Artigo 4

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Journal of South American Earth Sciences 129 (2023) 104490

Contents lists available at ScienceDirect

Journal of South American Earth Sciences


journal homepage: www.elsevier.com/locate/jsames

Tracking a magmatic arc within a confined orogen: New evidence from the
Araçuaí orogen (SE Brazil)
Cristina Araujo a, *, Antonio Pedrosa-Soares a, 1, Cristiano Lana b, 1, Mahyra Tedeschi a, 1,
Jorge Roncato a, Ivo Dussin c, 1, Paula Serrano b, 1, Elton Dantas d, 1
a
Universidade Federal de Minas Gerais, Programa de Pós-Graduação em Geologia, CPMTC-IGC, Campus Pampulha, Belo Horizonte, Brazil
b
Universidade Federal de Ouro Preto, Departamento de Geologia, Morro do Cruzeiro, Ouro Preto, Brazil
c
Universidade do Estado do Rio de Janeiro, Faculdade de Geologia, Rio de Janeiro, Brazil
d
Laboratório de Geocronologia, Instituto de Geociências, Universidade de Brasília, Brasília, DF, Brazil

A R T I C L E I N F O A B S T R A C T

Keywords: By comparison with igneous rock assemblages of well-exposed modern magmatic arcs and their typical calc-
U-Pb geochronology alkaline, magnesian, I-type signatures, many deeply eroded magmatic arcs have been characterized world­
Hf–Nd isotopes wide, unraveling ancient plate margins. Since the 1980’s, different assemblages of Ediacaran to Cambrian
Slab retreat
magmatic rocks have been systematically mapped and characterized according to their distinctive petrographic,
Active margin
Brasiliano orogeny
structural, lithochemical, geochronologic, and isotopic features, allowing to establish the pre-collisional, colli­
sional, and post-collisional stages of the Araçuaí orogen (SE Brazil). Among them, the tonalitic-granodioritic
batholiths and stocks with calc-alkaline, magnesian, I-type signatures, rich in dioritic to mafic enclaves, and
related metavolcano-sedimentary successions compose a magmatic arc formed in the early Ediacaran, the Rio
Doce magmatic arc. Extending for more than 1000 km along the Araçuaí-Ribeira Orogenic System (AROS), the
Rio Doce arc shows a hybrid isotopic signature and crustal inheritance, compatible with modern magmatic arcs
built on continental active margins. We present new field, petrographic, and lithochemical studies, coupled with
U–Pb (LA-ICPMS) geochronology and isotopic (in-zircon Lu–Hf and whole-rock Sm–Nd) analysis, unraveling the
least evolved isotopic signatures ever found for tonalitic-granodioritic plutons of the Rio Doce magmatic arc, and
a comprehensive data comparison with modern magmatic arcs. The studied Santa Tereza and Jequitibá suites
comprise tonalitic-granodioritic and Opx-bearing charnockitic rocks, with calcic-to-alkali-calcic-magnesian I-
type signature, associated with gabbroic rocks. Zircon U–Pb data yielded Concordia ages of 612 ± 3 Ma (Santa
Tereza) and 594 ± 3 Ma (Jequitibá) for magmatic crystallization, and 580 ± 5 Ma to 571 ± 3 Ma for meta­
morphic recrystallization, with associated εNd(612Ma) values from − 0.06 to − 7.05 (TDM model ages: 1.4–2.04
Ga), and εNd(594Ma) values from − 6.67 to − 8.8 (TDM model ages: 1.64–1.87 Ga). The zircon εHf(t) (+2.3 to − 4.8)
and Hf TDM model ages (1.33–1.65 Ga) assigned a moderately juvenile to evolved signature for the Santa Tereza
suite, while the Jequitibá suite (εHf(t): − 1.7 to − 36.4; TDM model ages: 1.33–3.21 Ga) shows a more evolved
signature. Altogether, data from the Rio Doce arc are similar to typical continental-margin magmatic arcs with
the involvement of mantle melts and/or melting of juvenile crust, suggesting a westward slab retreat during
ocean-floor subduction within the Araçuaí confined orogen.

1. Introduction accretionary orogens (Cawood and Kröner, 2009; Condie, 2007; Moyen
et al., 2017). Continental magmatic arcs provide key evidence to un­
Melts from differentiation of primitive mantle magmas contribute to derstand the formation and recycling of the continental crust through
continental crustal growth in subduction-related settings along plate time because they are outcomes of intricate thermal-magmatic in­
margins, either in intra-oceanic or continental magmatic arcs of teractions involving subducted oceanic lithosphere, the overlain

* Corresponding author. at: Universidade Federal de Minas Gerais, Programa de Pós-Graduação em Geologia, CPMTC-IGC, Av. Antônio Carlos, 6627, Belo
Horizonte, Brazil
E-mail address: crisgeo@ufmg.br (C. Araujo).
1
Fellow of the Brazilian Research Council (CNPq).

https://doi.org/10.1016/j.jsames.2023.104490
Received 17 April 2023; Received in revised form 13 June 2023; Accepted 11 July 2023
Available online 17 July 2023
0895-9811/© 2023 Elsevier Ltd. All rights reserved.
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

sub-continental mantle wedge, and the continental crust itself along an petrographic and geochemical features, including wide-ranging silica
active plate margin (Castro et al., 2010, 2011; Ducea et al., 2015a, contents (from acid and intermediate to basic terms), largely prevailing
2015b; Kearey et al., 2014; Moyen et al., 2017; Paterson and Ducea, I-type granitic rocks with calc-alkaline-magnesian signature, distinctive
2015). Thus, magmatic arcs comprise the most diversified associations trace element patterns (such as negative Nb–Ta–Ti anomaly), and hybrid
of igneous rocks on Earth. After their volcano-sedimentary covers have isotopic attributes and zircon inheritance, reflecting variable
been removed by erosion, continental magmatic arcs can be recognized thermal-magmatic interactions between continental crust and mantle
by: i) short-to-long-lived, linear-shaped belts of calc-alkaline plutons components (Best, 2003; Brown and Johnson, 2019; Castro et al., 2010;
relatively rich in silica, forming tonalite-granodiorite-rich batholiths, Castro, 2013; Ducea et al., 2015a, 2015b; Frost et al., 2001; Frost and
eventually associated with paired metamorphic belts and/or ophiolitic Frost, 2008; Kearey et al., 2014; Kelemen et al., 2014).
accretionary wedges; ii) common evidence of magma mingling-mixing By comparison with igneous rock assemblages of well-exposed
processes involving mafic (gabbroic, noritic), intermediate (dioritic, modern magmatic arcs and their typical lithochemical signature,
tonalitic) and felsic (granodioritic, granitic) terms; and iii) typical many deeply eroded magmatic arcs have been characterized worldwide,

Fig. 1. (A) Geotectonic setting of Araçuaí-Ribeira orogenic system (AROS), other Brasiliano-Pan-African orogenic systems, and the related cratons in West Gondwana
(modified from Alkmim et al., 2006). (B) Brasiliano orogenic systems (blue color) exposed in the Brazilian territory and the related cratons (pink color) in the South
American Platform context (modified from Cordani et al., 2016). Orogenic systems: Mantiqueira system - Araçuaí (Ar), Ribeira (Rb) and Dom Feliciano (DF) belts;
Tocantins system (T) - Araguaia, Paraguai and Brasília (B) belts; Borborema system (Bo) - Rio Preto, Riacho do Pontal and Sergipano belts; Pan-African system –
Damara (D), Gariep (G), Kaoko (K) and West Congo (WC). Cratons: Amazonian (Am), São Francisco (SFC), Paranapanema (Pp); Luís Alves (LA), Rio de La Plata (RP),
São Luiz (SL); Parnaíba (Pr). (C) Simplified geological map of the AROS (modified from Santiago et al., 2020, after Pedrosa-Soares et al., 2011, Tedeschi et al., 2016,
Degler et al., 2018; Peixoto et al., 2017). Legend: 1 - Cenozoic cover (TQ); 2 - post-collisional plutonism (ca. 540–480 Ma); 3: collisional plutonism (ca. 585–540 Ma);
4: Rio Doce magmatic arc (ca. 630–585 Ma); 5: Neoproterozoic ophiolite-bearing rock assemblages; 6: Rio Negro – Serra da Prata arc domain (Rio Negro and Serra da
Prata magmatic arcs, and related units; ca. 860 Ma–620 Ma); 7: Neoproterozoic metasedimentary and metavolcanic successions; 8: Tonian and Cryogenian
rift-related magmatic rocks; 9: Southern Brasília belt; 10: Pre-Neoproterozoic units; CTB: Central Tectonic Boundary; CFTD: Cabo Frio tectonic domain.

2
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

unraveling ancient plate margins. In the South American platform Providing new data for understanding the Rio Doce magmatic arc
(Fig. 1B) several orogenic belts of Brazil and Africa were shaped during evolution, we present additional bulk-rock chemistry and isotopic ana­
the amalgamation of the supercontinent Gondwana (Fig. 1A and B) as lyses (whole-rock Sm–Nd, and U–Pb and Lu–Hf in zircon) on samples
various plates converged and collided during Neoproterozoic/Cambrian collected from the Santa Tereza and Jequitibá suites in the Espírito Santo
times. The Brasiliano orogenic systems encompass distinct components state, disclosing the easternmost expression of pre-collisional magma­
of a Wilson cycle (rift-drift-subduction-collision-collapse) with the tism associated with the Araçuaí orogen (Fig. 1). The studied rocks
development of magmatic arcs from the Tonian to the early Ediacaran comprise charnockite, tonalite, and granodiorite, intruded in migmatitic
(Caxito et al., 2021, 2022; Cordani et al., 2016; Heilbron et al., 2004, paragneisses that represent an arc-related basin. The new results unravel
2017, 2020; Santiago et al., 2020; and references therein). the least evolved magmatism ever found in the Rio Doce magmatic arc.
Since the 1980’s, different assemblages of Ediacaran to Cambrian We also correlate our data with a thorough data compilation for arc-
magmatic rocks have been systematically mapped in the Araçuaí orogen related rocks found in the Araçuaí and Ribeira orogens and, alto­
of southeastern Brazil (Fig. 1), according to their distinctive field and gether, compare them with modern magmatic arcs. Finally, we discuss
structural relations, and petrographic, lithochemical, geochronologic, the geochronological and petrological implications of the Rio Doce
and isotopic features, resulting in the characterization of the major pre- magmatic arc evolution within the Araçuaí confined orogen in relation
collisional, collisional and post-collisional igneous and metaigneous to the Adamastor Ocean closing and Western Gondwana amalgamation.
supersuites (Table 1). Among them, the N–S-trending range of tonalitic-
granodioritic batholiths and stocks, crowded with dioritic to mafic en­ 2. Geological setting
claves, displaying calc-alkaline-magnesian I-type signatures, hybrid
isotopic attributes, and crustal inheritance, locally associated with The Araçuaí orogen of southeastern Brazil together with the West
metavolcano-sedimentary successions, compose the Ediacaran Rio Doce Congo Belt of central-southwestern Africa (Fig. 1) make up a Brasiliano –
magmatic arc (Fig. 1), developed from ca. 630 Ma to ca. 585 Ma Pan-African confined orogenic system (the AWCO) developed within an
(Gonçalves et al., 2018; Novo et al., 2018; Pedrosa-Soares et al., 2011, embayment of the São Francisco – Congo paleocontinent (Alkmim et al.,
2020; Tedeschi et al., 2016). 2006, 2017; Pedrosa-Soares et al., 2001, 2008; 2020). To the south along
Combined with petrographic and lithochemical studies, U–Pb ages the Mantiqueira Province (or South Atlantic Brasiliano Orogenic System;
coupled with Hf isotopic signatures from zircon grains are a powerful Heilbron et al., 2020; Caxito et al., 2022) (Fig. 1), the Araçuaí orogen
tool to unveil magmatic arc evolution (Ducea et al., 2015a; Milani et al., connects with the Ribeira Orogen (or Belt; Corrales et al., 2020; Heilbron
2015; Moyen et al., 2017; Kohanpour et al., 2019). Moreover, in addi­ et al., 2013, 2020; Tedeschi et al., 2016).
tion to magmatic arc rock studies, the sedimentary record of arc-related The Araçuaí orogen developed from a precursor basin system
basins can provide additional evidence for the history of a magmatic arc, evolved from continental rifts to passive margin and ophiolitic rock
since detrital zircon grains may provide complementary information on assemblages, indicating ocean-floor spreading before the generation of a
old and eroded arc-related sediments (Ducea et al., 2015a, 2015b; magmatic arc (the Rio Doce arc) triggered by subduction of the oceanic
Gradim et al., 2014; Novo et al., 2018; Paterson and Ducea, 2015; lithosphere during plate convergence, followed by continental collision
Pedrosa-Soares et al., 2020; Richter et al., 2016). and post-collisional gravitational collapse (Alkmim et al., 2006, 2017;

Table 1
Main characteristics of the G1 to G5 plutonic supersuites of the Araçuaí orogen (Sources of data compilation in Supplementary File A, the present paper; and references
therein).
Supersuite G1 G2 G3 G4 G5

U–Pb age (Ma) 630–585 585–540 540–500 530–480 530–480


Lithotypes mostly tonalite to granodiorite, mostly biotite-garnet alkali feldspar granite to garnet-bearing two- alkali feldspar granite to
minor diorite to gabbronorite, syenogranite to alkali feldspar syenogranite rich in mica granite, granodiorite, Opx-bearing
with biotite, amphibole and/or granite, minor garnet-rich cordierite and/or garnet, pegmatoid granite, charnockitic rocks, gabbro to
pyroxenes; generally garnet- monzogranite to tonalite and generally biotite-free minor biotite granite norite,
free garnet-two-mica granite ultramafic rock
Field relations tonalitic-granodioritic granitic batholiths, stratoid leucosomes in patches and intrusive plutons and intrusive balloon-shaped
batholiths and stocks generally bodies and stocks, showing solid- veins, and minor single stratoid bodies; plutons (stocks) to batholith-
rich in dioritic to mafic state deformation, stocks, with restites and minor batholiths; size composite bodies, free of
enclaves, showing solid-state metamorphism and schlieren of host rocks; free mostly free of regional ductile deformation,
deformation, metamorphism migmatization, with restites and of regional ductile regional ductile rich in magma mingling-
and migmatization, locally xenoliths of metasedimentary deformation; mainly hosted deformation mixing features, locally with
with well-preserved igneous rocks; locally with well- by G2 metagranites dioritic to mafic cores
fabrics; few single mafic preserved igneous fabrics
intrusions; associated with the
metavolcano-sedimentary Rio
Doce Group
Lithochemistry metaluminous to slightly peraluminous, peraluminous, peraluminous metaluminous to slightly
peraluminous (A/CNK <1.1), sub-alkaline to calc-alkaline sub-alkaline to high-K sub-alkaline (K > Na) peraluminous (A/CNK <1.1),
magnesian, calcic to alkali- alkaline to alkaline (Na > K) mostly ferroan, high-K calc-
calcic, medium- to high-K, alkaline to alkaline; minor
expanded calc-alkaline series tholeiite
87
Isotopic εNd(t): − 0.06 to − 13.6; εNd(t): − 4.05 to − 15; εHf(t): − 4 to − 10; Sr/86Sr(i): 0.713 εNd(t): − 2.5 to − 23.8;
signature Nd TDM: 1.19 to 2.26 Ga; Nd TDM ages: 1.33 to 2.29 Hf TDM: 1.3 to 1.7 Ga Nd TDM: 1.4 to 3.29
87
Sr/86Sr(i): 0.703 to 0.723; Ga;87Sr/86Sr(i): 0.707 to 0.713; Ga;87Sr/86Sr(t): 0.702 to 0.721;
εHf(t): +2.3 to − 36.4; εHf(t): − 1.8 to − 20.4; εHf(t): +13.5 to − 43.8
Hf TDM: 1.1 to 3.21 Ga Hf TDM: 1.3 to 1.9 Ga Hf TDM: 0.6 to 3.8 Ga
Genetic type mostly metaluminous I-type, mostly S-type, S-type S-type A-type and I-type
minor peraluminous I-type minor peraluminous I-type
Tectonic stage pre-collisional to early late pre-collisional to late late collisional to post- late collisional to post-collisional
and setting collisional, continental margin collisional collisional post-collisional
magmatic arc (Rio Doce arc)

3
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

Amaral et al., 2020; Araujo et al., 2020; Caxito et al., 2022; Pedrosa-­ calcic to alkali-calcic tonalitic to granodioritic rocks, generally with
Soares et al., 2001, 2008; 2020; Serrano et al., 2018). mafic to dioritic enclaves, representing granitic (s.l.) melts formed by
The basement comprises Archean and Paleoproterozoic infracrustal deep crustal anatexis with minor involvement of mantle magmas. The
assemblages, including gneissic-migmatitic, granitoid, and granulitic magmatic arc comprises multiple plutons composed of tonalite and
rocks of the Mantiqueira, Guanhães, Gouveia, Porteirinha, Juiz de Fora, granodiorite, locally represented by Opx-bearing rocks (enderbitic to
Quirino, and Pocrane complexes, as well as Statherian anorogenic charnockitic rocks), with variable amounts of mafic to dioritic enclaves,
granites, and supracrustal units, such as Archean greenstone belts, and some gabbroic-noritic intrusions (Tedeschi et al., 2016; Gonçalves
Siderian to Orosirian metavolcano-sedimentary and metasedimentary et al., 2018; Soares et al., 2020). All those rocks generally show ductile
units, and the continental rift (aulacogen) to sag units of the Statherian deformation and metamorphism, although well-preserved igneous fab­
to Stenian Espinhaço supergroup; all of them belonging to the São rics and magma mingling-mixing features are also found elsewhere
Francisco – Congo paleocontinent (Alkmim et al., 2017; Degler et al., along the magmatic arc (Gonçalves et al., 2016, 2018; Gouvêa et al.,
2018). 2020; Gradim et al., 2014; Narduzzi et al., 2017; Narduzzi, 2018;
The Araçuaí orogen shows three distinct tectonic stages, disclosed by Pedrosa-Soares and Wiedemann-Leonardos, 2000; Pedrosa-Soares et al.,
different rock assemblages: i) the pre-collisional stage (c. 630-585 Ma) 2011; Pereira et al., 2023; Soares et al., 2020; Tedeschi et al., 2016).
mainly marked by the I-type tonalitic-granodioritic batholiths rich in Most U–Pb ages from magmatic zircon grains range from ca. 630 Ma to
mafic enclaves of the Rio Doce magmatic arc; iii) the collisional stage (c. ca. 580 Ma, and zircon metamorphic ages cluster between ca. 592 Ma
585-540 Ma) associated with the regional tectonics and metamorphism, and ca. 560 Ma (Gonçalves et al., 2016, 2018; Gouvêa et al., 2020;
and production of a huge amount of S-type granites; and iii) the post- Narduzzi et al., 2017; Narduzzi, 2018; Pedrosa-Soares et al., 2011, 2020;
collisional stage (c. 530-480 Ma) associated with the generation of a Pereira et al., 2023; Tedeschi et al., 2016). Isotopic signatures, found in
myriad of granitic-charnockitic and gabbronoritic intrusions, cutting the literature prior to the present paper, show wide variation for
87
across the regional tectonic framework (Fig. 1 C). Accordingly, the Sr/86Sr(t) (0.703–0.723), εNd(t) (− 4.8 to − 13.6), Nd TDM model ages
orogenic plutonism has been grouped into five supersuites namely (1.36 Ga to 2,26 Ga), εHf(t) (− 2 to − 24.5), and Hf TDM (1.47–1.9 Ga)
(Table 1): the G1 supersuite (c. 630-585 Ma), comprising the plutonic (Gonçalves et al., 2016; Narduzzi, 2018; Soares et al., 2020; Tedeschi
igneous rocks formed from pre-collisional to early collisional stages in et al., 2016; Pedrosa-Soares et al., 2020, Pereira et al., 2023, and ref­
the Rio Doce arc; the G2 supersuite (c. 585-540 Ma), mostly including erences therein). Chemical signatures, zircon inheritances, and hybrid
the S-type, garnet – biotite ± cordierite ± sillimanite metagranites and isotopic attributes for the G1 rocks suggest interactions between the
granites formed from the late pre-collisional to tardi-collisional stages; Paleoproterozoic continental basement and variable mantle-derived
the G3 supersuite (540-500 Ma), represented by cordierite-garnet leu­ magmas in a continental margin setting.
cogranites formed after the partial melting on G2 metagranites in the In the study area (Fig. 2), the G1 supersuite rocks form elongated
late collisional to post-collisional stages; and the G4 and G5 supersuites, intrusions comprised in the Santa Tereza and Jequitibá suites (Féboli,
encompassing a number of intrusions emplaced in the post-collisional 1993; Figueiredo and Campos Neto, 1993; Leite et al., 2004; Silva et al.,
stage (Aranda et al., 2020; Araujo et al., 2020; De Campos et al., 2004; Tuller, 1993; Vieira et al., 2015), intruding the Nova Venécia
2016; Deluca, 2018; Gonçalves et al., 2016, 2018; Gradim et al., 2014; complex (Fig. 2). The Santa Tereza and Jequitibá suites consist of
Melo et al., 2017, 2020; Pedrosa-Soares et al., 2011, 2020; Serrano et al., gabbroic to granodioritic orthogneisses (Fig. 4D) with or without
2018; Soares et al., 2020; Tedeschi et al., 2016). orthopyroxene, garnet, and amphibole.
The Ediacaran Rio Doce magmatic arc and related orogenic basins The Nova Venécia complex covers a large area in the focused region
extend at least 800 km from the northern sector of the Araçuaí orogen, at (Figs. 1 and 2), representing the sedimentary infill of an arc-related
latitude 17◦ 30’ N, to the central Ribeira Orogen (Fig. 1). It includes basin, largely located on the back-arc zone of the Rio Doce magmatic
plutonic rock assemblages and related orogenic basins (Pedrosa-Soares arc (Noce et al., 2004; Gradim et al., 2014; Richter et al., 2016; Tedeschi
et al., 2020), represented by the G1 supersuite of the Araçuaí orogen et al., 2016). Consisting of migmatitic Al-rich paragneisses, with in­
(Gonçalves et al., 2014, 2016; 2018; Gouvêa et al., 2020; Pedrosa-Soares tercalations of cordierite granulites and calc-silicate rocks, the Nova
et al., 2011; Soares et al., 2020; Tedeschi et al., 2016), and the Venécia paragneisses represent pelitic to wacky sedimentary protoliths,
Marceleza-Leopoldina-Serra da Bolívia complex of the Ribeira orogen and the calcsilicate rocks formed from pelite-carbonate (marl) deposits
(Corrales et al., 2020; Heilbron et al., 2013, 2020), as well as the (Gradim et al., 2014). The cordierite granulites, generally bearing silli­
arc-related supracrustal successions included in the Rio Doce and manite, hercynite, and garnet, are poor to free in biotite, and represent
Andrelândia groups, the Nova Venécia complex, and mélange-related the highest-grade metamorphic rocks from the metamorphism of Nova
rock assemblages (Araujo et al., 2020; Degler et al., 2017; Gradim et al., Venécia paragneisses (Gradim et al., 2014; Richter et al., 2016). Quan­
2014; Noce et al., 2004; Novo et al., 2018; Peixoto et al., 2015; Richter titative data constrain the regional (collisional) metamorphism on the
et al., 2016; Schannor et al., 2019). In the Araçuaí orogen, an evolution Nova Venécia Complex from 712 ◦ C to 930 ◦ C at 5.0 kbar–7.5 kbar, i.e.,
model for the Rio Doce arc includes oceanic lithosphere subduction from high-grade amphibolite facies to granulite facies, between 580 and 550
west to east, inducing arc plutons rising, followed by slab retreating Ma (Gradim et al., 2014; Munhá et al., 2005; Richter et al., 2016). U–Pb
driven magmatic migration from east to west, subduction of the oceanic analyses on detrital zircon grains from rocks of the Nova Venécia com­
ridge and, finally, slab break-off after continental collision (Tedeschi plex yield ages from 2597 Ma to 582 Ma, outlining a rather unimodal age
et al., 2016). Comparing field, petrographic, geochemical, and isotopic spectrum with the main peak at 600 Ma, and εHf(t) from − 39 to +4 with
attributes, the Rio Doce magmatic arc shows great similarity with HfTDM from 3.5 Ga to 1.2 Ga (Araujo et al., 2020; Belém, 2014; Gradim
several continental magmatic arcs found elsewhere in the world (Gon­ et al., 2014; Noce et al., 2004; Pedrosa-Soares et al., 2020; Richter et al.,
çalves et al., 2018; Narduzzi et al., 2017; Narduzzi, 2018), such as the 2016). Whole-rock samples from the peraluminous paragneisses show
87
Famatinian magmatic arc (Chile-Argentina; Otamendi et al., 2012; Sr/86Sr(t) between 0.708 and 0.71938, and εNd(t) of − 8.14 with NdTDM
Pankhurst et al., 1998), North and South Patagonian Batholith (Chi­ at 1.88 Ga (De Campos et al., 2004; IBGE, 1987; Munhá et al., 2005;
le-Argentina; Castro et al., 2011; Hervé et al., 2007; Pankhurst et al., Tuller, 1993). The available dataset suggests that the Rio Doce arc is,
1999), Western and Eastern Peninsular Ranges (Western USA; Lee et al., indeed, the most important sediment source for the Nova Venécia
2007), Sierra Nevada Batholith (Western USA; Bateman et al., 1984; complex, with minor contributions from older sources, such as the Serra
Cecil et al., 2012; Chapman et al., 2012; Dodge et al., 1982), and Nar­ da Prata and Rio Negro magmatic arcs of the Ribeira Orogen (Heilbron
sajuaq terrane (Canadá; Dunphy and Ludden, 1998). and Machado, 2003; Tupinambá et al., 2012; Lobato et al., 2015; Peix­
The G1 supersuite mostly encompasses medium-to-high-K calc- oto et al., 2017, 2022; Heilbron et al., 2020), the Southern Bahia
alkaline, magnesian, metaluminous to slightly peraluminous I-type, Alkaline Province (Rosa et al., 2007), the Salto da Divisa Suite (ca.

4
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

Fig. 2. Simplified geological map of the study area, showing the analyzed samples locations (modified from Leite et al., 2004; Silva et al., 2004; Vieira et al., 2015).
*Zircon U–Pb and Lu–Hf isotopic analyses by Araujo et al. (2020).

915-880 Ma, Silva et al., 2008; Victoria et al., 2022) and basement 18◦ 30′ S), Central (from 18◦ 30′ to 20◦ S) and Southern (South of 20◦ S)
complexes (Degler et al., 2018). sectors, and Western (West of 41◦ W) and Eastern (East of 41◦ W) sec­
Under high-amphibolite to granulite facies conditions, anatexis tors. Geographical coordinates refer to the present-day position of con­
events on the Nova Venécia complex formed a huge amount of tinents, even in relation to paleocontinental descriptions. Located in the
autochthonous to allochthonous peraluminous (S-type) granitic melts, Eastern and Southern-Central sectors, the study area includes rocks of
crystallized as garnet-biotite granites comprised by the Ataléia and the pre-collisional, collisional, and post-collisional orogenic stages,
Carlos Chagas suites of the G2 supersuite, during the late pre-collisional intruded in the Nova Venécia complex (Fig. 2).
to late collisional stages of the Araçuaí orogen (Araujo et al., 2020; Our main focus is on intrusions comprised by the Santa Tereza and
Gradim et al., 2014; Melo et al., 2017; Mauri et al., 2021; Pedrosa-Soares Jequitibá suites (Féboli, 1993; Figueiredo and Campos Neto, 1993; Leite
et al., 2011, 2020; Richter et al., 2016). et al., 2004; Silva et al., 2004; Tuller, 1993; Vieira et al., 2015).
Representative samples of these suites were selected, after field and
3. Materials and methods petrographic studies, to be analyzed for bulk-rock chemistry and iso­
topic Sm–Nd, and zircon U–Pb and Lu–Hf methods. The terms granite
In this work, the Araçuaí orogen was divided into Northern (North of and granitic rocks are used in a general sense for felsic igneous rocks rich

5
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

Fig. 3. Photos and photomicrographs (under transmitted light and polarized light) showing mineralogical and textural features. CA-08 sample: A, B, C. (A) Greenish
gray, coarse-grained biotite metagranodiorite, from Santa Tereza suite. (B and C) Photomicrographs showing the concentration of felsic minerals elongated according
to bands and displaying undulatory extinction. Plagioclase displays irregular twinning and antiperthitic intergrowths (C); CA-14 sample: D, E, F. (A) Greenish gray,
coarse-grained garnet-orthopyroxene-biotite-bearing charnockitic metagranodiorite from Santa Tereza suite. (E and F) Photomicrographs showing local concen­
tration of biotite, orthopyroxene, amphibole, and garnet. Plagioclase displays irregular twinning and antiperthitic intergrowths (F). CA-03 sample: G, H, I. (G) Fine to
coarse-grained garnet-biotite metagranodiorite from Jequitibá suite; (H and I) Photomicrographs showing stretched felsic minerals and poikiloblastic garnet. CA-31
sample: J, K, L. (J) Coarse-grained garnet-biotite metagranodiorite from Jequitibá suite. (K and L) Photomicrographs showing plagioclase with irregular twinning and
antiperthitic intergrowths, and biotite oriented according to foliation.

in quartz unless otherwise specified. Mineral abbreviations are from analyzed by X-Ray fluorescence (XRF79C) after fusion with lithium
Whitney and Evans (2010). tetraborate, with detection limits of 0.01%–0.1%. Trace element con­
Our new analytical data are available in Supplementary Files A and centrations were analyzed via inductively coupled plasma mass spec­
B. The selected samples are homogenous and free of weathering coats, trometry (ICP-MS) after fusion with lithium tetraborate and digestion
fracture fillings, veins, and any hydrothermal alteration evidence. All of with a multi-acid solution (HCl, HNO3, HF, and HClO4), and detection
them were properly cleaned before the analytical procedures. The limits of 0.05–10 ppm. The loss on ignition (LOI) was calculated by
samples were crushed and milled at the SEPURA laboratory of the weighting difference after 1000 ◦ C heating. The analytical results were
CPMTC-UFMG research center, Federal University of Minas Gerais, Belo organized, evaluated, and interpreted by using the GCDKit software
Horizonte, Brazil. Bulk-rock geochemistry (major and trace elements (Janoušek et al., 2006). They are presented in Supplementary File A,
including Rare-Earth Elements - REE) analyses were performed in the Table 2, and Figs. 4–7.
SGS-Geosol Laboratories (Vespasiano, Brazil). Major elements were For isotopic analyses on zircon grains, about 3 kg of each rock sample

6
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

Fig. 4. Lithochemical data for the studied samples compared to the compiled dataset. A: FeOt/(FeOt + MgO) vs. SiO2 (wt%) (Frost et al., 2001); B: Na2O + K2O –
CaO vs. SiO2 (wt%) (Frost et al., 2001); C: Alumina saturation Index (Frost et al., 2001). Dashed line (ASI = 1.1) delimits the transition between I- and S-type granites
according to Chappell and White (2001); D: R1-R2 diagram (De la Roche et al., 1980). Sources of data compilation in Supplementary File A.

were prepared by crushing and milling, followed by heavy minerals Zircon U–Pb isotopic analyses were carried out using a Thermo­
concentration by conventional methods (e.g., dense liquid and magnetic Scientific Element II sector field coupled to a PhotonMachine ArF 193
separator) in the SEPURA laboratory (CPMTC-UFMG) and at the CPGeo mm laser ablation system (LA–SF–ICP-MS) at the Isotope Geochemistry
of the University of São Paulo. Disregarding color, shape, or size, zircon Laboratory of the DEGEO-UFOP. Laser ablations were performed in peak
grains were handpicked under a binocular microscope and mounted in jumping mode during 20s background measurement followed by 20s
epoxy mounts at the Isotopic Geochemistry Laboratory of UFOP (Federal sample ablation, spot-size 30 μm, 1,9 J/cm2 fluence, 6 Hz repetition
University of Ouro Preto). After polishing the mounts to expose grain rate, and carried out using He as a carrier gas mixed with Ar before
centers, they were imaged under cathodoluminescence (CL), secondary introduction into the ICP-MS. As the primary standard was used the BB-9
electron (SE), and the backscattered electron (BSE) in a JEOL 6510 zircon (Santos et al., 2017) and, for quality control, the Plesovice (Sláma
Scanning Electron Microscope hosted at Microscopy and Microanalysis et al., 2008) as a secondary standard. The software Glitter (van Ach­
Laboratory of the Federal University of Ouro Preto. These images were terbergh et al., 2001) was employed to reduce the signal data, and for
obtained to examine morphological features, internal textures, and the correct background signal, common Pb, laser-induced elemental frac­
presence of inclusions, cracks, and damaged areas to assist in the tionation, instrumental mass discrimination, and time-dependent
discrimination of grain regions for locating the U–Pb and Lu–Hf spot elemental fractionation of Pb/U was used a MS Excel spreadsheet pro­
analyses. gram (Gerdes and Zeh, 2006). The common-Pb correction was based on

7
C. Araujo et al.
Table 2
Summary of the main features of the studied from samples with whole rock Sm–Nd analyses.
Suite Santa Tereza Jequitibá

Sample CA- 08 CA -14 CA-17 CA-18 A CA-19 CA-20 A CA -03 CA-31 CA-32

Coordinates S19.935◦ / S19.858◦ /W40.516◦ S19.933◦ / S19.933◦ /W40.468◦ S19.937/W40.483 S19.934/ S20.412/W40.859 S20.047/W40.699 S20.045/W40.723
W40.499◦ W40.466◦ W40.501
Location Santa Teresa, Santa Teresa, Fundão, Espírito Fundão, Espírito Santo Santa Teresa, Espírito Santa Teresa, Marechal Floriano, Santa Leopoldina, Santa Maria de
Espírito Santo state Espírito Santo state Santo state state Santo state Espírito Santo Espírito Santo state Espírito Santo state Jetibá, Espírito
state Santo state
Field Feature Greenish gray rock, Greenish gray rock, Greenish gray rock, Greenish gray rock, Light gray, with fine to Greenish gray Foliated with Foliated with Foliated with
coarse-grained, coarse-grained, with coarse-grained, coarse-grained, with coarse-grained levels, rock, coarse- milimetric garnet, light milimetric garnet, milimetric garnet,
composed of felsic millimetric garnet. with high high concentration of with concentration of grained, gray, fine to coarse- light gray, coarse- light gray, coarse-
and mafic bands Locally, the minerals concentration of biotite. Locally, the mafic minerals, and composed of grained, showing local grained, showing grained, showing
are oriented biotite. Locally, the minerals are oriented. local small felsic felsic and mafic small felsic patches local small felsic local small felsic
minerals are patches very coarse- bands with core of garnet up patches patches
oriented grained. to 2 cm in diameter
Lithotype Bt Grt-Opx-Bt Opx-Bt metadiorite Opx-Cpx-Bt Bt metagranodiorite Grt-Opx-Bt Grt-Bt Grt-Bt Grt-Bt
sample metagranodiorite metagranodiorite metamonzogabbro metatonalite metagranodiorite metagranodiorite metatonalite
Lithochemical Magnesian, calcic, Magnesian, calcic, Magnesian, calc- Magnesian, alkali- Magnesian, calcic, Magnesian, Magnesian, calcic, Magnesian, calcic, Magnesian, calc-
signature slightly slightly alkalic, calcic, metaluminous, slightly peraluminous, calcic, slightly slightly peraluminous, slightly alkalic, slightly
peraluminous, calc- peraluminous, calc- metaluminous, high-K calc-alkaline calc-alkaline serie peraluminous, calc-alkaline serie peraluminous, calc- peraluminous,
alkaline serie alkaline serie high-K calc-alkaline serie calc-alkaline alkaline serie high-K calc-
serie serie alkaline serie
Main Pl-Qz-Bt-Kfs Pl-Qz-Bt-Opx-Kfs-Grt Pl-Qz-Bt-Opx-Kfs- Pl-Bt-Opx-Cpx-Qz- Pl-Qz-Bt-Kfs Pl-Qz-Bt-Opx-Grt Pl-Qz-Kfs-Bt-Grt Pl-Qz-Kfs-Bt-Grt Pl-Qz-Kfs-Bt-Grt
mineralogy* Amp Amp
Acessory Opq-Zrn-Ap-Mnz- Zrn-Ap-Mnz-Aln-Ep Zrn-Ap-Mnz Zrn-Ap-Kfs Zrn-Ap-Mnz Opq-Kfs-Zrn-Ap- Zrn-Ap-Mnz-Aln Zrn-Ap-Mnz-Aln- –
mineralogy Ru-Aln Mnz-Ttn Opq
8

Secondary Ser-Ms-Chl-Cb-Ru Amp-Opq-Sr-Ms-Bt- Amp-Opq-Sr-Ru-Bt- Amp-Opq-Bt-Chl-Cb Ser-Ms-Opq-Chl-Cb-Ru Amp-Opq-Bt-Chl Ser-Ms-Chl-Opq Ser-Ms-Chl –


mineralogy Chl-Cb Chl-Cb
Sm (ppm) 13.429 5.656 11.819 10.340 4.401 5.321 3.999 7.126 5.319
Nd (ppm) 61.513 27.983 65.545 50.155 30.817 21.642 24.131 39.061 24.374
147
Sm/144Nd 0.1320 0.1222 0.1090 0.1246 0.0863 0.1486 0.1002 0.1103 0.1319
143
Nd/144Nd 0.512103 0.511976 0.512075 0.512344 0.511995 0.512166 0.51181 0.511958 0.512030
± 2SE 5 5 8 7 6 6 5 3 8
TDM (Ga) 1.74 1.76 1.40 1.19 1.25 2.04 1.64 1.59 1.87
eNd (T1)** − 5.34 − 7.05 − 4.09 − 0.06 − 3.87 − 5.42 − 8.8 − 6.67 − 6.91
eNd (T2)*** − 5.15 − 6.84 − 3.83 +0.16 − 3.55 − 5.27 − 8.43 − 6.34 − 6.66
176
Hf/177Hf (T) 0.282457 to – – – – – 0.282359 to 0.282292 – –

Journal of South American Earth Sciences 129 (2023) 104490


0.282271 (from 545 (from 562 to 624 Ma)1
to 635 Ma) 0.282184 to 0.281375
(from 560 to 607 Ma)2
176
Hf/177Hf (T2) 0.282456 to – – – – – 0.282359 to 0.282282 – –
1
*** 0.282271
0.282183 to 0.281375
2

ƐHf(T) +2.3 to − 4.8 – – – – – − 1.7 to − 4.9 1 – –


− 8.0 to − 36.4 2
ƐHf(T2)*** +2.5 to − 4.6 – – – – – − 1.2 to − 3.9 1 – –
− 7.4 to − 36.0 2
T1 (Ma) 612 612 612 612 612 612 594 594 594
T2 (Ma) 635 635 635 635 635 635 624 624 624

* in decreasing order of abundance.


Aln: allanite; Amp: amphibole; Ap: apatite; Bt: biotite; Cb: carbonate; Chl: chlorite; Ep: epidote; Grt: garnet; Kfs: K-feldspar; Mnz: monazite; Ms: muscovite; Opq: opaques minerals; Opx: orthopyroxene; Cpx: clinopyroxene
Pl: plagioclase; Qz: quartz; Ser: sericite; Ru: rutile; Ttn: titanite; Zrn: zircon. Lu–Hf data in zircon **T1: Concordia age ***T2 estimated crystallization age 1 - group (1); 2- group (2).
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

Fig. 5. Lithochemical data for the studied samples


compared to available data from G1 and G2 super­
suites of Araçuaí orogen, Marceleza-Leopoldina-Serra
da Bolívia complex and Rio Negro arc system from
Ribeira orogen, and other classical continental
magmatic arcs. A and C: Geotectonic discriminating
diagrams by Pearce et al. (1984); B: AFM diagram
(Irvine and Baragar, 1971); D: Hf–Rb/30-Ta*3
ternary diagram (Harris et al., 1986); E: Na2O + K2O
vs. SiO2 (wt%); F: Source Diagram (Laurent et al.,
2014). Meaning fields: ORG, Ocean Ridge Granites;
VAG or VA, Volcanic Arc Granites; WPG or WP,
Within Plate Granites, COLG, Collision Granites;
Group 2, Syn-collision peraluminous intrusions (leu­
cogranites); Group 3, Late or post-collision calc-al­
kaline intrusions. Sources of data compilation in
Supplementary File A.

the interference- and background-corrected 204Pb signal and the Pb static mode during 60 s of ablation and the 50 μm spot-size laser was
composition model (Stacey and Kramers, 1975). U–Pb spots cover grain drilled with at 9 Hz repetition rate. The isotopes 172Yb, 173Yb, and 175Lu
rims and cores, with <1% of discordance, low common Pb, and suitable were simultaneously monitored during each analysis step to allow for
relations of isotopic ratios were selected for calculations and plotting to correction of isobaric interferences of Lu and Yb isotopes on 176 mass.
obtain the Concordia age distribution diagrams in Isoplot/Ex 4.15 Corrections for background signal, instrumental mass bias, and isobaric
(Ludwig, 2012). Unless otherwise indicated, we use the 206Pb/238U ages. interferences of Lu and Yb isotopes on 176 mass were done following the
The results are presented in Supplementary File B, Table 2, Figs. 8 and 9. methods by Gerdes and Zeh (2006, 2009). The decay constant for 176Lu
Representative grains with different U–Pb ages, morphologies, and in­ is 1.865 × 10− 11 a− 1 (Söderlund et al., 2004) and the 176Lu/177Hf ratio is
ternal textures were selected for Lu–Hf isotope analyses. 0.0113 to the average crust in the Paleoproterozoic to Archean (Scherer
Lu and Hf isotopic ratios were obtained at the Isotope Geochemistry et al., 2001). For accuracy and external reproducibility, the analyses
Laboratory of the DEGEO-UFOP using a Thermo-Scientific Finnigan were verified by BB-9 (Santos et al., 2017), GJ-1 (Jackson et al., 2004),
Neptune Multi-collector inductively coupled to a PhotonMachine ArF and Plesovice (Sláma et al., 2008) standard zircon, which yielded ratios
193 mm laser ablation system (LA-MC-ICP-MS). The zircon spots were are in good agreement with the accepted values, with an average of
176
drilled on top of the previous U–Pb laser spot in order to analyze zircon Hf/177Hf of 0.281675 ± 0.000018 (n = 17), 0.282000 ± 0.000018 (n
domains with the same textures. Where this was not possible, care was = 17), 0.282471 ± 0.000020 (n = 17) respectively (errors in ±2σ). The
taken to drill into the same zircon domain previously analyzed for U–Pb, results are presented in Supplementary File B, Table 2, Figs. 9–11.
characterized by the CL, SE, and BSE images. Data were collected in Sm–Nd isotopic rations were obtained at the Isotope Geochemistry

9
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

Fig. 6. A) Multi-elemental diagram normalized to primitive mantle, according


to Sun and McDonough (1989); B) Chondrite-normalized REE patterns,
normalized according to Boynton (1984). Sources of data compilation in Sup­
plementary File A.
Fig. 7. Average compositions from Santa Tereza and Jequitibá samples, and
compiled data with SiO2 > 54%. A) Multi-elemental diagram normalized to the
Laboratory of the University of Brasília, on a multi-collector TRITON –
primitive mantle, according to Sun and McDonough (1989); B)
Thermo Finnigan mass spectrometer in static mode and methodology as
Chondrite-normalized REE patterns, according to Boynton (1984). Sources of
described by Gioia and Pimentel (2000). Whole-rock samples (ca. 50 mg data compilation in Supplementary File A.
powdered) were mixed with 149Sm-150Nd spike solution and dissolved in
HF, HNO3, and HCl in Savillex capsules. Cation exchange techniques
In addition, we present a comparative study of our results with an
were employed for Sm and Nd extraction from dissolved whole-rock
updated and wide dataset, including isotopic U–Pb, Lu–Hf and Sm–Nd,
samples using Teflon columns containing LN-Spec resin (HDEHP-di-
and lithochemical data, compiled from previous studies on arc-related
fractions ethylhexil phosphoric acid supported on PTFE powder). Sm
rocks from the Araçuaí and Ribeira orogens, as well as with other clas­
and Nd were loaded in Re evaporation filaments in a double filament
sical continental magmatic arcs found elsewhere around the world
assembly. Uncertainties for Sm/Nd and 143Nd/144Nd ratios are better
(Figs. 4–12). The lithochemical from plutonic rocks of the Andean,
than ±0.5% (2σ) and ±0.005% (2σ), respectively, based on repeated
North American, Cascades, Central American, Sierra Nevada, and the
analyses of international rock standards BHVO-2 and
Trans-Mexican volcanic arcs were compiled from Castro et al. (2011),
BCR-1.143Nd/144Nd ratios were normalized to 146Nd/144Nd of 0.7219.
Cecil et al. (2012), Hervé et al. (2007), Pankhurst et al. (1999) and
The De Paolo (1981) model was used to calculate TDM ages. The results
Rodríguez et al. (2018), as well as from the PetDB Database (www.
are presented in Supplementary File B, Table 2, Fig. 9G and 9H.
earthchem.org/petdb) and GEOROC (http://georoc.mpch-mainz.gwdg.

10
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

Fig. 8. (A and B) Concordia diagrams, weighted mean ages, and representative zircon grains from Santa Tereza and Jequitibá samples. The U–Pb ages are reported as
206
Pb/238U age in Ma, with errors at 2σ.

de/georoc/). The compiled dataset of zircon U–Pb ages encompasses


spot zircon ages with less than 10% of concordance and excludes
inherited zircon grains. When spot ages were not available, published
Concordia ages were compiled.

11
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

Fig. 9. (A) Th vs U diagram, (B) Th/U vs U–Pb age (C and D) Lu–Hf data calculated for individual zircon ages, and (E and F) Lu–Hf data calculated for estimated
crystallization ages from Santa Tereza and Jequitibá samples. (G and H) Diagrams comparing the Nd isotopic signature from studied samples with the compiled
dataset. DM = depleted mantle; CHUR = chondrite uniform reservoir. T1: Concordia age; T2: Estimated crystallization ages. Sources of data compilation: De Campos
et al. (2004)1; Figueiredo (2009)1; Gonçalves (2009)1; Gonçalves et al. (2016)1; Heilbron et al. (2013)2; Martins et al. (2004)1; Nalini Jr. (1997)1; Nalini et al.
(2000)1; Novo et al. (2010)1; Pedrosa-Soares et al. (2011)1; Peixoto et al. (2017)3,4; Pereira et al. (2023)1; Santiago et al. (2020)4; Soares et al. (2020)1; Tedeschi et al.
(2016)1; Tupinambá et al. (2012)3.

12
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

Fig. 10. Diagrams comparing the U–Pb and Lu–Hf data from studied samples with the compiled dataset. Northern sector: North of 18◦ 30’; Central sector from 18◦ 30′
to 20◦ ; Southern sector: South of 20◦ ; Western sector: West of 41◦ ; Eastern sector: East of 41◦ ; * Jequitibá and Santa Tereza data, respectively; DM = depleted mantle;
CHUR = chondrite uniform reservoir. Black dashed lines classify fields of juvenile (0–5 ε-units below DM), moderately juvenile (5–12 ε-units below DM) and evolved
(>12 ε-units below DM) composition (Bahlburg et al., 2011). Sources of data compilation: Araujo et al. (2021)8; Araujo et al. (2020)1,5; Castro et al. (2020)7; Corrales
et al. (2020)3; Degler et al. (2017, 2018)7,8; Gonçalves et al. (2016)2; Kuribara et al. (2019)8; Narduzzi (2018)2; Pereira et al. (2023)2; Santiago et al. (2020)4,
Santiago et al.(2022)7; Schannor et al. (2019)2,6, (2020)2,8; Tedeschi et al. (2016)2. In A and B diagrams, only magmatic zircon grains were plotted.

4. Results or polygonised sometimes. Some plagioclase crystals exhibit myrmekitic


intergrowths, compositional zoning, and conspicuous to irregular
4.1. Petrography and textural relationships deformed twins (Fig. 3C). Antiperthitic texture is omnipresent in all
samples (Fig. 3C and F). Subgranulation and undulatory extinction are
4.1.1. Santa Tereza suite common in quartz and feldspars presented evidence of solid-state
The Santa Tereza rocks have a greenish and, if somewhat weathered, deformation. Commonly oriented, biotite presents dark brown to light
a light gray color. They are medium-to coarse-grained with local alter­ greenish-brown pleochroism and occasionally is a product of pyroxene
nations of compositional bands (Fig. 3A and D). Most samples are and/or amphibole breakdown (Fig. 3E). K-feldspar is a microcline and
garnet-bearing to garnet-free, locally biotite-bearing, foliated meta­ trichinized orthoclase with a frequent perthitic texture. Pyroxene has
charnockitic rocks mostly with tonalitic to granodioritic compositions, colorless to light greenish-brown or pale pink to light green pleochroism,
and minor dioritic and monzogabbroic terms. They consist essentially of presenting intergrows with biotite and amphibole, as well as moderate
plagioclase (40–55%), quartz (20–35%), biotite (10–20%), and K-feld­ uralitization (Fig. 3E). Amphibole presents brown to greenish pleoch­
spar (5–10%), including orthopyroxene (5–10%), clinopyroxene roism and occasionally is a product of pyroxene breakdown. Garnet
(0–10%), amphibole (0–5%) and garnet (0–5%) in the charnockitic shows anhedral to subhedral habitus, light pink color, alteration into
rocks. Accessory phases include opaque minerals, zircon, apatite, chlorite, and inclusions of quartz, biotite, zircon, plagioclase, and
monazite, allanite, rutile, epidote, and titanite. Carbonate, amphibole, apatite (Fig. 3E). There are also garnets free of inclusions.
biotite, opaque minerals, sericite, rutile, epidote, muscovite, chlorite,
and clay minerals are secondary minerals. Although magmatic features 4.1.2. Jequitibá suite
can be found, they mainly present granoblastic to granolepidoblastic The Jequitibá rocks are light gray to greenish, fine to coarse-grained,
textures, with the contacts between the grains being irregular, polygonal isotropic to foliated (Fig. 3G and J), and locally show small, coarse-
and straight, and interlobate. The foliation is given by the orientation of grained, felsic patches with garnet up to 2 cm in diameter (Fig. 3G). It
stretched minerals in both felsic and mafic minerals bands and laminae is also possible to observe alternations of compositional bands and the
(Fig. 3A–F). Quartz is anhedral and mostly interstitial, occurs in ribbons presence of mafic enclaves and xenoliths of migmatitic paragneiss. They

13
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

Fig. 11. Diagrams shown in-zircon Hf isotope data from Santa Tereza and
Jequitibá samples, compared with compiled data from Nova Venécia para­
gneisses. B) Detail outlook showing the Hf signature spreading along approxi­
mately horizontal lines and indicating progressive Pb loss.

Fig. 12. Histograms and probability density curves showing zircon U–Pb ages
comprise plagioclase (45–60%), quartz (15–30%), K-feldspar (5–15%), distribution. A: Studied samples and compiled dataset of the orogenic pluton­
biotite (10–15%), and garnet (0–1%). Accessory phases include opaque ism, metamorphism, and arc-related basins (Andrelândia and Rio Doce groups,
minerals, zircon, apatite, monazite, and allanite. Carbonate, biotite, Salinas formation, and Mpioka formation) from Araçuaí – West Conto orogen.
opaque minerals, sericite, muscovite, chlorite, and clay are secondary B: Igneous zircon U–Pb ages from Rio Doce magmatic arc (including new data)
alteration minerals. The samples consist of foliated diorite to granodi­ for Northern to Southern sectors. C: Igneous zircon U–Pb ages from Rio Doce arc
orite, with granoblastic to granolepidoblastic textures and locally pre­ in Araçuaí orogen (including new data) for Western and Eastern sectors.
served magmatic fabrics. The foliation is defined by the orientation of Northern sector: North of 18◦ 30’; Central sector from 18◦ 30′ to 20◦ ; Southern
stretched minerals along felsic and mafic bands (Fig. 3I and K), with the sector: South of 20◦ ; Western sector: West of 41◦ ; Eastern sector: East of 41◦ .
contacts between the grains being irregular, polygonal and straight, and Sources of data compilation in Supplementary File A.
interlobate. Commonly, quartz and feldspars show elongated crystals
with subgrain patterns and undulatory extinction (Fig. 3I). Plagioclase 4.2.1. Lithochemistry
exhibits antiperthitic and myrmekitic intergrowths, compositional The samples from the Santa Tereza and Jequitibá suites have
zoning, and conspicuous to irregular, deformed twins (Fig. 3L). K-feld­ granodioritic to monzogabbroic compositions, with calc-alkaline,
spar is microcline and trichinized orthoclase, with a typical perthitic magnesian, calcic to alkali-calcic, I-type, and slightly peraluminous
texture. Usually oriented along the foliation, the biotite presents dark (ASI <1.1) signatures, representing volcanic arc magmas related to
brown to light greenish-brown pleochroism. Garnet has an anhedral mafic sources (Figs. 4 and 5). They show enrichment in LILE (Large Ion
habitus, light pink color, and inclusions of rounded quartz and monazite, Lithophile Elements), and negative Ba, Nb, Ti, and P anomalies, with
and of elongated biotite and apatite (Fig. 3H and K). positive Th anomalies in silica-rich (acid) samples, and negative Th–U
and Ti anomalies in basic to intermediate samples (Fig. 6). One sample
with intermediate silica content from the Jequitibá suite presents a
4.2. Lithochemistry, geochronology, and isotope geochemistry positive Ba anomaly (Fig. 6). The REE patterns show LREE enrichment
(LaN/YbN (intermediate to basic) = 12 to 49; LaN/YbN (acid) = 8–87) with
Besides lithochemical analyses on thirteen (13) bulk-rock samples of slightly negative to positive Eu anomalies (Eu/Eu* 0.76 and 1.12),
the Santa Tereza and Jequitibá suites, we also present U–Pb geochro­ except for the CA-08 sample presenting a prominent negative Eu
nology and isotopic Lu–Hf data for two (02) zircon concentrates and anomaly (Eu/Eu* 0.32) (Fig. 6B; Supplementary File A).
isotopic Sm–Nd results from nine (09) whole-rock samples, of both
suites.

14
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

4.2.2. Zircon U–Th–Pb and Lu–Hf data 1.33 Ga to ca. 1.48 Ga. The other group shows greater values variation
with 176Hf/177Hf(t) ratios ranging from 0.282184 to 0.281375, εHf(t)
4.2.2.1. Santa Tereza suite. A sample of a biotite metagranodiorite was between − 8.0 and − 36.4, and HfTDM model ages from ca. 1.33 Ga to ca.
selected for U–Pb and Lu–Hf in-zircon analyses (CA-08; Figs. 2 and 3A). 1.48 Ga (Figs. 9–11). Hf isotopic analyses on older zircon grains (six
The zircon population is light brown to colorless and translucent, spots) yielded initial 176Hf/177Hf ratios ranging from 0.282348 to
exhibiting short to long prismatic habitus with subhedral to euhedral 0.281895, εHf(t) values between +0.6 and − 9.3, and HfTDM model ages
terminations (Fig. 8A). Inclusions of apatite and plagioclase are common from ca. 1.31 Ga to ca. 2.04 Ga (Figs. 9–11).
in zircon crystals. Grains range from 100 to 600 μm in length, with a
length/width ratio from 2 to 11. CL-images display non-zoned to thin 4.2.3. Sm–Nd data
and parallel magmatic oscillatory zoning, prevailing medium to high Nine whole-rock samples were analyzed to obtain isotopic Sm–Nd
luminescence (Fig. 8A). Some crystals have a low to dark luminescence compositions, including six from the Santa Tereza suite and three from
(Fig. 8A, zircon grains 008, 066). Several crystals are surrounded by thin the Jequitibá suite (Table 2). Initial 143Nd/144Nd ratios and εNd(t) values
structureless domains of dissolution–recrystallization and overgrowth were calculated using the Concordia age at 612 Ma for the Santa Tereza
(Fig. 8A, zircon grains 053, 059). Oscillatory zoning is absent in BSE rocks and 594 Ma for the Jequitibá. The results show 143Nd/144Nd ratios
images. between 0.51234 and 0.51181, and NdTDM model ages from 1.19 Ga and
Fifty-three U–Pb spot data plot from 635 ± 9 Ma to 545 ± 7 Ma in the 2.04 Ga with εNd(t) values ranging from − 3.88 to − 8.80 for metadiorite
Concordia diagram, comprising two main age groups (Figs. 8A, 9A and to metatonalite samples. The metamonzogabbro sample has εNd(t) of
9B). The older cluster furnished ages ranging from 630 ± 8 Ma to 593 ± − 0.06 and NdTDM model age of 1.19 Ga (Table 2). Garnet-bearing
8 Ma, with Th/U ratios from 1.8 to 0.1, resulting in a Concordia age of samples present more negative Nd signatures with εNd(t) from − 5.42
612 ± 3 Ma (39 zircon spots; MSWD = 0.71) and a mean age of 612 ± to − 8.8, and NdTDM model ages from 1.59 Ga to 2.04 Ga.
3.0 Ma (MSWD = 5.1). The younger cluster ages range from 587 ± 7 Ma
to 545 ± 7 Ma with Th/U ratio from 1.4 to 0.2 from non-zoned zircon 5. Discussion
grains with low to dark CL luminescence (metamorphic domains),
yielding a Concordia age of 580 Ma ±5 Ma (9 zircon spots, MSWD = 5.1. Tectonic correlations along the neoproterozoic brasiliano belts of
0.24) and a mean age of 580 ± 4 Ma (MSWD = 1.6). Unmix age Brazil
calculation routine (Isoplot/Ex 4.15, Ludwig, 2012) yields agreement
age populations with peaks at 577 ± 2 Ma to 615 ± 1 Ma (relative misfit The orogenic belts of the Brasiliano Cycle surround the few cratonic
0.436). One inherited zircon has a spot age of 796 ± 10 Ma. areas that are tectonically stable at least since Neoproterozoic times
Twenty-eight analyses on zircon crystals with ages from ca. 545 ± 7 when several continental masses were amalgamated to form West
Ma to 635 ± 9 show initial 176Hf/177Hf ratios ranging from 0.282457 to Gondwana (Fig. 1 A e B), involving periods of accretion and collision
0.282271, εHf(t) between +2.3 and - 4.8, and Hf TDM model ages from ca. (from the Tonian to the early Ediacaran) and including numerous arc-
1.13 Ga to ca. 1.5 Ga (Figs. 9–11). The inherited zircon (796 ± 10 Ma) related volcanic-plutonic suites, high-grade metamorphic sequences,
yield initial 176Hf/177Hf ratios of 0.282152, with εHf(t) − 4.6 and Hf TDM and occurrences of ophiolite slivers. Therefore, within all these orogenic
model ages of 1.65 Ga. belts, magmatic arcs were registered and presented rocks with typical
magmatic arc geochemical and isotopic signatures generated in the
4.2.2.2. Jequitibá suite. Zircon grains from a garnet-biotite meta­ supra-subduction region of the asthenospheric mantle from varied dif­
granodiorite were analyzed for U–Pb and Lu–Hf (Sample CA-03; Figs. 2 ferentiation, mixing, and contamination processes. The Brasiliano
and 3G). The zircon population is light brown to colorless, translucent, orogenic belts were formed from consumed of two large oceanic realms
exhibiting short to long prismatic and bipyramidal habitus with sub­ by subduction-to-collision tectonic episodes (Cordani et al., 2016;
hedral to euhedral terminations (Fig. 8B). Inclusions are scarce. The Heilbron et al., 2004, 2017, 2020; and references therein).
grain lengths range from 70 to 550 μm, with length/width ratios from 2 The huge Goiás-Pharusian Ocean was closed after a long duration
to 10. CL-images show magmatic oscillatory zoning, prevailing medium subduction phase, between ca. 900 and 600 Ma, that brought together
to low luminescence, with inherited core in few crystals (Fig. 8B). the Amazonian and the West African cratons and the São Francisco-
Several crystals are surrounded by domains of dis­ Congo, the Rio de La Plata and the African counterparts. As a result of
solution–recrystallization and overgrowth rims (Fig. 8B, zircon grains this collision, the Tocantins orogenic system, the Borborema tectonic
052, 257, and 303). In BSE-images, magmatic oscillatory zoning is also province, and the Trans-Saharan belts in Africa were developed. This
present but absent in some crystals. tectonic episode is related to the juvenile intra-oceanic island arcs roots
Forty analytical spots on igneous domains yield U–Pb ages from 624 of the Goiás Magmatic Arc, in Brasília belt (Tocantins orogenic system),
± 8 Ma to 574 ± 7 Ma, with Th/U ratios from 1.3 to 0.1 (Figs. 8B, 9A and and the magmatic arcs from the Santa Quitéria Arc and the Sergipano
9B). For better analytical accuracy, thirty-five spot data were selected to belt from Borborema tectonic province, where magmatic-arc related
be plotted, resulting in a Concordia age and a mean age of 594 ± 3 Ma rocks, retro-eclogites, and dismembered ophiolites provide evidence of
(MSWD = 0.55). Six spot analyses on inherited cores yield ages from 652 accretion (Cordani et al., 2016; Heilbron et al., 2004, 2017, 2020; Santos
± 8 Ma to 1051 ± 12 Ma. Twenty-seven spot analyses on structureless et al., 2022; Pimentel, 2016, and references therein) (Fig. 1 A and B).
rims (metamorphic domains) yield U–Pb ages from 586 ± 8 Ma to 557 On the other hand, the Adamastor Ocean occupied an extended
± 7 Ma, with Th/U ratios ranging from 0.87 to 0.05. Twenty-one spots oceanic region that become the site of the amalgamation of smaller
furnished a Concordia age and a mean age of 571 ± 3 Ma (MSWD = cratonic masses and massifs, such as the Rio de La Plata, Paranapanema,
0.45) (Figs. 8B, 9A and 9B). Unmix age calculation routine (Isoplot/Ex Luiz Alves, and Kalahari (Fig. 1 A and B). With the closure of one of the
4.15; Ludwig, 2012) for zircon grains younger than 624 Ma yield age Adamastor terminal segments, along the Brazilian and Uruguayan
populations with peaks at 595 ± 2 Ma and 574 ± 2 Ma (relative misfit Atlantic coast, the Araçuaí, Ribeira, and Dom Feliciano orogenic belts
0.622). were developed. The tectono-magmatic evolution was by the
Forty-one analyses on zircon crystals with ages from ca. 624 Ma to subduction-driven contraction phase taking place to a confined orogen
560 Ma show varied initial Hf isotopic signature, being separated into with a pre-collisional continental-margin Andean-type arc (Araçuaí),
two groups. One group presents more clustered and less evolved Hf connected, in turn, to complex accretionary-collisional orogens (Ribeira
signatures with 176Hf/177Hf(t) ratios ranging from 0.282359 to and Dom Feliciano) with accreted juvenile pre-collisional island arc
0.282292, εHf(t) between − 1.7 and − 4.9, and HfTDM model ages from ca. terranes (Caxito et al., 2021, 2022; Cordani et al., 2016; De Toni et al.,
2020; Heilbron et al., 2004, 2017, 2020; and references therein) (Fig. 1

15
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

A and B). The Rio Doce, Rio Negro, and Serra da Prata magmatic arcs are regional paragneisses of the Nova Venécia complex.
formed in the Ribeira-Araçuaí Orogenic System (Southeast Brazil). In several rocks of the Rio Doce magmatic arc, garnet is relatively
In this scenario, we discuss the Santa Tereza and Jequitibá suites common as an accessory mineral (e.g., Gonçalves et al., 2016, 2018;
from Araçuaí orogen through the correlation between new and compiled Gouvêa et al., 2020; Narduzzi et al., 2017; Novo et al., 2010; Soares
data, as well as their geological implications. et al., 2020). After a thorough review and detailed studies, Narduzzi
et al. (2017) concluded that garnet crystals found in tonalites and
5.2. Petrogenesis of Santa Tereza and Jequitibá suites granodiorites of the Galiléia (G1) suite have a magmatic origin consis­
tent with crystallization above 25 km depth and temperatures below
Geochemical analyses coupled with field mapping and petrographic 700 ◦ C. Gouvêa et al. (2020) described a pressure of 10 kbar for
studies provide robust evidence to recognize distinct granitic rock as­ calcium-rich garnet. The presence of magmatic orthopyroxene and the
semblages (cf. Best, 2003; Frost and Frost, 2008), such as those from high contents of calcium in garnet crystals of metamorphosed tonalites
magmatic arcs (mostly calcic to calc-alkaline, magnesian, metal­ and granodiorites of the Muriaé batholith suggest crystallization of
uminous), continental rifts (mostly ferroan, alkaline to alkali-calcic with arc-related magmas in the deep crust (Soares et al., 2020). Garnet can
anhydrous tendency), post-collisional settings (magnesian to ferroan, also be produced by contamination with peraluminous melts, intra­
high-K calc-alkaline to alkali-calcic, I- and A-type), and collisional leu­ magmatic metasomatism in late crystallization, and metamorphism by
cogranites (S-type, rich in peraluminous silicates, such as muscovite dehydration reactions involving biotite and hornblende, or plagioclase
and/or almandine and/or cordierite and/or sillimanite). Indeed, the breakdown plus Opx in high amphibolite to granulite facies (Ducea
Araçuaí orogen is a good example of the successful application of that et al., 2015a; Best, 2003). Indeed, samples from the Santa Tereza and
methodology (field mapping + petrographic studies + geochemistry) for Jequitibá suites show microstructures consistent with high equilibrium
clearly discriminating distinct anorogenic (e.g., Victoria et al., 2022) temperatures under upper amphibolite to granulite facies conditions,
and orogenic granite suites (e.g., Araujo et al., 2020; Gradim et al., 2014; such as coexistence of antiperthite and perthite, conspicuous deformed
Pedrosa-Soares et al., 2011; Tedeschi et al., 2016; and references to irregular twinning in feldspars, subgrain patterns and stretched fab­
therein), as well the whole Mantiqueira Province of southeast and rics (cf., Best, 2003; Kruhl, 1996; Yardley et al., 1996).
southern Brazil (Caxito et al., 2021, 2022; Heilbron et al., 2020; San­
tiago et al., 2020). 5.3. Isotopic correlations
Overall, the Santa Tereza and Jequitibá suites present similar lith­
ochemical signatures and fit well with those signatures recorded for A correlation of the presented geochronological data with the
rocks from the Rio Doce magmatic arc (Figs. 4–7). Altogether, their compiled Araçuaí dataset indicates that the Santa Tereza and Jequitibá
geochemical attributes suggest a magmatic arc origin, with a well- suites, with Concordia ages at 612 ± 3 Ma and 594 ± 3 Ma, respectively,
defined calc-alkaline trend in the AFM diagram, and magnesian, calcic represent the easternmost part of the Rio Doce magmatic arc already
to alkali-calcic, metaluminous to slightly peraluminous I-type signa­ found (Figs. 1 and 12). The metamorphic ages from the samples of both
tures, as well as negative Nb, Ta and, Ti anomalies, for a wide range of suites also have a close correlation with the compiled metamorphic data
silica contents (Figs. 4–7). The REE signatures are also compatible with from Araçuaí zircon grains (Fig. 12A). Indeed, the U–Pb data spreading
magmatic arc settings. They plot mainly in volcanic arc fields, with along the Concordia curve suggest post-crystallization disturbance
predominantly mafic sources (Fig. 5). All those chemical attributes also under high-temperature conditions (Vervoort and Kemp, 2016; Tedeschi
show remarkable similarities to signatures concerning other continental et al., 2018), which is consistent with the granulite facies features
magmatic arcs around the world, remarkably in spidergram patterns observed in the studied samples and described in previous articles on the
(Figs. 4–7). regional metamorphism (Gradim et al., 2014; Richter et al., 2016).
Petrography from Santa Tereza and Jequitibá rocks also shows sig­ Furthermore, the Hf signatures of each sample are distributed roughly
nificant similarities with G1 rock associations described in the literature horizontally (Fig. 11) indicating progressive Pb loss (Vervoort and
(Gonçalves et al., 2014, 2016; 2018; Gouvêa et al., 2020; Gradim et al., Kemp, 2016). Thus, the crystallization ages of these rocks may be older
2014; Narduzzi et al., 2017; Narduzzi, 2018; Pedrosa-Soares et al., 2011; than the Concordia ages, and εHf(t) may be underestimated (Table 2;
Pereira et al., 2023; Soares et al., 2020; Tedeschi et al., 2016). Fig. 9E and F). By considering the older grains within the same Hf ratio
Furthermore, charnockitic rocks found in the south and west of the Santa ranges, the estimated crystallization ages of 635 ± 8 Ma for the Santa
Tereza suite have been related to the granulitic root of the Rio Doce Tereza sample and 624 ± 8 for the Jequitibá sample seem to be reliable
magmatic arc (Novo et al., 2010; Soares et al., 2020). As observed in the (Fig. 11B). The recalculated Hf values are εHf(635) +2.5 to − 4.6 for the
Santa Tereza suite, those charnockitic rocks can contain garnet that can Santa Tereza suite and εHf(624) − 1.2 to − 36.0 for the Jequitibá suite
be related to crustal contamination (Gouvêa et al., 2020; Novo et al., (Table 2).
2010; Soares et al., 2020) and/or partial melting of the very deep crust The new Hf isotope data (εHf(t) +2.3 to − 4.8 for Santa Tereza suite;
(Narduzzi et al., 2017; Narduzzi, 2018). Geothermobarometric calcu­ εHf(t) − 1.7 to − 36.4 for Jequitibá suite) represent the most primitive
lations carried out on pre-collisional Opx-bearing metatonalite from the signatures already found in the Rio Doce magmatic arc (Fig. 9C–F, and
Araçuaí-Ribeira boundary records metamorphic conditions of 10). Nevertheless, the Jequitibá Hf signature shows the largest range,
high-amphibolite to granulite facies (670 ◦ C–860 ◦ C) over a wide pres­ suggesting mixing between partial melting from the Nova Venécia or
sure range from 5 kbar to 10 kbar (Gouvêa et al., 2020). basement rocks and juvenile components, with the data of εHf(t) from − 8
Crystallized from hot CO2-bearing magmas in deep arc levels, Opx- to − 36 imply in crustal contamination (Fig. 9C–F, 10). The presented
bearing granitic rocks form enderbitic to charnockitic series, generally isotopic data show a striking correlation with those from rocks of the
associated with basic (noritic) terms displaying magnesian, calcic to Marceleza-Leopoldina-Serra da Bolívia complex. Conversely, they show
calc-alkalic, metaluminous to slightly peraluminous I-type signatures minor isotopic correspondence with the migmatitic paragneiss of the
(Frost and Frost, 2008). Orthopyroxene along with garnet even in Nova Venécia complex (Figs. 10 and 11).
slightly peraluminous magmas implies an increasing Al2O3 content that Evaluating the overall Hf dataset in relation to samples locations
promotes the crystallization of Grt over Opx (Frost and Frost, 2008). from north to south and east to west in the Araçuaí orogen (Fig. 10A and
Therefore, according to the geochemical signature, the Santa Tereza B), it is noteworthy that the progressive increase in εHf(t) from north to
charnockitic rocks would have been crystallized from CO2-bearing south and from west to east, i.e., the Hf signatures become more prim­
magmas in deep arc levels after assimilation of a component enriched in itive to the south and east. The same tendencies are outlined by zircon
Al and H2O, producing garnet, amphibole, and more biotite. Such a U–Pb ages from the Rio Doce Arc which peaks of probability density
crustal component could be a partial melt fraction derived from the curves become increasingly older from north to south and from west to

16
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

east (Fig. 12B and C). continental rift and oceanic settings; cf. Amaral et al., 2020; Souza et al.,
Araujo et al. (2020) recorded Hf data from the border of a 2022; Victoria et al., 2022) of the Araçuaí orogen, the presented robust
post-collisional intrusion (the Mestre Álvaro pluton, CA-05 sample; dataset of lithochemical and isotopic signatures point to an active con­
Fig. 2) to the south and east of the Santa Tereza sampling localities, tinental margin for the Rio Doce magmatic arc development. In this
showing several inherited zircon grains with a Concordia age of 598 ± 1 scenario, the Marceleza-Leopoldina-Serra da Bolívia complex (650-590
Ma and εHf(t) between - 0.7 and − 6.2, fitting well with Hf signatures Ma; εNd(t) = − 5.5 to − 11; εHf(t) = − 0.9 to − 3.75; Heilbron et al., 2013;
from the Jequitibá, Santa Tereza, Marceleza, Leopoldina and Serra da Corrales et al., 2020), representing the Rio Doce arc in the northern
Bolívia samples (Fig. 10A, B, 10D). These data suggest the existence of Ribeira Orogen (Corrales et al., 2020; Tedeschi et al., 2016), have
(probably) non-exhumed arc-related intrusions further east of the geochemical similarities with studied samples (Figs. 4–7) and can be
studied area. considered the best correlation with the Santa Tereza and Jequitibá
The new whole-rock Nd signatures (εNd(t) from − 0.06 to − 8.80) are suites.
among the least negative values ever reported for Rio Doce magmatic The differences between Hf signatures of the Eastern and Western
arc and fit well with the compiled data from G1 supersuite samples, magmatic arc sectors are remarkable. The more juvenile signature in the
especially the sample 18A (εNd(t) = − 0.06), an Opx-Cpx-Bt meta­ Eastern sector indicates the involvement of mantle magma and/or ju­
monzogabbro (Fig. 9G and H). The Nd results recalculated for estimated venile crust in the melting processes that formed at least part of the
crystallization ages show εNd(635) from +0.16 to − 6.84 for the Santa Santa Tereza suite. Indeed, according to Hf data, the partial melting of
Tereza suite and εNd(624) from − 6.34 to − 8.43 for the Jequitibá suite basement or country supracrustal rocks would not explain the isotopic
(Table 2). These slightly positive to moderately negative εNd values are signature of Santa Tereza rocks without mixing with mantle melts
expected for granitic rocks from continental magmatic arcs where in­ (Fig. 10C). Moreover, NdTDM model ages (between 1.2 and 2.0 Ga) are
teractions of mantle magmas with country rocks are common. The more younger than the basement rocks of the Juiz de Fora, Quirino, and
negative signatures in samples with garnet (εNd(t) from − 5.42 to − 8.8) Pocrane complexes (cf., Degler et al., 2018; Novo, 2013; Schannor et al.,
also suggest the assimilation of a crustal melt component. 2020). In addition to basement units, juvenile Hf and Nd signatures are
From the compiled dataset, the G2 supersuite ages display a found in the Tonian to Cryogenian Serra da Prata and Rio Negro
remarkable overlap with G1 supersuite ages between 600 Ma to 560 Ma magmatic arcs of the Ribeira orogen (Heilbron and Machado, 2003;
(Fig. 12A). However, though I-type magmatism (G1 rocks) pre­ Tupinambá et al., 2012; Lobato et al., 2015; Peixoto et al., 2017, 2022;
dominates from 600 Ma to 580 Ma, S-type magmatism (G2 rocks) pre­ Santiago et al., 2020). Indeed, melts from juvenile rocks of the Serra da
dominates between 580 Ma and 560 Ma. The generation of I-type Prata magmatic arc (860–838 Ma; εNd(t) +6.4 to − 3.7; εHf(t) +14 to
granites before S-type granites is further evidence for an orogeny driven +10; cf. Peixoto et al., 2017; Santiago et al., 2020) can be a potential
by ocean subduction followed by continent collision (Best, 2003). The source for the less evolved Santa Tereza melts (Fig. 9G, H, 10C), if the
overlapping of age ranges is also consistent with the timing of magma Serra da Prata arc has a prolongation (not yet known) in the studied
generation in relation to tectonic stages which is pre-collisional to early region.
collisional for G1 rocks and late pre-collisional to collisional for G2 That interpretation of sources for the Santa Tereza juvenile magmas
rocks. However, the large number of zircon grains from G1 rocks aged can be also related to studies on the Rio Doce Group, the metavolcano-
between 580 Ma and 560 Ma may also be related to the opening of the sedimentary cover of the Rio Doce magmatic arc (Novo et al., 2018).
U–Pb system during the collisional stage, although there is no evidence Whole-rock data and U–Pb age spectra, juvenile nature, and trace
of this process in the CL images. element composition recorded in detrital zircon grains from the Rio
From the compiled literature is remarkable that zircon U–Pb ages Doce Group suggest sediment sources in juvenile crustal setting, such as
from the Nova Venécia complex show an expressive number of zircon an intra-oceanic island arc (Schannor et al., 2019). Those U–Pb ages and
grains aged between 660 Ma and 580 Ma, even if compared with other Hf isotopic signatures from detrital zircon grains of the Rio Doce Group
arc-related basins (Fig. 12A). However, detrital zircon grains with ages fit well with rocks from the Serra da Prata magmatic arc, which is typical
<630 Ma could also reflect Pb-loss not detected by microstructural of an intra-oceanic island arc setting (Peixoto et al., 2017, 2022; San­
features shown in CL and BSE images. In fact, the available Hf signature tiago et al., 2020) (Fig. 10C). Conversely, the isotopic signatures of
for these rocks (Araujo et al., 2020) shows scattering along approxi­ zircon grains and whole-rock samples from Nova Venécia paragneisses
mately horizontal lines (Fig. 11), which is expected due to the high indicate a continental setting for the basin, which has been largely
temperature and pressure conditions experimented for those samples. interpreted as a back-arc basin developed during most part of the Rio
Thus, the ages younger than 630 Ma for detrital zircon grains may have Doce Arc evolution (Araujo et al., 2020; Gradim et al., 2014; Noce et al.,
been partially or completely reset by processes that triggered U–Pb 2004; Pedrosa-Soares et al., 2011; Richter et al., 2016).
zircon disturbances and could represent arc-related metamorphism fol­ Altogether, the new and compiled data provide further evidence for
lowed by collisional metamorphism. Compared to other arc-related ocean subduction beneath a continental margin which includes the
basins, the Nova Venécia complex has a more negative Hf signature previous collage of juvenile crust, such as an island arc (the Serra da
and presents isotopic similarities with some zircon grains from the Prata magmatic arc), that could generate melts evolving from juvenile to
Andrelândia Group, while the Rio Doce Group shows great discrepancy, evolved crustal melts, producing the Hf signatures shown by the rocks of
with juvenile zircon grains, which is also found in the Andrelândia the Santa Tereza suite and the Marceleza-Leopoldina-Serra da Bolívia
Group (Fig. 10C). The more evolved signatures given by Nova Venécia complex (Degler et al., 2017; Corrales et al., 2020, Fig. 10C). In a broad
paragneisses indicate prevailing sediment provenance from sources rich scenario, the data presented here also corroborate the consumption of
in continental rocks, such as the Rio Doce magmatic arc and its basement the Neoproterozoic Adamastor Ocean, including its northern termina­
(Araujo et al., 2020; Gradim et al., 2014; Richter et al., 2016). tion within a confined basin (an inland-sea basin, cf. Condie, 2007),
namely the Araçuaí – West Congo basin system, through a complete
5.4. Implications for araçuaí geotectonic evolution Wilson Cycle supported by wide evidence from several fields of the
Geosciences (Caxito et al., 2022).
Together with a thorough information collection of field relations The new data provide robust evidence for a rather juvenile arc
and petrographic features from the orogenic magmatic supersuites (e.g., magmatism followed by more evolved magmatic activity in the east­
De Campos et al., 2016; Gonçalves et al., 2016, 2017; Melo et al., 2017; ernmost portion of the Araçuaí orogen. This is also consistent with the
Pedrosa-Soares et al., 2011, 2020; Serrano et al., 2018; Tedeschi et al., evolutionary model, which suggests that the oldest plutons of the Rio
2016; Wisniowski et al., 2021), besides their tectonic relations with Doce magmatic arc were emplaced to the east followed by migration of
other geotectonic components (e.g., anorogenic magmatism related to the magmatic front to the west (Tedeschi et al., 2016), owing to

17
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

subduction retreat (Fig. 13). The zircon U–Pb and Lu–Hf dataset record respectively, and estimated crystallization ages of 635 ± 8 Ma and 624
westward arc migration (Fig. 12C), with increasing continental crust ± 8 Ma, the Lu–Hf and whole-rock Sm–Nd data show the least evolved
involvement (Fig. 10A and B). Back-arc basin evolution in an initial and magmatism in the Rio Doce magmatic arc. Together with a wide data
extensional phase followed by a compressional phase would explain compilation, our dataset indicates the involvement of juvenile crust
older arc magmatism and less evolved isotopic compositions in the and/or mantle melts in the Eastern sector of the Rio Doce arc, supporting
easternmost section (cf. Ducea et al., 2015a). Indeed, the studied region a complete Wilson Cycle for the orogenic evolution of the Araçuaí oro­
is related to the development of a back-arc basin on a hot and thin gen, including the magmatic front migration owing to subduction
continental crust (Gradim et al., 2014; Richter et al., 2016). retreat (Fig. 13). The studied rocks can be correlated with the Marceleza-
Leopoldina-Serra da Bolívia complex of the Ribeira Orogen, providing
6. Conclusions further evidence of the connection between both orogens during the
evolution of the Rio Doce magmatic arc. The presented data also support
New zircon U–Pb and Hf isotopic data for Santa Tereza and Jequitibá the evolution of a continental-margin magmatic arc within a confined
rocks, coupled with whole-rock Sm–Nd results, demonstrate the east­ orogen, the Araçuaí - West Congo Orogen, formed by the closing of an
ernmost expression of pre-collisional, arc-related magmatism in the inland-sea basin, providing an example of a singular geotectonic setting
Araçuaí orogen. With Concordia ages of 612 ± 3 Ma and 594 ± 3 Ma, in the paleocontinental history, such as the Western Gondwana.

Fig. 13. Evolutionary model for the Rio Doce magmatic arc (modified from Tedeschi et al., 2016).

18
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

CRediT authorship contribution statement paleogeography of southern Peru and northern Bolivia. J. S. Am. Earth Sci. 32,
196–209. https://doi.org/10.1016/j.jsames.2011.07.002.
Bateman, P.C., Dodge, F.C.W., Bruggman, P.E., 1984. Major Oxide Analyses, CIPW
Cristina Araujo: Writing – review & editing, Writing – original draft, Norms, Modes, and Bulk Specific Gravities of Plutonic Rocks from the Mariposa 1
Methodology, Investigation, Formal analysis, Data curation, Conceptu­ Degrees by 2 Degrees Sheet, Central Sierra Nevada, California, vols. 84–0162. US
alization. Antonio Pedrosa-Soares: Writing – review & editing, Visu­ Geological Survey OF, p. 59. https://doi.org/10.3133/ofr84162.
Belém, J., 2014. Geoquímica, geocronologia e contexto geotectônico do magmatismo
alization, Validation, Supervision, Resources, Funding acquisition, máfico associado ao feixe de fraturas Colatina, Estado do Espírito Santo.
Conceptualization. Cristiano Lana: Methodology, Formal analysis, Data Universidade Federal de Minas Gerais, Belo Horizonte. MS Dissertation, p. 138 p).
curation. Mahyra Tedeschi: Writing – review & editing, Visualization, Best, M.G., 2003. Igneous and Metamorphic Petrology, 2◦ ed. Blackwell Science, Malden,
p. 758p.
Validation, Supervision. Jorge Roncato: Writing – review & editing, Boynton, W.V., 1984. Cosmochemistry of the rare earth elements: meteorites studies. In:
Visualization, Investigation. Ivo Dussin: Visualization, Methodology. Henderson, P. (Ed.), Rare Earth Element Geochemistry. Elsevier, pp. 63–114.
Paula Serrano: Writing – review & editing, Visualization. Elton Dan­ Brown, M., Johnson, T., 2019. Metamorphism and the evolution of subduction on Earth.
Am. Mineral. 104, 1065–1082. https://doi.org/10.2138/am-2019-6956.
tas: Methodology, Formal analysis, Data curation. Castro, A., 2013. Tonalite-granodiorite suites as cotectic systems: a review of
experimental studies with applications to granitoid petrogenesis. Earth Sci. Rev. 124,
68–95. https://doi.org/10.1016/j.earscirev.2013.05.006.
Declaration of competing interest Castro, A., Gerya, T., Garcia-Casco, A., Fernandez, C., Diaz-Alvarado, J., Moreno-
Ventas, I., Löw, I., 2010. Melting relations of MORB–sediment Mélanges in
The authors declare that they have no known competing financial underplated mantle wedge plumes; implications for the origin of Cordilleran-type
batholiths. J. Petrol. 51, 1267–1295. https://doi.org/10.1093/petrology/egq019.
interests or personal relationships that could have appeared to influence
Castro, A., Moreno-Ventas, I., Fernández, C., Vujovich, G., Gallastegui, G., Heredia, N.,
the work reported in this paper. Martino, R.D., Becchio, R., Corretgé, L.G., Díaz-Alvarado, J., Such, P., García-
Arias, M., Liu, D.Y., 2011. Petrology and SHRIMP U-Pb zircon geochronology of
Cordilleran granitoids of the Bariloche area, Argentina. J. S. Am. Earth Sci. 32,
Data availability
508–530. https://doi.org/10.1016/j.jsames.2011.03.011.
Castro, M.P., Queiroga, G.N., Martins, M., Pedrosa-Soares, A.C., Dias, L., Lana, C.,
Data will be made available on request. Babinski, M., Alkmim, A.R., Silva, M.A., 2020. Provenance shift through time in
superposed basins: from Early Cryogenian glaciomarine to Late Ediacaran orogenic
sedimentations (Araçuaí Orogen, SE Brazil). Gondwana Res. 87, 41–66. https://doi.
Acknowledgments org/10.1016/j.gr.2020.05.019.
Cawood, P.A., Kröner, A. (Eds.), 2009. Earth Accretionary Systems in Space and Time.
The authors acknowledge the financial support and scholarships Geological Society, London, Special Publications, p. 318. https://doi.org/10.1144/
SP318.
provided by Brazilian research and development agencies (CNPq, Caxito, F.A., Heilbron, M., Valeriano, C.M., Bruno, H., Pedrosa-Soares, A., Alkmim, F.F.,
CAPES, and CODEMIG-CODEMGE). We are also grateful to the Micro­ Chemale, F., Hartmann, L., Dantas, E., Basei, M.A.S., 2021. Integration of elemental
scopy and Microanalysis Laboratory (LMIc) of the Federal University of and isotope data supports a Neoproterozoic Adamastor ocean realm. Geochem.
Perspect. Lett. 17, 6–10.
Ouro Preto, a member of the Microscopy and Microanalysis Network of Caxito, F.A., Hartmann, L., Heilbron, M., Pedrosa-Soares, A., Bruno, H., Basei, M.A.S.,
Minas Gerais State/Brazil/FAPEMIG. For the use of the laboratories and Chemale, F., 2022. Multi-proxy evidence for subduction of the neoproterozoic
analytical facilities from Minas Gerais (UFMG), and Ouro Preto (UFOP) Adamastor Ocean and Wilson cycle tectonics in the South atlantic Brasiliano
orogenic system of western Gondwana. Precambrian Res. 376, 106678 https://doi.
universities and their staffs, we thank all people involved. Our gratitude org/10.1016/j.precamres.2022.106678.
to the editors and reviewers that greatly helped us to improve this paper. Cecil, M.R., Rotberg, G.L., Ducea, M.N., Saleeby, J.B., Gehrels, G.E., 2012. Magmatic
growth and batholithic root development in the Northern Sierra Nevada, California.
Geosphere 8, 592–606. https://doi.org/10.1130/GES00729.1.
Appendix A. Supplementary data Chapman, A.D., Saleeby, J.B., Wood, D.J., Piasecki, A., Kidder, S., Ducea, M.N., Farley, K.
A., 2012. Late cretaceous gravitational collapse of the southern Sierra Nevada
Supplementary data related to this article can be found at https batholith, California. Geosphere 8, 314–341. https://doi.org/10.1130/GES00740.1.
Chappell, B.W., White, A.J.R., 2001. Two contrasting granite types: 25 years later. Aust.
://doi.org/10.1016/j.jsames.2023.104490. J. Earth Sci. 48, 489–499. https://doi.org/10.1046/j.1440-0952.2001.00882.x.
Condie, K.C., 2007. Accretionary orogens in space and time. In: Hatcher, R.D.,
References Carlson, M.P., McBride, J.H., Catalán, J.R. (Eds.), 4-D Framework of Continental
Crust. Geological Society of America, Memoirs, pp. 1–14.
Cordani, U.G., Ramos, V.A., Fraga, L.M., Cegarra, M., Delgado, I., Souza, K.G., Gomes, F.
Alkmim, F.F., Marshak, S., Pedrosa-Soares, A.C., Peres, G.G., Cruz, S.C.P.,
E.M., Schobbenhaus, C., 2016. Tectonic map of south America. Population (Paris):
Whittington, A., 2006. Kinematic evolution of the araçuaí–west Congo orogen in
Commission for the Geological Map of the World, Second edition scale 1:5 000 000.
Brazil and Africa: nutcracker tectonics during the neoproterozoic assembly of
Corrales, F.F.P., Dussin, I.A., Heilbron, M., Bruno, H., Bersan, S.A., Valeriano, C.M.,
Gondwana. Precambrian Res. 149, 43–64. https://doi.org/10.1016/j.
Pedrosa-Soares, A.C., Tedeschi, M., 2020. Coeval high Ba-Sr arc-related and OIB
precamres.2006.06.007.
Neoproterozoic rocks linking pre-collisional magmatism of the Ribeira and Araçuaí
Alkmim, F.F., Kuchenbecker, M., Reis, H.L.S., Pedrosa-Soares, A.C., 2017. The Araçuaí
orogenic belts, SE-Brazil. Precambriam Res. 337, 105476 https://doi.org/10.1016/j.
belt. In: Heilbron, M., Cordani, U.G., Alkmim, F.F. (Eds.), São Francisco Craton,
precamres.2019.105476.
Eastern Brazil. Regional Geology Reviews. Springer International Publishing Co.,
De Campos, C.P., Mendes, J.C., Ludka, I.P., Medeiros, S.R., Moura, J.C., Wallfass, C.,
pp. 255–276. https://doi.org/10.1007/978-3-319-01715-0_14
2004. A review of the Brasiliano magmatism in Southern Espírito Santo, Brazil, with
Amaral, L., Caxito, F.A., Pedrosa-Soares, A., Queiroga, G., Babinski, M., Trindade, R.,
emphasis on post-collisional magmatism. J. Virtual Explor. 17, 1–35. https://doi.
Lana, C., Chemale, F., 2020. The Ribeirão da Folha ophiolite-bearing accretionary
org/10.3809/jvirtex.2004.00106.
wedge (Araçuaí orogen, SE Brazil): new data for Cryogenian plagiogranite and
De Campos, C.P., Medeiros, S.R., Mendes, J.C., Pedrosa-Soares, A.C., Dussin, I., Ludka, I.
metasedimentary rocks. Precambriam Res. 336, 105522 https://doi.org/10.1016/j.
P., Dantas, E.L., 2016. Cambro-ordovician magmatism in the Araçuaí belt (SE
precamres.2019.105522.
Brazil): snapshots from a post-collisional event. J. S. Am. Earth Sci. 68, 248–268.
Aranda, R.O., Chaves, A.O., Medeiros Júnior, E.B., Venturini Junior, R., 2020. Petrology
https://doi.org/10.1016/j.jsames.2015.11.016.
of the afonso cláudio intrusive complex: new insights for the cambro-ordovician
De la Roche, H., Leterrier, J., Grandclaude, P., Marchal, M., 1980. A classification of
post-collisional magmatism in the araçuaí-west Congo orogen, southeast Brazil. J. S.
volcanic and plutonic rocks using R1R2-diagram and major element analyses - its
Am. Earth Sci. 98, 102465 https://doi.org/10.1016/j.jsames.2019.102465.
relationships with current nomenclature. Chem. Geol. 29, 183–210. https://doi.org/
Araujo, C., Pedrosa-Soares, A., Lana, C., Dussin, I., Queiroga, G., Serrano, P., Medeiros-
10.1016/0009-2541(80)90020-0.
Júnior, E., 2020. Zircon in emplacement borders of post-collisional plutons
De Paolo, D.J., 1981. A neodymium and strontium isotopic study of the Mesozoic
compared to country rocks: a study on morphology, internal texture, U-Th-Pb
calcalkaline granitic batoliths of the Sierra Nevada and Peninsular Rangers,
geochronology and Hf isotopes (Araçuaí orogen, SE Brazil). Lithos 352–353, 105252.
California. J. Geophys. Res. 86, 10470–10488. https://doi.org/10.1029/
https://doi.org/10.1016/j.lithos.2019.105252.
JB086iB11p10470.
Araujo, L.E.A.B., Heilbron, M., Teixeira, W., Dussin, I., Valeriano, C.M., Bruno, Henrique,
De Toni, G.B., Bitencourt, M.F., Nardi, L.V.S., Florisbal, L.M., Almeida, B.S., Geraldes, M.,
Sato, K., Paravidini, G., Castro, M., 2021. Siderian to rhyacian evolution of the Juiz
2020. Dom Feliciano Belt orogenic cycle tracked by its pre-collisional magmatism:
de Fora complex: arc fingerprints and correlations within the minas-bahia orogen
the Tonian (ca. 800 Ma) Porto Belo Complex and its correlations in southern Brazil
and the western central Africa belt. Precambrian Res. 359, 106118 https://doi.org/
and Uruguay. Precambrian Res. 342, 105702 https://doi.org/10.1016/j.
10.1016/j.precamres.2021.106118.
precamres.2020.105702.
Bahlburg, H., Vervoort, J.D., DuFrane, S.A., Carlotto, V., Reimann, C., Cárdenas, J.,
Degler, R., Pedrosa-Soares, A., Dussin, I., Queiroga, G., Schulz, B., 2017. Contrasting
2011. The U-Pb and Hf isotope evidence of detrital zircons of the Ordovician
provenance and timing of metamorphism from paragneisses of the Araçuaí-Ribeira
Ollantaytambo Formation, southern Peru, and the Ordovician provenance and

19
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

orogenic system, Brazil: hints for Western Gondwana assembly. Gondwana Res. 51, Heilbron, M., Valeriano, C.M., Peixoto, C., Tupinambá, M., Neubauer, F., Dussin, I.,
30–50. https://doi.org/10.1016/j.gr.2017.07.004. Corrales, F., Bruno, H., Lobato, M., Almeida, J.C.A., Silva, L.G.E., 2020.
Degler, R., Pedrosa-Soares, A., Novo, T., Tedeschi, M., Silva, L.C., Dussin, I., Lana, C., Neoproterozoic magmatic arc systems of the central Ribeira belt, SE-Brazil, in the
2018. Rhyacian-Orosirian isotopic records from the basement of the Araçuaí-Ribeira context of the West-Gondwana pre-collisional history: a review. J. S. Am. Earth Sci.
orogenic system (SE Brazil): links in the Congo-São Francisco palaeocontinent. 103, 102710 https://doi.org/10.1016/j.jsames.2020.102710.
Precambrian Res. 317, 179–195. https://doi.org/10.1016/j.precamres.2018.08.018. Hervé, F., Pankurst, R.J., Fanning, C.M., Calderón, M., Yaxely, G.M., 2007. The South
Deluca, C., 2018. Evolução do Batólito Itaporé e rochas encaixantes, Orógeno Araçuaí Patagonian batholith: 150 my of granite magmatism on a plate margin. Lithos 97,
(MG): geoquímica, geocronologia e petrogênese. Universidade Federal de Minas 373–394. https://doi.org/10.1016/j.lithos.2007.01.007.
Gerais, Belo Horizonte (MS Dissertation, p. 116. IBGE, 1987. Geologia. In: Projeto RADAMBRASIL. Folha SE 24 Rio Doce. IBGE, Rio de
Dodge, F.C.W., Millard, H.T., Elsheimer, H.N., 1982. Compositional Variations and Janeiro, p. 548.
Abundances of Selected Elements in Granitoids Rocks and Constituent Minerals, Irvine, T.M., Baragar, W.R., 1971. A guide to the chemical classification of common
Central Sierra Nevada Batholiths. US Geological Survey Professional Paper, volcanic rocks. Can. J. Earth Sci. 8, 523–548. https://doi.org/10.1139/e71-055.
California, p. 1248. Jackson, S.E., Pearson, N.J., Griffin, W.L., Belousova, E.A., 2004. The application of laser
Ducea, M.N., Saleeby, J.B., Bergantz, G., 2015a. The architecture, chemistry, and ablation-inductively coupled plasma-mass spectrometry to in situ U–Pb zircon
evolution of continental magmatic arcs. Annu. Rev. Earth Planet Sci. 43, 10. https:// geochronology. Chem. Geol. 211, 47–69. https://doi.org/10.1016/j.
doi.org/10.1146/annurev-earth-060614-105049, 11-10.33. chemgeo.2004.06.017.
Ducea, M.N., Paterson, S.R., DeCelles, P.G., 2015b. High-volume magmatic events in Janoušek, V., Farrow, C.M., Erban, V., 2006. Interpretation of whole-rock geochemical
subduction systems. Elements 11, 99–104. https://doi.org/10.2113/ data in igneous geochemistry: introducing Geochemical Data Toolkit (GCDkit).
gselements.11.2.99. J. Petrol. 47, 1255–1259. https://doi.org/10.1093/petrology/egl013.
Dunphy, J.M., Ludden, J.N., 1998. Petrological and geochemical characteristics of a Kearey, P., Klepeis, K.A., Vine, F.J., 2014. Tectônica Global, 3◦ ed., p. 436p Bookman.
Paleoproterozoic magmatic arc (Narsajuaq terrane, Ungava Orogen, Canada) and Kelemen, P.B., Hanghoj, K., Greene, A.R., 2014. One view of the geochemistry of
comparisons to Superior Province granitoids. Precambriam Res. 91, 109–142. subduction-related magmatic arcs, with an emphasis on primitive andesite and lower
https://doi.org/10.1016/S0301-9268(98)00041-2. crust. Treatise Geochem. 4, 749–806. https://doi.org/10.1016/B978-0-08-095975-
Féboli, W.L., 1993. Geologia da Folha Domingos Martins. Programa de Levantamentos 7.00323-5.
Geológicos Básicos do Brasil, Folha SF.24-V-A-III – Domingos Martions. Brasília. Kohanpour, F., Kirkland, C.L., Gorczyk, C., Occhipinti, S., Lindsay, M.D., Mole, D.,
DNPM/CPRM. Vaillant, M., 2019. Hf isotopic fingerprinting of geodynamic settings: integrating
Figueiredo, M.C.H., Campos Neto, M.C., 1993. Geochemistry of the Rio doce magmatic isotopes and numerical models. Gondwana Res. 73, 190–199.
arc, Southeastern Brazil. An Acad. Bras Ciências 65, 63–81. Kruhl, J.H., 1996. Prism- and basal-plane parallel subgrain boundaries in quartz: a
Frost, B.R., Frost, C.D., 2008. On charnockites. Gondwana Res. 13, 30–44. https://doi. microstructural geothermobarometer. J. Metamorph. Geol. 14, 581–589. https://
org/10.1016/j.gr.2007.07.006. doi.org/10.1046/j.1525-1314.1996.00413.x.
Frost, B.R., Barnes, C.G., Collins, W.J., Arculus, R.J., Ellis, D.J., Frost, C.D., 2001. Kuribara, Y., Tsunogae, T., Santosh, M., Takamura, Y., Costa, A.G., Rosière, C.A., 2019.
A geochemical classification for granitic rocks. J. Petrol. 42, 2033–2048. https://doi. Eoarchean to Neoproterozoic crustal evolution of the Mantiqueira and the Juiz de
org/10.1093/petrology/42.11.2033. Fora Complexes, SE Brazil: petrology, geochemistry, zircon U-Pb geochronology and
Gerdes, A., Zeh, A., 2006. Combined U-Pb and Hf isotope LA-(MC)-ICP-MS analyses of Lu-Hf isotopes. Precambrian Res. 323, 82–101. https://doi.org/10.1016/j.
detrital zircons: comparison with SHRIMP and new constraints for the provenance precamres.2019.01.008.
and age of an Armorican metasediment in Central Germany. Earth Planet Sci. Lett. Laurent, O., Martin, H., Moyen, J.F., Doucelance, R., 2014. The diversity and evolution of
249, 47–61. https://doi.org/10.1016/j.epsl.2006.06.039. late-Archean granitoids: evidence for the onset of ’modern-style’ plate tectonics
Gerdes, A., Zeh, A., 2009. Zircon formation versus zircon alteration - new insights from between 3.0 and 2.5 Ga. Lithos 205, 208–235. https://doi.org/10.1016/j.
combined U-Pb and Lu-Hf in-situ LA-ICP-MS analyses, and consequences for the lithos.2014.06.012.
interpretation of Archean zircon from the Central Zone of the Limpopo Belt. Chem. Lee, C.-T.A., Morton, D.M., Kistler, R.W., Baird, A.K., 2007. Petrology and tectonics of
Geol. 261, 230–243. https://doi.org/10.1016/j.chemgeo.2008.03.005. Phanerozoic continent formation: from island arcs to accretion and continental arc
Gioia, S.M.C.L., Pimentel, M.M., 2000. The Sm-Nd isotopic method in the geochronology magmatism. Earth Planet Sci. Lett. 263, 370–387. https://doi.org/10.1016/j.
laboraty of the university of Brasília. Anais da Academia Brasília de Ciências 72 (2), epsl.2007.09.025.
219–245. Leite, C.A.S., Souza, J.D., Silva, S.L., Kosin, M., Silva, L.C., Bento, R.V., Santos Vieira, V.
Gonçalves, L.E.S., 2009. Características gerais e história deformacional da Suíte Granítica S., Camozzato, E., Paes, V.J.C., Netto, C., Junqueira, P.A., 2004. Folha rio doce
G1, entre Governador Valadares e Ipanema, MG. Universidade Federal de Ouro (SE.24), escala 1:1.000.000. In: Schobbenhaus, C., Gonçalves, J.H., Santos, J.O.S.,
Preto, Ouro Preto (MS Dissertation, p. 112. Abram, M.B., Leão Neto, R., Matos, G.M.M., Vidotti, R.M., Ramos, M.A.B., Jesus, J.D.
Gonçalves, L., Farina, F., Lana, C., Pedrosa-Soares, A.C., Alkmim, F., Nalini Jr., H.A., A. de (Eds.), Carta Geológica Do Brasil Ao Milionésimo. Programa Geologia Do
2014. New U-Pb ages and lithochemical attributes of the Ediacaran Rio Doce Brasil, CPRM–Serviço Geológico Do Brasil, Brasília.
magmatic arc, Araçuaí confined orogen, southeastern Brazil. J. S. Am. Earth Sci. 52, Lobato, M., Heilbron, M., Torós, T., Ragatky, D., Dantas, E., 2015. Provenance of the
129–148. https://doi.org/10.1016/j.jsames.2014.02.008. neoproterozoic high-grade metasedimentary rocks of the arcarc-related oriental
Gonçalves, L., Alkmim, F.F., Pedrosa-Soares, A.C., Dussin, I.A., Valeriano, C.M., Lana, C., terrane of the Ribeira belt: implications for Gondwana amalgamation. J. S. Am.
Tedeschi, M., 2016. Granites of the intracontinental termination of a magmatic arc: Earth Sci. 63, 260–278. https://doi.org/10.1016/j.jsames.2015.07.019.
an example from the Ediacaran Araçuaí orogen, southeastern Brazil. Gondwana Res. Ludwig, K.R., 2012. Isoplot/Ex Version 4.15: A Geochronological Toolkit for Microsoft
36, 439–458. https://doi.org/10.1016/j.gr.2015.07.015. Excel. Berkeley Geochronology Center, Berkeley, p. 72.
Gonçalves, L., Alkmim, F.F., Pedrosa-Soares, A.C., Gonçalves, C.C., Vieira, V., 2018. Martins, V.T.S., Teixeira, W., Noce, C.M., Pedrosa-Soares, A.C., 2004. Sr and Nd
From the plutonic root to the volcanic roof of a continental magmatic arc: a review characteristics of brasiliano/pan-african granitoid plutons of the araçuai orogen,
of the Neoproterozoic Araçuaí orogen, southeastern Brazil. Int. J. Earth Sci. 107, southeastern Brazil: tectonic implications. Gondwana Res. 7, 75–89.
337–358. https://doi.org/10.1007/s00531-017-1494-5. Mauri, S., Melo, M.G., Lana, C., Marques, R.A., 2021. Petrologic and geochronological
Gouvêa, L.P., Medeiros, S.R., Mendes, J.C., Soares, C., Marques, R., Melo, M., 2020. constraints on the polymetamorphic evolution of the collisional granites, Araçuaí
Magmatic activity period and estimation of P-T metamorphic conditions of pre- Orogen (SE Brazil). An Acad. Bras Ciências 93 (3), e20200639. https://doi.org/
collisional Opx-metatonalite from Araçuaí-Ribeira orogens boundary, SE Brazil. J. S. 10.1590/0001-3765202120200639.
Am. Earth Sci. 99, 102506 https://doi.org/10.1016/j.jsames.2020.102506. Melo, M.G., Lana, C., Stevens, G., Pedrosa-Soares, A.C., Gerdes, A., Alkmin, L.A.,
Gradim, C., Roncato, J., Pedrosa-Soares, A.C., Cordani, U., Dussin, I., Alkmim, F.F., Nalini Jr., H.A., Alkmim, F.F., 2017. Assessing the isotopic evolution of S-type
Queiroga, G., Jacobsohn, T., Silva, L.C., Babinski, M., 2014. The hot back-arc zone of granites of the Carlos Chagas batholith, SE Brazil: clues from U-Pb, Hf isotopes, Ti
the Araçuaí Orogen, Eastern Brazil: from sedimentation to granite generation. Braz. geothermometry and trace element composition of zircon. Lithos 284–285, 730–750.
J. Genet. 44, 155–180. https://doi.org/10.5327/Z2317-4889201400010012. https://doi.org/10.1016/j.lithos.2017.05.025.
Harris, N.B.W., Pearce, J.A., Tindle, A.G., 1986. Geochemical characteristics of collision- Melo, M.G., Lana, C., Stevens, G., Hartwig, M.E., Pimenta, M.E., Nalini Jr., H.A., 2020.
zone magmatism. In: Coward, M.P., Ries, A.C. (Eds.), Collision Tectonics, vol. 19. Deciphering the source of multiple U–Pb ages and complex Hf isotope composition in
Geological Society London Special Publication, pp. 67–81. https://doi.org/10.1144/ zircon from post-collisional charnockite-granite associations from the Araçuaí
GSL.SP.1986.019.01.04. orogen (southeastern Brazil). J. S. Am. Earth Sci. 103, 102792 https://doi.org/
Heilbron, M., Machado, N., 2003. Timing of terrane accretion in the neoproterozoic- 10.1016/j.jsames.2020.102792.
eopaleozoic Ribeira orogen (SE Brazil). Precambrian Res. 125, 87–112. https://doi. Milani, L., Kinnaird, J.A., Lehmann, J., Naydenov, K.V., Saalmann, K., Frei, D.,
org/10.1016/S0301-9268(03)00082-2. Gerdes, A., 2015. Role of crustal contribution in the early stage of the Damara
Heilbron, M., Cordani, U.G, Alkmim F.F. 2017, The São Francisco Craton and Its Margins. Orogen, Namibia: new constraints from combined U–Pb and Lu–Hf isotopes from the
In: Heilbron, M., Cordani, U.G, Alkmim F.F. (Eds.), São Francisco Craton, Eastern Goas Magmatic Complex. Gondwana Res. 28, 961–986. https://doi.org/10.1016/j.
Brazil. Regional Geology Reviews, Springer International Publishing Co., pp. gr.2014.08.007.
255–276. https://doi.org/10.1007/978-3-319-01715-0_14. Moyen, J.-F., Laurent, O., Chelle-Michou, C., Couzinié, S., Vanderhaeghe, O., Zeh, A.,
Heilbron, M., Pedrosa-Soares, A.C., Campos-Neto, M.C., Silva, L.C., Trouw, R., Janasi, V. Villaros, A., Gardien, V., 2017. Collision vs. subduction-related magmatism: two
A., 2004. Província Mantiqueira. In: Mantesso-Neto A, V., Bartorelli, C.D.R., contrasting ways of granite formation and implications for crustal growth. Lithos
Carneiro, B.B., Brito-Neves (Org (Eds.), Geologia Do Continente Sul-Americano. São 277, 154–177. https://doi.org/10.1016/j.lithos.2016.09.018.
Paulo, Editora Beca, XIII, pp. 203–234. Munhá, J.M.U., Cordani, U.G., Tassinari, C.C.G., Palácios, T., 2005. Petrologia e
Heilbron, M., Tupinambá, M., Valeriano, C., Armstrong, R., Silva, L.G.E., Melo, R.S., termocronologia de Gnaisses Migmatíticos da Faixa de Dobramentos Araçuaí
Simonetti, A., Pedrosa-Soares, A.C., Machado, N., 2013. The Serra da Bolívia (Espírito Santo, Brasil). Rev. Bras. Geociencias 35, 123–134.
complex: the record of a new Neoproterozoic arc-related unit at Ribeira belt. Nalini Jr., H.A., 1997. Etude géochimique et structurale des suítes Galiléia et Urucum et
Precambrian Res. 238, 158–175. https://doi.org/10.1016/j.precamres.2013.09.014. relations avec les pegmatites à éléments rares associées. In: Caractérisation des suítes

20
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

magmatiques néoprotérozoiques de la region de Conselheiro Pena et Galiléia (Minas costeiro basins: reconstructing the outer Magmatic Arc system of the Ribeira belt, SE
Gerais, Brésil). Ph.D. thesis. École des Mines de Saint Etienne et École des Mines de Brazil. Precambrian Res. 382, 106879 https://doi.org/10.1016/j.
Paris, France, p. 237. precamres.2022.106879.
Nalini Jr., H.A., Bilal, E., Correia Neves, J.M., 2000. Syn-collisional peraluminous Pereira, V.H.M., Mendes, J.C., Carvalho, M.Q.T., Coelho, V.S.A., Medeiros, S.R.,
magmatism in the Rio Doce region: mineralogy, geochemistry and isotopic data of Valeriano, M., 2023. Petrogenesis of estrela granitoid and implications for the
the Urucum suite (eastern Minas Gerais state, Brazil). Rev. Bras. Geociencias 30, evolution of the rio doce magmatic arc: araçuaí-Ribeira orogenic system, SE Brazil.
120–125. J. S. Am. Earth Sci. 126, 104337 https://doi.org/10.1016/j.jsames.2023.104337.
Narduzzi, F., 2018. Genesis and evolution of a Neoproterozoic magmatic arc: the Pimentel, M.M., 2016. The tectonic evolution of the Neoproterozoic Brasília Belt, central
Cordilleran-type granitoids of the Araçuaí Belt, Brazil. PhD Thesis, Escola de Minas, Brazil: a geochronological and isotopic approach. Braz. J. Genet. 46 (Suppl. 1),
Departamento de Geologia. Universidade Federal de Ouro Preto, p. 245p. 67–82. https://doi.org/10.1590/2317-4889201620150004.
Narduzzi, F., Farina, F., Stevens, G., Lana, C., Nalini Jr., H.A., 2017. Magmatic garnet in Richter, F., Lana, C., Steven, G., Buick, I., Pedrosa-Soares, A.C., Alkmim, F.F., Cutts, K.,
the Cordilleran-type Galiléia granitoids of the Araçuaí belt (Brazil): evidence for 2016. Sedimentation, metamorphism and granite generation in a back-arc region:
crystallization in the lower crust. Lithos 282–283, 82–97. https://doi.org/10.1016/j. records from the ediacaran Nova Venécia complex (Araçuaí orogen, southeastern
lithos.2017.02.017. Brazil). Precambrian Res. 272, 78–100. https://doi.org/10.1016/j.
Noce, C.M., Pedrosa-Soares, A.C., Piuzana, D., Armstrong, R., Laux, J.H., De Campos, C. precamres.2015.10.012.
M., Medeiros, S.R., 2004. Ages of sedimentation of the kinzigitic complex and of a Rodríguez, N., Díaz-Alvarado, J., Fernández, C., Fuentes, P., Breitkreuz, C., Tassinari, C.
late orogenic thermal episode of the Araçuaí Orogen, Northern Espírito Santo State, C.G., 2018. The significance of U-Pb zircon ages in zoned plutons – the case of the
Brazil: zircon and monazite U-Pb SHRIMP and ID–TIMS data. Rev. Bras. Geociencias Flamenco pluton, Coastal Range batholith, northern Chile. Geosci. Front. 10,
34, 587–592. 1073–1099. https://doi.org/10.1016/j.gsf.2018.06.003.
Novo, T.A., 2013. Caracterização do Complexo Pocrane, magmatismo básico Rosa, M.d.L.d.S., Conceição, H., Macambira, M.J.B., Galarza, M.A., Cunha, M.P.,
Mesoproterozóico e unidades neoproterozóicas do Sistema Araçuaí-Ribeira, com Menezes, R.C.L., Marinho, M.M., Filho, B.E.d.C., Rios, D.C., 2007. Neoproterozoic
ênfase em geocronologia U-Pb (SHRIMP e LA-ICP-MS). PhD Thesis. Instituto de anorogenic magmatism in the Southern Bahia Alkaline Province of NE Brazil: U–Pb
Geociências, Universidade Federal de Minas Gerais, p. 177. and Pb-Pb ages of the blue sodalite syenites. Lithos 97, 88–97. https://doi.org/
Novo, T.A., Pedrosa-Soares, A.C., Noce, C.M., Alkmim, F.F., Dussin, I.A., 2010. Rochas 10.1016/j.lithos.2006.12.011.
charnockíticas do sudeste de Minas Gerais: a raiz granulítica do arco magmático do Santiago, R., Caxito, F.A., Pedrosa-Soares, A.C., Neves, M.A., Dantas, E.L., 2020. Tonian
Orógeno Araçuaí. Rev. Bras. Geociencias 40, 573–592. island arc remnants in the northern Ribeira orogen of western Gondwana: the caxixe
Novo, T.A., Pedrosa-Soares, A., Vieira, V.S., Dussin, I., Silva, L.C., 2018. The Rio Doce batholith (ESPÍRITO SANTO, SE Brazil), precambrian research 105944. https://doi.
Group revisited: an Ediacaran arc-related volcano-sedimentary basin, Araçuaí org/10.1016/j.precamres.2020.105944.
orogen (SE Brazil). J. S. Am. Earth Sci. 85, 345–361. https://doi.org/10.1016/j. Santiago, R., Caxito, F.A., Pedrosa-Soares, A.C., Neves, M.A., Calegari, S.S., Lana, C.,
jsames.2018.05.013. 2022. Detrital zircon U–Pb and Lu–Hf constraints on the age, provenance and
Otamendi, J.E., Ducea, M.N., Bergantz, G.W., 2012. Geological, petrological and tectonic setting of arc-related high-grade units of the transition zone of the Araçuaí
geochemical evidence for progressive construction of an arc crustal section, Sierra de and Ribeira orogens (SE Brazil). J. S. Am. Earth Sci. 116, 103861 https://doi.org/
Valle Fértil, Famatinian Arc, Argentina. J. Petrol. 53, 761–800. https://doi.org/ 10.1016/j.jsames.2022.103861.
10.1093/petrology/egr079. Santos, M.M., Lana, C., Scholz, R., Buick, I., Schmitz, M.D., Kamo, S.L., Gerdes, A.,
Pankhurst, R.J., Rapela, C.W., Saavedra, J., Baldo, E.G., Dahlquist, J., Pascua, I., Corfu, F., Tapster, S., Lancaster, P., Storey, C.D., Basei, M.A.S., Tohver, E.,
Fanning, C.M., 1998. The Famatinian arc in the central Sierras Pampeanas: an early Alkmim, A., Nalini, H., Krambrock, K., Fantini, C., Wiedenbeck, M., 2017. A New
to mid-Ordovician continental arc on the Gondwana margin. In: Pankhurst, R.J., appraisal of Sri Lankan BB zircon as reference material for LA-ICP-MS U-Pb
Rapela, C.W. (Eds.), The Proto-Andean Margin of Gondwana, vol. 142. Geological geochronology and Lu-Hf isotope tracing. Geostand. Geoanal. Res. 41, 335–358.
Society of London, Special Publications, pp. 343–367. https://doi.org/10.1144/GSL. https://doi.org/10.1111/ggr.12167.
SP.1998.142.01.17. Santos, L.C.M.L., Caxito, F.A., Bouyo, M.H., Ouadahi, S., Araibia, K., Lages, G.A.,
Pankhurst, R.J., Weaver, S.D., Hervé, F., Larrondo, P., 1999. Mesozoic-cenozoic Santos, G.L., Pitombeira, J.P.A., Cawood, P.A., 2022. Relics of ophiolite-bearing
evolution of the north patagonian batholith in aysen, southern Chile. J. Geol. Soc., accretionary wedges in NE Brazil and NW Africa: connecting threads of western
Lond. 156, 673–694. https://doi.org/10.1144/gsjgs.156.4.0673. Gondwana’s ocean during Neoproterozoic times. Geosyst. Geoenviron. https://doi.
Paterson, S.R., Ducea, M.N., 2015. Arc magmatic tempos: gathering the evidence. org/10.1016/j.geogeo.2022.100148.
Elements 11, 91–98. https://doi.org/10.2113/gselements.11.2.91. Schannor, M., Lana, C., Fonseca, M., 2019. São Francisco-Congo Craton break-up
Pearce, J.A., Harris, N.B.W., Tindle, A.G., 1984. Trace element diagrams for the tectonic delimited by U-Pb-Hf isotopes and trace-elements of zircon from metasediments of
interpretation of granitic rocks. J. Petrol. 25, 956–983. https://doi.org/10.1093/ the Araçuaí Belt. Geosci. Front. 10, 611–628. https://doi.org/10.1016/j.
petrology/25.4.956. gsf.2018.02.011.
Pedrosa Soares, A.C., Wiedemann-Leonardos, C.M., 2000. Evolution of the Araçuaí belt Schannor, M., Lana, C., Mazoz, A., Narduzzi, F., Cutts, K., Fonseca, M., 2020.
and its connection to the Ribeira belt, eastern Brazil. In: Cordani, U.G., Milani, E.J., Paleoproterozoic sources for Cordilleran-type Neoproterozoic granitoids from the
Thomaz Filho, A., Campo, D.A. (Eds.), Tectonic Evolution of South America. Rio de Araçuaí orogen (SE Brazil): constraints from Hf isotope zircon composition. Lithos,
Janeiro, 31st International Geological Congress, pp. 265–285. 105815. https://doi.org/10.1016/j.lithos.2020.105815, 378–379.
Pedrosa-Soares, A.C., Noce, C.M., Wiedemann, C., Pinto, C.P., 2001. The Araçuaí- west- Scherer, E., Münker, C., Mezger, K., 2001. Calibration of the lutetium–hafnium clock.
Congo orogen in Brazil: an overview of a confined orogen formed during Science 293, 683–687. https://doi.org/10.1126/science.1061372.
gondwanaland assembly. Precambrian Res. 110, 307–323. https://doi.org/10.1016/ Serrano, P., Pedrosa-Soares, A., Medeiros-Junior, E., Fonte-Boa, T., Araujo, C., Dussin, I.,
S0301-9268(01)00174-7. Queiroga, G., Lana, C., 2018. A-type Medina batholith and post-collisional anatexis
Pedrosa-Soares, A.C., Alkmim, F.F., Tack, L., Noce, C.M., Babinski, M., Silva, L.C., in the Araçuaí orogen (SE Brazil). Lithos 320–321, 515–536. https://doi.org/
Martins-Neto, M.A., 2008. Similarities and differences between the Brazilian and 10.1016/j.lithos.2018.09.009.
african counterparts of the neoproterozoic araçuaí-west Congo orogen. In: Silva, M.A., Camozzato, E., Paes, V.J.C., Junqueira, P.A., Ramgrab, G.E., 2004. Folha
Pankhurst, R.J., Trouw, R.A.J., Brito Neves, B.B., De Wit, M.J. (Eds.), West SF.24-Vitoria. In: Schobbenhaus, C., Gonçalves, J.H., Santos, J.O.S., Abram, M.B.,
Gondwana: Pre-cenozoic Correlations across the South Atlantic Region, vol. 294. Leão Neto, R., Matos, G.M.M., Vidotti, R.M., Ramos, M.A.B., Jesus, J.D.A. de (Eds.),
Geological Society of London, Special Publication, London, pp. 153–172. https:// Carta Geológica do Brasil ao Milionésimo, Sistema de Informações Geográficas.
doi.org/10.1144/SP294.9. Programa Geologia do Brasil. CPRM, Brasília.
Pedrosa-Soares, A.C., De Campos, C.P., Noce, C.M., Silva, L.C., Novo, T., Roncato, J., Silva, L.C., Pedrosa-Soares, A.C., Teixeira, L.R., Armstrong, R., 2008. Tonian rift-related,
Medeiros, S., Castañeda, C., Queiroga, G., Dantas, E., Dussin, I., Alkmim, F., 2011. A-type continental plutonism in the Araçuaí Orogen, Eastern Brazil: new evidence
Late neoproterozoic-cambrian granitic magmatism in the Araçuaí orogen (Brazil), for the breakup stage of the São Francisco-Congo Paleocontinent. Gondwana Res. 13,
the eastern Brazilian pegmatite province and related mineral Resources. In: Sial, A. 527–537. https://doi.org/10.1016/j.gr.2007.06.002.
N., Bettencourt, J.S., De Campos, C.P., Ferreira, V.P. (Eds.), Granite-Related Ore Sláma, J., Košler, J., Condon, D.J., Crowley, J.L., Gerdes, A., Hanchar, J.M.,
Deposits, vol. 350. Geological Society of London, Special Publication, London, Horstwood, M.S.A., Morris, G.A., Nasdala, L., Norberg, N., Schaltegger, U.,
pp. 25–51. https://doi.org/10.1144/SP350.3. Schoene, B., Tubrett, M.N., Whitehouse, M.J., 2008. Plešovice zircon – a new natural
Pedrosa-Soares, A.C., Deluca, C., Araujo, C., Gradim, C., Lana, C., Dussin, I., Silva, L.C., reference material for U–Pb and Hf isotopic microanalysis. Chem. Geol. 249, 1–35.
Babinski, M., 2020. O Orógeno Araçuaí à luz da geocronologia: um tributo a https://doi.org/10.1016/j.chemgeo.2007.11.005.
Umberto Cordani. In: Bartorelli, A., Teixeira, W., Brito Neves, B.B. (Eds.), Soares, C.C.V., Queiroga, G., Pedrosa-Soares, A.C., Gouvêa, L.P., Valeriano, C.M.,
Geocronologia e Evolução Tectônica do Continente Sul-Americano: a contribuição de Melo, M.G., Marques, R.A., Freitas, R.D.A., 2020. The Ediacaran Rio Doce magmatic
Umberto Giuseppe Cordani. Solarias Edições Culturais, São Paulo, p. 728p. arc in the Araçuaí – Ribeira boundary sector, southeast Brazil: lithochemistry and
Peixoto, E., Pedrosa-Soares, A.C., Alkmim, F.F., Dussin, I.A., 2015. A suture-related isotopic (Sm–Nd and Sr) signatures. J. S. Am. Earth Sci., 102880 https://doi.org/
accretionary wedge formed in the Neoproterozoic Araçuaí orogen (SE Brazil) during 10.1016/j.jsames.2020.102880.
Western Gondwanaland assembly. Gondwana Res. 27, 878–896. https://doi.org/ Söderlund, U., Patchett, J.P., Vervoort, J.D., Isachsen, C.E., 2004. The 176Lu decay
10.1016/j.gr.2013.11.010. constant determined by Lu-Hf and U-Pb isotope systematics of Precambrian mafic
Peixoto, C.A., Heilbron, M., Ragatky, D., Armstrong, R., Dantas, E., Valeriano, C.M., intrusions. Earth Planet Sci. Lett. 219, 311–324. https://doi.org/10.1016/S0012-
Simonetti, A., 2017. Tectonic evolution of the Juvenile Tonian Serra da Prata 821X(04)00012-3.
magmatic arc in the Ribeira belt, SE Brazil: implications for early West Gondwana Souza, M.E., Martins, M.S., Queiroga, G.N., Pedrosa-Soares, A.C., Dussin, I., Castro, M.P.,
amalgamation. Precambrian Res. 302, 221–254. https://doi.org/10.1016/j. Serrano, P., 2022. Time and isotopic constraints for Early Tonian basaltic
precamres.2017.09.017. magmatism in a large igneous province of the São Francisco – Congo paleocontinent
Peixoto, C.A., Lobato, M., Freitas, N.C., Heilbron, M., Dussin, I., Dantas, E., 2022. (Macaúbas basin, Southeast Brazil). Precambrian Res. 373, 106621 https://doi.org/
Geochronological constraints of high-grade metasedimentary rocks of the italva and 10.1016/j.precamres.2022.106621.

21
C. Araujo et al. Journal of South American Earth Sciences 129 (2023) 104490

Stacey, J.S., Kramers, J.D., 1975. Approximation of terrestrial lead isotope evolution by a van Achterbergh, E., Ryan, C.G., Jackson, S.E., Griffin, W., 2001. Data reduction software
two-stage model. Earth Planet Sci. Lett. 26, 207–221. https://doi.org/10.1016/ for LA-ICP-MS. In: Sylvester, P. (Ed.), Laser-ablation-ICPMS in the Earth Sciences,
0012-821X(75)90088-6. Principles and Applications, Mineralogical Association of Canada, Short Course
Sun, S.S., McDonough, W.F., 1989. Chemical and isotopic systematics of oceanic basalts: Series, vol. 29, pp. 239–243.
implications for mantle composition and processes. In: Saunders, A.D., Norry, M. Vervoort, J.D., Kemp, A.I.S., 2016. Clarifying the zircon Hf isotope record of
(Eds.), Magmatism in Ocean Basins, vol. 42. Geological Society of London Special crust–mantle evolution. Chem. Geol. 425, 65–75. https://doi.org/10.1016/j.
Publications, pp. 313–345. https://doi.org/10.1144/GSL.SP.1989.042.01.19. chemgeo.2016.01.023.
Tedeschi, M., Novo, T., Pedrosa-Soares, A., Dussin, I., Tassinari, C., Silva, L.C., Victoria, A., Pedrosa-Soares, A., Cruz, S.C.P., Lana, C., Dantas, E., Dussin, I., Borges, R.,
Gonçalves, L., Alkmim, F., Lana, C., Figueiredo, C., Dantas, E., Medeiros, S., De 2022. Magmatic diversity in continental rifts: a case study on the Early Tonian,
Campos, C., Corrales, F., Heilbron, M., 2016. The Ediacaran Rio Doce magmatic arc plutono-volcanic Salto da Divisa complex, Araçuaí Orogen, Eastern Brazil. Lithos
revisited (Araçuaí-Ribeira orogenic system, SE Brazil). J. S. Am. Earth Sci. 68, 434–435, 106920. https://doi.org/10.1016/j.lithos.2022.106920.
167–186. https://doi.org/10.1016/j.jsames.2015.11.011. Vieira, V.S., Menezes, R.G., Orgs, 2015. Geologia e Recursos Minerais do Estado do
Tedeschi, M., Pedrosa-Soares, A., Dussin, I., Lanari, P., Novo, T., Pinheiro, M.A.P., Estado do Espírito Santo: texto explicativo do mapa geológico e de recursos minerais,
Lana, C., Peters, D., 2018. Protracted zircon geochronological record of UHT garnet- escala 1:400.000. In: Programa Geologia Do Brasil, CPRM - Serviço Geológico Do
free granulites in the Southern Brasília orogen (SE Brazil): petrochronological Brasil, Belo Horizonte.
constraints on magmatism and metamorphism. Precambrian Res. 316, 103–126. Whitney, D.L., Evans, B.W., 2010. Abbreviations for names of rock-forming minerals.
https://doi.org/10.1016/j.precamres.2018.07.023. Am. Mineral. 95, 185–187. https://doi.org/10.2138/am.2010.3371.
Tuller, M.P., 1993. Geologia da Folha Colatina. Programa de Levantamentos Geológicos Wisniowski, L., Pedrosa-Soares, A., Medeiros-Junior, E., Belém, J., Dussin, I.,
Básicos do Brasil, Folha SE.24-Y-C-VI - Colatina. Brasília, DNPM/CPRM. Queiroga, G., 2021. Ultra-high temperature, mid-crustal level, contact
Tupinambá, M., Heilbron, M., Valeriano, C.M., Porto Jr., R., Blanco de Dios, F., metamorphism imprinted on granulite facies paragneisses by a norite intrusion (São
Machado, N., Silva, L.G.E., Almeida, J.C.H., 2012. Juvenile contribution of the Gabriel da Baunilha, Araçuaí orogen, southeast Brazil). J. Metamorph. Geol. 1–29.
neoproterozoic rio Negro magmatic arc (Ribeira belt, Brazil): implications for Yardley, B.W.D., Mackenzie, W.S., Guildford, C., 1996. Atlas de rocas metamórficas y sus
western Gondwana amalgamation. Gondwana Res. 21, 422–438. https://doi.org/ texturas. Barcelona, Masson, p. 128.
10.1016/j.gr.2011.05.012.

22

You might also like