Buscar

Tavares_JSAES_2018

Prévia do material em texto

Contents lists available at ScienceDirect
Journal of South American Earth Sciences
journal homepage: www.elsevier.com/locate/jsames
The multistage tectonic evolution of the northeastern Carajás Province,
Amazonian Craton, Brazil: Revealing complex structural patterns
Felipe Mattos Tavaresa,b,∗, Rudolph Allard Johannes Trouwb, Cíntia Maria Gaia da Silvaa,
Ana Paula Justoa,c, Junny Kyley Mastop Oliveiraa
aGeological Survey of Brazil (CPRM/SGB), Brazil
b Instituto de Geociências, Universidade Federal do Rio de Janeiro (IGEO – UFRJ), Brazil
c Instituto de Geociências, Universidade de Brasília (IG – UnB), Brazil
A R T I C L E I N F O
Keywords:
Carajás
Amazonian Craton
Structural geology
Tectonic evolution
A B S T R A C T
Structural data collected in the Carajás Province region led to a new interpretation of the southeastern
Amazonian Craton geotectonic evolution. The purpose of this article is to present a new evolutionary proposal
for the region. A detailed analysis of the several extensional-compressional cycles that overprinted each other
from the Archean to the Neoproterozoic-Cambrian is presented. At about 2.87–2.83 Ga, collisional processes led
to the formation of a stable crustal substrate that supported the installation of an extensional basin at
2.76–2.70 Ga and the deposition of the Itacaiúnas Supergroup shallow marine volcanosedimentary sequences,
together with contemporary bimodal plutonism. Paleoproterozoic arc magmatism in the Bacajá Domain was
followed by collision with the Carajás Province between 2.09 and 2.06 Ga, resulting in expressive tectonic
thickening and low to high grade regional metamorphism, and in the deposition of the Águas Claras Formation.
A second Paleoproterozoic orogenic event affected the Carajás Province, which resulted in oblique tectonism and
regional counterclockwise rotation of previous associations, followed by late to post-orogenic sedimentation and
1.88 Ga anorogenic alkaline A-type magmatism. The eastern Carajás Province margin was extensionally re-
activated during the Neoproterozoic, in a rifting event, followed by tectonic inversion during the Ediacaran/
Cambrian.
1. Introduction
The accumulation of tectonic processes that led to the formation of
modern cratons is expressed by the structural complexity of many of
their Archean nuclei. The successive reworking of the margins of these
crustal blocks resulted in notable tectonic overprint fabrics, for ex-
ample, in the Yilgarn craton, Australia (e.g. Swager, 1997; Cawood and
Korsh, 2008), in the Bastar craton, peninsular India (e.g. Bhadra et al.,
2004), in the North China craton, eastern Asia (e.g. Kusky and Li, 2003;
Zhao et al., 2005; Zhang et al., 2009) in the Slave and Rae cratons,
Canada (e.g. Hoffman, 1988; Kusky, 1990; Kusky et al., 2014), or in the
São Francisco Craton, Brazil (e.g. Baltazar and Zuchetti, 2007).
Although the timing when plate tectonics began is still a con-
troversial matter (Condie and Pease, 2008), it is possible that it was
active since the Archean or even before, though episodic and not ne-
cessarily in a strictly uniformitarian way (e.g. Brown, 2006; Hopkins
et al., 2008; Korenga, 2013). It is also expected that repeated tectonic
reactivations, related to extensional-compressional processes of
different age and evolutionary context were superposed in regions
previously affected by Archean tectonics. Olsson et al. (2010), for ex-
ample, showed evidence of superposition of three archean-paleopro-
terozoic taphrogenic events in the Kaapvaal craton, while Kusky et al.
(2014) suggest that events correlated to the Wilson Cycle already oc-
curred before 2.5 Ga in northern China and Slave cratons.
Carajás Province (CP), in the southeastern Amazonian Craton,
Brazil, is an Archean crustal segment (Cordani et al., 1984; Teixeira
et al., 1989; Santos, 2003), located in a region until recently fully
covered by dense tropical forest. Although it is one of the largest me-
talliferous provinces in the world, with giant iron ore deposits and
many others of copper-gold, copper-zinc, manganese, nickel, REE and
PGE, available geological maps for the region are mostly on 1:250,000
and 1:1,000,000 scales (Araújo and Maia, 1991; Oliveira et al., 1994;
Vasquez et al., 2008a), which restricts basic knowledge to the regional
recognition level.
Available tectonic models for the northern part of CP are quite
simplified, based on regional observations and/or on localized field
https://doi.org/10.1016/j.jsames.2018.08.024
Received 11 April 2018; Received in revised form 30 August 2018; Accepted 31 August 2018
∗ Corresponding author. Geological Survey of Brazil (CPRM/SGB), Rio de Janeiro office, 404 Pasteur avenue, Urca, Rio de Janeiro, 22290-255, Brazil.
E-mail address: felipe.tavares@cprm.gov.br (F.M. Tavares).
Journal of South American Earth Sciences 88 (2018) 238–252
Available online 01 September 2018
0895-9811/ © 2018 Elsevier Ltd. All rights reserved.
T
http://www.sciencedirect.com/science/journal/08959811
https://www.elsevier.com/locate/jsames
https://doi.org/10.1016/j.jsames.2018.08.024
https://doi.org/10.1016/j.jsames.2018.08.024
mailto:felipe.tavares@cprm.gov.br
https://doi.org/10.1016/j.jsames.2018.08.024
http://crossmark.crossref.org/dialog/?doi=10.1016/j.jsames.2018.08.024&domain=pdf
data. They refer to a hypothetical evolutionary model of large in-
tracontinental strike-slip faults, subjected only to minor reactivation for
up to 900 million years between the Neoarchean and Paleoproterozoic
(Araújo et al., 1988; Costa et al., 1995; Pinheiro and Holdsworth, 1995,
1997a; 1997b, 2000; Holdsworth and Pinheiro, 2000).
Recent systematic geological mapping at scale 1:100,000, like
Tavares and Silva (2012) and others undertaken by the Geological
Survey of Brazil have reactivated the debate around the tectonostrati-
graphic evolution of CP's northeastern portion, evidencing tectonic
superposition not recognized in previous models. This article aims to
present a new evolutionary proposal for the region, based on new
structural data and supported by a critical review of CP's tectonos-
tratigraphy. We contribute to the scientific discussion of the extensional
and compressional (orogenic) cycles that overprinted each other from
the Archean to the Neoproterozoic-Cambrian.
2. Regional tectonostratigraphic context
The Amazonian Craton is a large crustal portion mostly constituted
and structured between the Archean and the Mesoproterozoic, tecto-
nically stabilized at around 1.0 Ga (Brito Neves and Cordani, 1991).
According to Vasquez et al. (2008a), based on the model of Santos
(2003), the craton's eastern portion is constituted by Carajás and
Transamazonas tectonic/geochronological provinces (Fig. 1), the first
representing a mostly Archean cratonic nucleus and the second a Pa-
leoproterozoic collage.
2.1. Carajás Province (CP)
Santos (2003) divided the CP into two tectonostratigraphic do-
mains, Rio Maria, to the south and Carajás, to the north. The Carajás
Domain presents Mesoarchean basement (pre-2.86 Ga), as well as
Neoarchean meta-volcanosedimentary cover and intrusions (2.76 –2.71
Ga e.g. Machado et al., 1991; Barros et al., 2004; Sardinha et al. 2006,
Feio et al., 2012). Detrital zircon ages up to 3.6 Ga and TDMs greater
than 3.20 Ga for some of the Carajás Domain units suggest Pa-
leoarchean contribution of its basement composition, while Rio Maria
Domain is composed of Mesoarchean juvenile crust (Mougeot et al.,
1996a; Macambira et al., 2001; Galarza and Macambira, 2002;
Dall'Agnol et al., 2005).
The known basement of the Carajás Domain is composed of infra-
crustal mafic granulite (Xicrim-Cateté Complex, Vasquez et al., 2008a)
and of many different lithotypes that are usually grouped into the Xingu
Complex. These are here informally subdivided into three different
Fig. 1. Tectonostratigraphic map of the southeastern Amazonian Craton, showing Rio Maria and Carajás Domains (Carajás Province), southern Bacajá Domain
(Transamazonas Province) and younger assemblages (Orosirianunits, to the west, belong to the Iriri-Xingu Domain, Central Amazonian Province, while
Neoproterozoic units, to the east, are part of the Araguaia Belt, Tocantins Province, delimiting the eastern cratonic boundary). The study area is marked by the dashed
polygon. Modified from Vásquez et al. (2008a), also including new data from Tavares and Silva (2012) and Feio et al. (2013).
F.M. Tavares et al. Journal of South American Earth Sciences 88 (2018) 238–252
239
associations: migmatitic orthogneisses of granodioritic to tonalitic
composition (Bom Jesus Orthogneiss, reinterpreted after Feio et al.,
2013), gneissified granitoids (Canaã dos Carajás, Campina Verde,
Cruzadão and Serra Dourada plutons, Feio et al., 2013, Moreto et al.,
2014) and lenticular greenstone fragments (Rio Novo Group, Hirata
et al., 1982; Sapucaia Group, Araújo & Maia, 1991).
The agglutination of the Carajás Domain basement units is related to
accretion and collision involving portions of consolidated continental
crust, in the intervals of 3.08–2.93 Ga and 2.87–2.83 Ga (Feio et al.,
2013; Silva, 2014). Moreto et al. (2014) and Silva (2014) presented
crystallization ages between 3.08 and 3.00 Ga for TTG orthogneisses of
the Xingu Complex (U-Pb, zircon), representative of the Bom Jesus
Orthogneiss. Moreto et al. (2014) also dated greenstone-related sub-
volcanic rocks with crystallization ages of around 2.97 Ga (U-Pb,
zircon), synchronous to the crystallization of Canaã dos Carajás Meta-
granite, of calc-alkaline affinity (Feio et al., 2013). These assemblages
were intruded by calc-alkaline to alkaline granitoids between 2.83 and
2.87 Ga (Campina Verde Metatonalite, Cruzadão Metagranite, Serra
Dourada Granite), during an apparent collisional event between Rio
Maria terrane and a supposed “Carajás paleocontinent”. Collision was
accompanied by regional high-grade metamorphism and migmatiza-
tion, with its peak dated at 2.86 Ga (Machado et al., 1991).
The basement units are covered by Neoarchean volcanosedimentary
rocks of the Itacaiúnas Supergroup (DOCEGEO, 1988). This unit can be
informally subdivided into lower, intermediate and upper associations.
The lower association is dominantly marked by volcanism dated be-
tween 2.76 and 2.73 Ga (U-Pb, zircon, e.g. Gibbs et al., 1986, Machado
et al., 1991, Trendall et al., 1998). The intermediate association is re-
presented by thick layers of banded iron formation, locally interlayered
with black shales (Cabral et al., 2013) or graphite schists, according to
the metamorphic grade. The upper association is clastic-sedimentary,
with (meta)sandstone/quartzite, (meta)conglomerates and (meta)pe-
lites, accompanied by rare levels of banded iron formation and pyrite-
rich black shales/graphite shists, as well as volcanic and volcanoclastic
rocks.
Neoarchean plutonism shows bimodal pattern. Mafic-ultramafic
bodies like the Luanga layered complex (e.g. Ferreira Filho et al., 2007)
with ages between 2.76 and 2.74 Ga intrude the basement units and the
Itacaiúnas Supergroup lower association. Similarly, many A-type
granites were emplaced between 2.76 and 2.70 Ga, like Estrela (Barros
et al., 2001, 2004), Gelado (Barbosa, 2004) and other plutons corre-
lated to the Planalto Suite (Feio et al., 2012). Moreover, late A-type
granitic magmatism is also recognized in the northern portion of the
Carajás Domain, represented by the Old Salobo Granite (2573 ± 7Ma
– Machado et al., 1991; 2547 ± 5.3Ma – Melo et al., 2013).
The Águas Claras Formation (Araújo and Maia, 1991; Nogueira
et al., 1995) represents a thick sedimentary cover in the central-eastern
part of the Carajás Domain, unconformably overlaying the Itacaiúnas
Supergroup. It consists of pelites, arenites and subordinated carbonatic
rocks of shallow marine environment at the base, followed by a thick
package of fluvial sandstones and conglomerates at the top. Its de-
positional age is uncertain, as well as whether it represents one or more
different sequences. U-Pb ages between 2.65 and 2.70 Ga for zircons
extracted from mafic to intermediate sills and dykes that crosscut the
unit were initially understood as representing their crystallization ages,
limiting maximum depositional age of Águas Claras Formation to the
Neoarchean (Mougeot et al., 1996a; Dias et al., 1996; Trendall et al.,
1998). However, Mougeot et al. (1996b) reported a 2.06 Ga Pb-Pb age
for the unit's diagenetic pyrite. Also, non-typical MIF (mass-in-
dependent fractionation) signature of early-diagenetic pyrite supports
their paleoproterozoic formation (Fabre et al. (2011). These data sug-
gest that Águas Claras Formation was deposited after 2.10 Ga, leading
to the reinterpretation of zircon found in dikes as inherited or re-
presentative of crustal contamination.
When observed on a regional scale, the Carajás Domain has an
elongated shape of E-W direction, marked by the alignment of several
structures (including distinct tectonic styles) and by the structurally
controlled distribution of the main units. This feature was defined by
Araújo et al. (1988) as the Itacaiúnas Belt and, for those authors, would
be the southern margin of a hypothetical collisional orogeny, affecting
TP's associations as well, which would have occurred during the
Neoarchean-Paleoproterozoic transition, and presenting late reactiva-
tions up to hundreds of millions of years after the main deformational
stage. This hypothesis is also supported by Costa et al. (1995) and
Pinheiro and Holdsworth (2000). Nevertheless, none of these authors
discussed in detail the origin, timing, chronostratigraphic positioning or
evolutionary context of this supposed collisional system or of the re-
activations.
Cordani et al. (1984), on the other hand, suggested through K-Ar
and Rb-Sr geochronology that CP-TP boundary (or Central Amazonia/
Maroni-Itacaiunas provinces boundary, for those authors) was a feature
related to the Transamazonian Orogeny (2.2–2.0 Ga), much younger
than the supposed age of the Itacaiúnas Belt. More recently, other au-
thors followed the same interpretation, using zircon U-Pb geochro-
nology (e.g. Macambira et al., 2007). Considering that, the Itacaiunas
Belt should be either an old feature truncated by Paleoproterozoic
structures, or a Paleoproterozoic feature related to the Transamazonian
Orogeny, or, more likely, a multistage feature affected by both Archean
and Paleoproterozoic deformation.
2.2. Transamazonas Province (TP)
To the north of the CP, Cordani et al. (1984) recognized the pre-
dominance of rocks reworked during the Transamazonian Orogeny
(Paleoproterozoic). Teixeira et al. (1989) also identified those assem-
blages as belonging to this event and included them into the Maroni-
Itacaiúnas Province, while Santos (2003) incorporated them into the
Transamazonas Province (TP), nomenclature also adopted by Vasquez
et al. (2008a) and in this paper.
The Bacajá Domain is the TP segment that crops out on CP's
northern limits, consisting of Archean-Paleoproterozoic terranes and by
large portions of Paleoproterozoic juvenile crust, as a result of Rhyacian
accretionary-collisional processes (Macambira et al., 2007, 2009;
Faraco et al., 2004; Vasquez et al., 2008b). The rocks in the contact
zone are granulites/retrogranulites from the Cajazeiras Complex
(Vasquez et al., 2008a) and Vila Santa Fé Complex (Tavares and Silva,
2012). The first includes mainly tonalitic to granitic orthogneisses,
strongly re-hydrated, which represent the most evolved and exhumed
Bacajá Domain infracrust, at least in part with Mesoarchean protoliths
(∼3.0 Ga). The second represents a younger crustal segment, which
includes Rhyacian calk-alkaline associations thought to be re-
presentative of a continental magmatic arc on the margin of the Bacajá
paleocontinent.
2.3. Uatumã Magmatism and contemporary assemblages
Orosirian sedimentary covers and intrusions occur widespread in
the CP and TP. Sedimentary successions from the Paredão Group
(Oliveira et al., 1994) and the Caninana Formation (reinterpreted after
Pereira etal., 2009) cover parts of the Carajás and Bacajá domains in
the study area, containing sedimentary breccias, polimictic to oligo-
mictic conglomerates and poorly selected, immature sandstones. These
sequences are interpreted as representing continental alluvial fans and
braided fluvial environments.
Moreover, several anorogenic intrusions correlated to the Uatumã
Magmatism (Silva et al., 1974) are distributed in the study area, re-
presented by alkaline A-type granites of 1.88 Ga and by felsic dikes
oriented in NNE-SSW and NW-SE directions, grouped in the Serra dos
Carajás Intrusive Suite (Dall'Agnol et al., 2005), as well as by acidic
tuffs interbedded at the top of the Paredão Group (Tavares and Silva,
2012).
Detrital zircon analyzed by Pereira et al. (2009) indicated a
F.M. Tavares et al. Journal of South American Earth Sciences 88 (2018) 238–252
240
maximum age of 2.01 Ga for the Caninana Formation, suggesting that
the sedimentation of these sequences may have started before the suite's
emplacement.
2.4. Eastern Amazonian Craton boundary
The Neoproterozoic N-S Araguaia orogenic belt limits the eastern
Amazonian Craton and the Parnaíba Block (Cordani et al., 1984; Castro
et al., 2014). Mafic sills (Rio da Onça Gabbro, Tavares & Silva, 2012)
and sheeted dikes oriented in NNW-SSE direction occur along the
eastern Amazonian Craton margin and were interpreted as Neoproter-
ozoic by DOCEGEO (1988). Furthermore, the shallow marine sedi-
mentary cover from the Baixo Araguaia Group (Hasui et al., 1977),
onlaps and eventually thrusts CP and TP associations to the east of the
study area.
3. Structural geology
The stratigraphic relations and geological map of the study area are
shown on Figs. 2 and 3, and several cross sections are presented on
Fig. 4. A great variety of ductile and brittle structures were recognized.
The many different structures that affect the region can be subdivided
in four compressional deformation phases (D1 to D4), and three main
extensional episodes (post-D1, post-D3 and pre-D4) on a regional scale
and in accordance with spatial distribution, orientation, overprinting
relations and tectonic style criteria.
3.1. D1 structures
The first group (D1) affects only the Mesoarchean basement units
(Bom Jesus Orthogneiss, Rio Novo and Sapucaia groups, Cruzadão,
Fig. 2. Schematic stratigraphic chart of the study area. Color patterns are the same as on the geological map (Fig. 3) while tectonic regime and regional meta-
morphism columns show events as related in this work. (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of
this article.)
F.M. Tavares et al. Journal of South American Earth Sciences 88 (2018) 238–252
241
Fig. 3. Geological map of the study area with the main structures. Geographic coordinate system: SIRGAS 2000.
F.M. Tavares et al. Journal of South American Earth Sciences 88 (2018) 238–252
242
Campina Verde and Serra Dourada granitoids). Its main elements are
ductile: continuous schistosity (S1), stretching and/or mineral lineation
(L1), folds, L-tectonites and medium to high grade mylonites (with fully
recrystalyzed quartz by grain boundary migration), the last related to
high angle reverse shear zones. D1 structures were developed under
moderate to high grade metamorphic conditions, in lower to upper
amphibolite facies.
The S1 foliation is a penetrative feature recognized in practically all
rocks affected by D1 to the south of the Cinzento Lineament (Bom Jesus
Orthogneiss, Rio Novo and Sapucaia groups, Cruzadão and Campina
Verde granitoids). To the North, S1 foliation is mostly overprinted by
D2 structures. It was described in the field as a fine to coarse schistosity,
depending on the lithotype (Fig. 5a) and in thin sections as a continuous
schistosity. However, some greenstone belts sedimentary and ultra-
mafic shists also present locally interbedded folds, suggesting that S1 is
a transposition foliation in those lithotypes, preceded by a pre-to early-
D1 deformational stage.
S1 measurements show some spreading in stereoplots (Fig. 6), re-
flecting intense reworking during later tectonic events. Data collected
in the southwestern portion of the study area are apparently less af-
fected by D2 and/or D3, dipping steeply to south-southwest. In other
areas, they tend to follow the direction of younger structures.
L1 was largely reoriented or transposed by D2/D3; however it was
possible to recognize it on some S1 planes, formed by oriented meta-
morphic minerals and/or by ductile stretching, especially in or near D1
shear zones, in Bom Jesus orthogneiss and in Cruzadão/Campina Verde
granitoids. Despite later reworking, L1 tends to plunge down dip to
slightly oblique in relation to S1 planes (Fig. 6). D1 folds are generally
tight to isoclinal, symmetrical and with axial planes parallel to S1. Fold
axes have variable orientation, reflecting post-D1 deformation.
D1 shear zones are reverse and of high angle, aligned in E-W di-
rection when not affected by later deformation, presenting medium to
Fig. 4. Geological cross-sections of the study area. Color patterns and structures are the same as on the geological map (Fig. 3). (For interpretation of the references to
colour in this figure legend, the reader is referred to the Web version of this article.)
F.M. Tavares et al. Journal of South American Earth Sciences 88 (2018) 238–252
243
high temperature mylonites. Mesoarchean metagranitoids (Cruzadão,
Campina Verde and Serra Dourada plutons) are syn-to late-tectonic in
relation to D1, spatially related to syn-D1 shear zones and mainly E-W
oriented (Fig. 3). When visible, magmatic foliation and S1 are parallel
in the plutons (Fig. 5b). Serra Dourada granite, the youngest pluton
from this association is, however, mostly isotropic.
It is likely that D1 structures were generated in a compressional
regime, with regional σ1 N-S oriented, considering the S1/L1 attitudes
and the reverse kinematics of the major shear zones. Tectonic transport
seems to be top-to-north, but this statement is not fully reliable due to
post-D1 structural reworking.
3.2. Post-D1 structures
The structures that conditioned the opening of the Carajás basin and
coeval magmatic and hydrothermal activities are post-D1/pre-D2 and
show brittle, extensional behavior. Very few structures related to this
event were observed, mainly cutting Itacaiúnas Supergroup lithotypes,
as the majority was obliterated by younger deformation. The exceptions
are the structures that host hydrothermal veins and Old Salobo-like
granitic dykes that are of late stage in relation to the basin develop-
ment, but related to magnetite-rich IOCG mineralizations, dated by
different authors between 2.7 and 2.5 Ga (Moreto et al., 2015 and
Fig. 5. a) S1 foliation in migmatitic orthogneiss from
Xingu Complex (Bom Jesus Orthogneiss;
49°50′1885″W, 6°27′4703″S); b) Campina Verde me-
tatonalite with mafic xenolith, presenting a magmatic
foliation parallel to S1 (49°51′32,899″W, 6°24′6958″S);
c) Neoarchean A-type-like metagranitic body showing
coarse S2 foliation (49°34′4042″W, 5°56′18,161″S); d)
S2a/S2b in Itacaiúnas Supergroup uppermost sequence
paraderivate lithotype (S2a: slaty cleavage; S2b: pro-
gressive crenulation cleavage; 50°14′30,636″W,
5°50′55,392″S); e) syn-D3 recumbent folds in Rio Novo
Group lithotypes (49°41′47,980″W, 5°50′34,490″S); f)
S3 crenulation in muscovite-graphite-quartz-schist of
Itacaiúnas Supergroup uppermost association
(49°32′37,011″W, 5°52′47,436″S); g) syn-D4 fracture
zone with brecciation and silicification
(49°30′6248″W, 5°59′14,622″S); h) syn-D4 reverse
faults affecting Paredão Group arenitic lithotypes
(49°29′13,481″W, 5°48′2700″S). Geographic co-
ordinate system: SIRGAS 2000.
F.M. Tavares et al. Journal of South American Earth Sciences 88 (2018) 238–252
244
references therein).
The late stage of post-D1 structural evolution is fluid-dominated,
with pervasive sodic and calc-sodic regional alterations. Syntaxial
veins,sometimes in stockwork, are also common, presenting hydro-
thermal assemblage usually composed of one or more of the following
minerals: hastingsite, quartz, albite, microcline, apatite, allanite, biotite
and tourmaline. The main oxide is magnetite, but ilmenite is also lo-
cally present, as well as the main sulfides chalcopyrite, pyrrhotite and
pyrite.
The faults that conditioned the opening of the basin have not been
directly observed, but the E-W elongated shape (Figs. 1 and 3) roughly
follows the previous tectonic trend (D1), suggesting its extensional re-
activation.
3.3. D2 structures
Structures attributed to D2 affect all Archean rocks in the study
area, representing the first ductile fabric to affect the Itacaiunas
Supergroup and other Neoarchean units. D2 also represents the main
tectonic features that delineate the contact between TP and CP and is
also recognized in the Rhyacian units of the Vila Santa Fé Complex. The
group is formed by S2 foliation, L2 stretching and/or mineral lineation,
folds and thrust/transpressional shear zones.
S2 is formed by amphibolitic to granulitic mineral fabrics to the
north and sub-greenschist to greenschist facies fabrics to the south,
dipping moderately to steeply to north-northeast when not reworked by
D3. The S2 foliation is either a continuous or a crenulation cleavage
Fig. 6. Equal area projection (lower hemisphere) stereoplots of structures from the different deformation phases, divided by D3-related domains. Planar data were
plotted as poles. Density isolines were calculated by cosine sums from zero to maximum, for plots with 25 or more data. D3 domains (I to V) and main structures are
named as described in this work.
F.M. Tavares et al. Journal of South American Earth Sciences 88 (2018) 238–252
245
(Fig. 5c and d). Granulites and retrogranulites from the Vila Santa Fé
Complex present S2 as a coarse, continuous to anastomosing schistosity.
In Neoarchean rocks (Itacaiúnas Supergroup, A-type granites, Mafic-
ultramafic intrusions), S2 varies from a fine slaty cleavage to a coarse
shistosity, depending on the affected lithotype and metamorphic grade,
progressively more intense to the north. In some outcrops of schistose
lithotypes from the Itacaiúnas Supergroup, in the northern sector, S2
was described as a crenulation cleavage; however, the crenulated fabric
was classified as early-D2, as it is formed by the same mineral assem-
blage that constitutes the crenulation cleavage (S2a/S2b relation). In
the southern Mesoarchean greenstone belt rocks (Rio Novo and Sapu-
caia groups), where D2 is less intense, S2 is a less penetrative crenu-
lation cleavage, marked by a retrograde mineral assemblage (chlorite,
sericite) and by flexural folding and dissolution of S1, with minor re-
crystallization of syn-D1 mineral fabrics. Mesoarchean gneisses from
the Xingu Complex that crop out in the northern sector usually show S1
folded and partially transposed by recrystallization to S2. Competent
lithotypes that crop out in the southern sector (Mesoarchean gneisses
and granites) usually do not present visible S2 features.
Stereoplots from the northern sector show very homogenous data,
with a peak of measurements dipping moderately to NNE and some
minor dispersion related to D3 deformation (Fig. 6). On the southern
sector, however, S2 measurements show considerable D3-related
spreading, with smooth peaks of measurements dipping steeply to SSE
or NNW.
L2 usually plunges down dip to slightly oblique in relation to S2 and
is mainly formed by preferred orientation of elongated metamorphic
minerals, such as amphiboles and quartz rods (Fig. 6). D2 folds are open
to isoclinal, with axial plane parallel to S2 and usually with axes
plunging smoothly to SE. In high strain zones, however, axes are par-
allel to L2.
The syn-D2 thrusts imbricate TP's Paleoproterozoic assemblages
over CP's Neoarchean rocks (Figs. 3 and 4a, b, c). They are associated
with mylonitic foliation and stretching lineation with up-dip kinematics
and amphibolite facies mineralogy. Locally, some shear zones present
reverse-sinistral movement. Mylonites present medium to high tem-
perature recrystallization mechanisms in the northern sector, such as
grain boundary migration, recrystallized quartz ribbons and grain size
reduction of hornblende by recrystallization. In the southern sector,
mylonite zones have sharper contacts and are of low to medium tem-
perature.
Considering the structural data presented in this section, D2 is un-
derstood as the result of a compressional (locally transpressive) regime,
with regional NE-SW to NNE-SSW oriented σ1 and with top-to-SW
tectonic transport.
3.4. D3 structures
These structures are present in the entire study area and affected all
Archean and Rhyacian units (Bom Jesus Orthogneiss, Rio Novo and
Sapucaia groups, Cruzadão, Campina Verde and Serra Dourada grani-
toids, Itacaiúnas Supegroup, Neoarchean A-type granites and mafic-
ultramafic intrusions, Vila Santa Fé Complex and Águas Claras
Formation). They are of ductile-brittle to ductile character, developed
in sub-greenschist facies conditions in the northwest, until upper
greenschist facies in the southeastern part of the area. D3 structures are
subdivided into five domains, according to the geographical distribu-
tion and regional structural arrangement (Fig. 6). In two of them, the
development of D3 structures is restricted and the continuity of earlier
tectonic fabrics is well preserved. The other three show different in-
tensity of reworking of the previous structures.
Domain I overlaps the region of occurrence of D2 structures, to the
north and northwest of the syn-D3 Sereno thrust front. Structural su-
perposition fabrics of D3 over D2 are recognizable in this domain, lo-
cally generating complex tectonic interference patterns. The S2 folia-
tion, although mostly parallel to its main regional direction (WNW-
ESE), presents asymmetric, normal to overturned open to tight D3 folds
(Fig. 5e), with axial planes dipping smoothly to steeply to south-
southeast and sub-horizontal fold axes. On the map scale, D3 folding led
to a slight counterclockwise rotation of older structures (Fig. 3). Some
less competent lithotypes from Rio Novo Group and Itacaiúnas Super-
group (micaceous schists) developed a weak S3 crenulation cleavage,
dipping smoothly to south-southeast (Fig. 6).
Dextral subvertical NW-SE strike-slip faults that crosscut Domain I
are also attributed to D3. These structures, with kilometric displace-
ment, dislocate tectonic contacts and locally rotate S2 in drag-like
patterns.
Domain II is localized between the Sereno thrust front to the north,
the Parauapebas fault zone to the west, and the Curionópolis thrust
front to the south, affecting CP's Meso-Neoarchean lithotypes and also
the Águas Claras Formation. Older tectonic fabrics were largely over-
printed by structures developed under lower greenschist facies meta-
morphic conditions. The S3 foliation dips on average 30° to south/
south-southeast, locally associated with L3 down dip stretching linea-
tion (Fig. 6). It appears with two distinct morphologies: in the Águas
Claras Formation lithotypes, it is a fine grained slaty cleavage (locally
anastomosed in coarse lithotypes), marked mainly by white mica; in the
Itacaiúnas Supergroup rocks and in ultramafic schists of the Luanga
Complex and Rio Novo Group, S3 occurs as an asymmetric crenulation
cleavage (Fig. 5f) associated with dissolution planes and sometimes
with oriented growth of chlorite or white mica. Where present, the S1
or S2 foliations were parallelized and/or refolded by D3. Close to D3
thrusts, the interference pattern between S2 and S3, together with
strong L3 streching lineation, produces a pencil-like structure in some
volcanosedimentary lithotypes.
The D3 folds are asymmetrical, open to tight and mostly overturned,
with axial planes dipping smoothly to moderately to south-southeast
and fold axes plunging either to east-northeast, or to west-southwest.
Their vergence isto north or northwest. The syn-D3 thrust faults are
usually coincident with D3 megafold limbs (Fig. 3), striking ENE-WSW
to E-W and dipping smoothly to steeply to south/south-southeast, lo-
cally associated with low temperature mylonitic foliation and strong L3
stretching lineation plunging down dip. This low angle set of faults,
here named as Serra Leste thrust system, imbricates Domain II asso-
ciations on top of Domain I, with tectonic transport to north-northwest
(Fig. 4e). Subvertical dextral oblique faults in the Domain II southern-
most portion, oriented NNW-SSE, present continuity with the thrust
fronts and are related to lateral movement between blocks. In the vi-
cinity of these structures, the S1 and S2 foliations follow the same strike
and dip steeply to west-southwest, overprinted by oblique L3 stretching
lineation, plunging smoothly to moderately to south-southeast.
Domain III corresponds to the region to the north of the Carajás fault
and to the south of the Parauapebas fault zone, where lithotypes belong
to the Itacaiúnas Supergroup, the Águas Claras Formation and, in the
far eastern portion, the Santa Inês Gabbro. The main structural features
are related to D2, however, they are strongly reoriented and reworked
by D3. The weak S2 foliation observed in some lithotypes of the
Itacaiúnas Supergroup, as well as D2 folds, present D3 refolding pat-
terns typical of shallow tectonics, like kink bands and box folds, with
average axial planes striking NE-SW and dipping moderately to north-
west or southeast. Altogether, S2 foliation describes an arc between the
westernmost portion of the study area and the Parauapebas River to the
east in an apparent syn-D3 mega-antiform with ENE-WSW to NE-SW
axial plane. On the southern limit of Domain III with Domain V, S2
apparently follows the strike of the Carajás Fault, dipping moderately to
south-southwest.
The Parauapebas fault zone is the limit between Domain III and
domains I and II. It is an anastomosing system of high angle reverse-
dextral oblique faults, oriented between NW-SE and E-W, associated
with an L3 stretching lineation with subhorizontal E-W to moderate SSE
plunge. At the northernmost limit of domain III, the truncation of the
structures and metamorphic grade with respect to Domain I can be
F.M. Tavares et al. Journal of South American Earth Sciences 88 (2018) 238–252
246
observed: to the north, portions of the Gelado metagranite and the
Itacaiúnas Supergroup are oriented WNW-ESE, with S2 dipping mod-
erately to steeply to east-northeast and showing amphibolite facies
metamorphic conditions; to the south, lithotypes of the Itacaiúnas
Supergroup are oriented ENE-WSW to N-S, dipping steeply to north-
northwest or west and metamorphosed under sub-greenschist facies
conditions (Fig. 3). This structure was interpreted as a ramp of a
shallow thrust system (Serra Norte thrust system), with inferred tec-
tonic transport to west-northwest (Fig. 4c).
Domain IV encompasses the southeastern portion of the study area,
limited with Domain II through the Curionópolis thrust system to the
north, and with Domain V, by a set of dextral-reverse oblique faults and
thrusts to the west. CP plutonic lithotypes are dominant, from both the
Mesoarchean and Neoarchean intrusive assemblages, associated with
relatively narrow imbricated segments of the Itacaiúnas Supergroup
lithotypes. D3 structures are dominant and developed under upper
greenschist facies metamorphic conditions. In Itacaiúnas Supergroup
lithotypes, S3 foliation is a tight crenulation cleavage, sometimes
transposed to a spaced schistosity, moderately to steeply dipping to
south or south-southeast. Plutonic lithotypes frequently present re-
folding and parallelization of S1 or S2 by D3. The L3 stretching linea-
tion, as in other areas, is regular and penetrative, plunging moderately
to southeast. D3 folds are normal, tight to isoclinal, with axial plane
dipping steeply to south-southeast. Fold axes plunge moderately to the
east or are parallel to L3 stretching lineation, especially in the vicinity
of D3 shear zones.
The northern boundary between domains II and IV is defined by a
set of curved thrusts (Curionópolis thrust front), which are inter-
connected with the Parauapebas fault zone. They overprint syn-D2
shear zones in basement and neoarchean lithotypes. Immediately to the
south, syn-D3 thrusts show higher linearity mainly with ENE-WSW
strike, presenting low-temperature mylonitic foliation, as in domain II.
However, they are steeply dipping to the south-southeast (Fig. 4d).
High-angle reverse-dextral oblique NW-SE to WNW-ESE faults are also
present.
Domain V is located to the south of the Carajás fault, where Meso
and Neoarchean units crop out and where the Mesoarchean basement is
better exposed. D1 and D2 fabrics are dominant and few D3 structures
were recognized. At its easternmost end, S1 foliation describes an open
megafold, with inferred NE-SW axial plane (Figs. 3 and 4c). In this
region, D1 and D2 structures follow the outline of the domain V limit,
the formation of L3 stretching lineation was observed, shallowly
plunging to south-southeast. D3 fault zones are very narrow and con-
tain low temperature mylonites, locally associated with the reorienta-
tion and reworking of D1 structures, especially the L1 lineation, rotated
to the average NW-SE direction. Syn-D3 thrust faults strike between E-
W to ENE-WSW and dip moderately to steeply to the south. There are
also reverse-oblique faults, like the Carajás Fault, along the northern
limit with Domain III, striking WNW-ESE and dipping moderately to
steeply to south-southwest. Drag folds associated with outcrop-scale
thrust fronts indicate up dip oblique movement with top-to-northwest
tectonic transport (reverse-dextral kinematics). In the region im-
mediately south of the fault zone, refolded fold interference patterns
were observed, understood as D3 interference over D2. Domain V ob-
lique faults and thrusts comprise the Serra Sul fault system.
The various fault systems attributed to D3 can be understood at map
scale as parts of a single oblique thrust belt. Serra Leste and Serra Norte
thrust systems, separated by the Parauapebas fault zone, in the north,
are low-angle and imbricate low-grade metamorphic rocks (domains II
and III) on top of Domain I areas (Fig. 4c, e), which were affected by
previous high grade deformation and metamorphism. In this context,
Domain I can be understood as para-autochthonous in relation to D3.
Curionopolis thrust system (domain IV), with higher dip angles
(Fig. 4d), is imbricated over the deformation front and is indented with
the Serra Sul fault system (domain V), resulting in rotation and im-
brication on top of domains II and III. It is also notable that several of
the D3 structures reactivate older shear zones, especially sin-D2 thrusts,
like Carajás, Parauapebas and Curionopolis faults.
In addition to the features described above, some structures un-
derstood as late-D3 were also identified. A weak ductile-brittle chevron-
like crenulation with subvertical NE-SW axial planes is present in se-
dimentary rocks of the Águas Claras Formation and in the Itacaiúnas
Supergroup, restricted to the Sereno and Curionópolis thrust fronts.
Less competent portions of domains II and IV (micaceous schists and
phyllites) may also present locally kink bands with axial plane in the
same NE-SW direction, with variable dip and fold axis.
D3 regional σ1 direction varies from NNW-SSE to NW-SE, usually
with top-to-NW tectonic transport. There are several antithetic struc-
tures, however, presenting vergence of top to SE, but sharing the same
regional σ1.
3.5. Post-D3 extensional structures
This group affects all Archean and Paleoproterozoic units in the
study area and was developed in a brittle, fluid-dominated extensional
environment. It comprises a set of subvertical faults and fractures, as-
sociated with a large volume of veins and widespread hydrothermal
alteration. Post-D3 structures occur preferably (butnot exclusively) in
the vicinity of previous shear zones, particularly along D3 structures,
but propagating into the Paredão Group/Caninana Formation and also
into Serra dos Carajás Intrusive Suite plutons.
Post-D3 fractures and faults are mostly subvertical, and sometimes
associated with pervasive silicification. On the map scale (Fig. 3), two
directions of fractures and veins are recognized: the first one of E-W
direction, varying between WNW-ESE, near the Carajás Fault region,
and ENE-WSW crosscutting the Serra Leste thrust system; the second
direction NNE-SSW to NE-SW crosscuts all regional structural trends.
The same X-shaped pattern is also recognized in stereoplots of post-D3
fractures and veins (Fig. 6). Regional σ1 is vertical, and σ2 and σ3 are
respectively ENE-WSW and WNW-ESE.
In some structures it was possible to recognize normal to normal-
dextral kinematics with very restricted displacement. Dilational brec-
cias with jigsaw puzzle texture occur in post-D3 fracture zones (Fig. 5g).
Coeval veins are sometimes forming stockwork and often present syn-
taxial mineral growth. Regional Post-D3 alteration assemblage includes
quartz, chlorite, epidote, albite, carbonate, actinolite, scapolite,
greenish biotite, sericite, tourmaline, fluorite, apatite and stilpnome-
lane. The main oxide is hematite, but occasionally magnetite is also
found, while sulphides include chalcopyrite, bornite and chalcocite.
3.6. Pre-D4 and D4 structures
The last recognized groups are of brittle character (locally brittle-
ductile), affecting all the Precambrian units in the study area. Pre-D4
structures are extensional and compose a joint system of NNW-SSE
direction, dozens of kilometers long, frequently filled with
Neoproterozoic mafic dikes (DOCEGEO, 1988). In some places, it was
possible to recognize that these structures were later reactivated as
high-angle reverse faults, during D4.
D4 compression affected marginal sectors of the Paredão Group
(Fig. 5h) and Caninana Formation, where the verticalization of sedi-
mentary bedding was observed and correlated to syn-D4 up-dip drag.
There is also much localized development of an anastomosing mylonitic
foliation, of very low temperature, that crosscuts some neoproterozoic
diabases, as well as a chevron-type crenulation cleavage (S4) that is
restricted to the immediate vicinity of syn-D4 reverse faults. Quartz
veins in the Curionópolis region, also oriented NNW-SSE, shallow to
steeply dipping to east-northeast, were similarly related to syn-D4 re-
verse faults, also associated with a restricted chevron crenulation
cleavage in incompetent host rocks (micaceous schists and phyllites).
Displacement along D4 reverse faults is minimal and cannot be shown
on the map.
F.M. Tavares et al. Journal of South American Earth Sciences 88 (2018) 238–252
247
The regional σ1 direction of D4 varies between E-W and ENE-WSW.
No tectonic transport could be inferred, due to the insignificance of
fault displacement and due to low representation of D4 structures.
4. Discussion
4.1. Redefinition of the Itacaiúnas Belt
As described in the previous sections, the CP's tectonic framework
comprises a variety of structures of different tectonic styles with com-
plex cutting and overprinting relations that can be relatively positioned
in the Mesoarchean (D1: compressional and ductile), Neoarchean (post-
D1: extensional and brittle), Paleoproterozoic (D2: compressional and
ductile; D3: compressional and ductile-brittle; post-D3: extensional and
brittle) and Neoproterozoic (pre-D4: extensional and brittle; D4: com-
pressional and brittle-ductile), as resumed in Table 1.
For several authors, however, the main structural control of the CP
is related to lateral shear due to NE-SW shortening at about 2.5 Ga,
associated with major strike-slip and oblique fault systems from the
Itacaiúnas Belt, of E-W mean direction (Araujo et al., 1988; Machado
et al., 1991; Costa et al., 1995; Pinheiro and Holdsworth, 2000). For
those authors, the Itacaiúnas Belt followed previous structural patterns
from the Mesoarchean basement with an insignificant Paleoproterozoic
tectonic overprint, while the only notable reactivation of the shear
system would be brittle and related to the 1.88 Ga Uatumã Magmatism.
Although we agree that younger structures of the Itacaiúnas Belt
follow the Mesoarchean structural framework (D1) and with the
1.88 Ga brittle reactivation (post-D3), the alleged Neoarchean de-
formation of the CP, as supported by those authors, is not consistent
with data presented in this work in three ways. Firstly, the study area
presents very few strike slip faults and seems to be dominated by
multistage compressional shearing rather than lateral shearing.
Secondly, Neoarchean deformation seems to be extensional (locally
transtensional) and brittle rather than transpressional and ductile.
Furthermore, the main deformational phase to affect CP's Neoarchean
units (D2) also affects Rhyacian assemblages along the CP-TP boundary
and, thus, restricts the maximum age of D2 to the Paleoproterozoic.
4.2. Carajás basin setting
The Mesoarchean accretionary and collisional processes that as-
sembled CP's basement (including the collision between Rio Maria ju-
venile terrane and Carajás paleocontinent) resulted in a relatively stable
continental substrate at around 2.83 Ga (Fig. 7a). Collisional infra-
structure was conditioned by the D1 tectonic framework and developed
under high metamorphic grade, together with the emplacement of syn-
tectonic (syn-to late-collisional) granitoids in major syn-D1 shear zones
(Fig. 5a and b).
The exact limit between the Rio Maria terrane and the Carajás pa-
leocontinent is yet unknown, but the area that was most affected by the
orogenic processes is coincident with the Carajás Domain basement
(Fig. 3), which can be understood as a deep tectonic discontinuity zone,
favorable to later reactivation.
We believe that the timing of Carajás basin installation is somehow
related to the Mesoarchean orogen collapse, although the opening
process remains unclear: the basin could either be installed by rift-re-
lated extension over the orogenic infrastructure or by post-collisional
crustal flexure, associated with tectonic burden. Evidence from the
literature points to the installation of the basin at around 2.76 Ga
(Gibbs et al., 1986), associated with bimodal magmatism and shallow
marine clastic-chemical sedimentation (Fig. 7b), yet those character-
istics are plausible in both evolutionary models.
The basin installation was coeval with a major mantellic thermal
anomaly, probably related to mantle upwelling, as suggested by
Ferreira Filho et al. (2007) to the magmatic origin of the Archean mafic-
ultramafic complexes. Whether this process was induced by the raise of
a hypothetical mantle plume, by mechanic upwelling through pure
extension of the continental crust or, more likely, as a result of slab
break-of associated with a late stage of the Mesoarchean orogeny, is left
undefined, though those hypotheses are not mutually exclusive.
Several authors like Gibbs et al. (1986) and DOCEGEO (1988) agree
about a rift-related origin to the Carajás basin. Hartlaub et al. (2004)
listed the features that are apparently common in Archean rift systems,
like the presence of a major crystalline substrate, crustal contamination
evidenced, for example, by xenocrystic zircons in volcanic rocks, as well
as the similar stratigraphy and, unavoidably, bimodal magmatism. At
the base, there are usually sedimentary successions of continental to
shallow marine environment, associated with dominantly mafic vol-
canism (and secondarily ultramafic and/or felsic), interbedded with
banded iron formations. At the top, sedimentary sequences of shallow
marine environment, including carbonate sequences occasionally in-
terbedded with (sub)volcanic rocks, are described. These characteristics
are quite similar to the observed in the Carajás basin, but the lack of a
known basal continental-like sedimentation in the ItacaiúnasSu-
pegroup raises serious doubts with respect to this model.
The Neoarchean bimodal plutonism was expressive and related to a
considerable volume of underplating-related crustal melting, as sug-
gested by Feio et al. (2012). The A-type-like granitic bodies of the
Planalto Suite and correlates, such as the Estrela and Gelado meta-
granites, usually present spatial relation with E-W Mesoarchean ductile
structures (Fig. 3) and are commonly understood as syntectonic to some
stage of activity of these shear zones. The mafic counterparts of these A-
type granites are understood as the Luanga layered Complex and cor-
relates.
For Barros et al. (2001, 2009), the Estrela metagranite and other A-
type-like plutons emplacement is synchronous to the inversion of the
Neoarchean volcanossedimentary sequences. This proposition is in-
compatible with the structural observations and tectonostratigraphy
presented in this article, as many of the arguments used by those au-
thors for the allegedly syntectonic emplacement (supposedly by bal-
looning) can also be explained in the study area by the multistage de-
formation to which the plutons were submitted. Besides, the existence
of a “foliated hornfels” halo in its surroundings, as stated by Barros
et al. (2001, 2009) was not confirmed by our geological mapping, since
both the metamorphism and the deformation have a regional character
and are not restricted to shear zones or to the vicinity of the Neoarchean
granitic bodies. In addition, regional foliation deflection around these
bodies was not confirmed. In fact, regional foliation and metamorphic
Table 1
Tectonic events that affected the northern Carajás Domain.
Event Age Tectonic Regime σ1 Tectonic transport Metamorphism Observations
D1 2.87–2.83 Ga compressive N-S top-to-N (?) lower amphibolite to granulite Carajás-Rio Maria Collision
post-D1 2.76–2.55 Ga extensional vertical Carajás basin opening
D2 2.10–2.06 Ga compressive NE-SW top-to-SW lower greenshist (S) to granulite (N) Carajás-Bacajá Collision (Transamazonian Orogeny)
D3 2.00–1.98 Ga compressive NNW-SSE top-to-NW lower to upper greenshist Sereno Tectono-thermal Event (intracontinental)
post-D3 1.90–1.85 Ga extensional vertical Paleoproterozoic orogenic collapse
pre-D4 ∼0.75 Ga extensional vertical Araguaia basin opening
D4 ∼0.55 Ga compressive E-W top-to-W (?) Araguaia basin inversion (Araguaia Belt/Brasiliano Orogeny)
F.M. Tavares et al. Journal of South American Earth Sciences 88 (2018) 238–252
248
Fig. 7. Tectonic evolution model for the northeastern Carajás Domain, as proposed in this work. Black arrows indicate compressional or extensional tectonic regimes.
F.M. Tavares et al. Journal of South American Earth Sciences 88 (2018) 238–252
249
zoning crosscut the plutons. Similarly, mafic xenoliths in upper am-
phibolite facies were understood as host rocks relicts rather than frag-
ments of Itacaiúnas Supergroup affected by contact metamorphism. The
existence of an alleged igneous banded pattern, reworked by a late,
submagmatic to subsolidus tectonic foliation was alternatively under-
stood as S2/S3 relation, consistent with other regional observations and
admitting that the bodies were gneissified after their emplacement,
during D2. Finally, Barros et al. (2001, 2009) understood that the de-
position of the volcanosedimentary units was prior to 2.76 Ga, which
would be an age that supposedly marked the beginning of basin in-
version under transpressive deformation, not consistent with the var-
ious U-Pb ages presented for the Itacaiúnas Supergroup by different
authors (e.g. Gibbs et al., 1986).
Alternatively to the proposition of Barros et al. (2001, 2009), the A-
type-like granites could have been emplaced into Mesoarchean shear
zones that are related to basement agglutination, but synchronous to
extensional reactivation during the Neoarchean, which would be nat-
ural pathways for magma migration. The flattened shape of many
bodies and different strain levels assimilated by them probably reflect
distinct reworking levels during D2 and D3.
Post-D1 extensional tectonics, magnetite-rich IOCG-related hydro-
thermalism and late-stage magmatism persisted until 2.5 Ga, when the
Old Salobo granite and other correlate felsic dykes were emplaced.
Fragmentation and drifting of the Carajás paleocontinent might
have taken place and is thought to be a post-D1, pre-D2 event, leaving
in the CP only a fraction of its former margin, but this is a highly
speculative (yet reasonable) event, which timing remains unclear.
4.3. Paleoproterozoic reworking of the CP
4.3.1. Rhyacian deformation (Transamazonian Orogeny)
On a global scale, Condie et al. (2009) stated that the beginning of
the Paleoproterozoic is of great tectonic, sedimentary and magmatic
monotony, which is corroborated by CP's virtual lack of known tecto-
nostratigraphic record between 2.5 and 2.1 Ga. The D2 structures re-
present the first event after that time gap in the CP. D2 structurally
modeled the CP-TP boundary and is understood as the record of the
Transamazonian Orogeny in the study area (Fig. 7c, d). It is assumed as
a collisional event between the Carajás-Rio Maria block and the Bacajá
paleocontinent (Carajás-Bacajá Collision), associated with important
tectonic thickening and evidenced by the regional metamorphic zoning:
in the CP, D2 structural framework is related to low greenschist facies
(south) to high amphibolite facies (north) metamorphic fabrics and
structures; to the north, in the TP, lithotypes are in granulite facies,
partially retrograded to amphibolite facies (Figs. 3 and 5c, d).
The syn-D2 granulitization observed in TP was dated by Macambira
et al. (2007) at about 2.06 Ga, similar to zircon U-Pb metamorphic ages
between 2.09 and 2.07 Ga (crystal rims and oscillatory-zoned migma-
titic crystals) reported by Tavares and Silva (2012). The metamorphic
peak apparently predated the deformation peak in those lithotypes, as
correlate mylonitization is connected to retrograde amphibolite facies
fabrics that are overprinted on the granulites.
Transamazonian structural and thermal influence over the studied
portion of the CP is remarkable, reaching at least 100 km to the south of
the CP-TP boundary. The Águas Claras Formation, however, barely
registers the Transamazonian deformation. Only a few brittle structures
and some bedding undulation are thought to be related (at least in part)
to the Carajás-Bacajá Collision. This suggests that the unit is late-to
post-tectonic in relation to D2, concordant with the estimated deposi-
tional age of Fabre et al. (2011).
4.3.2. Orosirian deformation (Sereno Event and later extensional
structures)
D3 structures are related to a third compressive event that was
dominated by thrusts and oblique reverse-dextral faults (Fig. 3). Araújo
and Maia (1991) and Oliveira et al. (1994) considered most structures
here classified as sin-D3 as components of the Carajás and Cinzento
strike-slip systems, relating them to positive flower and hemiflower
structures and supposedly related to the Itacaiúnas Belt Archean evo-
lution. However, in this paper, syn-D3 structures are interpreted as
forming an independent, superimposed oblique-thrust system, younger
than the Transamazonian structural framework, here named as the
Sereno Event.
D3 structural patterns and coeval metamorphism are of very low
grade (Fig. 5e and f) and, in the study area, not related to any sig-
nificant magmatism. Thus, it is supposed that Sereno Event represents
an intracontinental orogeny correlated to Orosirian accretionary-colli-
sional belts that surrounded the Amazonian protocraton at 2.00 to
1.98 Ga, like the Tapajós and Cauarane-Coeroni belts (Vásquez et al.,
2008a; Fraga et al., 2009). Considering the synthetic tectonic transport
direction (top-to-NW), it is supposed that an Orosirian collisional front
also affected the southeastern margin or the protocraton, prior to the
Araguaia Belt installation (Fig. 7e).
The age of the Sereno Event can be inferred from cutting relationsas
Orosirian, between 2.06 and 1.88 Ga, considering that D2 is superposed
by D3, as well as the intrusion of post-D3 granitic bodies of the Serra
dos Carajás Intrusive Suite. Arcanjo and Moura (2000) and Arcanjo
et al. (2013) described juvenile paleoproterozoic orthogneisses in the
southern basement of the Araguaia Belt, with crystallization ages be-
tween 2.08 and 2.01 Ga (Rio dos Mangues Complex). Additionally,
40Ar/39Ar plateau ages of Renne et al. (1988) for biotite crystals of the
Carajás Domain basement, collected immediately to the south of the
study area, indicate regional cooling at about 1.96–1.98 Ga, compatible
with the greenschist facies metamorphic conditions related to syn-D3
crustal thickening. These ages are also concordant with the cutting
relations previously described.
To the west of the study area, D3 deformation propagates by re-
verse-dextral oblique faults with related counterclockwise megablock
rotation, which generates a tectonic indentation system notably re-
presented by the Carajás Fault and other parallel structures. Syn-D3
strike-slip faults that crosscut the northern segment of the study area
are also interpreted in this way.
Late-D3 structures are probably related to the local rotation of σ1 to
NW-SE position due to the accommodation of late stage strain, selec-
tively registered in less competent lithotypes.
The continental-related sedimentary deposits of the Paredão Group
and Caninana Formation are spatially related to the Sereno Event thrust
fronts, locally covering D3 structures (Fig. 3). However, according to
Pereira et al. (2009), they are also affected by late thrusting of the
Carajás Fault along the western basin limit. The same authors provided
the previously mentioned detrital zircon ages that show maximum de-
position age of 2.01 Ga for the unit and notable absence of 1.88 Ga
zircon grains. These factors suggest that those sequences represent late-
to post-D3 foreland basins.
Post-D3 structures (Fig. 5g) are associated with hematite-rich IOCG
deposits, which yielded ages around 1.88 Ga (e.g. Moreto et al., 2014,
2015 and references therein), some tens of millions of years after the
collapse of the paleoproterozoic orogenies and synchronous to the an-
orogenic emplacement of the Serra dos Carajás Intrusive Suite bodies,
during Uatumã magmatism (Fig. 7f).
4.4. Araguaia Belt influence on CP
As mentioned above, the pre-D4 fracture network is orthogonal to
all main CP structural directions and subparallel to the main orientation
of the eastern Amazonian Craton boundary, coincident with the
Neoproterozoic-Cambrian Araguaia Belt. The initial development of
pre-D4 structures and concomitant dyke swarm emplacement is prob-
ably related to the opening of the Araguaia basin (Fig. 7g), while D4
structures represent the basin inversion. These structures (Fig. 5h) are
understood as the most distal expression of that orogeny inside the
cratonic region, a few dozen kilometers to the west of the previously
F.M. Tavares et al. Journal of South American Earth Sciences 88 (2018) 238–252
250
recognized limit (Fig. 7h).
5. Conclusions
The complex structural arrangement here described and discussed
shows the successive accumulation of several tectonic events in the
northeastern Carajás Province, between the Mesoarchean and the
Neoproterozoic-Cambrian, subjected to different regional stress fields.
The three main sedimentation ages, represented by the deposition of
the Neoarchean Itacaiúnas Supergroup, the Rhyacian Águas Claras
Formation and the Orosirian Caninana Formation/Paredão Group, are
late to post-tectonic in relation to the three major orogenic events re-
cognized in the area (D1, D2 and D3, respectively). We understand that
this relation corresponds to the superposition of three compressional-
extensional cycles in the region, which resulted in the current geotec-
tonic configuration of the southeastern Amazonian Craton. The distal
effects of a fourth extensional-compressional cycle, related to the
Araguaia Belt tectonostratigraphic evolution, were also recognized.
The present proposal is consistent with observations by different
authors in other Archean cratonic nuclei margins around the world,
diluting the current idea that in the northern Carajás Province some
special case of intracontinental tectonics guided by major strike-slip
faults would occur. The Archean-Paleoproterozoic deformation af-
fecting the region can well be explained by uniformitarian processes
related to usual plate tectonics.
Acknowledgments
All the Brazilian Geological Survey geological maps are available in
the institution's website (http://www.cprm.gov.br). Felipe Mattos
Tavares was a PhD a student of the geology graduate program at UFRJ
during the development of this research and acknowledges a PhD
scholarship from CAPES – Coordenação de Aperfeiçoamento de Pessoal
de Nível Superior. CPRM/SGB – Brazilian Geological Survey – spon-
sored fieldwork and all other research activities. Rudolph Allard
Johannes Trouw acknowledges a research scholarship from CNPq –
Conselho Nacional de Pesquisa.
Appendix A. Supplementary data
Supplementary data related to this article can be found at https://
doi.org/10.1016/j.jsames.2018.08.024.
References
Araújo, O.J.B., Maia, R.G.N., 1991. Serra dos Carajás, folha SB.22-ZA, Estado do Pará. In:
Programa Levantamentos Geológicos Básicos do Brasil. Companhia de Pesquisa de
Recursos Minerais, pp. 164p.
Araújo, O.J.B., Maia, R.G.N., Silva, J.J.X., Costa, J.B.S., 1988. A megaestruturação da
folha Serra dos Carajás. In: Congresso Latino Americano de Geologia. vol. 7. pp.
324–333.
Arcanjo, S.H.S., Moura, C.A.V., 2000. Geocronologia das rochas do embasamento do setor
meridional do Cinturão Araguaia. Região de Paraíso do Tocantins (TO). Rev. Bras.
Geociencias 30 (4), 665–670.
Arcanjo, S.H.S., Abreu, F.A.M., Moura, C.A.V., 2013. Evolução geológica das sequências
do embasamento do Cinturão Araguaia na região de Paraíso do Tocantins (TO),
Brasil. Braz. J. Genet. 43, 501–514.
Baltazar, O.F., Zuchetti, M., 2007. Lithofacies associations and structural evolution of the
Archean Rio das Velhas greenstone belt, Quadrilátero Ferrífero, Brazil: a review of
the setting of gold deposits. Ore Geol. Rev. 32, 471–499.
Barbosa, J.P.O., 2004. Unpublished Master Thesis In: Geologia estrutural, geoquímica,
petrografia e geocronologia de granitóides da região do Igarapé Gelado, norte da
Província Mineral de Carajás. Univesidade Federal do Pará, Belém, pp. 112.
Barros, C.E.M., Macambira, M.J.B., Barbey, P., 2001. Idade de zircão do Complexo
Granítico Estrela: relações entre magmatismo, deformação e metamorfismo na
Província Mineral de Carajás. In: SIMPÓSIO DE GEOLOGIA DA AMAZÔNIA, 7,
Belém. Resumos Expandidos. Sociedade Brasileira de Geologia, Belém, pp. 17–20.
Barros, C.E.M., Macambira, M.J.B., Barbey, P., Scheller, T., 2004. Dados isotópicos Pb–Pb
em zircão (evaporação) e Sm–Nd do Complexo Granítico Estrela, Província Mineral
de Carajás, Brasil: implicações petrológicas e tectônicas. Rev. Bras. Geociencias 34,
531–538.
Barros, C.E.M., Sardinha, A.S., Barbosa, J.P.O., Macambira, M.J.B., Barbey, P., Boullier,
A.M., 2009. Structure, petrology, geochemistry and zircon U/Pb and Pb/Pb geo-
chronology of the synkinematic archean (2.7 ga) A-type granites from the Carajás
Metallogenic Province, northern Brazil. Can. Mineral. 47, 1423–1440.
Bhadra, S., Gupta, S., Banerjee, M., 2004. Structural evolution across the Eastern Ghats
Mobile Belt–Bastar craton boundary, India: hot over cold thrusting in an ancient
collision zone. J. Struct. Geol. 26, 233–245.
Brito Neves, B.B., Cordani, U.G., 1991. Tectonic evolution of south America during the
late proterozoic. Precambrian Res. 53, 23–40.
Brown, M., 2006. Duality of thermal regimes is the distinctive characteristic of plate
tectonics since the Neoarchean. Geology 34, 961–964.
Cabral, A.R., Creaser, R.A., Nagler, T., Lehmann, B., Voegelin, A.R., Belyatsky, B., Pasava,
J., Seabra Gomes Jr., A.A., Galbiatti, H., Bottcher, M.E.,Escher, P., 2013. Trace-
element and multi-isotope geochemistry of late-archean black shales in the carajás
iron-ore district, Brazil. Chem. Geol. https://doi.org/10.1016/j.chemgeo.2013.08.
041.
Castro, D.L., Fuck, R.A., Phillips, J.D., Vidotti, R.M., Bezerra, F.H.R., Dantas, E.L., 2014.
Crustal structure beneath the Paleozoic Parnaíba Basin revealed by airborne gravity
and magnetic data, Brazil. Tectonophysics 614, 128–145.
Condie, K.C., O'Neill, C., Aster, R.C., 2009. Evidence and implications for a widespread
magmatic shutdown for 250 My on Earth. Earth Planet Sci. Lett. 282, 294–298.
Condie, K.C., Pease, V. (Eds.), 2008. When Did Plate Tectonics Begin on Planet Earth?
Geol. Soc. Am. Spec. Pap., vol 440. pp. 294 Boulder, CO: Geol. Soc. Am.
Cordani, U.G., Brito Neves, B.B., Fuck, R.A., Porto, R., Thomaz Filho, A., Cunha, F.M.B.,
1984. Estudo preliminar de integração do Pré-Cambriano com os eventos tectônicos
das bacias sedimentares brasileiras, vol 15. Ciência Técnica Petróleo, Seção
Exploração de Petróleo, Rio de Janeiro, pp. p70.
Costa, J.B.S., Araújo, O.J.B., Santos, A., Jorge João, X.S., Macambira, M.J.B., Lafon, J.M.,
1995. A Província Mineral de Carajás: aspectos tectono-estruturais, estratigráficos e
geocronológicos. Bol. Mus. Para. Emilio Goeldi 7, 199–235.
Cawood, P.A., Korsch, R.J., 2008. Assembling Australia: proterozoic building of a con-
tinent. Precambrian Res. 166, 1–38.
Dall'Agnol, R., Teixeira, N.P., Rämö, O.T., Moura, C.A.V., Macambira, M.J.B., Oliveira,
D.C., 2005. Petrogenesis of the paleoproterozoic, rapakivi, A-type granites of the
archean Carajás metallogenic province, Brazil. Lithos 80, 101–129.
Dias, G.S., Macambira, M.J.B., Dall'Agnol, R., Soares, A.D.V., Barros, C.E.M., 1996.
Datações de zircões de sill de metagabro: comprovação de idade arqueana da
Formação Águas Claras, Carajás, Pará. In: Simpósio de Geologia da Amazônia, 5.
Sociedade Brasileira de Geologia, Belém, pp. 376–378.
DOCEGEO, 1988. Revisão litoestratigráfica da Província Mineral de Carajás – litoestra-
tigrafia e principais depósitos minerais. XXXV Congresso Brasileiro de Geologia.
Belém, SBG, Proceedings 11–54.
Fabre, S., Nédélec, A., Poitrasson, F., Strauss, H., Thomazo, C., Nogueira, A., 2011. Iron
and sulphur isotopes from the Carajás mining province (Pará, Brazil): implications for
the oxidation of the ocean and the atmosphere across the Archaean–Proterozoic
transition. Chem. Geol. 289, 124–139.
Faraco, M.T.L., Marinho, P.A.C., Vale, A.G., Costa, E.J.S., Maia, R.G.N., Ferreira, A.L.,
Valente, C.R., Lacerda Filho, J.V., Moreton, L.C., Camargo, M.A., Vasconcelos, A.M.,
Oliveira, M., Oliveira, I.W.B., Abreu Filho, W.A., Gomes, I.P., 2004. Folha SB.22-
Araguaia. In: Schobbenhaus, C., Gonçalves, J.H., Santos, J.O.S., Abram, M.B., Leão
Neto, R., Matos, G.M.M., Vidotti, R.M., Ramos, M.A.B., Jesus, J.D.A. (Eds.), Carta
Geológica do Brasil ao Milionésimo, Sistema de Informações Geográficas. Programa
Geologia do Brasil. CPRM, Brasília (CD ROM).
Feio, G.R.L., Dall'Agnol, R., Dantas, E.L., Macambira, M.J.B., Gomes, A.C.B., Sardinha,
A.S., Oliveira, D.C., Santos, R.D., Santos, P.A., 2012. Geochemistry, geochronology,
and origin of the Neoarchean Planalto Granite suite, Carajás, Amazonian craton: a-
type or hydrated charnockitic granites? Lithos 151, 57–73.
Feio, G.R.L., Dall'Agnol, R., Dantas, E.L., Macambira, M.J.B., Santos, J.O.S., Althoff, F.J.,
Soares, J.E.B., 2013. Archean granitoid magmatism in the Canaã dos Carajás area:
implications for crustal evolution of the Carajás province, Amazonian craton, Brazil.
Precambrian Res. 227, 157–186.
Ferreira Filho, C.F., Cançado, F., Correa, C., Macambira, E.M.B., Junqueira-Brod, T.C.,
Siepierski, L., 2007. Mineralizações estratiformes de PGE-Ni associadas a complexos
acamadados em Carajás: os exemplos de Luanga e Serra da Onça. In: Rosa-Costa, L.T.,
Klein, E.L., Viglio, E.P. (Eds.), Contribuições à geologia da Amazônia. vol. 5.
Sociedade Brasileira de Geologia, Belém, pp. 1–14.
Fraga, L.M., Macambira, M.J.B., Dall'Agnol, R., Costa, J.B.S., 2009. 1.94-1.93 Ga char-
nockitic magmatism from the central part of the Guyana Shield, Roraima, Brazil:
single zircon evaporation data and tectonic implications. J. S. Am. Earth Sci. 27,
247–257 2009.
Galarza, M.A., Macambira, M.J.B., 2002. Geocronologia e evolução crustal da área do
depósito de Cu-Au Gameleira, Província Mineral de Carajás (Pará), Brasil. Revista do
Instituto de Geociências da USP 2, 143–159.
Gibbs, A.K., Wirth, K.R., Hirata, W.K., Olszewski Jr., W.J., 1986. Age and composition of
the Grão Pará group volcanics, Serra dos Carajás. Rev. Bras. Geociencias 16,
201–211.
Hartlaub, R.P., Heamana, L.M., Ashton, K.E., Chacko, T., 2004. The archean murmac bay
group: evidence for a giant Archean rift in the Rae province, Canada. Precambrian
Res. 131, 345–372.
Hasui, Y., Abreu, F.A.M., Silva, J.M.R., 1977. Estratigrafia da Faixa de Dobramentos
Paraguai-Araguaia no centro-norte do Brasil, vol. 8. Boletim do Instituto de
Geociências, São Paulo, pp. 107–118.
Hirata, W.K., Rigon, J.C., Kadekaru, K., Cordeiro, A.A.C., Meireles, E.A., 1982. Geologia
Regional da Província Mineral de Carajás. In: Simpósio de Geologia da Amazônia, 1.
Belém, Sociedade Brasileira de Geologia, pp. 100–110.
Hoffman, P.F., 1988. United plates of America: the birth of a craton. Early Proterozoic
assembly and growth of Laurentia. Annu. Rev. Earth Planet. Sci. Lett. 60, 543–603.
F.M. Tavares et al. Journal of South American Earth Sciences 88 (2018) 238–252
251
http://www.cprm.gov.br/
https://doi.org/10.1016/j.jsames.2018.08.024
https://doi.org/10.1016/j.jsames.2018.08.024
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref1
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref1
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref1
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref2
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref2
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref2
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref3
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref3
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref3
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref4
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref4
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref4
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref5
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref5
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref5
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref6
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref6
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref6
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref7
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref7
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref7
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref7
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref8
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref8
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref8
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref8
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref9
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref9
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref9
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref9
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref10
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref10
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref10
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref11
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref11
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref12
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref12
https://doi.org/10.1016/j.chemgeo.2013.08.041
https://doi.org/10.1016/j.chemgeo.2013.08.041
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref14
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref14
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref14
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref15
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref15http://refhub.elsevier.com/S0895-9811(18)30159-7/sref16
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref16
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref17
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref17
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref17
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref17
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref18
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref18
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref18
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref19
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref19
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref20
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref20
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref20
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref21
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref21
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref21
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref21
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref22
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref22
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref22
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref23
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref23
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref23
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref23
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref24
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref24
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref24
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref24
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref24
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref24
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref24
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref25
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref25
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref25
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref25
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref26
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref26
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref26
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref26
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref27
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref27
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref27
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref27
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref27
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref28
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref28
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref28
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref28
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref29
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref29
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref29
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref30
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref30
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref30
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref31
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref31
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref31
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref32
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref32
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref32
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref33
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref33
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref33
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref34
http://refhub.elsevier.com/S0895-9811(18)30159-7/sref34
Holdsworth, R.E., Pinheiro, R.V.L., 2000. The anatomy of shallow-crustal transpressional
structures: insights from the Archaean Carajás fault zone, Amazon, Brazil. J. Struct.
Geol. 22 1105-1023.
Hopkins, M., Harrison, T.M., Manning, C.E., 2008. Low heat flow inferred from>4Gyr
zircons suggests Hadean plate boundary interactions. Nature 456, 493–496.
Korenaga, J., 2013. Initiation and evolution of plate tectonics on Earth: theories and
observations. Annu. Rev. Earth Planet Sci. 41, 117–151. https://doi.org/10.1146/
annurev-earth-050212-124208.
Kusky, T.M., 1990. Evidence for archean ocean opening and closing in the southern Slave
province. Tectonics 9, 1533–1563. https://doi.org/10.1029/TC009i006p01533.
Kusky, T.M., Li, X., Wang, Z., Fu, J., Ze, L., Zhu, P., 2014. Are Wilson cycles preserved in
archean cratons? A comparison of the north China and Slave crátons. Can. J. Earth
Sci. 51, 297–311. https://doi.org/10.1139/cjes-2013-0163.
Kusky, T.M., Li, J.H., 2003. Paleoproterozoic tectonic evolution of the north China craton.
J. Asian Earth Sci. 22, 383–397. https://doi.org/10.1016/S1367-9120(03)00071-3.
Macambira, M.J.B., Barros, C.E.M., Silva, D.C.C., Santos, M.C.C., 2001. Novos dados
geológicos e geocronológicos para a região ao norte da Província de Carajás,
evidências para o estabelecimento do limite Arqueano-Paleoproterozóico no sudeste
do Cráton Amazônico. In: Simpósio de Geologia da Amazônia, 7, Belém, Brazil,
Sociedade Brasileira de Geologia Anais, CD-ROM.
Macambira, M.J.B., Pinheiro, R.V.L., Armstrong, R.A., 2007. A fronteira Arqueano-
Paleoproterozóico no SE do Cráton Amazônico; abrupta no tempo, suave na tectô-
nica? In: Simpósio de Geologia da Amazônia, 10, Porto Velho. Sociedade Brasileira de
Geologia, Anais, Brazil, pp. 105–108.
Macambira, M.J.B., Vasquez, M.L., Silva, D.C.C., Galarza, M.A., Barros, C.E.M., Camelo,
J.F., 2009. Crustal growth of the central-eastern Paleoproterozoic domain, SW
Amazonian craton: juvenile accretion vs. reworking. J. S. Am. Earth Sci. 27, 235–246.
Machado, N., Lindenmayer, D.H., Krough, T.E., Lindenmayer, Z.G., 1991. U-Pb geo-
chronology of Archean magmatism and basementreactivation in the Carajás area,
Amazon Shield, Brazil. Precambrian Res. 49, 329–354.
Melo, G.H.C., Monteiro, L.V.S., Xavier, R.P., Santiago, E.S.B., Santos, A.F.F., Torres, A.,
Aires, B., 2013. A new outlook on the giant Salobo IOCG deposit: a mesoarchean
basement-hosted deposit, Carajás Province. In: Simpósio de Geologia da Amazônia,
13. Sociedade Brasileira de Geologia, Belém, pp. 1052–1055.
Moreto, C.P.N., Monteiro, L.V.S., Xavier, R.P., Creaser, R.A., Dufrane, A., Melo, G.H.C.,
Delinardo Silva, M.A., Tassinari, C.C.G., Sato, K., 2014. Timing of multiple hydro-
thermal events in the iron oxide copper gold deposits of the Southern Copper Belt,
Carajás Province, Brazil. Miner. Deposita X, 1.
Moreto, C.P.N., Monteiro, L.V.S., Xavier, R.P., Creaser, R.A., DuFrane, S.A., Melo, G.H.C.,
Delinardo da Silva, M.A., Tassinari, C.C.G., Sato, K., 2015. Timing of multiple hy-
drothermal events in the iron oxide-copper-gold deposits of the Southern Copper Belt,
Carajás Province, Brazil. Miner. Deposits 50, 517–546. https://doi.org/10.1007/
s00126-014-0549-9.
Mougeot, R., Respaut, J.P., Briqueu, L., Ledru, P., Milesi, J.P., Macambira, M.J.B., Huhn,
S.B., 1996a. Geochronological constrains for the age of the Águas Claras Formation
(Carajás province, Pará, Brazil). In: Congresso Brasileiro de Geologia, 39, Salvador,
1996. Anais. Salvador, SBG 6. pp. 579–581.
Mougeot, R., Respaut, J.P., Briqueu, L., Ledru, P., Milesi, J.P., Lerouge, C., Marcoux, E.,
Huhn, S.B., Macambira, M.J.B., 1996b. Isotope geochemistry constrains for Cu, Au
mineralizations and evolution of the Carajás Province (Para, Brazil). In: Congresso
Brasileiro de Geologia, 39, Salvador, 1996, Anais, Salvador, SBG, 7:321–324. Moura
C.A.V. & Souza S.H.P. 1996. Síntese dos dados Geocronológicos das rochas do
Embasamento do Cinturão Araguaia e suas Implicações Estratigráficas. In: SBG, Cong.
Bras. Geol., 39, Anais, 6, p. 31-34.
Nogueira, A.C.R., Truckenbrodt, W., Pinheiro, R.V.L., 1995. Formação Águas Claras, Pré-
Cambriano da Serra dos Carajás: redescrição e redefinição litoestratigráfica. Boletim
Museu Paraense Emílio Goeldi 7, 177–277.
Oliveira, J.R., Silva Neto, C.S., Costa, E.J.S., 1994. Serra Pelada;

Continue navegando