Mitu Group in the East-
coincides with the important metamorphism and igne- ern Cordillera or the Ene Formation and its equivalent
ous activity recorded in the Amotape–Tahuin domain units in the sub-­Andean basins is the expression of this
(Figure 3) along the northwestern coast of Peru and main Gondwanide deformation. Some authors have
southern Ecuador. The main metamorphism of tonal- referred to this unconformity as the Juruá orogeny
itic gneisses of the Illescas Massif was dated by U-Pb (Rosas et al., 2007; Carlotto et al., 2009b), but recent
SHRIMP around 257 Ma by Cardona et al. (2008), as studies in the Solimoes and Acre Basins in adjacent
well as the San Miguel x-1 metadiorite in 259 Ma by foreland of Brazil demonstrated that this orogeny is
Romero et al. (2013). much younger in the Juruá type locality (Late Jurassic
The alternation of extensional periods with intru- according to Caputo, 2014).
sion of granitoids with periods of compressive defor- It is important to remark that in the Marañon and
mation, orogenic uplift, and metamorphic phases is Ucayali Basins, as well as in the Huallaga Basin, the
related to the absolute motion of Gondwana. The re- initial sedimentation was driven by foreland basin de-
treat of the trench is linked to extension as observed velopment in the Early Permian (Erlich et al., 2018).
in the Sierras Pampeanas in the Carboniferous (Astini The provenance of the Late Permian synorogenic
et al., 2009; Martina et al., 2011) and compression to deposits clearly show a maximum depositional age
the continent overriding the oceanic plate, as de- younger than 265 Ma, in agreement with the main
scribed for the Phanerozoic along the South American Gondwanide orogeny at ∼260 Ma, derived from the
margin (Cawood, 2005; Ramos, 2008, 2010b). Based on uplift of the Eastern Cordillera (see figure 12 of E
­ rlich
these processes the Paleozoic evolution of the Gond- et al., 2018). An important unconformity ­s eparates
wanide orogeny in the Eastern Cordillera of Peru as these deposits from the Mitu rift sequences.
described by Chew et al. (2016) can be explained by
this alternation of tectonic regimes. However, the oc-
currence of two separated metamorphic belts, one The Pangaea Breakup
along the coast and other in the Eastern Cordillera,
needs some additional explanation. The Gondwa- The tectonic activity that followed the Gonwanide
nide deformation is intense along the coast in the orogeny in the Central Andes of Peru and Bolivia is
Amotape–Tahuin domain and in the Eastern Cordil- characterized by several extensional episodes. The
lera north of the Abancay deflection and almost disap- oldest and more widespread was the product of the
pears to the south. The metamorphic peak at 260 Ma collapse of the Gondwanide orogen that produced
is common to both domains, and there are no medi- the rift system of the Mitu Group (Mégard, 1978,
um-grade metamorphic rocks of late Paleozoic age 1984; Carlotto et al., 2008; Spikings et al., 2016). The
south of Abancay deflection in southern Peru, neither red-beds and alkaline basalts of the Mitu Group
in Bolivia nor in the Sierras Pampeanas of Argentina. are separated by an angular unconformity from the

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 12 11/05/18 12:18 PM


Tectonic Evolution of the Central Andes: From Terrane Accretion to Crustal Delamination  13

underlying late Paleozoic Copacabana limestones extensional activity continued up to the Sinemurian,
(Rosas et al., 2007). The second episode that con- but from that time on it is linked to the beginning of
trolled the rifting of the Pucará Group in the Late subduction, as volcanic arc rocks are preserved in the
Triassic presents variations in the different seg- northwest of the basin. The Late Triassic–Early Juras-
ments and is the consequence of the final breakup of sic marine deposits developed eastward of the Pa-
Pangaea. Several authors have emphasized that the racas terrane, which was a positive area at that time
location of the Late Permian–Triassic granites and (Figure 5). The Norian Sea entered by the Río La Leche
rhyolitic rocks, as well as the associated Triassic– area, north of Chiclayo to the central trough that is
Early Jurassic rifting, coincides with the axis of the now along the present Eastern Cordillera (Carlotto et
Eastern Cordillera of Peru and Bolivia (­ Dalmayrac al., 2009b), controlled by east-northeast-trending faults
et al., 1980; Sempere et al., 1999, 2002; Jaillard et al., parallel to the Tahuín suture.
2000). The Norian Sea of the Pucará Basin did not go fur-
Hf isotope data indicate that crustal reworking ther south than the Totos-Paras High, which was the
was a dominant process during arc magmatism, but northern boundary of the Arequipa Basin (Carlotto
it shows a decreasing in intensity through time. Ini- et al., 2009b). This high was part of a series of horsts
tial eHf values for the Ordovician granitoids during and grabens parallel to the Abancay deflection that
the Famatinian cycle are 26.73, increase to 22.43 in reactivated the old suture between the Paracas and
­Carboniferous–Permian times, and reach 21.57 in the ­Arequipa terranes (Figure 5).
Late Triassic (Miškovic’ and Schaltegger, 2009). eHf val-
ues of 20.7 to 18.0 for Cenozoic back-arc rocks indi- Southern Segment of Eastern Cordillera of Peru
cate that the extent of crustal reworking decreased with and Cordillera Real of Bolivia  The extension along
time and was associated with larger inputs of juvenile this segment also started with the sedimentation in
magma in an extensional regime during subduction a rift setting of the Mitu Group along the Eastern
(Miškovic’ and Schaltegger, 2009; Ramos, 2010a). Cordillera (Figure 5), far away from the continen-
The tectonic settings of the extensional regimes also tal margin (Mégard, 1978; Jaillard et al., 1990). This
change from north to south and are linked to the dif- ­r ifting is also controlled by the suture between the
ferent Paleozoic orogenic segments. This is also seen Arequipa terrane and the Amazonia craton (Ramos,
in the late Paleozoic–Triassic granitoids, which are 2009). New ages available from these rocks indicate a
I-type in the north, I-S-A transitional in the central time span between 245–240 and 220 Ma (Spikings et
part, to A-type in the south (Miškovic’ et al., 2009b). To al., 2016), much younger that the Permian–Early Tri-
describe their differences, several sectors were identi- assic age assigned in previous works (Mégard, 1978;
fied (Figure 5). Dalmayrac et al., 1980; Kontak et al., 1990).
Triassic granitoids and related volcanic rocks were
Nor thern Segment of Eastern Cordillera of emplaced in an extensional setting along this segment
Peru  This segment corresponds to the Eastern Cor- of the Eastern Cordillera of Peru between 220 and 200
dillera north of the Abancay deflection, where on Ma. The Triassic magmatism was dominated by alka-
top of the red sandstones and minor alkali basalts of line lavas of within-plate affinities, and peraluminous
the Mitu Group, continental and marine carbonates granites, which are characteristic of continental rift
of the Pucará Group were deposited (Rosas et al., settings (Spikings et al., 2016).
2007). Most of the rift system of the Mitu Group, as These Triassic granites are well exposed along the
well as the extensional faults that controlled the Pu- Cordillera Real of Bolivia (Figures 4, 5), where granitic
cará depocenters, are concentrated along the Eastern stocks such as the Huato (2252218 Ma), ­Illampu (218–
Cordillera and developed in the hanging-wall of the 202 Ma), Zongo (219–201 Ma), and Huayna Potosí
previous sutures as noticed by Ramos (2009, figure (2182196 Ma), Taquesi (2142211 Ma), among others
11) and Chew et al. (2016). (Jiménez et al., 2009), were emplaced in an extensional
The distribution of the marine Late Triassic– setting (Ávila Salinas, 1991; Ramos and Jiménez,
Early Jurassic rocks between Ayacucho and north 2014).
of Jaén (Figure 5) shows that they were deposited in South of the Abancay deflection, the Arequipa B­ asin
a fault-controlled depression, which deepens to the developed since Sinemurian times, as part of the vol-
east. According to Rosas et al. (2007) the deposits of canic arc–retro-arc basin system. According to ­Carlotto
the Pucará Group represent the sag phases of the Mitu et al. (2009b), the northern limit coincides with the
synrift, but Carlotto et al. (2009b) interpreted them ­Totos-Paras High, whereas the eastern boundary is
as synextensional facies related to rifting, based on represented by the Cusco-Puno High along the foot-
the great thickness variation of these deposits. The hills of the Eastern Cordillera.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 13 11/05/18 12:18 PM


14 Ramos

Figure 5. Triassic–Jurassic rift basins and Early Jurassic magmatic arc (based on Sempere et al., 2002, 2004; Rosas
et al., 2007; Jiménez et al., 2009; Carlotto et al., 2009b; Spikings et al., 2016). Note the location of the Totos-Paras
High along the Abancay deflection and the control of Paleozoic sutures on the distribution of the rift systems.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 14 11/05/18 12:18 PM


Tectonic Evolution of the Central Andes: From Terrane Accretion to Crustal Delamination  15

The Arequipa Basin starts in the Early Jurassic THE ANDEAN EVOLUTION
with sedimentation of the Yura Group with volcanic
and detrital sequences in the Chocolate Formation There are several reviews of the Andean tectonic evo-
­(Hettangian?–Sinemurian) at the base (Boekhout et al., lution of this sector of the Central Andes (Mathalone
2013), and carbonates (Toarcian–Bajocian), turbidites and Montoya, 1995; Sempere, 1994, 1995; Ramos and
(Bathonian–Oxfordian), clastic sequences (Oxfordian– Alemán, 2000; Carlotto et al., 2009a; Carlotto, 2013;
Kimmeridgian), and finally ending with carbonates Garzione et al., 2017, among others). Several articles
(early Tithonian). New ages of the Chocolate Forma- have described the regional sedimentation patterns of
tion and calc-alkaline plutons along the continental the Andean and sub-Andean basins including Bena-
margin indicate the onset of subduction between 196 vides Cáceres (1962, 1999), Jaillard et al. (1993) , Jacay
and 176 Ma (Boekhout et al., 2013). (1994), Jaillard (1994) , Martinez et al. (1994), Baby et
There is a clear shifting of the locus of the extension al. (1995), Carlotto (1998, 2013), Carlotto et al. (2005),
and associated magmatism. Rifting started inboard in among many others. Since the detailed reviews of
the Eastern Cordillera at 245 Ma with the Mitu Group, Jaillard et al. (2000, and references therein) and Car-
continued in this region with the intrusion of peralka- lotto (2013) the importance of the role of inherited
line granites between 220 and 200 Ma, and moved to pre-Andean structures in the Mesozoic and Cenozoic
the continental margin with arc-related rocks at about evolution of the Andean basins and their structures
195 Ma, when subduction controlled by an extensional has become evident. In this chapter the main regional
regime started (Ramos, 2010a; Spikings et al., 2016). paleogeography will be discussed to understand the
controls that the inherited basement structures exerted
The Eastern Cordillera of Bolivia  The Eastern in the different Andean processes, such as crustal
Cordillera of Bolivia, south of the Cordillera Real, delamination and lithospheric removal as described
has a series of basaltic dikes emplaced after 196 Ma by James and Sacks (1999) and Kay et al. (1999).
such as in Tarabuco area (Figure 4). Younger with-
in-plate basalts are spread along the Eastern Cor-
dillera up to the border with Argentina (Sempere The Arc-Retro-Arc Extensional Regime
et al., 2002; Jiménez et al., 2009). There is a pro-
nounced and continuous decrease in the amount of The threefold division of a subduction system in this
extension from northern Peru to southern Bolivia part of the Central Andes is difficult to recognize.
during Triassic and Early Jurassic times in compar- More than 150 km (93 mi) of the forearc has been
ison with previous segments. At these latitudes, removed by tectonic erosion during Cenozoic times
there are some basaltic fields in the sub-Andean (see von Huene et al., 1996; Clift, 2003) and the forearc
system, east of the Eastern Cordillera, such as the is poorly preserved. Subduction of the Nazca Ridge
Entre Ríos and Camiri basaltic fields, near the bor- led to uplift of the forearc (Macharé and Ortlieb, 1992).
der with Argentina (Sempere, 1995; Suárez Soruco, The magmatic arc in some areas is directly exposed
2000). Previous K-Ar dating yielded ages around along the present continental margin (Pitcher et al.,
235 Ma for these basalts (Sempere et al., 2002), 1985; Boekhout et al., 2015). The successive magmatic
but new Ar-Ar geochronological data yield pre- arcs and their retro-arc basins are well exposed north
cise ages of 204.03 { 1.38 Ma for a Camiri basaltic of Arequipa, where late Cenozoic arc volcanic rocks
dike and 180.55 { 1.52 Ma for the Entre Ríos basalts are not due to present flat-slab subduction (Soler and
(Kusiak et al., 2014). These values are similar to the Bonhomme, 1990).
Bertrand et al. (2005) Ar-Ar ages from the Tarabuco Subduction in the Central Andes developed under
Basalts between 201.2 { 10.6 Ma and 185.5 { 6.3 Ma an extensional regime from Early Jurassic to Early
and from 200.0 { 2.7 Ma to 189.5 { 1.4 Ma for the Cretaceous times (Auboin et al., 1973; Charrier, 1984;
Camiri basalts. These authors based on the study Mpodozis and Ramos, 1989; Ramos, 1999), and the
of Marzoli et al. (2006), include these basalts in the magmatic arc products were juvenile poorly evolved
Central Atlantic Magmatic Province (CAMP) based in intra-arc basins up to 9 km (6 mi) thick along the
on their geochemistry and age. The CAMP basalts continental margin (Petford and Atherton, 1995;
exposed in the northeastern margin of South Amer- Romero et al., 2013). This regime was controlled by
ica were correlated with typical 200 Ma basalts of the absolute motion of the upper plate, when South
the east coast of North America and western mar- ­America—as a single block with Africa—moved to the
gin of Africa (Bertrand et al., 2005; Kusiak et al., east, and the trench rollback produced extension in
2014). This large igneous province is related to the upper plate (Daly, 1989; Ramos, 1999, 2000; Ramos
breakup of Pangea (Marzoli et al., 1999). and Alemán, 2000).

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 15 11/05/18 12:18 PM


16 Ramos

Retro-arc basins were developed in an extensional these authors, this collapse occurred at about 90289
regime during most of the Jurassic and Lower Creta- Ma in Turonian–Coniacian times. Its origin has proved
ceous in Peru and Bolivia (Sempere et al., 2002), and controversial through the years, varying from subaer-
subsidence mechanisms alternated between mechan- ial to submarine landslides controlled by contractional
ical extension and thermal sag. The locus of the ex- or extensional tectonics, or exclusively gravitational
tension was concentrated along the weakness zone mechanism. However, detailed studies of Callot (2008)
between Amazonia and the Arequipa Massif (Sempere and Callot et al. (2008a, b) allowed reconstructing the
et al., 2002), as well as between the Paracas block and distribution of their facies and their precise deposi-
Amazonia (Ramos, 2008). Normal faults were devel- tional environments (Figure 7). The area covered by
oped mainly along the upper plate of the sutures not the collapse is larger than 80,000 km2 with an estimated
only with the Amazonia craton but also on the suture volume of more than 10,000 km 3 (2399 mi 3; Callot
between the Paracas and Arequipa terranes along the et al., 2008a).
set of faults that define the Abancay deflection (Maro- Recent global analysis of the different kinds of
cco, 1978; Carlotto, 1998; Carlotto et al., 2009b). olistostromes produced during the evolution of the
orogenic belts shows different contractional settings
where orogenic uplift produced large collapses (Festa
The Arc and Foreland Basin System et al., 2016). The local control of some of these large
olistostromes is triggered by normal faults. The Ayab-
Important paleogeographic, structural, and magmatic acas c­ ollapse is one of the largest ancient submarine
changes occurred associated with the Mochica and landslides on a global scale (see figure 8 of Festa et al.,
Peruvian orogenies (Steinmann, 1929; Mégard, 1984) 2016).
at the end of the Early Cretaceous. It is well estab- The studies of Callot et al. (2008a, b) indicate
lished that the initiation of the westward shift of the ­important and rapid changes in thickness of the Ayab-
South American plate in the late Albian coincided acas Formation related to synsedimentary deforma-
with the beginning of contractional deformation in the tion associated with normal faults. The initiation of
Peruvian Andes (Jaillard and Soler, 1996). This change deformation along the Coastal Cordillera of northern
in the tectonic regime is recorded all along the conti- and southern Peru was related to the Mochica and
nental margin of South America and is related to the Peruvian phases between 100 and 90 Ma (Benavides
final opening of the central sector of the South Atlan- Cáceres, 1999). It is important to remark that large sec-
tic and the beginning of the absolute westward drift of tors of the forearc have been removed by subduction
the South American continent (Ramos, 2010a). erosion, and only part of the original magmatic arc is
The widespread Late Cretaceous regression in the presently preserved (Figure 7b). The beginning of con-
western basins was interpreted by Jaillard et al. (1993) tractional deformation produced important flexural
as the result of the Peruvian orogenic episode. The onset loading and the early subsidence of the adjacent fore-
of this deformation is recorded by fine-grained clastic land basin. The buckling associated with initial con-
sedimentation in the eastern Peruvian basins in the mid traction was linked to normal faults, responsible of the
Turonian–earliest Coniacian (Jaillard et al., 2005). How- Ayabacas collapse. The ­Puno-Cusco High, ­although
ever, one of the main synorogenic episodes produced as covered by the sea during the Late Cretaceous, acted
a consequence of the Peruvian phase is the development as a positive swell during the sedimentation of the
of the Ayabacas giant submarine collapse (Figure 6). Ayabacas Formation (Carlotto, 2002).
This giant mélange is known since the early work of The distribution of the Ayabacas collapse is outlined
Heim (1948a), who described these rocks as a chaotic by the Totos-Paras and Puno-Cusco Highs (Figure 6),
broken formation. It is worth to mention here that simi- which may indicate that the old sutures of Paracas
lar deposits in the Precordillera of San Juan, Argentina, and Arequipa terranes behaved as unstable weakness
were also described by Heim (1948b). Studies of these zones. The structures could be reactivated either by
mélanges interpreted their genesis as submarine chaotic incipient contractional structures by inversion of pre-
synorogenic deposits of a foreland basin as result of the vious normal faults, or the uplift could be controlled
contractional deformation during the Famatinian orog- by a peripheral bulge linked to thrust loading in the
eny (Ramos et al., 1996; Festa et al., 2016). Western Cordillera further west. The paleo-tectonic re-
The Ayabacas mélange is an extraordinary unit constructions of Carlotto (2013) identified the periph-
formed by the giant submarine collapse of the eral bulge in the Puno-Cusco High at the beginning
mid-Cretaceous carbonate platform of the western of the Paleocene as a consequence of thrust loading in
Peru back-arc basin (Callot et al., 2008a). ­According to the Western Cordillera. The continental deposits of the

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 16 11/05/18 12:18 PM


Tectonic Evolution of the Central Andes: From Terrane Accretion to Crustal Delamination  17

Figure 6. The beginning of the Andean deformation in the Late Cretaceous in the Central Andes. The Mochica deforma-
tion in northern Peru seems to be earlier than the Peruvian orogeny in southern Peru and northern Chile according to
Benavides Cáceres (1999). The outline of the giant submarine collapse of Ayabacas is based on Callot (2008), but it is
interpreted following Carlotto et al. (2002, 2009a) as result of the Late Cretaceous orogeny (see discussion in the text).
Note the inherited control of the previous sutures.

Vilquechico Group were associated with the Peruvian The basin foreland evolution indicates that the dep-
phase by Jaillard et al. (1993). ocenters in the Paleocene were located in the central
The main orogenic deformation in the Central western Altiplano and shifted to the east during the
Andes of Peru and Bolivia corresponds to the Incaic Eocene. The main orogenic deformation phase took
orogeny (Steinmann, 1929, 1930; Mégard, 1984; Jail- place between ∼43 and ∼30 Ma ­(Carlotto, 2013), and
lard et al., 2000). As a result of that, the Eastern Cor- exceptional growth strata developed in the Ayaviri re-
dillera was uplifted and different foreland basins gion (Figure 8) at 29226 Ma in the central Altiplano of
developed between 52 and 30 Ma (Carlotto, 2013). southern Peru (Perez and ­Horton, 2014).

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 17 11/05/18 12:18 PM


18 Ramos

Figure 7. The Ayabacas giant submarine collapse. A. Ayabacas Formation in southern Peru Altiplano (courtesy of V ­ ictor
­Carlotto, whose permissions is required for further use). B. Scheme showing the collapse during the Peruvian deformation
­controlled by the highs related to the sutures. C. Broken mélange near Santa Rosa (after Callot et al., 2008a). D. Scheme
of the western back-arc basin of southern Peru, and the potential control of the Peruvian phase on the collapse (based on
Carlotto, 2013 and Alonso et al., 2015). Note that main structures within the basin are extensional faults.

Flat Slab Subduction Leading to Crustal Delamination parallel to the continental margin (Carlotto et al.,
2009a,b). These calc-alkaline rocks were uplifted by
It is well established that a series of tonalities and tectonic inversion during the Late Cretaceous, with
granodiorites were emplaced during the Late Cre- important contraction produced by faults like the
taceous (82286 Ma) along the Arequipa segment Cincha-Lluta thrust described by Vicente (1989),

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 18 11/05/18 12:18 PM


Tectonic Evolution of the Central Andes: From Terrane Accretion to Crustal Delamination  19

Figure 8. Growth strata of the Puno Group of Late Oligocene age in the foreland deposits of central Altiplano near
­Ayaviri, southern Peru (courtesy of Nicholas Pérez, whose permission is required for further use).

Figure 9. Maximum extension of the main Paleogene arc with the subduction related magmatic rocks and a ­ ssociated
­porphyries of Paleocene and middle Eocene-early Oligocene age (after Perelló et al., 2003). The Z­ ongo-San G
­ abán
­“inner arc” rocks are exposed in the Cordillera de Carabaya and Cordillera Real according to Sandeman et al. (1995);
terrane boundaries after Ramos (2009). Numbers within brackets indicate mineralization age.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 19 11/05/18 12:18 PM


20 Ramos

where the basement overthrusts the sediments of the shifted westward toward the trench, as defined by the
Arequipa Basin. location of successive arcs: the Tacaza (30–24 Ma), the
The magmatic arc located in the western slope of the Huaylillas (24210 Ma), the Lower Barroso (1023 Ma),
Western Cordillera shifted to the east in the latest Cre- the Upper Barroso (321 Ma), and the active volcanic
taceous–Paleocene, emplacing a belt of important calc-­ arc ( 61 Ma), located at ∼230 km 1 ∼143 mi2 northeast
alkaline igneous rocks and associated porphyries along of the trench and 115 { 5 km (71 { 3 mi) above the
the western margin of the Arequipa basin in southern Benioff zone (Mamani et al., 2010). The volcanic rocks
Peru (Figure 9) known as the Toquepala arc (85245 Ma; erupted in the Tacaza arc consist of primary mantle
Mamani et al., 2010). These calc-alkaline rocks are rep- melts represented by shoshonitic and high-K mafic
resented by porphyries such as Cuajone, Quellaveco, lavas, and subordinate andesitic calc-alkaline mag-
Toquepala, Cerro Colorado, and Spence (57252 Ma), mas, followed by voluminous ignimbrites between
a belt of important mineralization of Paleocene age in 29 and 26 Ma as result of crustal melting (Mamani et
southern Peru and northern Chile (Clark et al., 1990; al., 2010). Large ignimbritic plateaux were developed
­Carlotto et al., 2009a). during the Huaylillas arc further to the west, ending
An important change in the subduction angle of the with the Barroso arcs along the Western Cordillera,
Toquepala arc occurred at about 53 Ma, followed by close to the present volcanic arc. The time variation of
an important eastward shift recorded between 50 and the flattening of the subduction zone is indicated in
42 Ma. A phase of normal subduction, with the arc lo- Figure 10.
cated about 300 km (186 ft) from the trench, gave way to It is important to note that after the shifting and
the onset of flat-slab subduction, as proposed by Sande- the expansion of the magmatic arc, Clark et al. (1990)
man et al. (1995), James and Sacks (1999), and Perelló et identified “inner arc” magmatism on geochemical
al. (2003). The main arc was characterized by the Anta grounds related to crustal melt with localized ex-
volcanic rocks and the Andahuaylas-Yauri calc-alkaline tension. The rocks of the “inner arc” are not sub-
batholith, emplaced further to the east between 48 and duction related, and they are similar to previously
34 Ma (Carlier et al., 1996) and bounded by the Abancay described Triassic weakly peraluminous granitic
deflection in the northeastern corner of the Arequipa rocks of the Cordillera de Carabaya (see Figure 4).
terrane (Figure 9). An important phase of mineraliza- Peraluminous monzogranitic stocks were generated
tion is related to dacites and andesites in the last phases at 29228 Ma by high-temperature crustal melt-
of batholith assembly and defines a belt of porphyry ing after the broadening of the main Paleogene arc
copper and skarn mineralization of middle–late Eocene along the Zongo-San Gabán area in southern Peru
age (42232 Ma) according to Perelló et al. (2003). and northern Bolivia (Clark et al., 1990; Sandeman
The Paleogene shift in the position of the main arc et al., 1995). These Oligocene granites are emplaced
was associated with the most important orogenic phase along the Zongo-San Gabán belt in the Cordillera de
of this part of the Andes of Peru and Bolivia: the Incaic Carabaya and in the Cordillera Real (Figures 10, 11),
orogeny (Steinmann, 1929; Megárd et al., 1984). Intense similar to the previously described peraluminous
deformation took place in the Western Cordillera and granites associated with the Arequipa–Amazonia
in the Altiplano, shifted to the Eastern Cordillera, and boundary. This boundary was described as a first or-
was related to the shallowing of the subducted slab be- der weakness zone, an inherited basement feature,
tween 43 and 30 Ma according to Carlotto, 2013, and which controlled the location of crustal extension
references therein. This deformation was linked to (Ávila Salinas, 1991; Sempere et al., 2002; Carlier et
northeast-verging thrusts that produced important up- al., 2005; Jiménez et al., 2009). The structural set-
lift, crustal shortening, and exhumation leading to the ting of the Huarinas fold belt, a Famatinian series of
formation of thick synorogenic deposits in the foreland backthrusts located west of the Cordillera Real fault
basins. The reactivation of the Huarinas backthrust zone, and the reactivation of this belt during Andean
belt also contributed to crustal shortening west of the times, led Ramos and Jiménez (2014) to propose that
Cordillera Real in Bolivia (Jiménez et al., 2009) and the suture between these cratonic blocks was the lo-
west of the Cordillera de Carabaya (Perez and Horton, cus of lower crustal extension. The “inner arc” mag-
2014) during the Oligocene. matism continued with the mantle-derived Picotani
The geochemical, isotopic, and geophysical studies (26222 Ma) and the peraluminous silicic Quenamari
of Mamani (2006) and Mamani et al. (2008, 2010) con- (1726.5 Ma) volcanic rocks in the Eastern Cordillera
tributed to understand the petrological processes as- of southern Peru. The voluminous 530 km3 (127 mi3)
sociated with the trench retreat of the main magmatic Quenamari rhyolites, although coeval with the main
arc related to the steepening of the Benioff zone after arc, were crustal melts not related to subduction
30 Ma. These authors recognized the volcanic front (Sandeman et al., 1997).

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 20 11/05/18 12:18 PM


Tectonic Evolution of the Central Andes: From Terrane Accretion to Crustal Delamination  21

Figure 10. Variation through time of the flattening of the subduction zone in southern Peru and Bolivia from the Oligo-
cene to the Pliocene (based on Jiménez et al., 2009; Kay and Coira, 2009; Mamani et al., 2010; Garzione et al., 2017,
and cites therein). There is also indicated the location of the “inner arc,” which is developed coevally with the steepen-
ing of the subduction and it is getting younger to the south. The position of the lower crustal melts associated with the
“inner arc” approximately coincides with the Arequipa–Amazonia suture weakness zone.

Further south, Jiménez et al. (2009) described Crustal and Lithosphere Delamination
in Bolivia (Figure 10) voluminous ignimbritic
eruptions such as Morococala (826 Ma), Lichiv- The need for mantle removal and the loss of lower crust
ico ( 827 Ma ), Condomasa ( 7 Ma ), Los Frailes were quantified by the pioneering work of Roeder and
(925 Ma to 322 Ma), among others. Note that Cerro Chamberlain (1995, and references therein) based on
Rico de Potosí (13 Ma) lies within the “inner arc” restoration of regional structural cross sections in the
zone. These eruptions are along the “inner arc” Central Andes (Figure 12). A similar conclusion was
according to Kay and Coira (2009), and their ages obtained by Kay and Kay (1993) and Kay et al. (1994)
indicate that this ignimbritic phase youngs to the who identified on petrological grounds evidence for
south. lithosphere delamination in the magmatic products.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 21 11/05/18 12:18 PM


22 Ramos

Figure 11. Main Paleogene structures in northern Bolivia based on Martinez et al. (1994) and the extensional intra-
plate granites of the “inner arc” after Jiménez et al. (2009). Note that intrusive age is decreasing to the south: Illimani
(28227 Ma); Quimsa Cruz (26224 Ma); Santa Vera Cruz (23 Ma).

The recent synthesis of Garzione et al. (2017) em- and the formation of a high-density eclogitic crustal
phasizes two different end-member processes for the root. The delaminated block would consist of litho-
removal of mantle lithosphere during Andean defor- spheric mantle and eclogitic lower continental crust
mation. The removal of dense lower lithosphere be- as inferred in the Central Andes by Kay et al. (1994).
neath the Central Andean Altiplano can be explained ­Evidence for this process was imaged through differ-
by ablative subduction of mantle lithosphere on the ent seismic experiments (Beck and Zandt, 2002; Schurr
foreland side of the system; this is a continuous pro- et al., 2006; Calixto et al., 2013, among others). This
cess in which the lower lithosphere is removed during fast block delamination produces a rapid surface up-
flattening of the subducted slab, associated with lift of more than one kilometer over a short period of
crustal thickening and gradual plateau uplift over tens time (Garzione et al., 2006, 2007).
of millions of years. This process is equivalent to the The application of these concepts to the Cenozoic
mechanical removal of pieces of lithosphere proposed tectonic evolution of the Central Andes of Peru and
by Kay et al. (1999), during shallowing of the seismic Bolivia clearly shows that shortening during the
zone and thickening of the crust, to explain the mass Late Cretaceous and Paleogene was accommodated
balance between crustal shortening and the excess vol- by ablative subduction. The crustal thickening and
ume of mantle lithosphere. uplift the Western Cordillera and western Altiplano
The second end-member process described by was followed by the uplift of the Eastern Cordillera
Garzione et al. (2017) is the rapid removal of a thick after the shallowing of the oceanic slab, as envisaged
portion of the lower lithosphere as proposed by Bird by Sandeman et al. (1995). A remarkable feature is
(1979) and Sobolev and Babeyko (2005), among oth- the giant Ayabacas submarine collapse as result of
ers. The rapid removal of lower crust and mantle the Peruvian orogeny controlled by the reactivation
lithosphere occurs after significant crustal thickening of previous sutures. The shortening of the Western

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 22 11/05/18 12:18 PM


Tectonic Evolution of the Central Andes: From Terrane Accretion to Crustal Delamination  23

Figure 12. Balanced structural crustal sections of the Eastern Cordillera and the sub-Andean belt of northern Boliv-
ian Andes. A. Present state showing the double-wedge style of Huarinas and sub-Andean belts and the required
lithospheric delamination for accounting the orogenic shortening. B. Restored section prior to the beginning of Neo-
gene deformation (after Roeder and Chamberlain, 1995). Note the partial melting of the lower crust associated with
lithospheric delamination. The initial stage is similar to the hypothesis of Isacks (1988), which needs a thin mantle
lithosphere for the thermal uplift of the Altiplano.

Cordillera, the Altiplano, and the Eastern Cordillera suture between Amazonia and Arequipa within the
amounts to about 185 km (115 mi) according to re- Cordillera Real fault zone. It is interpreted as the re-
cent estimates of Anderson et al. (2017), and it was sult of the interaction of hot asthenosphere with hy-
mainly developed during this time interval of abla- drated lithospheric mantle. The voluminous silicic
tive subduction. volcanism would be the expression of delamination
The steepening of the subducted slab after 30 Ma along the Zongo–San Gabán area. In a similar setting
produced the retreat of the main arc toward the across the Altiplano–Eastern Cordillera, the crustal
trench and almost synchronically the “inner arc” in roots have been delaminated according to Beck and
the Zongo–San Gabán area. Several authors empha- Zandt (2002), and the ignimbrite field of Los Frailes
size that the “inner arc” magmatism was the result of would be the evidence of such delamination. The
lower crustal melting. It was explained by James and “inner arc” magmatism is younger to the south and
Sacks (1999) as hydration of the metasomatized litho- several domes and ignimbrite fields are described
spheric mantle in contact with hot asthenospheric between 18 and 20°S by Kay and Coira (2009). These
flow during shallowing of the slab at about 40 Ma in authors include in the “inner arc” the Los Frailes, Mo-
the Incaic phase. The new available ages constrain the rococala, and Vila Vila volcanic rocks varying in age
“inner arc” magmatism as coeval or postdating the between 25 and 5 Ma, being the most voluminous
last stages of Incaic (including Aymará) deformation. Los Frailes field with 7,500 km2 (2896 mi2) erupted at
The most voluminous volcanism is associated with about 925 Ma.
the peraluminous Quenamari ignimbrites of middle– The partial melting of the lower crust associ-
late Miocene age, erupted in the hanging-wall of the ated with localized crustal melting along the suture

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 23 11/05/18 12:18 PM


24 Ramos

Figure 13. Structural evolution of the sub- fold-and-thrust belt in southern Bolivia (after Baby et al., 1994, and Moretti
et al., 1996). Note the rapid deformation after 6 Ma in coincidence with the main volcanic episodes in the “inner arc”
at these latitudes.

between Amazonia and Arequipa triggered defor- the Andean structural styles, ore mineralization, and
mation in the sub-Andean belt and important short- in many other geological processes. A brief appraisal
ening occurred after middle–late Miocene times. This of the different orogens will be made following the
second wave of deformation is related to the Quechua main inherited properties in each stage.
orogeny with its many different phases (see Fig. 13),
responsible of the development of the sub-Andean
fold-and-thrust belt that accounts for about 152 km Grenvillian Orogen
(94 mi) of shortening (Anderson et al., 2017). The early
work of Baby et al. (1994) and Moretti et al. (1996) and There are at least two orogens developed along the
subsequent fission track data indicate that major short- continental margin, the autochthonous Sunsas oro-
ening occurred mainly in late Miocene–present times gen along the Amazonian margin being the best
(see Garzione et al., 2017, and references therein). This known (see Cordani et al., 2000; Cordani and Teixeira,
shortening may have been enhanced by the change 2007). The eastern Sunsas belt is patchily exposed in
from arid/semi-arid climate to hyperaridity, occurred Querobamba and Río Picharí at the Eastern Cordillera
between 4 and 3 Ma as proposed by Hartley (2003). of Peru, and in the Serranías Chiquitanas of Bolivia,
east of the sub-Andean belt. A second Grenvillian
orogenic belt was developed in the continental frag-
CONCLUDING REMARKS ments amalgamated to Amazonia during the forma-
tion of Rodinia (Ramos, 2010b). Basement remnants of
The analysis of the different tectonic events that have the magmatic arc of this orogen have been identified
affected the continental margin of the Central Andes of in the northern offshore of Peru, as part of the outer
Peru and Bolivia shows a protracted evolution, where shelf high of the Paracas terrane (Romero et al., 2013),
their tectonic framework is strongly influenced by and all along the western margin of Arequipa–Anto-
previous inherited structures. The sutures developed falla terrane of southern Peru, Bolivia, and northern
during the Grenvillian orogeny have a primary control Chile (Tosdal, 1996; Loewy et al., 2004). Large parts of
in the formation of younger oceans, the inception of this Grenvillian basement have been eroded away by
sedimentary basins, the successive hanging-wall rifts, subduction e­ rosion. The sutures that bound these two

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 24 11/05/18 12:18 PM


Tectonic Evolution of the Central Andes: From Terrane Accretion to Crustal Delamination  25

terranes from the Amazonian craton were recognized granitic orthogneisses represent Carboniferous and
on geophysical and geological grounds (Carlier et al., Permian batholiths inboard of the continental margin
2005; Mamani et al., 2008; Ramos, 2009, and references and can be explained like in the Sierras Pampeanas,
therein). as a response to the collapse of the previous orogenic
deformation with extension of the crust and mantle
upwelling (Grosse et al., 2009). The only Late Paleozoic
Famatinian Orogen tonalitic orthogneisses that may represent a magmatic
arc are in the Tahuín terrane along the continental
The orogenic deformation that took place during the margin of northwestern Peru. This magmatic belt has
Ordovician has been identified all along the Gond- been interpreted by Ramos (2009) following Feininger
wanan proto-margin of the Andes. The metamor- (1987) as an allochthonous terrane, whereas Bellido et
phic and magmatic rocks are well exposed from the al. (2009) and Timoteo et al. (2012) considered these
Mérida Andes in Venezuela to central Chile–Argen- metamorphic rocks as part of Gondwana. Both alter-
tina Andes and are well dated in this central segment natives fail to explain the occurrence in the Cordillera
(Chew et al., 2007, 2016; Cardona et al., 2008, 2009). de Marañón, inboard of the continental margin, of
The important changes along-strike in metamorphic another belt of metamorphic rocks with similar ages
grade and the crustal structural level of the Famatin- more than 350 km (217 mi) from the present continen-
ian orogen, parallel to the proto-margin, are directly tal margin. If a conservative assessment of the crustal
related to the collision of allochthonous and paraau- erosion by subduction is considered, the Late Paleo-
tochthonous terranes. The occurrence of paired meta- zoic rocks were formed near 500 km (311 mi) inboard
morphic belts, ophiolitic assemblages, and magmatic of the margin.
arcs shows that these terranes have been detached The regional distribution of these late Paleozoic
from Amazonia, probably reactivating old Grenvil- metamorphic belts shows that they disappear to the
lian sutures, and later collided again in Ordovician south of central Peru, where either in southern Peru
times. Some of the continental blocks such as Areq- or in Bolivia the late Paleozoic rocks are preserved in
uipa–Antofalla were only partly detached, but oth- sedimentary facies. North of Peru metamorphic rocks
ers like the Paracas terrane developed oceanic crust of these ages are exposed in the Cordillera Real of Ec-
between this terrane and Amazonia. In the first case uador, the Central Cordillera of Colombia, and in the
where the terrane was partly detached, the Ordo- Merida Andes of Venezuela. This pattern may suggest
vician rocks in the Amazonian side are preserved that these metamorphic belts are more closely related
in sedimentary facies, whereas in the second case to the Alleghenide orogeny than to the Gondwanides,
when oceanic rocks are consumed and the terrane formed during the Late Carboniferous–Early Permian
collided against the craton, middle to high metamor- collision of Laurentia and the northern part of South
phic grade is dominant. Among different segments America.
a transitional domain is identified as in the Cordil- The Late Paleozoic granitic rocks, now preserved
lera of Vilcabamba, where these rocks are preserved as orthogneisses, were emplaced in an extensional
in greenschist facies. Based on these data Famatinian regime not related to subduction (Clark et al., 1990),
deformation should be more pervasive in the north- and the uplift of middle crustal levels at the surface
ern sub-Andean basins where the collision of Para- could be related to the Laurentia–Western Gondwana
cas terranes is recorded than in the south where there collision in the Late Paleozoic. The tonalitic metamor-
was no collision at that time. phic rocks of the Tahuin terrane (Romero et al., 2013)
may represent the magmatic arc of this collision, ei-
ther as a detached part of Laurentia or as autochtho-
The Gondwanide Orogen nous part of Gondwana. The east vergence of these
rocks and the tectonic contact with sedimentary rocks
The Gondwanides are well defined in the southern of the same age in the Amotape area favor the first
part of South America along the Patagonian margin interpretation.
after the classical proposal of Keidel (1921) followed The Late Paleozoic rocks of the Cordillera de Cara-
by Du Toit (1927) and many others. The term was pro- baya and its continuation in the Cordillera Real of Bolivia
posed for the Central Andes of Peru by Chew et al. were defined as the “inner arc” of an eroded away mag-
(2016), identifying two different episodes at ~305 Ma matic arc belt along the margin (Boekhout et al., 2015).
and ~260 Ma, as early and late Gondwanide, respec- These acidic rocks could represent an early delamination
tively. It is important to note that none of these epi- of the lower crust during an extensional period in Late
sodes is related to consumption of oceanic crust. The Permian–Early Triassic times, as recognized in many

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 25 11/05/18 12:18 PM


26 Ramos

areas along the continental margin of South America mantle, coincides with the suture between Arequipa
(Ramos, 2009; Sato et al, 2015; Spikings et al., 2016). This and Amazonia and was later exploited by potassic-­
delamination again took place along the hanging-wall of ultrapotassic mafic rocks, which delineated the two
the suture between Arequipa and Amazonia. lithospheric mantle blocks (Carlier et al., 2005).
The shallowing of the oceanic slab migrated from
southern Peru to the south into Bolivia and northern
The Pangea Breakup Argentina (James and Sacks, 1999; Kay et al., 1999; Kay
and Coira, 2009). This wave of flattening of the oceanic
The development of the Pucará Basin during Late slab produced not only a foreland expansion and mi-
Triassic–Early Jurassic times and the associated earlier gration of the magmatic arc but also a continued shift-
Mitu Triassic hanging-wall rift was clearly controlled ing of the “inner arc” magmatism as depicted by Kay
by Famatinian sutures (Ramos, 2009), which are in and Coira (2009).
turn an inherited feature from the Grenville amalga- Several geophysical studies have shown that lower
mation of Rodinia. The boundary of this basin and crustal delamination and mantle removal that con-
their deposits matches the suture between the Pucará trolled this “inner arc” magmatism along the suture
terrane and the Amazonian craton, as recognized by of Arequipa and Amazonia. Important “inner arc”
several authors (Sempere et al., 2002; Carlotto et al., magmatic rocks as in Potosí during the Miocene and
2009a, b, among others). Los Frailes in the Pliocene are related to world-class
The development of the Arequipa basin was con- ore deposits (Petersen, 1999), emplaced close to the
trolled by the Abancay fault zone, coinciding with the suture. Similarly the late Paleozoic Cerro de Pasco ore
northern suture between Paracas and Arequipa, as deposits are related to analogous geological process.
well as with the suture with the Amazonian craton to The occurrence of delamination in the lower crust
the east. Two highs (the Totos-Paras and Puno Cusco triggered lower crustal ductile shortening as orig-
Highs) developed along these sutures (Carlotto et al., inally proposed by Isacks (1988) and simple shear
2009a, b) and controlled extension within the basin. shortening in the upper crust in the latest stages the
These highs were active until the beginning of the Late sub-Andean fold-and-thrust belt. Many authors have
Cretaceous, when they controlled the Ayabacas giant described the migration of the thin-skinned deforma-
submarine collapse. tion in the sub-Andean belt from north to south (Perez
There are no doubts that during the extensional and Horton, 2014, as well as in different chapters of
subduction regime, these sutures were an important this volume).
structural control in the extent of the basins and the lo- As a final remark, it should be stated that the com-
cation of the magmatic rocks, as described by Sempere bined location of the “inner arc” magmatism, lower
et al. (2002) and Jiménez et al. (2009). crustal delamination along the sutures, and removal
of the lithospheric mantle led to pronounced deforma-
tion in the sub-Andean fold-and-thrust belt. However,
The Andean Evolution processes well exhibited in the Andean deformation
phases were also active during the Paleozoic and
The early work of Marocco (1978) and Carlotto (2002) controlled the early deformation in the sub-Andean
emphasized the control of the Abancay deflection in ­basins of the Central Andes.
the emplacement of the different igneous rocks, in the
tectonic inversion of the Andean structures, and in the
sedimentary synorogenic basins developed in southern ACKNOWLEDGMENTS
Peru (Carlotto, 2013). Several other workers studied the
control of the eastern contact of Arequipa terrane with The author wants to acknowledge several colleagues as
the Amazonian craton (Carlier et al., 2005; Mamani et Victor Carlotto, José Macharé, Darwin Romero in Peru,
al., 2008, Carlotto, 2013) in the magmatic rocks, the syn- and Néstor Jiménez, Ramiro Suárez Sorucco, Ramiro
orogenic deposits, and their structural styles. Matos in Bolivia, among many others, for many years
The changes in the subduction zone during Paleo- of interchange of ideas and discussions that signifi-
gene times, mainly in the extensional periods of retreat cantly contributed to the present manuscript. The man-
of the magmatic arc during steepening of the oceanic uscript was improved by the careful reviews of David
slab, again reactivated the old suture that controlled M. Chew of Trinity College Dublin and José Macharé
the “inner arc” magmatism in almost the same areas del INGEMET (Instituto Geológico, Minero y Metalúr-
as in the late Paleozoic. The delamination of the lower gico del Perú). This is the contribution R@237 of the
crust, associated with removal of the lithospheric ­Instituto de Estudios Andinos Don Pablo Groeber.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 26 11/05/18 12:18 PM


Tectonic Evolution of the Central Andes: From Terrane Accretion to Crustal Delamination  27

REFERENCES ­ eridionales: Revue de Géographie Physique et de


M
­Geologie Dynamique, v. 15, no. 1–2, p. 11–71.
Aceñolaza, F. G., and A. Toselli, 1976, Consideraciones Aumaitre, R., G. Grandin, and J. Guillon, 1977, Données lith-
­estratigráficas y tectónicas sobre el Paleozoico inferior ologiques et structurales relatives à un bloc précambrien
del Noroeste Argentino: 2° Congreso Latinoamericano de surélevé de la Cordillère andine orientale (Pérou central).
Geología (1973), Actas 2, p. 755–763, Caracas. Les corps de roches ultrabasiques qui y sont présents:
Alemán, A., and V. A. Ramos, 2000, The Northern An- Bulletin Société géologique de France, v. 19, no. 5,
des, in U. J. Cordani, E. J. Milani, A. Thomaz Filho, and p. 983–989.
D. A. Campos, eds., Tectonic evolution of South America: Ávila Salinas, W., 1991, Petrologic and Cenozoic evolution of
Río de Janeiro, 31st International Geological Congress, the Cenozoic volcanism in the Bolivian Western Andes,
p. 453–480. in R. S. Harmon, and C. W. Rapela, eds., Andean mag-
Alemán, A. M., D. Valasek, C. Ardiles, G. D. Wood, G. P. matism and its tectonic setting: Geological Society of
Wahlman, and J. R. Groves, 2003, Petroleum systems and ­America, Special Paper 265, p. 245–257.
tectono-stratigraphic evolution of the Madre de Dios Ba- Baby, P., B. Guillier, J. Oller, E. Mendez, G. Montemurro,
sin and its associated thrust-belt in Perú and Bolivia, in and D. Zubieta, 1994, Sintesis estructural del subandino
8th Simposio Bolivariano Exploracion Petrolera en las boliviano, in Memorias del 11° Congreso Geológico de
Cuencas Subandinas, p. 177–200. ­Bolivia, p. 161–169.
Alonso, J. L., A. Marcos, E. Villa, A. Súarez, O. A. Merino- Baby, P., I. Moretti, B. Guillier, R. Limachi, E. Méndez, J. Oller,
Tomé, and L. P. Fernández, 2015, Mélanges and other and M. Specht, 1995, Petroleum system of the northern
types of block-in-matrix formations in the Cantabrian and central Bolivian Subandean zone, in A. J. Tankard,
Zone (Variscan Orogen, northwest Spain): Origin and R. Suárez Soruco, and H. J. Welsink, eds., Petroleum ba-
significance: International Geology Review, v. 57, no. 5–8, sins of South America: AAPG Memoir 62, p. 445–458.
DOI: 10.1080/00206814.2014.950608. Baby, P. et al., 2018, The Peruvian sub-Andean foreland basin
Anderson, R. B., S. P. Long, B. K. Horton, A. Z. Calle, and V. system: Structural overview, geochronologic constraints,
Ramirez, 2017, Shortening and structural architecture of and unexplored plays, in G. Zamora, K. R. McClay,
the Andean fold-thrust belt of southern Bolivia (21°S): Im- and V. A. Ramos, eds., Petroleum basins and hydrocar-
plications for kinematic development and crustal thicken- bon p ­ otential of the Andes of Peru and Bolivia: AAPG
ing of the Central Andes: Geosphere, v. 13, no. 2, p. 1–21. ­Memoir 117, p. 91–120.
Aspden, J. A., and M. Litherland, 1992, The geology and Bahlburg, H., and F. Hervé, 1997, Geodynamic evolution
Mesozoic collisional history of the Cordillera Real, Ecua- and tectonostratigraphic terranes of northwestern Argen-
dor: Tectonophysics, v. 205, p. 187–204. tina and northern Chile: Geological Society of America
Aspden, J. A., N. Fortey, M. Litherland, F. Viteri, and S. M. Bulletin, v. 109, p. 869–884.
Harrison, 1992, Regional S-type granites in the Ecuado- Bahlburg, H., V. Carlotto, and J. Cárdenas, 2006, Evidence
rian Andes: Possible remnants of the breakup of western of Early to Middle Ordovician arc volcanism in the Cor-
Gondwana: Journal of South American Earth Sciences, dillera Oriental and Altiplano of southern Peru, Ollan-
v. 6, p. 123–132. taytambo formation and Umachiri beds: Journal South
Aspden, J. A., W. Bonilla, and P. Duque, 1995, The El Oro American Earth Sciences, v. 22, p. 52–65.
metamorphic complex, Ecuador: Geology and eco- Bahlburg, H., J. D. Vervoort, S. A. DuFrane, V. Carlotto,
nomic mineral deposits: British Geological Survey, C. Reimann, and J. Cárdenas, 2011, The U-Pb and Hf
Overseas Geology and Mineral Resources, v. 67, p. 1–63, isotope evidence of detrital zircons of the Ordovician
Nottingham. ­Ollantaytambo Formation, southern Peru, and the Ordo-
Astini, R. A., and J. L. Benedetto, 1993, A collisional model for vician provenance and paleogeography of southern Peru
the stratigraphic evolution of the Argentine Precordillera and northern Bolivia: Journal of South American Earth
during the early Paleozoic: 2° International Symposium Sciences, v. 32, p. 196–209.
on Andean Geodynamics (Oxford), p. 501–504, Paris. Bahlburg, H., J. Berndt, and A. Gerdes, 2016, The ages and
Astini, R. A., J. L. Benedetto, and N. E. Vaccari, 1995, The tectonic setting of the Faja Eruptiva de la Puna Ori-
early Paleozoic evolution of the Argentina Precordillera ental, Ordovician, NW Argentina: Lithos, v. 256–257,
as a Laurentian rifted, drifted, and collided terrane: A ge- p. 41–54.
odynamic model: Geological Society of America Bulletin, Baldo, E., C. Casquet, R. J. Pankhurst, C. Galindo, C. W.
v. 107, p. 253–273. Rapela, C. M. Fanning, J. Dahlquist, and J. Murra, 2006,
Astini, R. A., F. Martina, M. Ezpeleta, F. Dávila, and Neoproterozoic A-type magmatism in the Western Si-
P. ­Cawood, 2009, Chronology from rifting to foreland erras Pampeanas (Argentina): Evidence for Rodinia
­basin in the Paganzo Basin (Argentina), and a reappraisal break-up along a proto-Iapetus rift?: Terra Nova, v. 18,
on the “Eo- and Neohercynian” tectonics along West- p. 388–394.
ern Gondwana: 12° Congreso Geológico Chileno, Actas Beck, S. L., and G. Zandt, 2002, The nature of orogenic crust
S9(010), p. 1–4, Santiago de Chile. in the Central Andes: Journal of Geophysical Research,
Auboin, J. A., A. V. Borrello, G. Cecione, R. Charrier, P. v. 107, no. B10, p. 2230, DOI: 10.1029/2000JB000124.
Chotin, J. Frutos, R. Thiele, and J. C. Vicente, 1973, Bellido, F., P. Valverde, F. Jaimes, V. Carlotto, and E. Díaz-
­E squisse paleogeographique et structurale des Andes Martínez, 2009, Dating and geochemical characterization

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 27 11/05/18 12:18 PM


28 Ramos

of the peraluminous granitoids of Amotape, Illescas and Cardona, A., 2006, Reconhecimento da evolução tectônica da
Paita massifs (northwestern Peru): Boletín de la Sociedad proto-margem Andina do centro-norte Peruano, baseada
Geológica del Perú, v. 103, p. 197–213. em dados geoquímicos e isotópicos do embasamento da
Bellizzia, A., and N. Pimentel, 1994, Terreno Mérida: un Cordilheira Oriental na região de Huánuco-La Unión,
cinturón alóctono Herciniano en la Cordillera de Los Ph.D. Thesis, Universidade de São Paulo, São Paulo, 196 p.
Andes de Venezuela, in 5° Simposio Bolivariano Explo- Cardona, A., U. G. Cordani, J. Ruiz, V. Valencia, A. P. Nut-
ración Petrolera en las Cuencas Subandinas, Memoria, man, and A. W. Sanchez, 2005a, U/Pb detrital zircon
p. 271–299. geochronology and Nd isotopes from Paleozoic metased-
Benavides Cáceres, V., 1962, Estratigrafía pre-terciaria de la imentary rocks of the Marañon complex: Insights on the
región de Arequipa: Boletín de la Sociedad Geológica del proto-Andean tectonic evolution of the eastern Peruvian
Perú, v. 38, p. 5–63. Andes: in 5° South American Symposium on Isotope Ge-
Benavides Cáceres, V., 1999, Orogenic evolution of the ology, Proceedings, Punta del Este, p. 208–211.
­Peruvian Andes: The Andean cycle, in B. Skinner, ed., Cardona, A., U. G. Cordani, A. P. Nutman, U. Zimmermann,
­Geology and mineral deposits of Central Andes: Society and A. W. Sánchez, 2005b, Tectonic setting and geochro-
of Economic Geology Special Publication 7, p. 61–107. nology of Pre-Llanvirnian metamorphic rocks of the
Bertrand, H., M. Fornari, A. Marzoli, T. Sempere, and G. Marañon Complex (E. Peru): From rifting to collision
Feraud, 2005, Early Mesozoic rift-related magmatism in to form the Rheic Ocean: Gondwana 12 Abstracts, Aca-
the Bolivian Andes and Subandes: The southernmost re- demia Nacional de Ciencias, Mendoza, Argentina, p. 87.
cord of the Central Atlantic Magmatic Province, in 6th In- Cardona, A., U. G. Cordani, and A. W. Sánchez, 2007, Meta-
ternational Symposium on Andean Geodynamics (ISAG morphic, geochronological and geochemical constraints
2005, Barcelona), Extended Abstracts, p. 111–114. from the pre-Permian basement of the eastern Peruvian
Bird, P., 1979, Continental delamination and the Colorado Andes (10°S): A Paleozoic extensional-accretionary oro-
Plateau: Journal of Geophysical Research, v. 84, no. B13, gen?, in 20th Colloquium Latin American Earth Sci-
p. 7561–7571. ences: Kiel, Germany, Deutsche Forschungsgemeinschaft,
Boekhout, F., T. Sempere, R. Spikings, and U. Schaltegger, p. 29–30.
2013, Late Paleozoic to Jurassic chronostratigraphy of Cardona, A., U. G. Cordani, and A. P. Nutman, 2008, U.Pb
coastal southern Peru: Temporal evolution of sedimenta- SHRIMP zircon, 40Ar/39Ar geochronology and Nd iso-
tion along an active margin: Journal of South American topes from granitoid rocks of the Illescas Massif, Peru:
Earth Sciences, v. 47, p. 179–200. A southern extension of a fragmented Late Paleozoic oro-
Boekhout, F., N. M. W. Roberts, A. Gerdes and U. Schalteg- gen?, in 6th South American Symposium on Isotope Ge-
ger, 2015, A Hf-isotope perspective on continental forma- ology, Proceedings: Bariloche, Argentina, Abstracts, p. 78.
tion in the south Peruvian Andes, in N. M. W. Roberts, Cardona, A., U. G. Cordani, J. Ruiz, V. A. Valencia, R. Arm-
M. Van Kranendonk, S. Parman, S. Shirey, and P. D. Clift, strong, D. Chew, A. Nutman, and A. W. Sanchez, 2009, U‐
eds., Continent formation through time: London, Geolog- Pb zircon geochronology and Nd isotopic signatures of
ical Society, Special Publications 389, p. 305–321. the pre‐mesozoic metamorphic basement of the Eastern
Calderon, Y., P. Baby, C. Hurtado, and S. Brusset, 2017, Thrust Peruvian Andes: Growth and provenance of a Late Neo-
tectonics in the Andean retro foreland basin of northern proterozoic to Carboniferous accretionary orogen on the
Peru: Permian inheritances and petroleum implications: Northwest Margin of Gondwana: The Journal of Geology,
Marine and Petroleum Geology, v. 82, p. 238–250. v. 117, p. 285–305.
Calixto, F. J., E. Sandvol, S. M. Kay, P. Mulcahy, B. Heit, Cardona, A., D. Chew, V. A. Valencia, G. Bayona,
X. Yuan, B. L. Coira, D. Comte, and P. Alvarado, 2013, A. Miškovic’, and M. Ibañez-Mejía, 2010, Grenvillian rem-
Velocity structure beneath the southern Puna plateau: nants in the Northern Andes: Rodinian and Phanerozoic
Evidence for delamination: Geochemistry, Geophysics, paleogeographic perspectives: Journal of South American
Geosystems, G3, v. 14, DOI: 10.1002/ggge.20266. Earth Sciences, v. 29, p. 92–104.
Callot, P., 2008, La Formation Ayabacas (limite Turonien- Carlier, G., G. Grandin, G. Laubacher, R. Marocco, and
Coniacien, Sud-Pérou): collapse sous-marin en réponse à F. Mégard, 1982, Present knowledge of the magmatic
l’amorce de l’orogenèse andine: Doctorat de l’Université evolution of the Eastern Cordillera of Peru: Earth-Science
de Toulouse, Toulouse, 251 p. ­Reviews, v. 18, p. 253–283.
Callot, P., T. Sempere, F. Odonne, and E. Robert, 2008a, Carlier, G., J. P. Lorand, M. Bonhomme, and V. Carlotto,
­Giant submarine collapse of a carbonate platform at the 1996, A reappraisal of the Cenozoic inner arc magmatism
Turonian-Coniacian transition: The Ayabacas Formation, in the Southern Peru: Consequences for the evolution of
southern Peru: Basin Research, v. 20, p. 1–24. the Central Andes for the past 50 Ma, in 3rd Symposium
Callot, P., F. Odonne, T. Sempere, 2008b, Liquification and International sur la Géodynamique Andine, Saint Malo,
soft-sediment deformation in a limestone megabreccia: p. 551–554.
The Ayabacas giant collapse, Cretaceous, southern Peru: Carlier, G., J. P. Lorand, J. P. Liégeois, M. Fornari, P. Soler,
Sedimentary Geology, v. 212, p. 49–69. V. Carlotto, and J. Cárdenas, 2005, Potassic-ultrapotassic
Caputo, M. V., 2014, Juruá Orogeny: Brazil and Andean mafic rocks delineate two lithospheric mantle blocks be-
Countries: Brazilian Journal of Geology, v. 44, no. 2, neath the southern Peruvian Altiplano: Geology, v. 33,
p. 181–190. p. 601–604.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 28 11/05/18 12:18 PM


Tectonic Evolution of the Central Andes: From Terrane Accretion to Crustal Delamination  29

Carlotto, V., 1998, Evolution andine et raccourcissement au Resúmenes Extendidos: Sociedad Geológica del Perú,
niveau de Cusco (13-6°S), Pérou: enregistrement sédi- Cusco, p. 54–62.
mentaire, chronologie, contrôles paléogéographiques, Cawood, P. A., 2005, Terra Australis orogen: Rodinia breakup
évolution cinémematique, Thèse Doctorat, Université and development of the Pacific and Iapetus margins of
Joseph Fourier, Grenoble, France, 159 p. Gondwana during the Neoproterozoic and Paleozoic:
Carlotto, V., 2002, Evolution andine et raccourcissement Earth-Science Reviews, v. 69, p. 249–279.
au niveau de Cusco (13-16°S) Pérou: Thèse de Doctorat, Cediel, F., R. P. Shaw, and C. Cáceres, 2003, Tectonic as-
Université Joseph Fourier, Grenoble. Géologie Alpine, sembly of the northern Andean block, in C. Bartolini,
­Mémoire Hors Série, v. 39, 203 p. R. T. Buffler, and J. Blickwede, eds., The Circum-Gulf of
Carlotto, V., 2013, Paleogeographic and tectonic controls on Mexico and Caribbean: Hydrocarbon habitats, basin for-
the evolution of Cenozoic basins in the Altiplano and mation, and plate tectonics: American Association of Pe-
Western Cordillera of southern Peru: Tectonophysics, troleum Geologists Memoir 79, p. 815–848.
v. 589, p. 195–219. Charrier, R., 1984, Áreas subsidentes en el borde occidental
Carlotto, V., E. Jaillard, G. Carlier, J. Cárdenas, L. Cerpa, T. Flo- de la cuenca de trasarco Jurásico-Cretácica, Cordillera
res, O. Latorre, and I. Ibarra, 2005, Las cuencas terciarias Principal chilena entre 34° y 3430°S, in 9th Congreso Ge-
sinorogénicas en el Altiplano y en la Cordillera Occidental ológico Argentino, S.C. de Bariloche, v. 2, p. 107–124.
del sur del Perú, in J. Arce, ed., Sociedad Geológica del Perú: Chew, D. M., U. Schaltegger, J. Košler, M. J. Whitehouse,
Alberto Giesecke Matto Volumen Especial 6, p. 103–126. M. Gutjahr, R. A. Spikings, and A. Miškovic’, 2007, U-Pb
Carlotto, V., J. Cárdenas, and G. Carlier, 2008, The litho- geochronologic evidence for the evolution of the Gond-
sphere of Southern Peru: A result of the accretion of al- wanan margin of the north Central Andes: Bulletin of the
lochthonous blocks during the Mesoproterozoic, in 7th Geological Society of America, v. 119, no. 5–6, p. 697–711.
International Symposium on Andean Geodynamics
Chew, D. M., T. Magna, C. L. Kirkland, A. Miškovi c’ ,
(ISAG 2008, Nice), Extended Abstracts, p. 105–108.
A. Cardona, R. Spikings, and U. Schaltegger, 2008, Detri-
Carlotto, V. et al., 2009a, Dominios geotectónicos y metalo-
tal zircon fingerprint of the Proto-Andes: Evidence for a
génesis del Perú: Boletín Sociedad Geológica del Perú,
Neoproterozoic active margin?: Precambrian Research,
v. 103, p. 1–89.
v. 167, p. 186–200.
Carlotto, V., R. Rodríguez, H. Acosta, J. Cárdenas, and
E. Jaillard, 2009b, Totos-Paras (Ayacucho) structural high: Chew, D. M., A. Cardona, and A. Miškovic’, 2011, Tectonic
A paleogeographic boundary in the Mesozoic evolu- evolution of western Amazonia from the assembly of Ro-
tion of the Pucará (Late Triassic-Liassic) and Arequipa dinia to its break-up: International Geology Review, v. 53,
­( Jurassic-Cretaceous) basins: Sociedad Geológica del no. 11–12, p. 1280–1296.
Perú, Volumen Especial, v. 7, p. 1–46. Chew, D. M., G. Pedemonte, and E. Corbett, 2016, Proto-­
Carlotto, V., H. Acosta, M. Mamani, L. Cerpa, R. Rodríguez, Andean evolution of the Eastern Cordillera of Peru:
F. Jaimes, P. Navarro, and C. Checaltana, 2011, Los do- Gondwana Research, v. 35, p. 59–78.
minios geotectónicos del Perú: Aportes para la interpre- Clark, A. H., D. J. Kontak, and E. Farrar, 1984, A comparative
tación de la evolución de los Andes, in 14th Congreso study of the metallogenetic and geochronological relation-
Latinoamericano de Geología and 13th Congreso Colom- ships in the northern part of the Central Andean tin belt,
biano de Geología, p. 109, Medellín. SE Peru and NW Bolivia, in T. V. Janelidze, and A. G. Tval-
Castroviejo, R., J. F. Rodrigues, J. Acosta, E. Pereira, chralidze, eds., 6th IAGOD Symposium: Stuttgart, p. 269–279.
D. Romero, J. Quispe, and J. A. Espi, 2009a, Geología Clark, A. H. et al., 1990, Geologic and geochronologic con-
de las ultramafitas pre-andinas de Tapo y Acobamba, straints on the metallogenic evolution of the Andes of
Tarma, Cordillera Oriental del Perú: Geogaceta, v. 46, southeastern Peru: Economic Geology, v. 85, p. 1520–1583.
p. 7–10. Clift, P. D., 2003, Tectonic erosion of the Peruvian forearc,
Castroviejo, R., J. F. Rodrigues, R. Carrascal, H. Chirif, Lima Basin, by subduction and Nazca Ridge collision: Tec-
J. Acosta, and R. Uribe, 2009b, Mineralogía del Complejo tonics, v. 22, no. 3, p. 1023, DOI: 10.1029/2002TC001386.
Ultramáfico de Tapo (Cordillera Oriental Andina, Perú): Coira, B. L., J. D. Davidson, C. Mpodozis, and V. A. Ramos,
Revista de la Sociedad Española de Mineralogía, Macla, 1982, Tectonic and magmatic evolution of the Andes of
v. 11, p. 55–56. northern Argentina and Chile: Earth-Science Reviews,
Castroviejo, R., J. Macharé, P. Castro, E. Pereira, J. F. Rod- v. 18, p. 303–332.
rigues, C. G. Tassinari, A. Willner, and J. Acosta, 2010a, Cordani, U. G., and W. Teixeira, 2007, Proterozoic accretion-
Significado de las ofiolitas neoproterozoicas de la Cor- ary belts in the Amazonian Craton, in R. D. Hatcher Jr.,
dillera Oriental del Perú (9°30’–11°30’), in 15th Congreso M. P. Carlson, J. H. McBride, J. R. Martínez-Catalán, eds.,
Peruano de Geología, Resúmenes Extendidos, Sociedad 4-D Framework of Continental Crust: Geological Society
Geológica del Perú, Cusco, p. 51–53. of America Memoir 200, p. 297–320.
Castroviejo, R., R. Carrascal, H. Chirif, J. F. Rodrigues, Cordani U. G., K. Sato, W. Teixeira, C. C. G. Tassinari, and M. A.
J. Acosta, and H. J. Bernhardt, 2010b, Metalogenia aso- Basei, 2000, Crustal evolution of the South American plat-
ciada a los segmentos ofiolíticos de la Cordillera Oriental form, in U. G. Cordani, E. J. Milani, A. Thomaz Filho, and D.
del Perú Central. Simposio S3 Bloques Litosféricos de la A. Campos, eds., Tectonic evolution of South America: 31st
Cadena Andina, in 15th Congreso Peruano de Geología, International geological Congress, Rio de Janeiro, p. 19–40.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 29 11/05/18 12:18 PM


30 Ramos

Dalmayrac, B., 1978, Géologie de la Cordillére orientale de G a r z i o n e , C . N . , P. M o l n a r, J . C . L i b a r k i n , a n d


la région de Huanuco: sa place dans une transversale B. J. ­MacFadden, 2006, Rapid late Miocene rise of the
des ­Andes du Pérou central (9°S à 10°30’ S): ­Géologie Bolivian ­Altiplano: Evidence for removal of mantle litho-
des ­A ndes péruviennes: Travaux et documents de sphere: Earth Planetary Science Letters, v. 241, p. 543–556.
l’O.R.S.T.O.M 1, Office de la Recherche Scientifique et Garzione, C. N., P. Molnar, J. C. Libarkin, and B. J. MacFadden,
Technique Outre-mer, Paris, Mémoire, v. 93, 162 p. 2007, Reply to Comment on “Rapid late Miocene rise of the
Dalmayrac, B., J. R. Lancelot, and A. Leyreloup, 1977, Bolivian Altiplano: Evidence for removal of mantle litho-
­Two-billion-year granulites in the late Precambrian meta- sphere” by Garzione et al. (2006), Earth Planet. Sci. Lett.
morphic basement along the southern Peruvian coast: 241 (2006) 5432556: Earth and Planetary Science Letters,
Science, v. 198, p. 49–51. v. 259, p. 630–633.
Dalmayrac, B., G. Laubacher, R. Marocco, C. Martinez, and Garzione, C. N. et al., 2017, Tectonic evolution of the Cen-
P. Tomasi, 1980, La Chaine Hercynienne d’Amerique du tral Andean Plateau and implications for the growth of
Sud. Structure et evolution d’une orogène intracrato- Plateaus: Annual Review of Earth and Planetary Sciences,
nique: Geologische Rundschau, v. 69, p. 1–21. v. 45, p. 529–559.
Daly, M., 1989, Correlations between Nazca/Farallon plate Grandin, G., and J. Zegarra, 1979, Las rocas ultrabásicas en
kinematics and forearc evolution in Ecuador: Tectonics, el Perú, las intrusiones lenticulares y los sills de la región
v, 8, no. 4, p. 769–790. Huanuco-Monzon: Boletín de la Sociedad Geológica del
Dalziel, I. W. D., and R. D. Forsythe, 1985, Andean evolu- Perú, v. 63, p. 99–115.
tion and the terrane concept, in D. G. Howell, ed., Tec- Grosse, P., F. Söllner, M. Báez, A. J. Toselli, J. N. Rossi, and
tonostratigraphic terranes of the Circum-Pacific-Region: J. de la Rosa, 2009, Lower Carboniferous post-orogenic
Circum-Pacific Council of Energy and Mineral Resources, granites in central-eastern Sierra de Velasco, Sierras Pam-
Earth Science Series 1, p. 565–581. peanas, Argentina: U-Pb monazite geochronology, geo-
Dorbath, C., M. Granet, G. Poupinet, and C. Martínez, 1993, chemistry and Sr-Nd isotopes: International Journal of
A teleseismic study of the Altiplano and the Eastern Earth Sciences, v. 98, p. 1001–1025.
­Cordillera and northern Bolivia: New constraints on a lith- Haller, M. A., and V. A. Ramos, 1984, Las ofiolitas fam-
ospheric model: Journal of Geophysical Research, v. 98, atinianas (Eopaleozoico) de las provincias de San Juan
p. 9825–9844. y Mendoza, in 9th Congreso Geológico Argentino (S.C.
Du Toit, A. L., 1927, A geological comparison of South America Bariloche), Actas, v. 2, p. 66–83.
with South Africa: Publications Carnegie Institute, 157 p. Hartley, A. J., 2003, Andean uplift and climate change: Jour-
Egeler, C. G., and T. de Booy, 1961, Preliminary note on the nal of the Geological Society, London, v. 160, p. 7–10.
geology of the Cordillera Vilcabamba (SE Peru), with em- Hatcher, R. D., 2010, The Appalachian orogen: A brief sum-
phasis on the essentially pre-Andean origin of the struc- mary, in R. P. Tollo, M. J. Bartholomew, J. P. Hibbard,
ture: Geologie en Mijnbouw, v. 40 Jaargang, p. 319–325. and P. M. Karabinos, eds., From Rodinia to Pangea:
Erlich, R. N., J. Fallon, and P. O’Sullivan, 2018, Stratigra- The Lithotectonic Record of the Appalachian Region:
phy and LA-ICP-MS zircon U-Pb provenance of Middle ­G eological Society of America Memoir 206, p. 1–19,
Permian to Maastrichtian sandstones from outcrop and DOI: 10.1130/2010.1206(01).
subsurface control in the sub-Andean basins of Peru, in Heim, A., 1948a, Geología de los rios Apurimac y Urubamba:
G. Zamora, K. R. McClay, and V. A. Ramos, eds., Petro- Boletín del Instituto Geológico del Perú, v. 10, p. 1–25.
leum basins and hydrocarbon potential of the ­Andes of Heim, A., 1948b, Observaciones tectónicas en la Rincon-
Peru and Bolivia: AAPG Memoir 117 p. 179–226. ada, Precordillera de San Juan, in Dirección de Minas y
Escayola, M. P., C. R. van Staal, and W. J. Davis, 2011, The ­Geología, Buenos Aires, Boletín 64, 47 p.
age and tectonic setting of the Puncoviscana Formation in Isacks, B., 1988, Uplift of the Central Andean plateau and
northwestern Argentina: An accretionary complex related bending of the Bolivian oroclino: Journal Geophysical
to Early Cambrian closure of the Puncoviscana Ocean ­Research, v. 93, p. 3211–3231.
and accretion of the Arequipa-Antofalla block: Journal of Jacay, J., 1994, Evolution sédimentaire de la marge andine:
South American Earth Sciences, v. 32, no. 4, p. 438–459. le Crétacé supérieur des Andes du Pérou central et sep-
Fanlo, I., F. Gervilla, R. Castroviejo, J. F. Rodrigues, tentrional: Memoires de DEA, Université Joseph Fourier,
E. Pereira, J. Acosta, and R. Uribe, 2009, Metamorphism Grenoble, 97 p.
of chromitites in the Tapo ultramafic massif, Eastern Cor- Jaillard, E., 1994, Kimmeridgian to Paleocene tectonic and
dillera, Peru, in Proceedings 10th Biennial SGA Meeting, geodynamic evolution of the Peruvian (and Ecuadorian)
Townsville, Australia, p. 161–163. margin, in J. A. Salfity, ed., Cretaceous tectonics in the
Feininger, T., 1978, Geologic Map of Western El Oro Prov- Andes: Earth Evolution Sciences, Braunschweig/Wies-
ince: Quito, Politécnica Nacional, scale 1:50,000. baden, Fried. Vieweg & Sohn, p. 101–167.
Feininger, T., 1987, Allochthonous terranes in the Andes of Jaillard, E., P. Soler, G. Carlier, and T. Mourier, 1990, Geody-
Ecuador and northwestern Peru: Canadian Journal of namic evolution of the northern and Central Andes dur-
Earth Sciences, v. 24, p. 266–278. ing early to middle Mesozoic times: a Tethyan model:
Festa, A., K. Ogata, G. A. Pini, Y. Dilek, and J. L. Alonso, Journal of the Geological Society, v. 147, p. 1009–1022.
2016, Origin and significance of olistostromes in the evo- Jaillard, E., H. Cappetta, P. Ellenberger, M. Feist, N.
lution of orogenic belts: A global synthesis: Gondwana ­G rambast-Fessard, J. P. Lefranc, and B. Sigé, 1993,
Research, v. 39, p. 180–203. The Late Cretaceous Vilquechico Group of southern

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 30 11/05/18 12:18 PM


Tectonic Evolution of the Central Andes: From Terrane Accretion to Crustal Delamination  31

Peru. Sedimentology, paleontology, biostratigraphy, ­ xploración y D


E ­ esarrollo de Hidrocarburos, Mendoza,
­correlations: Cretaceous Research, v. 14, p. 623–661. Trabajos ­Técnicos, v. 2, p. 639–658.
Jaillard, E., and P. Soler, 1996, Cretaceous to early Paleogene Litherland, M., J. A. Aspden, and R. A. Jemielita, 1994, The
tectonic evolution of the northern Central Andes (0–18°S) metamorphic belts of Ecuador: British Geological Survey
and its relations to geodynamics: Tectonophysics, v. 259, Overseas Memoir, v. 11, p. 1–146.
p. 41–53. Loewy, S. L., J. N. Connelly, I. W. D. Dalziel, and C. F. Gower,
Jaillard, E., G. Hérail, T. Monfret, E. Díaz-Martínez, P. Baby, 2003, Eastern Laurentia in Rodinia: Constraints from
A. Lavenu, and J. F. Dumon, 2000, Tectonic evolution whole-rock Pb and U/Pb geochronology: Tectonophys-
of the Andes of Ecuador, Peru, Bolivia y Northernmost ics, v. 375, p. 169–197.
Chile, in U. J. Cordani, E. J. Milani, A. Thomaz Filho, and Loewy, S. L., J. N. Connelly, and I. W. D. Dalziel, 2004, An
D. A. Campos, eds., Tectonic evolution of South America: orphaned basement block: The Arequipa-Antofalla base-
31st International Geological Congress, Río de Janeiro, ment of the central Andean margin of South America:
p. 481–559. Geological Society of America Bulletin, v. 116, p. 171–187.
Jaillard, E., P. Bengtson, and A. Dhondt, 2005, Late Creta- Macharé, J., and L. Ortlieb, 1992, Plio-Quaternary vertical
ceous marine transgressions in Ecuador and northern motions and the subduction of the Nazca Ridge, central
Peru: a refined stratigraphic framework: Journal of South coast of Peru: Tectonophysics, v. 205, p. 97–108.
American Earth Sciences, v. 19, p. 307–323. Mamani, M., 2006, Variations in magma composition in time
James, D. E., and I. S. Sacks, 1999, Cenozoic formation of the and space along the Central Andes ( 12°S228aS ): PhD
Central Andes: A geophysical perspective, in B. Skinner, Dissertation, Universität zu Göttingen, Göttingen, 123 p.
ed., Geology and mineral deposits of Central Andes: So- Mamani, M., A. Tassara, and G. Wörner, 2008, Composition
ciety of Economic Geology Special Publication 7, p. 1–25. and structural control of crustal domains in the Central
Jiménez, N., and S. López-Velásquez, 2008, Magmatism in Andes: Geochemistry, Geophysics, Geosystems, G3, v. 9,
the Huarina belt, Bolivia, and its geotectonic implica- no. 3, p. Q03006.
tions: Tectonophysics, v. 459, p. 85–106. Mamani, M., G. Wörner, and T. Sempere, 2010, Geochemi-
Jiménez, N., S. López-Velásquez, and R. Santivañez, 2009, Evolu- cal variations in igneous rocks of the Central Andean
ción tectonomagmática de los Andes bolivianos: Revista de orocline (13°S to 18°S): Tracing crustal thickening and
la Asociación Geológica Argentina, v. 65, no. 1, p. 36–67. magma generation through time: Geological Society of
Kay, R. W., and S. M. Kay, 1993, Delamination and delamina- America Bulletin, v. 122, p. 162–182.
tion magmatism: Tectonophysics, v. 219, p. 177–189. Marocco, R., 1978, Un segment E-W de la chaine des Andes
Kay, S. M., and B. Coira, 2009, Shallowing and steepening péruviennes: la déflexion d’Abancay. Etude géologique
subduction zones, continental lithosphere loss, mag- de la Cordillére orientale et des hauts plateaux entre
matism and crustal flow under the Central Andean Cuzco et San Miguel, sud du Pérou (12°30’S–14°00’S):
Altiplano-Puna Plateau, in S. M. Kay, V. A. Ramos, and Géologie des Andes péruviennes, Travaux et documents
W. Dickinson, eds., Backbone of the Americas: Geological de l’O.R.S.T.O.M., Paris, v. 94, 196 p.
Society of America Memoir 204, p. 229–259. Martina, F., J. M. Viramonte, R. A. Astini, M. M. Pimen-
Kay, S. M., B. Coira, and J. Viramonte, 1994, Young mafic tel, and E. Dantas, 2011, Mississippian volcanism in the
backarc volcanic rocks as indicators of continental lith- south-Central Andes: New U-Pb SHRIMP zircon geo-
ospheric delamination beneath the Argentina Puna Pla- chronology and whole-rock geochemistry: Gondwana
teau, Central Andes: Journal Geophysical Research, v. 99, Research, v. 19, no. 2, p. 524–534.
p. 24323–24339. Martinez, C., C. Dorbath, and A. Lavenu, 1994, La cuenca
Kay, S. M., C. Mpodozis, and B. Coira, 1999, Neogene mag- subsidente cenozoica noraltiplanica y sus relaciones con
matism, tectonism, and mineral deposits of the Central una subducción transcurrente continental, in Memorias
Andes (22°S to 33°S), in B. Skinner, ed., Geology and min- 12th Congreso Geológico de Bolivia, Tarija, p. 3–8.
eral deposits of Central Andes: Society of Economic Geol- Martin-Gombojav, N., and W. Winkler, 2008, Recycling of
ogy, Special Publication 7, p. 27–59. Proterozoic crust in the Andean Amazon foreland of
Keidel, J., 1921, Sobre la distribución de los depósitos glaci- Ecuador: Implications for orogenic development of the
ares del Pérmico conocidos en la Argentina y su signifi- Northern Andes: Terra Nova, v. 20, p. 22–31.
cación para la estratigrafía de la serie del Gondwana y la Marzoli, A., P. R. Renne, E. M. Piccirillo, M. Ernesto, G. Bell-
paleogeografía del Hemisferio Austral: Boletín Academia ini, and A. De Min, 1999, Extensive 200-million-year-old
Nacional de Ciencias, Córdoba, v. 25, p. 239–368. continental flood basalts of the Central Atlantic Mag-
Kontak, D. J., A. H. Clark, E. Farrar, D. A. Archibald, and matic Province: Science, v. 284, p. 616–618.
H.  Baadsgaard, 1990, Late Paleozoic–early Mesozoic Marzoli, A., H. Bertrand, M. Chiaradia, D. Fontignie, N.
magmatism in the Cordillera de Carabaya, Puno, south- Youbi, and G. Bellieni, 2006, The genesis of CAMP basalts
eastern Peru: Geochronology and Geochemistry: Journal (Morocco) from enriched lithosphere to late asthenosphere
of South American Earth Sciences, v. 3, p. 213–230. mantle sources: Geochimica et Cosmochimica Acta, Gold-
Kusiak, M. A., G. Mascle, and V. A. Ramos, 2014, El mag- schmidt Conference Abstracts, v. 70, no. 18, p. A396.
matismo mesozoico asociado a los procesos extension- Mathalone, J. M. P., and M. Montoya, 1995, Petroleum geol-
ales en las cuencas de rift del Subandino Boliviano y ogy of the Subandean basins of Peru, in A. J. Tankard, R.
la importancia de su ubicación geocronológica en la Suárez, and H. J. Welsink, eds., Petroleum basins of South
paleogeografía de América del Sur, in 9th Congreso de America: AAPG Memoir 62, p. 423–444.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 31 11/05/18 12:18 PM


32 Ramos

McLeod, C. L., J. P. Davidson, G. M. Nowell, S. L. de Silva, Pankhurst, R., C. W. Rapela, J. Saavedra, E. Baldo,
and A. K. Schmit, 2013, Characterizing the continental J. ­Dahlquist, and I. Pascua, 1996, Sierra de Los Llanos,
basement of the Central Andes: Constraints from Bolivian Malanzán, and Chepes: Ordovician I and S-type granitic
crustal xenoliths: Geological Society of America ­Bulletin, magmatism in the Famatinuan orogen, in Actas 13th Con-
v. 125, no. 5/6, p. 985–997. greso ­Geológico Argentino and 3rd Congreso de Explo-
Mégard, F., 1978, Etude Géologique du Pérou Central: ración de Hidrocarburos, Buenos Aires, v. 5, p. 415.
Contribution à l’Etude Géologique des Andes : Géolo- Pankhurst, R. J., C. W. Rapela, J. Saavedra, E. Baldo, J.
gie des Andes péruviennes 1, Travaux et documents de ­Dahlquist, I. Pascua, and C. M. Fanning, 1998, The Fam-
l’O.R.S.T.O.M, Office de la Recherche Scientifique et Tech- atinian magmatic arc in the Central Sierras Pampeanas:
nique Outre-mer, Paris, Mémoire 86, 301 p. An early to mid-Ordovician continental arc on the Gond-
Mégard, F., 1979, Estudio geológico de los Andes del Perú wana margin, in R. J. Pankhurst and C. W. Rapela, eds.,
Central: Instituto Geología, Minería y Metalogenia, Lima, The Proto-Andean Margin of Gondwana: Geological
Boletín Serie D: Estudios Especiales, v. 8, 227 p. ­Society of London Special Publication 142, p. 343–367.
Mégard, F., 1984, The Andean orogenic period and its ma- Perelló, J., V. Carlotto, A. Zárate, P. Ramos, H. Posso,
jor structures in central and northern Peru: Journal of the C. Neyra, A. Caballero, N. Fuster, and R. Muhr, 2003, Por-
Geological Society of London, v. 141, no. 5, p. 893–900. phyry-style alteration and mineralization of the middle
Mégard, F., Noble, D. C, McKee, E. H., and Bellon H., 1984. Eocene to early Oligocene Andahuaylas–Yauri belt, Cuzco
Multiple pulses of Neogene compressive deformation region, Peru: Economic Geology, v. 98, p. 1575–1605.
in the Ayacucho intermontane basin, Andes of cen- Perelló, J., Carlotto, V., Zárate, A., Ramos, P., Posso, H., Neyra,
tral Peru. Geological Society of America, Bulletin v. 95, C., Caballero, A., Fuster, N., And Muhr, R.,2003, Por-
p. 1108–1117. phyry-Style Alteration and Mineralization of the Middle
Miškovic’, A., and U. Schaltegger, 2009, Crustal growth along Eocene to Early Oligocene Andahuaylas-Yauri Belt, Cuzco
a non-collisional cratonic margin: A Lu–Hf isotopic sur- Region, Peru, Economic Geology, v. 98, p. 1575–1605.
vey of the Eastern Cordilleran granitoids of Peru: Earth Perez, N. D., and B. K. Horton, 2014, Oligocene-Miocene de-
and Planetary Science Letters, v. 279, p. 303–315. formational and depositional history of the Andean hin-
Miškovi c’ , A., R. A. Spikings, D. M. Chew, J. Kosler, A. terland basin in the northern Altiplano plateau southern
­U lianov, and U. Schaltegger, 2009a, Tectonomagmatic Peru: Tectonics, v. 33, p. 1819–1847.
evolution of Western Amazonia: Geochemical characteri- Petersen, U., 1999, Magmatic and metallogenic evolution of
zation and zircon UePb geochronologic constraints from the Central Andes, in B. Skinner, ed., Geology and min-
the Peruvian Eastern Cordilleran granitoids: Geological eral deposits of Central Andes: Society of Economic Geol-
Society of America Bulletin, v. 121, no. 9/10, p. 1298–1324. ogy, Special Publication 7, p. 109–153.
Miškovic’, A., U. Schaltegger, and D. Chew, 2009b, Carbon- Petford, N., and M. P. Atherton, 1995, Cretaceous-Tertiary
iferous plutonism along the Eastern Peruvian Cordillera: volcanism and syn-subduction crustal extension in north-
Implications for the Late Paleozoic to Early Mesozoic ern central Peru, in J. L. Smelie, ed., Volcanism associated
Gondwanan tectonics, in 6th International Symposium on with extension at consuming plate margins: Geological
Andean Geodynamics (ISAG 2005, Barcelona), E ­ xtended Society London, Special Publication 81, p. 233–248.
Abstracts, p. 508–511. Pitcher, W. S., M. P. Atherton, E. J. Cobbing, and R. D. Beck-
Mitouard, P., C. Kissel, and C. Laj, 1990, Post-Oligocene rotations insale, eds., 1985, Magmatism at a plate edge: The Peru-
in southern Ecuador and northern Peru and the formation of vian Andes: Glasgow (Blackie) and New York (Halsted
the Huancabamba deflection in the Andean Cordillera: Earth Press), 328 p.
and Planetary Science Letters, v, 98, no. 3–4, p. 329–339. Raimondi, A., 1929, El Perú, Itinerarios de viaje: Lima,
Moretti, I., P. Baby, E. Mendez, and D. Zubieta, 1996, ­Imprenta Torres Aguirre, 371 p.
­Hydrocarbon generation in relation to thrusting in the Ramos, V. A., 1988, The tectonics of the Central Andes: 30° to
Sub Andean Zone from 18 to 22°S—Bolivia: Petroleum 33°S latitude, in S. Clark, and D. Burchfiel, eds., Processes
Geoscience, v. 2, p. 17–28. in continental lithospheric deformation: Geological Soci-
Mourier, T., C. Laj, F. Mégard, P. Roperch, P. Mitouard, and M. ety of America Special Paper 218, p. 31–54.
Farfan, 1988, An accreted continental terrane in northwestern Ramos, V. A., 1999, Plate tectonic setting of the Andean
Peru: Earth and Planetary Science Letters, v. 88, p. 182–192. ­Cordillera: Episodes, v. 22, no. 3, p. 183–190.
Mpodozis, C., and V. A. Ramos, 1989, The Andes of Chile Ramos, V. A., 2000, Evolución tectónica de la Argentina, in
and Argentina, in G. E. Ericksen, M. T. Cañas Pinochet, R. Caminos, ed., Geología Argentina: Buenos Aires,
and J. A. Reinemud, eds., Geology of the Andes and its re- ­Anales Instituto de Geología y Recursos Minerales, v. 29,
lation to Hydrocarbon and Mineral Resources: Houston, p. 715–784.
Circumpacific Council for Energy and Mineral ­Resources, Ramos, V. A., 2008, The basement of the Central Andes: The
Earth Sciences Series 11, p. 59–90. Arequipa and related terranes: Annual Review of Earth
Otamendi, J. E., A. M. Tibaldi, G. I. Vujovich, and G. A. and Planetary Sciences, v. 36, p. 289–324,
Viñao, 2008, Metamorphic evolution of migmatites from Ramos, V. A., 2009, Anatomy and global context of the
the deep Famatinian arc crust exposed in Sierras Valle ­Andes: Main geologic features and the Andean orogenic
Fértil e La Huerta, San Juan, Argentina: Journal of South cycle, in S. M. Kay, V. A. Ramos, and W. Dickinson, eds.,
American Earth Sciences, v. 25, p. 313–335. Backbone of the Americas: Shallow subduction, plateau

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 32 11/05/18 12:18 PM


Tectonic Evolution of the Central Andes: From Terrane Accretion to Crustal Delamination  33

uplift, and ridge and terrane collision: Geological Society ­ lmos y Pomahuaca: Lima, Instituto Geología, Min-
O
of America Memoir 204, p. 31–65. ería y Metalogenia, Boletín Serie A: Carta Geológica
Ramos, V. A., 2010a, The tectonic regime along the ­Andes: ­Nacional, v. 39, 83 p.
Present settings as a key for the Mesozoic regimes: Riel, N. et al., 2013, A metamorphic and geochronogical
­Geological Journal, v. 45, p. 2–25. study of the Triassic El Oro metamorphic Complex in Ec-
Ramos, V. A., 2010b, The Grenville-age basement of the uador: Implications for high-temperature metamorphism
­Andes: Journal of South American Earth Sciences, v. 29, in a forearc zone: Lithos, v. 156–159, p. 41–68.
no. 1, p. 77–91. Riel, N., J.-E. Martelat, S. Guillot, E. Jaillard, P. Monié,
Ramos, V. A., 2010c, La free-board hypothesis versus una J. Yuquilema, G. Duclaux, and J. Mercier, 2015, Forearc
colisión eopaleozoica: el ejemplo del norte del Perú, in tectonothermal evolution of the El Oro metamorphic
15th Congreso Peruano de Geología, Sociedad Geológica province (Ecuador) during the Mesozoic: Tectonics,
del Perú, Cusco, Resúmenes Extendidos, Publicación Es- v. 680, p. 174–191.
pecial 9, p. 71–74. Rodrigues, J., J. Acosta, R. Castroviejo, J. Quispe, D. Romero,
Ramos, V. A., 2018, The Famatinian orogen along the pro- R. Uribe, M. Campián, 2010a, Geología y estructura de
tomargin of Western Gondwana: Evidence for a nearly las ultramafitas de Tapo y Acobamba (Tarma, Perú),
continuous Ordovician magmatic arc between Venezuela ­removilización tectónica andina de un segmento ofi-
and Argentina, in Folguera et al., eds., The evolution of olítico pre-andino: Sociedad Geológica del Perú, Publi-
the Chilean‐Argentinean Andes, Chapter 6: Springer cación ­Especial, v. 9, p. 79–82.
Earth System Sciences, p. 154–184. Rodrigues, J., J. Acosta, J. Macharé, E. Pereira, R. Castroviejo,
Ramos, V. A., and A. Alemán, 2000, Tectonic evolution of the 2010b, Evidencias estructurales de aloctonía de los cuer-
Andes, in U. J. Cordani, E. J. Milani, A. Thomaz Filho, and pos ultramáficos y máficos de la Cordillera Oriental del
D. A. Campos, eds., Tectonic evolution of South America: Perú en la región de Huánuco: Sociedad Geológica del
31st. International Geological Congress, Río de Janeiro, Perú, Publicación Especial, v. 9, p. 75–78.
p. 635–685. Roeder, D., and R. L. Chamberlain, 1995, Structural geol-
Ramos, V. A., and N. Jiménez, 2014, Extensión oriental del ogy of Sub-Andean fold and thrust belt in northwestern
macizo de Arequipa en los Andes bolivianos: Sus im- ­Bolivia, in A. J. Tankard, S. R. Suárez, and H. J. Welsink,
plicancias tectónicas, in Simposio Tectónica preandina, eds., Petroleum basins of South America: AAPG Memoir
Actas 19° Congreso Geológico Argentino, Córdoba, v. 62, p. 459–479.
S21–50, p. 2. Romero, D., K. Valencia, P. Alarcón, D. Peña, and V. A.
Ramos, V. A., G. I. Vujovich, and R. D. Dallmeyer, 1996, Los ­Ramos, 2013, The offshore basement of Perú: Evidence
klippes y ventanas tectónicas de la estructura preándica for different igneous and metamorphic domains in the
de la Sierra de Pie de Palo (San Juan): edad e implica- forearc: Journal of South American Earth Sciences, v. 42,
ciones tectónicas, in Actas 13rd Congreso Geológico Ar- p. 47–60.
gentino and 3rd Congreso Exploración de Hidrocarburos, Romero Fernández, D., K. Valencia, P. Alarcón, and V. A. Ra-
Buenos Aires, v. 5, p. 377–392. mos, 2011, Geología de la costa pacífica del Perú central
Reimann, C. R., H. Bahlburg, E. Kooijman, J. Berndt, A. Ger- entre Chiclayo y Paracas (7°214° sur), in Memorias 14th
des, V. Carlotto, and S. López, 2010, Geodynamic evolu- Congreso Latinoamericano de Geología and 13th Con-
tion of the early Paleozoic Western Gondwana margin greso Colombiano de Geología, Medellín, p. 110–111.
14°217°S reflected by the detritus of the Devonian and Rosas, S., L. Fontboté, and A. Tankard, 2007, Tectonic evolu-
Ordovician basins of southern Peru and northern Bolivia: tion and paleogeography of the Mesozoic Pucará Basin,
Gondwana Research, v. 18, p. 370–384. central Peru: Journal of South American Earth Sciences,
Reimann, C. R., H. Bahlburg, V. Carlotto, E. Kooijman, v. 24, no. 1, p. 1–24.
F. Boekhout, J. Berndt, and S. López, 2015, Multi-method Sánchez, J., O. Palacios, T. Feininger, V. Carlotto, and
provenance model for early Paleozoic sedimentary basins L. Quispesivana, 2006, Puesta en evidencia de grani-
of southern Peru and northern Bolivia (13°218°S): Jour- toides Triásicos en los Amotapes-Tahuín: Deflexión de
nal of South American Earth Sciences, v. 64, p. 94–115. Huancabamba, in 13th Congreso Peruano de Geología,
Reitsma, M. J., 2012, Reconstructing the Late Paleozoic-Early Resúmenes Extendidos: Lima, Perú, Sociedad Geológica
Mesozoic plutonic and sedimentary record of southeast del Perú, p. 312–315.
Peru: Orphaned back-arcs along the western margin of Sandeman, H. A., A. H. Clark, and E. Farrar, 1995, An inte-
Gondwana, Terre & Environment, v. 11, p. 1–228. grated tectono-magmatic model for the evolution of the
Restrepo-Pace, P. A., J. Ruiz, G. Gehrels, and M. Cosca, 1997, southern Peruvian Andes (13220°S) since 55 Ma: Inter-
Geochronology and Nd isotopic data of Grenville-age national Geology Review, v. 37, p. 1039–1073.
rocks in the Colombian Andes: New constraints for Late Sandeman, H. A., A. H. Clark, E. Farrar, and G. Arroyo
Proterozoic–early Paleozoic paleocontinental reconstruc- Pauca, 1997, Lithostratigraphy, petrology and 40Ar-39Ar
tions of the Americas: Earth and Planetary Science Let- geochronology of the Crucero Supergroup, Puno Depart-
ters, v. 150, p. 427–441. ment, SE Peru: Journal South American Earth Sciences,
Reyes, L., and J. Caldas, 1987, Geología de los cuadrán- v. 10, no. 3–4, p. 223–245.
gulos de Las Playas, La Tina, Las Lomas, Ayabaca, Sato, A. M., E. J. Llambías, M. A. S. Basei, C. E. Castro, 2015,
San Antonio, Chulucanas, Morropón, Huancabamba, Three stages in the Late Paleozoic to Triassic magmatism

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 33 11/05/18 12:18 PM


34 Ramos

of southwestern Gondwana, and the relationships with Suárez Soruco, R., 1999, Bolivia: Late Proterozoic—Early
the volcanogenic events in coeval basins: Journal of South Paleozoic, in 4 th International Symposium on Andean
American Earth Sciences, v. 63, p. 48–69. Geodynamics, Goettingen, p. 716–718.
Schurr, B., A. Rietbrock, G. Asch, R. Kind, and O. Oncken, Suárez Soruco, R., 2000, Compendio de geología de B ­ olivia:
2006, Evidence for lithospheric detachment in the Central Revista Técnica de Yacimientos Petrolíferos Fiscales
Andes from local earthquake tomography: Tectonophys- ­Bolivianos, v. 18, p. 1–166.
ics, v. 415, no. 1–4, p. 203–223. Tankard, A. J., 1986, On the depositional response to thrusting
Sempere, T., 1994, Kimmeridgian? To Paleocene tectonic and lithosphere flexure: Example from the ­Appalachian
evolution of Bolivia, in J. A. Salfity, ed., Cretaceous tccton- and Rocky Mountain basins, in P. A. Allen, and P. Home-
ics in the Andcs: Wiesbaden, Vieweg, p. 168–212. wood, eds., Foreland basins: International ­Association of
Sempere, T., 1995, Phanerozoic evolution of Bolivia and adja- Sedimentologists, Special Publication 8, p. 369–392.
cent regions, in A. Tankard, S. R. Suárez, and H. Welsink, Timoteo, D., D. Romero, and K. Valencia, 2012, Posible ex-
eds., Petroleum basins of South America: AAPG Memoir istencia de la Faja Alleghanide? en el Noroeste del Perú:
62, p. 207–230. Recientes dataciones Ar/Ar y U-Pb, e implicancias en
Sempere, T., G. Carlier, V. Carlotto, J. Jacay, N. Jiménez, un Sistema Petrolero de la Cuenca Talara: Sociedad
S. Rosas, P. Soler, J. Cardenas, and N. Boudesseu, ­Geológica del Perú, 16th Congreso Peruano de Geología.
1999, Late Permian-early Mesozoic rifts in Peru and Lima, (digital files).
­B olivia, and their bearing on Andean-age tectonics, in Tosdal, R. M., 1996, The Amazon-Laurentian connection as
4th Symposium on Andean Geodynamic, Goettingen, viewed from the Middle Proterozoic rocks in the Central
p. 680–685. Andes, western Bolivia and northern Chile: Tectonics,
Sempere, T., J. Jacay, A. Pino, H. Bertrand, V. Carlotto, v. 15, no. 4, p. 827–842.
M. Fornary, R. García, N. Jiménez, A. Marzoli, C. A. Troëng, B., E. Soria, H. Claure, R. Mobarec, and F. Murillo,
­Meter, S. Rosas, and P. Soler, 2004, Estiramiento litosférico 1994, Descubrimiento de Basamento Precámbrico en la Cor-
del Paleozoico superior al Cretáceo medio en el Perú y dillera Occidental Altiplano de los Andes Bolivianos, in Ac-
Bolivia: Sociedad Geológica del Perú, Publicación Espe- tas 11th Congreso Geológico de Bolivia, La Paz, p. 231–236.
cial, v. 5, p. 45–79. Vicente, J. C., 1989, Early late Cretaceous overthrusting in the
Sempere, T. et al., 2002, Late Permian-Middle Jurassic lith- Western Cordillera of southern Peru, in G. E. Ericksen, M.
ospheric thinning in Peru and Bolivia, and its bearing on Cañas, and J. A. Reinemund, eds., Circum-Pacific Council
Andean-age tectonics: Tectonophysics, v. 345, no. 124, for energy and mineral resources: Houston, Earth Science
p. 153–181. Series 11, p. 91–117.
Sobolev, S. V., and A. Y. Babeyko, 2005, What drives orogeny Vinasco, C., 2004, Evolucao Crustal e História Tectònica Dos
in the Andes?: Geology, v. 33, no. 8, p. 617–620. Granitoides Permo-Triásicos Dos Andes Do Norte: Ph.D.
Soler, P., and M. Bonhomme, 1990, Relations of magmatic thesis, Universidade de São Paulo, São Paulo, 121 p.
activity top late dynamics in central Peru from Late Cre- von Huene, R., I. A. Pecher, and M. A. Gutscher, 1996, Devel-
taceous to present, in S. Kay, and C. W. Rapela, eds., Plu- opment of the accretionary prism along Perú and mate-
tonism from Antarctica to Alaska: Geological Society of rial flux after subduction of Nazca Ridge: Tectonics, v. 15,
America Special Paper 241, p. 173–191. no. 1, p. 19–33.
Spikings, R., R. Cochrane, D. Villagomez, R. Van der Lelij, Wasteneys, A. H., A. H. Clark, E. Farrar, and R. J.
C. Vallejo, W. Winkler, and B. Beate, 2015, The geological L a n g r i d g e , 1 9 9 5 , G re n v i l l i a n g r a n u l i t e - f a c i e s
history of northwestern South America: From Pangaea metamorphism in the Arequipa massif, Peru: A Lau-
to the early collision of the Caribbean Large ­I gneous rentia-Gondwana link: Earth Planetary and Science
Province ( 290275 Ma ): Gondwana Research, v. 27, Letters, v. 132, p. 63–73.
p. 95–139. Willner, A. P., C. C. G. Tassinari, J. F. Rodrigues, J. Acosta,
Spikings R., Reitsma, M.J., Boekhout, F., Miškovic’, A., R. Castroviejo, and M. Rivera, 2014, Contrasting Ordovi-
­U lianov, A., Chiaradia, M., Gerdes, A., and Schalteg- cian high- and low-pressure metamorphism related to a
ger, U., 2016, Characterisation of Triassic rifting in Peru microcontinent-arc collision in the Eastern Cordillera of
and implications for the early disassembly of western Perú (Tarma province): Journal of South American Earth
­Pangaea. Gondwana Research, v. 35, p. 124–143. Sciences, v. 54, p. 71–81.
Starck, D., 1995, Silurian Jurassic stratigraphy and basin Witt, C., M. Rivadeneira, M. Poujol, D. Barba, D. Beida,
evolution of Northwestern Argentina, in A. J. Tankard, G. Beseme, and G. Montenegro, 2017, Tracking ancient
R. Suárez S., and H. J. Welsink, eds., Petroleum basins magmatism and Cenozoic topographic growth within the
of South America: American Association of Petroleum Northern Andes forearc: Constraints from detrital U-Pb
­Geologists Memoir 62, p. 251–267. zircon ages: Geological Society of America Bulletin (on-
Steinmann, G., 1929, Geologie von Peru, Karl Winter, line), DOI: 10.1130/B31530.1.
­Heidelberg, 448 p. Wörner, G., J. Lezaun, J. Beck, V. Heber, F. Lucassen,
Steinmann, G., 1930, Geología del Perú, Karl Winter, E. Zinngrebe, R. Rößling, and H. G. Wilke, 2000, Pre-
­Heidelberg, 251 p. cambrian and early Paleozoic evolution of the Andean
Steinmann, G., R. Stappenbeck, A. H. Sieberg, and basement at Belen (northern Chile) and Cerro Uyarani
C. I. Lissón, 1930, Geología del Perú, Karl Winter, (western Bolivia Altiplano): Journal of South American
Universitätsbuchhandlung, Heidelberg, 448 p. Earth Sciences, v. 13, p. 717–737.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/4658069/13972_ch01_ptg01_001-034.pdf


by guest
on 09 May 2020

13972_ch01_ptg01_001-034.indd 34 11/05/18 12:18 PM

You might also like