Você está na página 1de 166

UNIVERSIDADE ESTADUAL DE CAMPINAS

INSTITUTO DE GEOCIÊNCIAS

MARCO ANTÔNIO DELINARDO DA SILVA

“EVOLUÇÃO TECTONO-METAMÓRFICA DO EMBASAMENTO


MESOARQUEANO DO DOMÍNIO CARAJÁS, PROVÍNCIA CARAJÁS”

CAMPINAS
2018
MARCO ANTÔNIO DELINARDO DA SILVA

“EVOLUÇÃO TECTONO-METAMÓRFICA DO EMBASAMENTO


MESOARQUEANO DO DOMÍNIO CARAJÁS, PROVÍNCIA CARAJÁS”

TESE DE DOUTORADO APRESENTADA AO


INSTITUTO DE GEOCIÊNCIAS DA
UNICAMP PARA OBTENÇÃO DO TÍTULO
DE DOUTOR EM CIÊNCIAS, NA ÁREA DE
GEOLOGIA E RECURSOS NATURAIS.

ORIENTADORA: PROFA. DRA. LENA VIRGÍNIA SOARES MONTEIRO


CO-ORIENTADOR: PROF. DR. TICIANO JOSÉ SARAIVA DOS SANTOS

ESTE EXEMPLAR CORRESPONDE À VERSÃO FINAL


DA TESE DEFENDIDA PELO ALUNO MARCO ANTÔNIO
DELINARDO DA SILVA, ORIENTADO PELA PROFA.
DRA. LENA VIRGÍNIA SOARES MONTEIRO.

CAMPINAS
2018
Agência(s) de fomento e nº(s) de processo(s): CAPES
ORCID: https://orcid.org/0000-0003-0365-6729

Ficha catalográfica
Universidade Estadual de Campinas
Biblioteca do Instituto de Geociências
Marta dos Santos - CRB 8/5892

Silva, Marco Antônio Delinardo da, 1984-


Si38e SilEvolução tectono-metamórfica do embasamento mesoarqueano do Domínio
Carajás, Província Carajás / Marco Antônio Delinardo da Silva. – Campinas,
SP : [s.n.], 2018.

SilOrientador: Lena Virgínia Soares Monteiro.


SilCoorientador: Ticiano José Saraiva dos Santos.
SilTese (doutorado) – Universidade Estadual de Campinas, Instituto de
Geociências.

Sil1. Geologia estratigráfica - Arqueano. 2. Geocronologia. 3. Geoquímica. 4.


Carajás, Serra dos (PA). I. Monteiro, Lena Virgínia Soares, 1970-. II. Santos,
Ticiano José Saraiva dos, 1964-. III. Universidade Estadual de Campinas.
Instituto de Geociências. IV. Título.

Informações para Biblioteca Digital

Título em outro idioma: Tectonic-metamorphic evolution of the Mesoarchean basement of


the Carajás Domain, Carajás Province
Palavras-chave em inglês:
Geology, Straitigraphic - Archaen
Geochronology
Geochemistry
Carajas, Serra dos (PA)
Área de concentração: Geologia e Recursos Naturais
Titulação: Doutor em Geociências
Banca examinadora:
Lena Virgínia Soares Monteiro [Orientador]
Elton Luiz Dantas
Davis Carvalho de Oliveira
Vinícius Tieppo Meira
Wanilson Luiz Silva
Data de defesa: 10-08-2018
Programa de Pós-Graduação: Geociências

Powered by TCPDF (www.tcpdf.org)


UNIVERSIDADE ESTADUAL DE CAMPINAS
INSTITUTO DE GEOCIÊNCIAS

AUTOR: Marco Antônio Delinardo da Silva

“EVOLUÇÃO TECTONO-METAMÓRFICA DO EMBASAMENTO


MESOARQUEANO DO DOMÍNIO CARAJÁS, PROVÍNCIA CARAJÁS”

ORIENTADORA: Profa. Dra. Lena Virgínia Soares Monteiro


CO-ORIENTADOR: Prof. Dr. Ticiano José Saraiva dos Santos

Aprovado em: 10 / 08 / 2018

EXAMINADORES:

Profa. Dra. Lena Virgínia Soares Monteiro - Presidente

Prof. Dr. Davis Carvalho de Oliveira

Prof. Dr. Elton Luiz Dantas

Prof. Dr. Vinícius Tieppo Meira

Prof. Dr. Wanilson Luiz Silva

A Ata de Defesa assinada pelos membros da Comissão Examinadora,


consta no processo de vida acadêmica do aluno.

Campinas, 10 de agosto de 2018.


AGRADECIMENTOS

Eu gostaria de começar agradecendo minha família que sempre me incentivou a estudar, o que
me permitiu chegar até aqui. Na qualidade de homem branco, que não fez mais que sua
obrigação, eu lembro de parceiros da minha rua, da escola pública, mulheres, homens, negros
e brancos que também nasceram ali na zona leste de São Paulo. Boa parte deles está, neste
exato momento, bem distante da realidade que eu vivo atualmente. Uma realidade de classe
média, de uma vida digna. É só isso que eles queriam também. Mas o gargalo é muito
apertado pra quem nasce naquela região. Ainda mais se você não é branco e homem (como se
espera que você seja).
Eu não vi e nem vivi essa meritocracia de que tanto se fala. Eu vi e vivi uma exceção a regra.
Quanto mais distoante do status dominante, mais fácil essa exceção ela se vende como regra.
As pessoas acreditam nisso, elas sonham… De acordo com Yuval Harari no livro Sapiens –
Uma breve história da humanidade, o grande passo da nossa espécie se deu por meio da
revolução cognitiva, que permitiu aos homo-sapiens, através de crenças em comum, confiar
uns nos outros. Isso explica, dentre outros paradigmas modernos, o mito da meritocracia… As
pessoas, compartilhando dessa crença, sonham… Um sonho que nunca para de ser sonhado
até que se torna um pesadelo. Em um país como o Brasil, é possível que todos tenham ao
menos uma vida digna, como a que eu julgo ter agora. A função social exercida por um gari, é
tão importante quanto a função social exercida por um médico. Por que a diferença no
tratamento social? Uma ideia que martela constantemente minha cabeça é: O que fazer com o
que eu aprendi até aqui? Do ponto de vista social… Por vezes até ocupando posições que
podem ajudar a mudar a regra, mesmo que localmente. Eu vejo oposição a política de cotas na
universidade, burocratização a educação continuada de egressos que vem de uma condição
social dificil e precisam de mais suporte. Enfim, a pergunta fica ali, martelando, espero que
sempre fique ali, porque são muitas barreiras.
Focando nos agradecimentos, eu não poderia deixar de destacar a gigantesca dedicação da
Profa. Lena ao seus alunos. Ela, definitivamente, faz o possível, o impossível pra que a gente
consiga desenvolver os nossos trabalhos. Não há como avaliar a gratidão que eu tenho por ela,
por ter me ajudado a construir um pedaço enorme da minha vida; só posso dizer, muito
obrigado Lena… Divivindo a responsabilidade pela construção da minha carreira como
geólogo e pesquisador está o cabra mais estiloso da geociências, Prof. Ticiano. Um grande
parceiro e grande conselheiro, muito obrigado Tici! Nesse contexto agradeço a todos os
docentes do IGe que, por vezes me ensinaram mais do que geologia. Agradeço também aos
queridos funcionários do nossa universidade, que ficam ali nos bastidores, fazendo tudo
funcionar. Prefiro não citar nomes para não cometer uma injustiça. Agradeço aos parceiros da
minha turma de geologia, da moradia, da rep, do Lava Junkies, da escalada. Para finalizar,
agradeço a minha atual namorida e grande companheira Erica Santos de Sousa, que viveu
comigo todos os momentos do desenvolvimento deste trabalho e compreendeu a necessidade
de sua realização, mesmo em momentos que tivemos que abdicar passar tempo juntos para
que eu pudesse me dedicar ao trabalho. Retroceder nunca, desistir jamais…
SÚMULA CURRICULAR

Marco Antônio Delinardo da Silva


Formado em geologia pela Universidade Estadual de Campinas (Unicamp) concluiu o
mestrado e desenvolveu o doutorado no Programa de Pós-Graduação em Geociências pela
mesma universidade. Tem experiência na área de Geociências com enfase em Geologia com
particular interesse nas áreas de mineralogia, petrologia ignea e metamórfica, mapeamento
geológico, geologia econômica e geotectônica. Participou, durante a graduação, do programas
de iniciação científica, com projetos nas areas de geologia ambiental e metalogenese, e do
Programa de Apoio Didático (PAD). Na pós-graduação participou do Programa de Estágio
Docente (PED), presidiu o Student Chapter vinculado a Society of Economic Geology no ano
de 2012 e participou, como representante discente, da Comissão do Departamento de
Geologia e Recursos Naturais e da Comissão da Biblioteca.
No periodo, publicou três capitulos de livro, um artigo em periódico internacional e
um artigo em periódico nacional. Submete dois artigos para publicação em periódicos
internacionais.
RESUMO

O Domino Carajás, Província Carajás (PA; Cráton do Amazonas), representa um


singular exemplo de rochas de afinidade TTG mesoarqueanas (ca. 3,06–2,93 Ga) que
atingiram o metamorfismo de facies granulito de ultra-alta temperatura (> 900 ºC). As rochas
ortoderivadas estão divididas nas unidades: (i) Ortogranulito Xicrim-Cateté (cristalização: ca.
3,06–2,93 Ga; metamorfismo: 2,89-2,85 Ga; T = 907–1128 ºC e P = 8,1–13,7 kbar) e
Complexo Xingu (cristalização: ca. 2,97–2,93 Ga; metamorfismo: 2,86 Ga; T = 785 ºC e P =
8.8 kbar). O Ortogranulito Xicrim-Cateté é composto de ortopiroxênio-diopsídio gnaisses
metatexiticos de composição tonalítica a granodioritica, diopsídio-pargasita granulitos,
granulitos máficos metatexíticos e metanoritos. O Complexo Xingu inclui hornblenda-biotita
gnaisses metatexíticos de composição tonalítica a granodiorítica e anfibolitos. As rochas
félsicas apresentam as seguintes características químicas: (i) são rochas ricas em silica (65,58
< wt% SiO2 < 72,90); (ii) com alto conteúdo de Na2O (4,36 – 6,10 wt%); (iii) pobres em
elementos ferromagnesianos (FeOt+MgO+MnO+TiO2 < 4wt%); (iv) que apresentam
anomalias negativas de Nb e Ta; (v) padrão fracionado de elementos terras raras e baixo
conteúdo de elementos terras raras pesados (e.g. 0,10 < Yb ppm < 1,90; 5,48 LaN/YbN ppm <
138,33); (vi) baixo conteúdo de Y (1,10 < Y ppm < 14,40); e (vii) alta razão Sr/Y (20,93 –
731,82 ppm). As rochas máficas são caracterizadas por: (i) afinidade cálcio-alcalina; (ii)
índice de saturação metaluminoso a levemente peraluminoso; (iii) conteúdo de Ti (3477 < Ti
ppm < 7734), Zr (11 < Zr ppm < 29), (Y 11 < Y ppm < 29) e V (114 < V ppm < 371)
semelhantes ao de basaltos de arco de ilhas; e (iv) anomalias negativas de Nb, Ta e Ti. O
conjunto de dados indica que estas rochas se formaram em um ambiente de arco magmático
entre ca. 3,06 e 2,93 Ga. O sistema de arco magmático evolui para um sistema colisional com
a aproximação do Domínio Rio Maria (porção sul da Província Carajás). A colisão dos blocos
Carajás e Rio Maria culminou em metamorfismo de ultra-alta temperatura (ca. 2,89–2,85 Ga)
associado a fusão por quebra de minerais hidratados e por adição de fluidos. A zona de sutura
é marcada por serpentinitos e peridotidos de alto grau metamórfico entre as microplacas dos
Domínios Carajás e Rio Maria.

Palavras-Chave: TTG, UHT, Mesoarqueano, Província Carajás


ABSTRACT

The Carajás Domain, Carajás Province (PA; Amazon Craton), represents a singular
example of Mesoarchean TTG rocks (3.06-2.93Ga) that reached the ultra-high temperature
metamorphism (> 900 ºC). The orthoderived rocks are divided into: (i) Xicrim-Cateté
Orthogranulite (crystallization: ca. 3.06–2.93 Ga; metamorphism: 2.89-2.86Ga; T = 907–1128
ºC and P = 8.1–13.7 kbar); and (ii) Xingu Complex (cristalization: ca. 2.97–2.93 Ga;
metamorphism: 2.86Ga; T = 785 ºC and P = 8.8 kbar). The Xicrim-Cateté Orthogranulite
encompasses metatexitic orthopyroxene-diopside tonalite to granodiorite gneisses, metatexitic
diopside-pargasite and mafic granulites, and metanorites. The Xingu Complex is composed of
hornblende-biotite and biotite tonalite to granodiorite gneisses and amphibolites. The felsic
rocks show the following geochemical characteristics: (i) silica-rich rocks (65.58 < wt% SiO2
< 72.90), (ii) with high Na2O contents (4.36 – 6.10 wt%), (iii) are poor in ferromagnesian
elements (FeOt + MgO + MnO + TiO2 < 4wt.%), (iv) have Nb and Ta negative anomalies, (v)
fractionated pattern due to low HREE concentrations (0.10 < Yb ppm < 1.90; 5.48 < LaN/YbN
ppm < 138.33), (vi) low Y contents (1.10 < Y ppm < 14.40) and (vii) high Sr/Y ratios (20.93
– 731.82 ppm). The mafic magmatism is characterized by: (i) calc-alkaline affinity; (ii)
metaluminous to slightly peraluminous saturation index; (iii) Ti (3477 < Ti ppm < 7734), Zr
(11 < Zr ppm < 29), Y (11 < Y ppm < 29) and V (114 < V ppm < 371) contents of volcanic
arc basalts (Pearce, 1996); and (iv) Nb, Ta and Ti negative anomalies. The database indicate
that these rocks were formed in a magmatic arc setting between ca. 3.06 and 2.93Ga. The arc
setting evolved into a collisional setting with the approximation of the Rio Maria Domain (the
southern portion of the Carajás Province). The collision between these two blocks lead to the
ultra-high temperature metamorphism (ca. 2.89–2.85 Ga) related to the dehydration melting
and water-fluxed melting. The suture zone is maked by serpentinites and high-grade
peridotites between the Carajás and Rio Maria domain microplates.

Keywords: TTG, UHT, Mesoarchean, Carajás Province


INDICE DE FIGURAS

FIGURA 1. DIVISÃO DAS PROVÍNCIAS GEOCRONOLÓGICAS DO CRÁTON AMAZÔNICO NO ESTADO DO PARÁ


(VASQUEZ & ROSA-COSTA, 2008)................................................................................................................ 28
FIGURA 2. MAPA GEOLÓGICO DO DOMÍNIO CARAJÁS (MODIFICADO DE COSTA ET AL., 2016)................................ 30

Anexo 01: Artigo “Unraveling the Mesoarchean metamorphic evolution of the Carajás
Province, Amazon Craton: constrains from petrological evidences, mineral chemistry and U-
Pb geochronology”

FIGURE 1. A) THE DISTRIBUTION OF THE CARAJÁS PROVINCE IN THE PARÁ ESTATE, IN THE NORTH OF BRAZIL, AT
RIGHT, AND ITS SUBDIVISION INTO THE CARAJÁS (CD) AND RIO MARIA DOMAIN (RM), AT LEFT. THE BLACK

RECTANGLE SHOWS THE MAP AREA. B) GEOLOGICAL MAP OF THE NORTHERN PORTION OF THE CARAJÁS
PROVINCE, MODIFIED FROM COSTA ET AL. (2016). ....................................................................................... 39
FIGURE 2. GEOLOGICAL MAP OF THE GOIABA CREEK AREA. THE MAP RESULTS FROM THE FIELD AND
PETROGRAPHIC STUDIES INTEGRATED WITH THE COMPILED DATA FROM THE PREVIOUS WORKS OF

DALL’AGNOL ET AL. (2005; 1), FEIO ET AL. (2012; 2), SANTOS ET AL. (2013A; 3), SOUSA ET AL. (2015; 4),
LEITE-SANTOS (2016; 5) FEITO ET AL. (2013; 6) GABRIEL AND OLIVEIRA (2013; 7); AND ALMEIDA ET AL.
(2011; 8), ARAÚJO AND MAIA (1991), MARANGOANHA (2016), COSTA ET AL. (2016)................................. 45
FIGURE 3. FIELD AND PETROGRAPHIC FEATURES OS THE METATEXITIC MAFIC GRANULITE AND DIOPSIDE-
PARGASITE-PLAGIOCLASE GRANULITE. A) METER SCALE BOULDERS OF THE METATEXITIC MAFIC
GRANULITES. THE MIGMATIZATION IS PARALLEL TO THEIR FOLIATION. B) BOUDINS OF THE METATEXITIC
MAFIC GRANULITE INSIDE THE METATEXITIC ORTHOPYROXENE-DIOPSIDE GNEISS. THE BLACK DOTS AND THE

MILK QUARTZ VEINS IN THE GNEISSIC FOLIATION HIGHLIGHT THE BOUDINAGE PROCESS; C) A CLOSER LOOK
IN THE CONTACT BETWEEN THE MAFIC AND FELSIC GRANULITES SUGGESTS THAT THE FIRST MAY BE

INTRUSIVE. D) BLASTO-SUBOPHITIC TEXTURE IN THE METANORITES. THE PLAGIOCLASE CRYSTALS SHOW


THE DEFORMATIONAL AND WEDGE-SHEPED TWINS AND LOCALLY SUBGRAIN ROTATION

RECRYSTALLIZATION. E) THE TEXTURE TRANSFORMATION DURING THE REPLACEMENT OF ENSTATITE (EN)


AND CA-RICH PLAGIOCLASE BY NA-RICH PLAGIOCLASE (PL) + PARGASITE (PRG) + QUARTZ (QTZ) +
MAGNETITE (MGT) WHICH IS RELATED TO THE MELT TEXTURES SHOWED BELOW; F) PLAGIOCLASE CRYSTALS

WITH CA-RICH (RED) NUCLEI AND NA-RICH (BLUE) BOUNDARY IN THE SEM COMPOSITIONAL MAPS. THE NA-

RICH BOUNDARY OCCURS TOGETHER WITH THE HORNBLENDE (HBL) CRYSTALS IN BETWEEN PLAGIOCLASE
AND ENSTATITE. THE ENSTATITE MAY SHOW DIOPSIDE (DI) CRYSTALS INSIDE IT. G) RUTILE (RT)
EXOLUTIONS INSIDE THE ENSTATITE. THE INCLUSIONS ARE REPLACED BY MAGNETITE. H) THE PARAGENESIS
OF THE METANORITES REPRESENTED BY PLAGIOCLASE + DIOPSIDE + ENSTATITE. THE GRANOBLASTIC
TEXTURE IS DEFINED BY INTERLOBATE CONTACTS DUE TO GRAIN BOUNDARY MIGRATION, BULGING AND

SUBGRAIN ROTATION RECRYSTALLIZATION; I) THE SEM IMAGE SHOWS THE PARAGENESIS OF THE MAFIC
GRANULITE (PLAGIOCLASE + DIOPSIDE + ENSTATITE + PARGASITE). THE UPDATED LIST OF WHITNEY AND
EVANS (2010) FOR THE ABBREVIATIONS OF THE ROCK-FORMING MINERALS WAS USED. .............................. 48
FIGURE 4. FIELD AND PETROGRAPHIC FEATURES RELATED TO THE MIGMATIZATION IN THE METATEXITIC MAFIC
GRANULITES. A) CHARACTERISTIC ASPECT OF THE BOULDERS OF METATEXITIC MAFIC GRANULITE WITH THE
STROMATIC AND NET STRUCTURES. B) DETAIL IN THE LEUCOSOME-RESIDUMM INTERFACE HIGHLIGHTING
THE HORNBLENDE-RICH MELANOSOME; C) SHOLLEN DIATEXITES HIGHLIGHTING THE ASSIMILATION OF THE
MAFIC GRANULITE BY THE LEUCOSOMES. D AND E) THE CHARACTERISTICS OF THE ORTHPYROXENE
PHENOCYRSTS IN THE “JGUAR SKIN” DIATEXITE (D) AND IN THE LEUCOSOMES OF THE METATEXITE (E); F
AND G) ORTHOPYROXENE (OPX) AND PLAGIOCLASE (PL) MEGACRYSTS IN THE LEUCOSOMES. THE
ORTHOPYROXENE CRYSTALS SHOW UNDULATORY EXTINCTION AND THE PLAGIOCLASE CRYSTALS SHOW

CORROSION-LIKE TEXTURES; H) PLAGIOCLASE (PL) CRYSTALS IN THE METATEXITIC MAFIC GRANULITES


SHOWING AMOEBOID FILM OF MELT. THE IMAGE HIGHLIGHTS THE RELATION BETWEEN HORNBLENDE (HBL)
CRYSTALS AND MELT POCKETS; I) PLAGIOCLASE-BEARING VEINS THAT OCCUR IN THE METATEXITIC MAFIC
GRANULITES, ESPECIALLY IN THE METANORITES. THE HORNBLENDE CRYSTALS GENERALLY OCCUR IN THE
BOUNDARIES OF THE VEINS; I). THE UPDATED LIST OF WHITNEY AND EVANS (2010) FOR THE
ABBREVIATIONS OF THE ROCK-FORMING MINERALS WAS USED. ................................................................... 50

FIGURE 5. FIELD AND PETROGRAPHIC FEATURES OF THE METATEXITIC ORTHOPYROXENE-DIOPSIDE GNEISS. A) THE
MASSIVE GRANULITIC TEXTURE OF THE METATEXITIC ORTHOPYROXENE-DIOPSIDE GNEISS. THE RED CIRCLE
SHOWS THE ENSTATITE CRYSTAL SURROUNDED BY BIOTITE; B) THE PEAK PARAGENESIS OF THE METATEXITIC

ORTHOPYROXENE-DIOPSIDE GNEISS. IN THE IMAGE THE UNDULATORY EXTINCTION, BULGING AND SUBGRAIN

RECRYSTALLIZATION AND GRAIN BOUNDARY MIGRATION CAN BE OBSERVED IN THE PLAGIOCLASE (PL),
QUARTZ (QTZ), K-FELDSPAR (KFS), DIOPSIDE (DI) AND ENSTATITE (EN) CRYSTALS; C) THE IMAGE
HIGHLIGHT THE CHARACTERISTIC DEFORMATIONAL TWINS OF THE PLAGIOCLASE AND ITS WEDGE-SHAPE

ASPECT; D) PLAGIOCLASE AND ENSTATITE CRYSTALS WITH PRESERVED GROWTH CRYSTALLINE FACES; E)
THE SITES OF WELL-DEVELOPED GRANOBLASTIC TEXTURE; F) THE EW SOUTH-DIPPING GNESSIC FOLIATION
DEVELOPED ALONG THE SHEAR ZONES; G) THE RETROGRADE PARAGENESIS WITH BIOTITE (BT) + QUARTZ +
MAGNETITE (MGT) IN THE METATEXITIC ORTHOPYROXENE-DIOPSIDE GNEISS WITH MASSIVE TEXTURE; H)
THE LEPIDOBLASTIC AND DECUSSATE TEXTURE OF BIOTITE (BT) IN THE METATEXITIC ORTHOPYROXENE-
DIOPSIDE GNEISS WITH GNEISSIC FOLIATION. I) BIOTITE-RICH DOMAINS WITH HORNBLENDE CRYSTALS
REPLACING ENSTATITE. THE BLUE DOTS, SHOWS THE GRANOBLASTIC TEXTURE LOCALLY DEVELOPED. THE
UPDATED LIST OF WHITNEY AND EVANS (2010) FOR THE ABBREVIATIONS OF THE ROCK-FORMING MINERALS

WAS USED. .................................................................................................................................................... 52

FIGURE 6. FIELD AND PETROGRAPHIC FEATURES OF THE MELT-RELATED FEATURES IN THE METATEXITIC
ORTHOPYROXENE-DIOPSIDE GNEISS. A) THE STROMATIC STRUCTURE IN THE METATEXITIC

ORTHOPYROXENE-DIOPSIDE GNEISS WITH MASSIVE GRANULITE TEXTURE AND IN (B) THE ROCKS WITH
GNEISSIC FOLIATION. THE LEUCOSOME HAS A COARSE-GRAINED TEXTURE; C) THE MELT-RELATED TEXTURE
SHOWED BY K-FELDSPAR CRYSTAL CROSSCUTTING A PLAGIOCLASE CRYSTAL; D) THE AMOIEBOID FORM OF
K-FELDSPAR, PLAGIOCLASE AND QUARTZ CRYSTALS IN THE BOUNDARY BETWEEN THE METATEXITIC
ORTHOPYROXENE-DIOPSIDE GNEISS AND THE LEUCOSOMES; E) THE COARSE-GRAINED LEUCOSOMES WITH
PLAGIOCLASE MEGACRYSTS, K-FELSPAR WITH MYMERKITES AND POST-CRYSTALLIZATION DEFORMATION,
ASSOCITATED WITH BULGING AND SUBGRAIN ROTATION RECRYSTALLIZATION AND GRAIN BOUNDARY

MIGRATION; F) THE PLAGIOCLASE CRYSTALS SHOW GROWTH CRYSTAL FACES IN THE LEUCOSOMES. THE
UPDATED LIST OF WHITNEY AND EVANS (2010) FOR THE ABBREVIATIONS OF THE ROCK-FORMING MINERALS

WAS USED. .................................................................................................................................................... 53


FIGURE 7. FIELD AND PETROGRAPHY FEATURES OF THE METATEXITIC HORNBLENDE-BIOTITE AND BIOTITE
GNEISSES OF THE XINGU COMPLEX. A) OUTCROP OF THE METATEXITIC BIOTITE GNEISS WITH THE LESSER
MELT-FRACTION OBSERVED IN THE GOIABA CREEK AREA; B) METATEXITIC PARGASITE-BIOTITE GNEISS WITH

COARSE PARGASITE CRYSTALS WITH PLAGIOCLASE INCLUSIONS (POIKILOBLASTIC TEXTURE) BEING


REPLACED BY BIOTITE. THE IMAGE ALSO SHOWS THE APATITE + PISTACITE + QUARTZ REPLACING BOTH,
PARGASITE AND BIOTITE. C) METATEXITIC BIOTITE GNEISS WITH THE GRANOBLASTIC AND DECUSSATE
TEXTURES. THE APATITE + PISTACITE + QUARTZ PARAGENESIS IS ALSO PRESENT OVER BIOTITE AND
PLAGIOCLASE. D) AN OUTCROP OF THE METATEXITIC BIOTITE GNEISS INSIDE THE NOVA CANADÁ
LEUCOGRANITE. THE GNEISS SHOWS A HIGH MELT-FRACTION AND THE SEGREGATION OF THE NEOSOME INTO
QUARTZ-FELDSPHATIC LEUCOSOMES AND BIOTITE-RICH MELANOSOME. ADDITIONALY, THE LOW ANGLE
FOLIATION IS WELL-DEFINED AND THE MELT-RELATED STRUCTURES ARE PARALLEL TO IT. E AND F) SHEATH

AND DRAG FOLDS OBSERVED IN THE METATEXITIC HORNEBLENDE-BIOTITE AND BIOTITE GNEISSES. THE
LEUCOSOMES HIGHLIGHT THE STRUCTURES. IN THE CASE OF THE SHEATH FOLDS (E) THE STREACHING
LINEATION, PARALLEL TO THE AXIAL PLANES OF THE FOLDS, IS WELL-DEFINED IN THE LEUCOSOMES. G) THE

BIOTITE-RICH LEPIDOBLASTIC AND QUARTZ-FELDSPHATIC DOMAINS THAT DEFINES THE GNEISSIC FOLIATION

OF THESE ROCKS; H) THE RED ARROWS SHOW THE MAIN RECRYSTALLIZATION FEATURES IN THE
GRANOBLASTIC INTERLOBATE DOMAINS: BULGING AND SUBGRAIN ROTATION RECRYSTALLIZATION AND

GRAIN BOUNDARY MIGRATION. I) PISTACITE + QUARTZ SYMPLECTITES OVER THE BIOTITE, PLAGIOCLASE, K-
FELDSPAR AND QUARTZ CRYSTALS. THE UPDATED LIST OF WHITNEY AND EVANS (2010) FOR THE
ABBREVIATIONS OF THE ROCK-FORMING MINERALS WAS USED. ................................................................... 55

FIGURE 8. FIELD AND PETROGRAPHIC FEATURES OF THE AMPHIBOLITES OBSERVED WITHIN THE METATEXITIC
PARGASITE-BIOTITE AND BIOTITE GNEISSES OF THE XINGU COMPLEX AND IN THE NOVA CANADÁ
LEUCOGRANITE. A AND B) THE OCCURRENCE MODE OF THE AMPHIBOLITE REPRESENTED BY ELONGATED
BOUDINS IN THE FOLIATION OF THE HORNBLENDE-BIOTITE AND BIOTITE GNEISSES (A). FOLIATED AND
FOLDED AMPHIBOLITE XENOLITHS INSIDE THE NOVA CANADÁ LEUCOGRANITE (B). C AND D) MICRO-
ESTRUTURAL ASPECTS OF THE AMPHIBOLITES., REPRESENTED BY INTERLOBATE GRANOBLASTIC TEXTURE,

WITH EVIDENT GRAIN BOUDARY MIGRATION (C). INTRA-CRYSTALLINE DEFORMATION OF ALBITE, SHOWING
DEFORMATIONAL TWINS IN THE CORNER OF THE CRYSTALS. E AND F) PARAGENETIC EVOLUTION OF THE
AMPHIBOLITES EVIDENCED BY REPLACEMENT OF THE MAGNESIO-HORNBLENDE BY BROWN BIOTITE

CRYSTALS FOLLOWED BY THE REPLACEMENT OF BOTH BY TITANITE AND EPIDOTE. ..................................... 56

FIGURE 9. FIELD AND PETROGRAPHIC CHARACTERISTIC OF THE MELT-RELATED FEATURES OF THE XINGU
COMPLEX ROCKS. A AND B) THE MOST TYPICAL MELT-RELATED MICROSTRUCTURE OF THE METATEXITIC
HORNBLENDE-BIOTITE AND BIOTITE TONALITE GNEISSES. A) ELONGATE FILMS WITH CUSPATE SHAPE
GROWING FROM THE K-FELDSPAR (KFS) CRYSTALS SURROUNDED BY QUARTZ (QTZ) AND ALBITE (PL); B)
GRANITIC VEINLETS WITH K-FELDSPAR, QUARTZ AND PLAGIOCLASE; C AND D) MICROTEXTURES OF K-
FELDSPAR CRYSTALS IN THE LEUCOSOMES, WHICH INCLUDE THE MESOPERTHITIC TEXTURE (C) AND THE
DEVELOPMENT OF TARTAME TWINING (D) IN THE CORNERS OF THE CRYSTALS; E) RELATIONSHIP BETWEEN
THE LEUCOSOMES OF THE XINGU METATEXITIC BIOTITE TONALITE GNEISS AND THE NOVA CANADÁ
LEUCOGRANITE WHOSE CONNECTION COULD BE OBSERVED IN THE FIELD; F) BIOTITE CLUSTERS AND THE
BIOTITE STRIPES THAT DEFINE THE SCHILIEREN STRUCTURE OF THE NOVA CANADÁ LEUCOGRANITE; G) THE
RELATIONSHIP BETWEEN THE XINGU COMPLEX ROCKS AND THE ÁGUA LIMPA GRANODIORITE. THE
INTRUSIVE CONTACT IS MARKED BY THE CONTEMPORANEOUS LEUCOSOME FORMATION. IN THESE CASES,

THE LEUCOSOME ASSIMILATE THE K-FELDSPAR PHEOCRYSTS FROM THE ÁGUA LIMPA GRANODIORITE. H
AND I) FIELD AND PETROGRAPHIC CHARACTERISTIC OF THE SPACED HIGH ANGLE MYLONITIC FOLIATION THA

TRANSPOSES THE PREVIOUS STRUCTURES. IT ROTATES THE PREVIOUS STRUCTURES (H) AND IS RESPONSIBLE
FOR THE RECRYSTALLIZATION OF QUARTZ CRYSTALS (I). ............................................................................. 58

FIGURE 10. MINERAL CHEMISTRY DIAGRAMS FOR THE METATEXITIC ORTHOPYROXENE-DIOPSIDE GRANODIORITE
GNEISS (SM11P), METATEXITIC DIOPSIDE-PARGASITE-PLAGIOCLASE GRANULITE (SM41R), METATEXITIC
MAFIC GRANULITE (XA15B), METANORITES (SM42; XS111B), METATEXITIC PARGASITE-BIOTITE TONALITE
GNEISS (XS105P) AND THE AMPHIBOLITE (XS33). A) RIETMEIJER (1983) DIAGRAM FOR CHEMICAL
DIFFERENTIATION OF IGNEOUS AND METAMORPHIC ORTHOPYROXENE; B) SEM MAP FOR THE TERNARY MG–

FE–CA COMPOSITION SHOWING THE VARIATION IN THE 100CA/(FE+2+MG+CA) PARAMETER FROM CORE TO
RIM OF THE ENSTATITE CRYSTALS IN THE METANORITE; C) MN VS. MG BINARY DIAGRAM DISTINGUISHING
THE ENSTATITE COMPOSITION OF THE METATEXITIC ORTHOPYROXENE-DIOPSIDE GRANODIORITE GNEISS

(SM11P), MAFIC GRANULITE (XA15B) AND METANORITES (SM42; XS111B); D) AMPHIBOLE


CLASSIFICATION DIAGRAM OF HAWTHORNE ET AL. (2012); E) F/(F+CL+OH) VS. MG/(FE+MG) DIAGRAM
SHOWING THE DISTINCT HALOGEN COMPOSITION OF THE AMPHIBOLES IN THE GRANULITES (XA15B;
SM41R), METANORITES (SM42; XS111B), PARGASITE-BIOTITE GNEISS (XS105P) AND AMPHIBOLITE
(XS33); F) AL VS. SI DIAGRAM FOR THE FELDSPAR CRYSTALS ALLOWING THE DISTINCTION BETWEEN THE
FELDSPAR CRYSTALS IN THE XICRIM-CATETÉ ORTHOGRANULITE AND XINGU COMPLEX. ........................... 72

FIGURE 11. A) CL IMAGES OF THE MAIN ZIRCON TYPES OF THE MAFIC GRANULITE (XA15B3) ANALYSED BY U–PB
LA–ICP–MS WITH RESPECTIVE 207PB/206PB AGES; B) CL IMAGES OF THE MAIN ZIRCON GRAIN-TYPES OF THE
SCHOLLEN DIATEXITE LEUCOSOME (XS112A) ANALYZED BY U-PB SHRIMP IIE WITH RESPECTIVE
207 206
PB/ PB AGES; C) CL IMAGES OF THE MAIN ZIRCON GRAIN-TYPES OF THE SCHOLLEN DIATEXITE
LEUCOSOMES (XS112A) ANALYZED BY U–PB LA–ICP–MS WITH RESPECTIVE 207PB/206PB AGES; D, E AND
206
F) PB/238U VS. 207
PB/235U DIAGRAM OF THE MAFIC GRANULITE AND THE SCHOLLEN DIATEXITE
LEUCOSOME. D) RESULTS OF THE ANALYSES OF ZIRCON CORES IN THE MAFIC GRANULITE ANALYZED BY U–
PB LA–ICP–MS; E) RESULTS OF THE ANALYSES OF ZIRCON CORES IN THE LEUCOSOME OF THE SCHOLLEN
DIATEXITE ANALYZED BY U–PB SHRIMP IIE; F) RESULTS OF THE ANALYSES OF ZIRCON CORES IN THE
SCHOLLEN DIATEXITE LEUCOSOME ANALYZED BY U–PB LA–ICP–MS. ....................................................... 78

FIGURE 12. CL IMAGES OF ZIRCON FROM THE METATEXITIC ORTHOPYROXENE-DIOPSIDE GNEISSES OF THE XICRIM-
CATETÉ ORTHOGRANULITE. A) CL IMAGES OF THE XA15A1 SAMPLE ANALYZED BY LA-ICP-MS. THE
207
ARROWS SHOW THE SPOTS ANALYZED WITH THE RESPECTIVE PB/206PB AGES. B) CL IMAGES OF THE
XA15A1 SAMPLE ANALYZED BY SHRIMP IIE. THE ARROWS SHOW THE ANALYZED SPOTS WITH THE
207
RESPECTIVE PB/206PB AGES; C) CL IMAGES OF THE SM46N SAMPLE. THE WHITE BLACK CIRCLES SHOW
THE PLACE WHERE THE GRAINS WERE ANALYZED BY THE SHRIMP IIE METHOD WITH THE RESPECTIVE
207 206
PB/ PB AGES. D) CL IMAGES OF THE CMS22P SAMPLE. THE WHITE CIRCLES SHOW THE PLACE WHERE
THE GRAINS WERE ANALYZED BY THE SHRIMP IIE METHOD WITH THE RESPECTIVE 207PB/206PB AGES. E) CL
IMAGES OF THE TS23A SAMPLE. THE WHITE CIRCLES SHOW THE PLACE WHERE THE GRAINS WERE ANALYZED
207
BY THE SHRIMP IIE METHOD WITH THE RESPECTIVE PB/206PB AGES. F) CL IMAGES OF THE XA12A
SAMPLE. THE WHITE CIRCLES SHOW THE PLACE WHERE THE GRAINS WERE ANALYZED BY THE SHRIMP IIE
207
METHOD WITH THE RESPECTIVE PB/206PB AGES. THE ZIRCON GRAINS IN ALL THE SAMPLES SHARE SIMILAR
FEATURES, SUCH AS PRESERVED OSCILLATORY ZONING AND HIGH LUMINESCENCE RIMS. THE THICKNESS OF
THIS RIMS ARE VARIABLE. IN SOME CASES IT CAN GROW ALL OVER THE GRAIN, WHICH IS THE CASE OF THE

FIRST GRAIN IN THE B. .................................................................................................................................. 81

FIGURE 13. 206PB/238U VS. 207PB/235U DIAGRAM FOR THE METATEXITIC ORTHOPYROXENE-DIOPSIDE GNEISSES
OF THE XICRIM-CATETÉ ORTOGRANULITE. A) RESULTS OF THE ANALYSES OF THE ZIRCON CORES IN THE
XA15A1 SAMPLE ANALYZED BY LA-ICP-MS. B) ANALYSES OF ZIRCON CORES AND RIMS IN THE SAMPLE
XA15A1 (SHRIMP IIE). C) ANALYSES OF ZIRCON CORES AND RIMS IN THE SAMPLE SM46N (SHRIMP IIE).
D) ANALYSES OF ZIRCON CORES IN THE CMS22P SAMPLE (SHRIMP IIE), WHICH SHOWED TWO AGE
GROUPS; E) RESULTS OF THE ANALYSES OF ZIRCON CORES IN THE TS23A SAMPLE (SHRIMP IIE); F)
RESULTS OF THE ANALYSES OF ZIRCON CORES IN THE XA12A SAMPLE (SHRIMP IIE). ............................... 82
FIGURE 14. A) CL IMAGES OF THE MAIN ZIRCON TYPES OF THE BIOTITE METATEXITIC GNEISS OF THE XINGU
COMPLEX (XS04P), SHOWING ANALYZED SPOTS AND CORRESPONDING 207PB/206PB AGES. B) CL IMAGES OF
THE MOST CONCORDANT ZIRCON GRAINS OF THE PARGASITE-BIOTITE METATEXITIC GNEISS (XS105P) AND
207
THEIR PB/206PB AGES. ............................................................................................................................... 83
206
FIGURE 15. PB/238U VS. 207
PB/235U DIAGRAMS FOR THE METATEXITIC GNEISSES OF THE XINGU COMPLEX. A)
CONCORDIA DIAGRAM FOR THE BIOTITE METATEXITIC GNEISS OF THE XINGU COMPLEX (SAMPLE XS04P); B)
CONCORDIA DIAGRAM FOR THE HORNBLENDE-BIOTITE METATEXITIC GNEISS OF THE XINGU COMPLEX
(SAMPLE XS105P). ....................................................................................................................................... 84

Anexo 02: Artigo “Mesoarchean arc-related system between 3.0–2.9 Ga in the Carajás
Domain, Carajás Province, PA”

FIGURE 1. GEOLOGICAL MAP OF THE CARAJÁS DOMAIN, MODIFIED FROM COSTA ET AL. (2016). INCLUDE THE
GEOCHRONOLOGICAL DATA FROM MACHADO ET AL. (1991), MORETO ET AL., (2011), FEIO ET AL., (2013),
RODRIGUES ET AL. (2014), SOUSA ET AL., (2010), LEITE-SANTOS (2016), PIMENTEL AND MACHADO (1994),
SILVA ET AL., (2010) AND LEITE ET AL., (2004). .......................................................................................... 118
FIGURE 2. GEOLOGICAL MAP OF GOIABA CREEK AREA. THE MAP RESULTS FROM THE FIELD AND PETROGRAPHIC
STUDIES INTEGRATED WITH THE COMPILED DATA FROM THE PREVIOUS WORKS OF DALL’AGNOL ET AL.
(2005; 1), FEIO ET AL. (2012; 2), SANTOS ET AL. (2013; 3), SOUZA ET AL. (2015; 4), LEITE-SANTOS (2016; 5),
FEIO ET AL. (2013; 6) GABRIEL AND OLIVEIRA (2013; 7); AND ALMEIDA ET AL. (2011; 8). ......................... 121
FIGURE 3. FIELD AND PETROGRAPHIC FEATURES OF THE XICRIM-CATETÉ ORTHOGRANULITE. A) COARSE-GRAINED
METATEXITIC ORTHOPYROXENE-DIOPSIDE GRANODIORITE GNEISS WITH GRANULITIC TEXTURE AND

ANASTOMOSED SN FOLIATION; B) METATEXITIC ORTHOPYROXENE-DIOPSIDE TONALITE GNEISS WITH ITS


TYPICAL COARSE-GRAINED GRANOBLASTIC TEXTURE AND THE PLAGIOCLASE (PL) + QUARTZ (QTZ) + K-
FELDSPAR (KFS) + ORTHOPYROXENE (OPX) ASSEMBLAGE, WHICH LOCALLY INCLUDE DIOPSIDE AND
PARGASITE ; C) RETROGRADE METAMORPHISM IN THE GRANULITIC GNEISSES: ORTHOPYROXENE + DIOPSIDE
 BIOTITE + MAGNETITE; D) THE ABRUPT CONTACT BETWEEN THE MAFIC GRANULITES AND THE
GRANULITIC GNEISSES; E) METER SCALE FRAGMENTS OF MAFIC GRANULITES HIGHLY DEFORMED INSIDE OF
A LEUCOSOME POCKET ASSOCIATED WITH AN METATEXITIC ORTHOPYROXENE-DIOPSIDE TONALITE GNEISS;
F) COARSE-GRAINED MAFIC GRANULITE WITH LOCAL PATCH OF MELTING AND MELT MOBILIZATION
EVIDENCES; G) MAFIC GRANULITES WITH COARSE-GRAINED GRANOBLASTIC TEXTURE ASSOCIATED WITH
THE ASSEMBLAGE WITH PLAGIOCLASE + ORTHOPYROXENE + DIOPSIDE; H) METANORITE WITH RELIC OF
BLASTO-SUBOPHIC TEXTURE LOCALLY PRESERVED IN THE THIN SECTION; I) THE STROMATIC AND NET
STRUCTURES TYPICAL OF THE METATEXITIC MAFIC GRANULITES. .............................................................. 124

FIGURE 4. FIELD AND PETROGRAPHIC FEATURES OF THE XINGU COMPLEX. A) THE INTRUSIVE CONTACT BETWEEN
THE HIGH-MG GRANODIORITES FROM ÁGUA LIMPA SANUKITOID SUITE AND THE XINGU COMPLEX (XC),
CONTROLLED BY ITS LOW ANGLE FOLIATION; B) METATEXITIC ORTHOGNEISS FROM XINGU COMPLEX,
SHOWING ITS COMMON FIELD CHARACTERISTICS: LIGHT GREY, MEDIUM-GRAINED PALEOSOME (P), WHITE-
PINK COARSE-GRAINED LEUCOSOMES (L) AND SLENDER STRIPES OF BLACK MEDIUM-GRAINED MELANOSOME

(M); C) SCHOLLEN FRAGMENT OF THE XINGU COMPLEX (XC) INVOLVED BY THE NOVA CANADÁ
LEUCOGRANITE (NC). THE IN-SOURCE LEUCOSOMES IN THE METATEXITIC GNEISS OF XINGU COMPLEX
SHOWS A GRADUAL CONTACT WITH THE NOVA CANADÁ LEUCOGRANITE (NC); D) TYPICAL ASSEMBLAGE OF

THE GNEISS: HORNBLENDE (HBL)+ PLAGIOCLASE (PL)+ K-FELDSPAR (KFS)+ QUARTZ, BEING REPLACED BY
BIOTITE (BT)  EPIDOTE (EP); E) VEINS OF LEUCOSOMES COMPOSED OF QUARTZ + PLAGIOCLASE + K-
FELDSPAR CROSSCUTTING PLAGIOCLASE MEGACRYSTALS; F) CUSPATE FILMS OF K-FELDSPAR ALONG THE
BOUNDARY OF PLAGIOCLASE + QUARTZ + K-FELDSPAR; G) SYMPLECTITIC TEXTURE WITH EPIDOTE (EP),
QUARTZ AND BIOTITE (BT) RELATED TO THE RETROGRADE METAMORPHISM; H) THE BOUDIN-LIKE
OCCURRENCE OF AMPHIBOLITE ENCLAVES IN THE XINGU COMPLEX ORTHOGNEISS; I) AMPHIBOLITE WITH A
DECUSSATE TEXTURE AND THE PARAGENESIS WITH HORNBLENDE (HBL) + PLAGIOCLASE (PL) WHICH IS
REPLACED BY BIOTITE  EPIDOTE. ............................................................................................................. 126

FIGURE 5. A) NORMATIVE CLASSIFICATION OF THE GNEISSES OF THE XICRIM-CATETÉ ORTHOGRANULITE AND


XINGU COMPLEX IN THE DIAGRAM OF BARKER (1979); B) K-NA-CA TRIANGLE COMPARING THE EVOLUTION
OF THE SAMPLES FROM XICRIM-CATETÉ ORTHOGRANULITE AND XINGU COMPLEX WITH THE MODERN CALC-

ALKALINE MAGMAS (BLACK CA TREND). ................................................................................................... 132

FIGURE 6. A-B) GRANITE CLASSIFICATION OF FROST ET AL. (2001) APPLIED TO THE ORTHOGNEISSES OF XICRIM-
CATETÉ ORTHOGRANULITE E XINGU COMPLEX; C) SCHAND (1943) DIAGRAM FOR THE ALUMINIUM
SATURATION INDEX; D) INDIVIDUALIZATION OF ARCHEAN GRANITOIDS COMPILED IN A SINGLE DIAGRAM,
ACCORDING TO LAURENT ET AL. (2014). ..................................................................................................... 133

FIGURE 7. SPIDER PLOTS OF THE XICRIM-CATETÉ ORTHOGRANULITE AND XINGU COMPLEX GNEISSES. THE
AVERAGE VALUES OF HIGH, MEDIUM AND LOW PRESSURE TTG FROM MOYEN (2011) WERE PLOTTED
TOGETHER WITH THE SAMPLE DATA. .......................................................................................................... 134
FIGURE 8. DIAGRAMS OF MOYEN (2011) FOR HP (HIGH PRESSURE), MP (MEDIUM PRESSURE) AND LP (LOW
PRESSURE) TTG. ........................................................................................................................................ 135
FIGURE 9. PLOTS OF TRACE ELEMENT VS. ZR, EXEMPLIFYING THE GOOD CORRELATION OF THE CONSIDERED LEAST
CONTAMINATED MAFIC GRANULITES. AVERAGE VALUES OF THE LOWER AND UPPER CRUST OF RUDNICK AND

GAO (2014) ARE ALSO PLOTTED FOR COMPARISON. .................................................................................... 136


FIGURE 10. THE DIAGRAMS FOR THE CLASSIFICATION OF MAGMATIC AFFINITY (LEFT; AFM; IRVINE AND
BARAGAR, 1971) AND ALUMINIUM SATURATION INDEX (RIGHT; SCHAND, 1943) FOR THE ROCKS OF XICRIM-
CATETÉ ORTHOGRANULITE AND XINGU COMPLEX. ................................................................................... 141
FIGURE 11. DIAGRAMS OF PEARCE AND CANN (1973; A) AND WINCHESTER AND FLOYD (1977; B) MODIFIED BY
PEARCE (1996). VAB: VOLCANIC ARC BASALTS; MORB: MID-RIDGE OCEANIC BASALT; WPB: WITHIN
PLATE BASALTS; FRACTIONAL CRYSTALLIZATION VECTORS: PLAGIOCLASE+OLIVINE+AUGITE

(POA);PLAGIOCLASE+OLIVINE+AUGITE+MAGNETITE (POAM); PLAGIOCLASE+OLIVINE+AUGITE+

HORNBLENDE+MAGNETITE (POAHM). PETROGENETIC VECTORS: SPINEL LHERZOLITE (SP. LHERZ), GARNET


LHERZOLITE (GT. LHERZ). MM: MORB MANTLE SOURCE............................................................................ 142
FIGURE 12. SPIDER PLOTS OF SAMPLES FROM XICRIM-CATETÉ ORTHOGRANULITES (MAFIC GRANULITES) AND
XINGU COMPLEX AMPHIBOLITES. THE PRIMITIVE MANTLE NORMALIZATION VALUES COME FROM
MCDONOUGH AND SUN (1995). ................................................................................................................. 142
FIGURE 13. PLOTS OF TECTONIC SETTING OF PEARCE (1996) AND PEARCE (2014) FOR THE MAFIC ROCKS OF
XICRIM-CATETÉ ORTHOGRANULITE AND XINGU COMPLEX ROCKS. OIB: OCEAN ISLAND BASALT; E-MORB:
ENRICHED MIDDLE OCEAN RIDGE BASALT; N-MORB: NORMAL MIDDLE OCEAN RIDGE BASALT; WPB: WITHIN-
PLATE BASALT; VAB: VOLCANIC ARC BASALT; IAT: ISLAND ARC THOLEIITE; BABB: BACK ARC BASIN BASALT;
FAB: FRONT ARC BASALT. ............................................................................................................................ 143
FIGURE 14. SPIDER PLOTS OF THE UNCONTAMINATED PROTOLITHS OF THE MAFIC GRANULITES COMPARED TO
OCEAN ISLAND BASALTS (OIB), CONTINENTAL ARC BASALTS (CAB), VOLCANIC RIFTED MARGIN BASALTS
(VRMB), NORMAL MIDDLE OCEAN RIDGE BASALTS (N-MORB) AND ENRICHED MIDDLE OCEAN RIDGE
BASALTS (E-MORB). OIB AND MORB VALUES ARE FROM SUN AND MCDONOUGH (1989), PRIMITIVE
MANTLE VALUES FROM THE PYROLITE MODEL OF MCDONOUGH AND SUN (1995), CAB VALUES ARE FROM
TAMURA ET AL. (2014), VRMB VALUES COME OF HOOPER ET AL. (1999)................................................... 143
ÍNDICE DE TABELAS

Anexo 01: Artigo “Unraveling the Mesoarchean metamorphic evolution of the Carajás
Province, Amazon Craton: constrains from petrological evidences, mineral chemistry and U-
Pb geochronology”

TABLE 1. COUNTING TIME DURING THE ANALYSES OF THE MINERALS USING THE ELECTRON MICROPROBE. .......... 41
TABLE 2. MINERAL CHEMISTRY DATA OF THE FELDSPAR CRYSTALS FROM XICRIM-CATETÉ ORTHOGRANULITE AND
XINGU COMPLEX ROCKS. ............................................................................................................................. 63
TABLE 3. MINERAL CHEMISTRY DATA OF THE AMPHIBOLES FROM THE XICRIM-CATETÉ ORTHOGRANULITE AND
XINGU COMPLEX ROCKS............................................................................................................................... 66
TABLE 4. MINERAL CHEMISTRY DATA OF THE PYROXENE CRYSTALS FROM THE XICRIM-CATETÉ ORTHOGRANULITE
AND X INGU COMPLEX ROCKS. ...................................................................................................................... 69

TABLE 5. SUMMARY OF THE GEOTHERMOBAROMETRIC DATA OF THE XICRIM-CATETÉ ORTHOGRANULITE AND


XINGU COMPLEX. ......................................................................................................................................... 76

SUPPLEMENTARY TABLE 1. SUMMARY OF U-PB LA-ICPMS ZIRCON DATA FROM THE XICRIM-CATETÉ
ORTHOGRANULITE AND XINGU COMPLEX. ................................................................................................. 103
SUPPLEMENTARY TABLE 2. SUMMARY OF U-PB SHRIMP IIE ZIRCON DATA FROM THE XICRIM-CATETÉ
ORTHOGRANULITE AND XINGU COMPLEX. ................................................................................................. 106

Anexo 02: Artigo “Mesoarchean arc-related system between 3.0–2.9 Ga in the Carajás
Domain, Carajás Province, PA”

TABLE 1. THE MESOARCHEAN (CA. 3.0-2.8GA) GRANITOIDS OF CARAJÁS DOMAIN. ........................................... 117
TABLE 2. LITHOGEOCHEMICAL DATA OF THE ORTHOGNEISSES FROM XICRIM-CATETÉ ORTOGRANULITE AND
XINGU COMPLEX. THE ACRONYMS HP, MP AND LP TTG CORRESPOND TO THE HIGH, MEDIUM AND LOW
PRESSURE TTG OF MOYEN (2011). THE NORMALIZATION WAS CARRIED OUT USING THE SUN AND
MCDOUNOUGH VALUES OF THE C1 CONDRITE1. THE EU/EU* RATIO WAS CALCULATED BY THE FORMULAE:
𝐸𝑢𝑁 ÷ (𝑆𝑚𝑁 ∗ 𝐺𝑑𝑁)2 .............................................................................................................................. 129
TABLE 3. CORRELATION MATRICES FOR THE LEAST CONTAMINATED SAMPLES OF THE MAFIC GRANULITES. THE
CORRELATION IS CARRIED OUT USING THE PEARSON CORRELATION COEFFICIENT: COS𝛼 = 𝑖 = 1𝑁(𝑥𝑖 − 𝑥 ) ×
(𝑦𝑖 − 𝑦)𝑖 = 1𝑁𝑥𝑖 − 𝑥2 × 𝑖 = 1𝑁𝑦𝑖 − 𝑦2 ................................................................................................ 137
TABLE 4. LITHOGEOCHEMICAL DATA OF THE XICRIM-CATETÉ ORTHOGRANULITE MAFIC GRANULITES AND XINGU
COMPLEX AMPHIBOLITES. UGP: UNCONTAMINATED GRANULITE PROTOLITHS; CGP: CONTAMINATED
GRANULITE PROTOLITHS; CAP: CONTAMINATED AMPHIBOLITE PROTOLITHS. .......................................... 140
SUMÁRIO

1. INTRODUÇÃO ...................................................................................................................................... 19
2. OBJETIVOS ........................................................................................................................................... 22
3. MÉTODOS ............................................................................................................................................. 23
3.1. Trabalho de Campo e Petrografia ................................................................................................. 23
3.2. Microscopia de Varredura Eletronica ........................................................................................... 23
3.3. Quimica Mineral e Geotermobarometria ...................................................................................... 23
3.4. Geocronologia ............................................................................................................................... 24
3.5. Litogeoquímica .............................................................................................................................. 25
4. CONTEXTO GEOLÓGICO REGIONAL .............................................................................................. 27
5. ESTRUTURAÇÃO DA TESE ................................................................................................................ 31
ANEXO 01 ARTIGO: “UNRAVELING THE MESOARCHEAN METAMOPHIC EVOLUTION OF THE CARAJÁS PROVINCE,
AMAZON CRATION: CONSTRAINS FROM PETROLOGICAL EVIDENCES, MINERAL CHEMISTRY AND U – PB
GEOCHRONOLOGY”............................................................................................................................................. 32
ABSTRACT ......................................................................................................................................................... 33
1. INTRODUCTION ........................................................................................................................................ 34
2. GEOLOGICAL SETTING ............................................................................................................................. 35
3. ANALYTICAL PROCEDURES...................................................................................................................... 40
3.1. Field and Petrography................................................................................................................... 40
3.2. Mineral Chemistry and Geothermobarometry ............................................................................... 40
3.3. Geochronology .............................................................................................................................. 41
4. GEOLOGY OF THE HIGH-GRADE ROCKS OF THE GOIABA CREEK AREA ..................................................... 43
4.1. Xicrim-Cateté Orthogranulite ....................................................................................................... 46
4.2. Xingu Complex .............................................................................................................................. 54
5. MINERAL CHEMISTRY AND GEOTHERMOBAROMETRY ............................................................................. 59
5.1. Xicrim-Cateté Orthogranulite ....................................................................................................... 59
5.2. Xingu Complex .............................................................................................................................. 61
5.3. Geothermobarometry..................................................................................................................... 72
6. U-PB GEOCHRONOLOGY .......................................................................................................................... 76
6.1. Xicrim-Cateté Orthogranulite ....................................................................................................... 76
6.2. Xingu Complex .............................................................................................................................. 83
7. DISCUSSION ............................................................................................................................................. 84
7.1. Periods of crust formation and reworking during the Mesoarchean in the Carajás Domain ....... 84
7.2. New constrains on metamorphic evolution of the Carajás Domain. ............................................. 87
8. CONCLUSION ........................................................................................................................................... 92
REFERENCE ....................................................................................................................................................... 92
APPEDIX .......................................................................................................................................................... 103
ANEXO 02 ARTIGO: “MESOARCHEAN ARC-RELATED SYSTEM BETWEEN 3.0 – 2.9 GA IN THE CARAJÁS DOMAIN,
CARAJÁS PROVINCE, PA”.................................................................................................................................. 112
ABSTRACT ....................................................................................................................................................... 113
KEYWORDS: TTG, MESOARCHEAN, GEOCHEMISTRY, CARAJÁS PROVINCE .................................................... 113
1. INTRODUCTION ...................................................................................................................................... 114
2. GEOLOGICAL SETTING ........................................................................................................................... 115
3. METHODS .............................................................................................................................................. 119
3.1. Field and Petrography................................................................................................................. 119
3.2. Lithogeochemistry ....................................................................................................................... 119
4. GEOLOGY OF THE GOIABA CREEK AREA ................................................................................................ 119
4.1. Xicrim-Cateté Orthogranulite ..................................................................................................... 122
4.2. Xingu Complex ............................................................................................................................ 125
5. GEOCHEMISTRY ..................................................................................................................................... 127
5.1. The orthogneisses of the Xicrim-Cateté Orthogranulite and Xingu Complex ............................. 127
5.2. The mafic rocks of the Xicrim-Cateté Orthogranulite and Xingu Complex ................................. 135
6. DISCUSSION ........................................................................................................................................... 144
6.1. The fingerprints of the Mesoarchean arc magmatism ................................................................. 144
7. CONCLUSIONS ........................................................................................................................................ 148
REFERENCES.................................................................................................................................................... 149
6. CONSIDERAÇÕES FINAIS ................................................................................................................ 158
REFERÊNCIAS BIBLIOGRÁFICAS ............................................................................................................ 159
19

1. INTRODUÇÃO

Os terrenos arqueanos são compostos majoritariamente de diversos tipos de


granitoides (Batuk Joshi et al., 2017; Laurent et al., 2014). Estes estão classificados nos
seguintes grupos: (i) tonalitos, tromdhjemitos e granodioritos (TTG), que são
volumetricamente dominantes; (ii) (monzo) dioritos e granodioritos metaluminosos ricos em
Mg, Fe e K, conhecidos como sanukitoides sensu lato; (iii) biotita granitos e granitos a duas
micas ricos em K; e (iv) granitos hibridos ricos em K, que combinam características dos três
primeiros grupos (Laurent et al., 2014). Em determinados terrenos arqueanos, conhecidos
como terrenos de alto grau, estas rochas ocorrem e representam exposições da crosta inferior
primitiva da Terra (Perchuk and Gerya, 2011). A formação, evolução e exumação destes
terrenos de alto grau metamórfico influencia diretamente o debate sobre o estilo de tectônica
na Terra primitiva (Brown, 2009; Moyen, 2011; Perchuk and Gerya, 2011; Sajeev et al.,
2009; van Hunen and van den Berg, 2008).
Os dados geoquímicos disponíveis sugerem que o magmatismo TTG, dominante
volumetricamente, se deveu à fusão de rochas máficas em alta pressão, com onfacita e
granada residual, durante a subducção da placa oceânica (Moyen, 2011; Moyen and Martin,
2012; Moyen and Stevens, 2006). Os sanukitoides sensu lato, representam a evidência do
metassomatismo do manto durante os eventos magmáticos prévios associados à geração dos
TTG devido a dualidade contrastante de características geoquímicas, como o alto conteúdo de
Mg e Cr vs. alto conteúdo de LILE e K (Martin et al., 2009; Moyen and Martin, 2012). Os
granitos a duas micas e os biotita granitos, por sua vez, seriam relacionados à fusão de
litotipos crustais (Laurent et al., 2014; Moyen, 2011). Adicionalmente, o registro de eclogitos
e granulitos de alta pressão no Arqueano e de cinturões pareados de alta temperatura e de alta
pressão também favorece o modelo subducção-colisão (Brown, 2006; O’Brien and Rötzler,
2003; Sajeev et al., 2013; Sizova et al., 2010). No entanto, a ausência de sedimentos
relacionados a sistemas orogênicos do Precambriano e o formato diapírico (harpolith shape)
de alguns terrenos de alto grau são usados como evidência para argumentar a favor de
modelos de redistribuição gravitacional e diapirismo (Perchuk and Gerya, 2011).
Os modelos numericos petrológicos-termomecânicos sugerem que o processo de
subducção está diretamente relacionado à temperatura do manto superior, devido a sua
influência na estabilidade da crosta (Sizova et al., 2010). Estes modelos indicam que a
transição para o regime tectônico moderno ocorreu quando as temperaturas do manto
atingiram valores entre 175-160 K acima da temperatura atual (Sizova et al., 2010). De
20

acordo com os cálculos da variação secular da temperatura mantélica, isto teria ocorrido em
torno de ca. 3 Ga (Abbott et al., 1994; Labrosse and Jaupart, 2007). A ocorrência de cinturões
pareados com granulitos de ultra pressão-alta temperatura, eclogitos e granulitos de alta
pressão entre ca. 3,2–2,5 Ga também reforça essa premissa (Brown, 2009, 2006; Sajeev et al.,
2013; Sizova et al., 2010).
Os dados geocronológicos e isotópicos permitem a interpretação da evolução de
terrenos arqueanos de alto grau em que os cinturões pareados de alta temperatura/alta pressão
ainda não foram encontrados ou foram desconfigurados pela evolução geológica (Amaldev et
al., 2016; Guitreau et al., 2017; LaFlamme et al., 2014; Laurent and Zeh, 2015; Oriolo et al.,
2016). Na Zona de Cisalhamento Mercara, no oeste da India, os dados petrológicos e de U–Pb
e Lu–Hf, ajudaram na individualição do período de crescimento crustal (entre ca. 3,2–3,1 Ga)
e retrabalhamento (3,0 Ga) durante a evolução de um sistema de subducção-colisão que levou
à amalgamação do Craton Dharwar Oeste e do Terreno Granulítico do Sul da India (Amaldev
et al., 2016). Os granulitos máficos da Zona de Cisalhamento Mercara registraram condições
de pico metamórfico em 700–900 ºC e 10–12 kbar (Amaldev et al., 2016). Adicionalmente,
os autores concluíram, usando dados isotrópicos Lu–Hf e U–Pb, que a Zona de Cisalhamento
Mercara representa a borda retrabalhada do Craton Dharwar Oeste. No Rae Craton, Canadá,
os gnaisses arqueanos de assinatura TTG do bloco Repulse Bay, revelaram, a partir da
assinatura de Lu–Hf e elementos terras raras, processos de reciclagem e adição de material do
manto metassomatizado na formação da crosta inferior e média deste terreno no Neoarqueano
(~2.6Ga; LaFlamme et al., 2014).
No Domínio Carajás, segmento norte da Província Carajás (Cráton do Amazonas),
ocorrem rochas de fácies granulito, migmatitos e ortognaisses mesoarqueanos (ca. 3,00–2,97
Ga; Araújo and Maia, 1991; Avelar et al., 1999; Pidgeon et al., 2000; Santos, 2003; Vasquez
and Rosa-Costa, 2008). Estas rochas, agrupadas nas unidades Ortogranulito Xicrim-Cateté e
Complexo Xingu, estão inseridas em um contexto de uma zona de cisalhamento
mesoarqueana (ca. 2,8 Ga) com quilômetros de extensão, denominada de Cinturão Itacaiúnas
(Pinheiro and Holdsworth, 2000). O metamorfismo de alto grau e a migmatização foram
reconhecidos nesse domínio em 2,86 Ga (Machado et al., 1991; Pidgeon et al., 2000). Nesse
contexto, a construção de um cenário integrado para a evolução do Domínio Carajás apresenta
as seguintes lacunas: (i) distribuição das ocorrências das rochas do Ortogranulito Xicrim-
Cateté e do Complexo Xingu, que representam as rochas metamórficas do embasamento; (ii)
idades, características geoquímicas dos protólitos destas rochas metamórficas; (iii) ambiente
tectônico de formação desses protólitos; (iii) relação entre o metamorfismo destas rochas e
21

colocação de granitoides (ex. Feio et al., 2013; Gabriel and Oliveira, 2013; Leite-Santos,
2016; Moreto et al., 2011; Silva et al., 2014); (iv) história metamórfica da assembleia que
compõe o embasamento.
A área do Córrego da Goiaba, uma sessão norte–sul ao longo do embasamento do
Domínio Carajás, foi escolhida para esse estudo. A individualização e caracterização das
rochas metamórficas do embasamento foi feita por meio de mapeamento geológico de semi-
detalhe, petrografia combinada com química mineral e geotermobarometria, geocronologia de
alta resolução e litoquímica. Além da contribuição importante para a evolução da Província
Carajás, este trabalho tem desdobramentos interessantes para a discussão sobre a evolução
crustal e tectônica global durante o Arqueano (Gerya et al., 2015; Hamilton, 2011; Kerrich
and Polat, 2006; Laurent and Zeh, 2015; Martin et al., 2009; Moyen, 2011).
22

2. OBJETIVOS

Essa Tese de Doutorado tem como objetivo caracterizar a evolução petrogenética,


metamórfica e tectônica do Complexo Xingu e do Ortogranulito Xicrim-Cateté, que
representam parte do núcleo arqueano mais extenso do Cráton do Amazonas, a Província
Carajás, localizada na porção sudeste do estado do Pará. Esse estudo considera as seguintes
metas:
a) Caracterizar a evolução magmática dos protólitos dos litotipos metamórficos
que compõem o Complexo Xingu e Ortogranulito Xicrim-Cateté a partir de
estudos litoquímicos dos elementos maiores, menores e traço e das assinaturas
isotópicas de Lu–Hf;
b) Definir a evolução metamórfica do Complexo Xingu e do Ortogranulito
Xicrim-Cateté de forma relativa por meio da caracterização das paragêneses
minerais e sua relação com a estruturação regional e a partir da quantificação
de parâmetros P–T a partir de cálculos geotermobarométricos com base na
composição química das paragêneses minerais;
c) Determinar as idades dos eventos magmáticos e metamórficos no Complexo
Xingu e no Ortogranulito Xicrim-Cateté;
d) Propor um modelo evolutivo para o embasamento mesoarqueano do Domínio
Carajás;
23

3. MÉTODOS
3.1. Trabalho de Campo e Petrografia

Trabalhos de campo, visando o levantamento de secções geológico-estruturais e coleta


de amostras, foram realizados em duas etapas de 15 dias e uma etapa de 5 dias. O mapa
geológico resultou dos levantamentos em campo e da compilação dos trabalhos de Araújo &
Maia (1991), Vasquez & Rosa-Costa (2008), Leite-Santos (2016), Marangoanha (2016) e
Costa et al. (2016). Os mapeamentos foram focados no detalhamento das ocorrências de
rochas de alto grau e sua relação com as estruturas regionais. Nas etapas de campo 200
afloramentos foram visitados e 129 amostras foram coletadas.
Estudos petrográficos em luz transmitida e refletida foram realizados a partir de 74
secções delgadas e delgadas-polidas no Laboratório de Microtermometria do Instituto de
Geociências da Universidade de Campinas.

3.2. Microscopia de Varredura Eletronica

O microscópio de varredura eletrônica (MEV) foi utilizado na caracterização de


minerais e suas estruturas e texturas internas e sua morfologia e na confecção de mapas
composicionais de texturas relacionadas a paragêneses minerais identificadas nas seções
delgadas e delgadas-polidas. O microscópio de varredura eletrônica (MEV) está acoplado a
um Espectroscópio de Energia Dispersiva (Energy Dispersive Spectroscopy; EDS) e pertence
ao Instituto de Geociências da Universidade Estadual de Campinas (UNICAMP).

3.3. Quimica Mineral e Geotermobarometria

A caracterização química dos minerais, especialmente de pares e associações de


diferentes domínios microestruturais e distintos estágios de reequilíbrio, foi realizada com uso
de Microssonda Eletrônica. Foram determinados alguns zonamentos químicos de
porfiroblastos, com suporte de mapas composicionais detalhados, por meio da caracterização
da distribuição dos elementos maiores. As análises foram realizadas usando microssonda
eletrônica JEOL JXA-FE-8530, com canhão eletrônico suportado por Field Emission (FE),
que permite alta resolução espacial, provida com cinco espectrômetros WDS e um
espectrômetro EDS, no Laboratório de Microssonda Eletrônica da USP.
As condições analíticas são: (i) voltagem de aceleração de 15 kV; (ii) corrente de 20
nA; (iii) feixe de elétrons de 5 µm de diâmetro. Na análise dos cristais de feldspato o feixe de
elétrons de 10 µm foi utilizado. Os padrões de calibração utilizados na microssonda eletrônica
foram: albita (Si; feldspato), diopsidio (Si, piroxênios e anfibólios), anortita (Al), fayalita (Fe;
24

Mn), wollastonita (Ca), microclinio (K; feldspato), orthoclásio (K; piroxênio e anfibólio),
estroncianita (Sr), beinitonita (Ba), rutilo (Ti), albita (Na), diopsidio (Mg), cromita (Cr),
sodalita (Cl), fluorapatita (F).
Os cálculos geotermobarométricos foram efetuados a partir da composição química
dos pares ou associações de minerais presentes nos litotipos estudados, utilizando-se de
métodos e/ou calibrações dos principais geotermômetros e geobarômetros.
Nas rochas metabásicas foram aplicados os seguintes geotermômetros e
geobarômetros: (i) ortopiroxênio-clinopiroxênio; (ii) clinopiroxênio-hornblenda; (iii)
hornblenda-plagioclásio; (iv) hornblenda-plagioclásio-clinopiroxênio-quartzo. Em relação aos
gnaisses de composição tonalítica-granodiorítica o par hornblenda-plagioclásio foi usado
como geotermômetro e geobarômetro (Blundy and Holland, 1990; Molina et al., 2015). Estas
estimativas foram feitas com uso de planilhas de cálculos e de softwares, tais como
THERMOCALC (Holland and Powell, 1998) e TWQEEU (Berman, 1991), que incluem o uso
de bancos de dados termodinâmicos internamente consistentes.

3.4. Geocronologia

Nove amostras foram selecionadas para análises geocronológicas após a caracterização


petrográfica dos litotipos. Duas amostras representam biotita e hornblenda ortognaisses do
Complexo Xingu (XS04P e XS105P), e as demais rochas são do Ortogranulito Xicrim-Cateté.
Destas, cinco amostras são de orthopyroxene-diopside gneisses (CMS22P, SM46N, XA12A,
XA15A1, TS23A), uma amostra refere-se a um granulito máfico (XA15B3) e outra representa
um leucossoma associado a um granulito máfico (XS112A). Os concentrados de zircão foram
obtidos a partir do uso de métodos gravimétricos e magnéticos convencionais. As frações do
mineral foram coletadas a mão com o auxílio de lupa binocular. Todos os procedimentos
foram realizados no Laboratório de Geologia Isotópica da Universidade de Campinas.
Dois métodos foram usados nas análises dos grãos de zircão: (i) U–Pb LA–ICP–MS e
(ii) U–Pb SHRIMP IIe. As análises pelo método U–Pb LA–ICP–MS foram feitas no
Laboratório de Geologia Isotópica da Universidade de Campinas. O equipamento de ablação a
laser usa o sistema Phanton Machine Excite 193, e a célula de ablação HelEx, acoplada a um
ICP–MS Thermo Scientific Element XR, com as seguintes caracteristicas: (i) diâmetro do feixe
do laser de 25µm, (ii) taxa de repetição de 10Hz e (iii) fluência de 4.74J/cm2. O equipamento
capta oito isótopos de cada amostra (202,204,206,207,206Pb, 232
Th, 235,238U). Todos os isótopos são
232 238
medidos pela contagem de ions, exceto os isotopos Th e U, medidos pelo sistema
anterior combinado a um sistema análogo. O procedimento analítico no equipamento LA-
25

ICP-MS usa os padrões de zircão 91500 (Wiedenbeck et al., 1995) e peixe (Geherls, não
publicado). A redução dos dados foi feita com uso do software VizualAge (Petrus & Kamber,
2012) do pacote IOLITE (Paton et al., 2010). O método envolve a subtração do branco
seguida da correção e comparação com o padrão do zircão 91500. Outros critérios envolvem a
avaliação da qualidade do dado: (i) conteúdo anômalo de Pb comum, relacionado à razão
206
Pb/204Pb menor que 3000 e alta porcentagem do 206
Pb no total de Pb (f206>1), (ii) alto
conteúdo de U, valores maiores que 500µg/g.
As análises pelo médoto SHRIMP IIe foram feitas no Laboratório de Geocronologia
de Alta Resolução do Centro de Pesquisas Geocronológicas da Universidade de São Paulo. Os
grãos de zircão foram enviados ao LGAR, onde foram montados junto com o zircão padrão
TEMORA. A montagem foi polida até expor seções quasi-centrais dos cristais de zircão.
Após a metalização dom ouro, as montagens polidas foram examinadas no microscopio de
varredura eletronica FEI-QUANTA 250, equipado com detectores de eletrons secundários e
de catodo-luminescencia. As condições de operação na análise de catodo-luminescencia
foram: (i) corrente de emissão de 60µA; (ii) voltagem de aceleração de 15 kV; (iii) feixe de
abertura de 7µm; (iv) tempo de aquisição de 200 µs and (v) e resolução de 1024x884 pixels.
A montagem com as amostras foi analisada pelo método U–Pb no equipamento SHRIMP Iie,
seguindo os procedimentos de Willians (1998). A correção da concentração de Pb comum foi
feita pelo 204Pb medido. O erro da razão 206Pb/238U foi inferior a 2% e a abundância de urânio
e das razões U/Pb foram calibrados pelo TEMORA. A análise estatística e o cálculo das
idades foram feitos no software Isoplot 4.1 (Ludwig, 2003).

3.5. Litogeoquímica

As amostras foram coletadas em ocorrências de rochas das unidades Ortogranulito


Xicrim-Cateté e Complexo Xingu respeitando uma distribuição geográfica que permitisse
uma informação representativa da evolução do magmatismo nestas unidades. Como as rochas
estão variavelmente migmatizadas, a amostragem foi cuidadosamente feita com foco na
separação das partes mais preservadas da anatexia. As 27 amostras foram moídas (< 75 µm)
para análise química usando britador de mandíbulas (Pulverisette 2, Fritsch, Germany) e um
moinho de ágata planetário e vibratório (Pulverisette 5 and 7, Fritsch, Germany) no
Laboratorio de Geoquímica da Universidade de Campinas. Os elementos maiores e menores
foram analisados por um espectrometro de emissão com indução de plasma acoplado (ICP–
ES), no qual as linhas atômicas e espectrais dos elementos foram excitadas por um feixe de
átomos argônio e detectados por fotomultiplicadores, comparados a linhas de calibração e
26

convertidos em concentração (Rollinson, 1993). Os elementos traço foram analisados por


espectrometria de massa com indução de plasma acoplada (ICP–MS). O material foi
dissociado no plasma de argônio e levado a um tubo curvo por um ima eletromagnético que
divide os átomos de acordo com suas massas (Rollinson, 1993). Estas análises foram feitas no
Acme Labs e no ALS Global Analytics, Canadá. Os dados foram tratados no software
Microsoft Excell 2010.
27

4. CONTEXTO GEOLÓGICO REGIONAL

A Província Carajás (Santos et al., 2000) é subdividida nos domínios Rio Maria, a sul,
e Carajás, a norte (Figura 1; Vasquez & Rosa-Costa, 2008). O Domínio Rio Maria representa
um terreno granito-greenstone Mesoarqueano (2,98 a 2,86 Ga; Almeida et al., 2011)
composto por sequências metavulcano-sedimentares, granitoides da série TTG, sanukitoides e
leucogranitos potássicos (Almeida et al., 2016, 2011). O Domínio Carajás apresenta uma
associação de rochas mesoarqueanas do embasamento (3,00–2,83Ga) recoberto por rochas
supracrustais do Supergrupo Itacaiúnas e da Formação Águas Claras e recortado por suites
félsicas, máficas e máfico-ultramáficas do Neoarqueano (2,76–2,57 Ga; Araújo & Maia,
1991; DOCEGEO, 1988; Feio et al., 2013, 2012; Gibbs et al., 1986; Machado et al., 1991;
Moreto, 2013; Pidgeon et al., 2000; Wirth et al., 1986). O conjunto arqueano é recortado por
granitoides de tipo A do Paleoproterozoico (ca. 1.88 Ga; Dall’Agnol et al., 2005; Santos et
al., 2013). O embasamento mesoarqueano é subdividido em: (i) ortognaisses, ortogranulitos,
migamtitios e greenstone belts do Mesoarqueano Médio (3,00–2,97 Ga; Araújo & Maia,
1991; Avelar et al., 1999; DOCEGEO, 1988; Pidgeon et al., 2000) and; (ii) granitoides TTG,
cálcio-alcalinos sódicos, cálcio-alcalinos potássicos, de alto Ba–Sr, de alto Mg do
Mesoarqueano Médio e Inferior (3,00–2,83 Ga; (Feio et al., 2013; Gabriel & Oliveira, 2013;
Leite-Santos, 2016; Moreto, 2013; Rodrigues et al., 2014; Silva et al., 2014, 2018). O limite
entre estes domínios foi definido por uma feição geofísica que segue, aproximadamente, a
direção E–W acompanhando a área de exposição das rochas do greenstone belt do Grupo
Sapucaia, ao sul da cidade de Água Azul do Norte (Figura 1; Dall'Agnol et al., 2000; Santos
et al., 2000; Sousa et al., 2015; Vasquez & Rosa-Costa, 2008).
28

Figura 1. Divisão das Províncias Geocronológicas do Cráton Amazônico no estado do Pará (Vasquez & Rosa-Costa, 2008).

No Arqueano, o embasamento foi afetado por duas grandes fases de deformação: (i) a
primeira é relacionada a uma foliação penetrativa de baixo ângulo com direção E–W que
culminou no desenvolvimento de grandes zonas de cisalhamento no Mesoarqueano (ca. 2,8
Ga; Cinturão de Cisalhamento Itacaiúnas) e; (ii) a segunda relacionada ao desenvolvimento de
uma foliação espaçada de alto ângulo com direção E–W a WNW–ESE relacionada à
reativação das zonas de cisalhamento regionais no Neoarqueano (ca. 2,76–2,74 Ga; Araújo &
Maia, 1991; Pinheiro and Holdsworth, 2000). Os ortognaisses, granulitos e migmatitos do
embasamento foram inicialmente agrupados nos complexos Xingu (Silva et al., 1974) e Pium
(Araújo & Maia, 1991). O primeiro compreendia granulitos, gnaisses, migmatitos,
granitoides, greenstone belts e complexos máfico-ultramáficos (Silva et al. 1974; DOCEGEO,
1988). O Complexo Pium, como definido originalmente, reunia granulitos máficos,
charnockitos, charno-enderbitos e enderbitos (Araújo & Maia, 1991).
Atualmente, o Complexo Xingu (em rosa na Fig. 2) compreende ortognaisses e
migmatitos, com encraves de anfibolitos, que representam o embasamento ou as rochas
encaixantes dos greenstone belt e dos granitoides neoarquenos do Domínio Carajás (Vasquez
& Rosa-Costa, 2008). Entre esse gnaisses, os de composição tonalítica predominam com
subordinadas ocorrências de rochas granodioríticas e trondhjemíticas (Araújo & Maia, 1991;
Vasquez & Rosa-Costa, 2008). Estas rochas apresentam uma trama fortemente assimétrica e
29

anastomosada definida por um regime tectônico imbricado. De acordo com Araújo & Maia
(1991), o Complexo Xingu apresenta migmatização intensa, mas pequena geração de
mobilizados graníticos. Avelar et al. (1999) obtiveram a idade de cristalização de 2.974 ±15
Ma (Pb–Pb em zircão) em um ortognaisse granodiorítico do Complexo Xingu na região de
São Félix do Xingu. No vilarejo de Curionópolis (Figura 2), Machado et al. (1991) obtiveram
a idade de 2.859 ±2 Ma (U–Pb em zircão) em leucossoma, o que foi interpretado como gerado
no último evento de migmatização. Recentemente, grande esforço vêm sendo empregado
visando a distinção entre granitoides e ortognaisses antes atribuídos ao Complexo Xingu.
O Complexo Pium foi definido nas margens dos rios Pium e Cateté como uma unidade
que compreendia um conjunto de rochas granulíticas de composição máfica e félsica, sem
granada em sua assembleia mineral (Araújo & Maia, 1991; Figura 2). De acordo com os
autores, as rochas máficas predominariam entre os rios Pium e Parauapebas e as rochas
félsicas seriam abundantes perto do Rio Cateté. Ricci & Carvalho (2006) observaram que as
rochas máficas localizadas nas proximidades do Rio Pium eram noritos com diopsídio,
correspondendo, portanto a rochas essencialmente ígneas. Segundo esses autores, rochas
granulíticas poderiam ser encontradas exclusivamente às margens do Rio Cateté, nas
vizinhanças da vila indígena de Xicrim (Figura 2). Nesse contexto, a ocorrência de granulitos
na Província Carajás se restringiu aos gnaisses charnockíticos a enderbíticos com encraves de
granulitos máficos da região de Xicrim-Cateté (Araújo & Maia, 1991; Ricci & Carvalho,
2006). Baseado nestas novas evidências, Vasquez & Rosa-Costa (2008) propuseram o
abandono do termo Complexo Pium, agrupando os granulitos da província na unidade
Ortogranulito Xicrim-Cateté e os noritos na unidade Diopsídio Norito Pium. No entanto,
Delinardo da Silva et al. (2015) encontraram litotipos metamórficos do Ortogranulito Xicrim-
Cateté a norte das rochas máficas do Diopsidio Norito Pium. Os autores sugeriram que tais
rochas têm assinatura TTG e apresentam uma série de texturas e estruturas relacionadas a
fusão parcial assistida por fluidos.
Previamente, Pidgeon et al. (2000) obtiveram idades de núcleo e borda de cristais de
zircão extraídos de um “enderbito” amostrado perto do Rio Paraúapebas. Foram obtidas
idades de 3.002 ±14 Ma para o núcleo e 2.859 ±9 Ma para a borda dos cristais de zircão. Tais
idades foram interpretadas como idade de cristalização e de metamorfismo em fácies
granulito, respectivamente (Pidgeon et al., 2000). A amostra datada por esses autores foi
coletada na porção norte da área de estudo, logo a sul das rochas mapeadas como Granito
Bom Jesus (BJ na Figura 2) por Feio et al. (2013). Vasquez & Rosa-Costa (2008)
consideraram que o gnaisse enderbítico seria parte do Ortogranulito Xicrim-Cateté, sugerindo
30

que a rocha poderia ser um xenólito assimilado pelo Diopsídio Norito Pium, recentemente
interpretado como uma intrusão máfica neoarqueana (2,74 Ga; Pb–Pb em zircão; Santos et al.,
2013).

Figura 2. Mapa geológico do Domínio Carajás (modificado de Costa et al., 2016).


31

5. ESTRUTURAÇÃO DA TESE

Essa tese reunirá dois artigos cinentíficos. O primeiro detalha as ocorrências de rochas
de alto grau metamórfico em uma faixa de direção norte–sul na região sul do Domínio
Carajás, entre o limite sul das unidades supracrustais do Supergrupo Itacaiúnas e o limite
norte do Domínio Rio Maria. O trabalho envolveu a caracterização de campo e petrográfica
destas rochas junto à obtenção de dados de química mineral, para fins de cálculos
geotermobarométricos, e dados geocronológicos U–Pb, visando a determinação de idades
absolutas dos eventos magmáticos e metamórficos impressos nessas rochas no Mesoaqueano.
O segundo artigo apresenta a caracterização litogeoquímica das protólitos dos gnaisses
tonalíticos a granodioríticos de alto grau metamórfico do Ortogranulito Xicrim-Cateté e do
Complexo Xingu e dos enclaves máficos encontrados nestes gnaisses. Neste segundo
trabalho, o foco é a compreensão das características do magmatismo mesoarqueano e suas
implicações para a configuração do Domínio Carajás e da província como um todo, neste
periodo. Este artigo deverá ser complementado com os dados de Lu-Hf que serão obtidos no
Laboratório de Geocronologia da UFOP.
32

Anexo 01
Artigo:
Unraveling the Mesoarchean metamorphic evolution of the Carajás Province, Amazon
Craton: constrains from petrological evidences, mineral chemistry and U-Pb
geochronology

Delinardo da Silva et al.


33

Abstract

The high-grade rocks of the Carajás Domain (ca. 3.06–2.93 Ga) represent a singular
evidence of the ultra-high temperature granulite metamorphism in the Mesoarchean (ca. 2.89–
2.85 Ga). These high-grade rocks are represented by the Xicrim-Cateté Orthogranulite (2.99–
2.94Ga) and the Xingu Complex (2.97–2.93 Ga). The Xicrim-Cateté Orthogranulite is
composed of metatexitic orthopyroxene-diopside gneisses of tonalitic to granodioritic
composition, metatexitic diopside-pargasite-plagioclase granulites, metatexitic mafic
granulites and metanorites. The zircon crystals of ortopyroxene-diopside gneisses show a
complex geochronological evolution related to (i) the assimilation of 3,066 ±6.6Ma (MSWD
= 0.072) xenocrystic zircon; (ii) envolved in oscillatory zoning overgrowths with 2,987
±18Ma (MSWD = 7.0), 2,979 ±31Ma (MSWD = 4.4), 2,955 ±15Ma (MSWD = 0.32), 2,954
±71Ma (MSWD = 22), 2,953 ±18Ma (MSWD = 4.4) and 2,935 ±8Ma (MSWD = 0.27) ages;
(iii) which are recrystallized at 2,853 ±19Ma (MSWD = 0.64). The Xingu Complex comprises
metatexitic pargasite-biotite and biotite gneisses of tonalitic to granodioritic composition and
amphibolites. The zircon crystals of the Xingu Complex orthogneisses showed the
characteristic oscillatory zoning with 2,936 ±6Ma (MSWD = 2.5) and 2,928 ±15Ga (MSWD
= 14) ages, and thinner recrystallized rims. High-temperature conditions (907–1128 ºC at 8.1–
13.7 kbar) were registered in the mafic rocks of the Xicrim-Cateté Orthogranulite. The
metamorphic zircon in the mafic granulite yielded a concordia age of 2,890 ± 7Ma (MSWD =
0.60). The Xingu Complex rocks have reached the upper amphibolite facies conditions at 785
ºC and 8.8 kbar. The data indicate that the dehydration melting reactions (i.e. Amp + Qtz =
Opx + Cpx + Melt) related to the ultra-high temperature metamorphism coupled with the
water-influx melting (Pl + Qtz + Kfs = Melt) allowing the widespread anatexis of the lower
and middle crust. The medium- to high-grade metamorphism, accompanied by migmatization
(ca. 2.89 –2.85 Ga) and granite emplacement (ca. 2.87–2.83 Ga), was related to the collision
between the Carajás and Rio Maria Domain in the Carajás Province.

Keywords: UHT metamorphism, U-Pb geochronology, Carajás Province


34

1. Introduction

The Archean high-grade terrains are the oldest examples of exposed lower crustal
sections related to former Earth’s crust (Perchuk and Gerya, 2011). They comprise
amphibolite to granulite-facies rocks, ultra-high temperature and high-pressure granulites,
eclogites and migmatites (Brown, 2006; Morfin et al., 2013; O’Brien and Rötzler, 2003;
Perchuk and Gerya, 2011; Sajeev et al., 2013). Models for their formation, evolution and
exhumation influence directly the debate on the tectonic style of the early Earth (Brown,
2006; Perchuk and Gerya, 2011). The Archean eclogite and high-pressure granulite represent
faborable evidence for the subduction-collisional models (O’Brien and Rötzler, 2003; Sajeev
et al., 2013). However, these rocks are not widespread in the Archean high-grade terrains
(Brown, 2006; Perchuk and Gerya, 2011). In addiction, the lack of sediments related to
nearby Precambrian orogenic systems and the harpolith (intrusive-like) geometry of some
granulite facies terrains were used as evidence for diapiric or gravitational redistribution
numerical models (Perchuk and Gerya, 2011). Petrological-thermomechanical numerical
models indicate that the subduction process is intrinsincally related to the upper mantle
temperature due its influence in the lithosphere stability (Sizova et al., 2010). These models
suggest that the transition to modern tectonic regime would have occurred at mantle
temperatures of about 160-175 K from the present day temperatures, which roughly
correspond with the temperature at ca. 3.0 Ga (Abbott et al., 1994; Labrosse and Jaupart,
2007; Sizova et al., 2010). In addition, the occurrence of paired metamorphic belts with ultra-
high temperature (UHT) granulites, eclogites and high-pressure (HP) granulites in the
Mesoarchean–Neaorchean (ca. 3.2-2.5 Ga) time also reinforces this assumption (Brown,
2008, 2006; Sizova et al., 2010).
Mesoarchean (3.00–2.97 Ga) granulite- and amphibolite-facies rocks and migmatites
were reported in the Carajás Domain, the northern portion of Carajás Province (Amazon
Craton; Araújo and Maia, 1991; Avelar et al., 1999; Pidgeon et al., 2000; Santos, 2003;
Vasquez and Rosa-Costa, 2008). These rocks are inserted in the context of a kilometre-scale
Mesoarchean (ca. 2.8 Ga) E–W Itacaiúnas shear zone (Holdsworth and Pinheiro, 2000). The
high-grade metamorphism and migmatization were locally dated in ca. 2.86 Ga (Machado et
al., 1991; Pidgeon et al., 2000). To build an integrated model for the geodynamic evolution
for the Mesoarchean in the Carajás Domain, the following pieces of this puzzle must be
founded: (i) the distribution of the occurrences of the metamorphic basement; (ii) the ages of
the protoliths and the characteristics of their sources; (iii) the metamorphic history of these
35

basement rocks; and (iv) the relationship between the metamorphic rocks and the
Mesoarchean pre- to syn-orogenic magmatism (i.e. Feio et al., 2013, 2012; Gabriel and
Oliveira, 2013; Leite-Santos and Oliveira, 2014; Moreto et al., 2011). The whole picture of
the evolution of the Carajás Domain and its connection with the evolution of the Rio Maria
Domain (the southern portion of the Carajás Province), will give an integrated view of the
Carajás Province evolution in the Archean. However, it demands the characterization of the
metamorphic evolution of the basement assemblage.
For this task, a north to south geological section in the southern portion of the Carajás
Domain, following the Goiaba Creek banks (black polygon in Figure 1), was chosen. In this
area, occurrences of orthoderived high-grade metamorphic rocks were previously reported
(Araújo and Maia, 1991; Delinardo da Silva, 2014; Vasquez and Rosa-Costa, 2008). The
well-preserved relationships between the mineral phases that indicate prograde and retrograde
metamorphism indentified in the petrography were mapped in the scanning electron
microscope (SEM). After that, the mineral phases were analysed in the microprobe.
Additionaly, zircon grains were collected from these high-grade rocks allowing the
characterization of the crystallization, metamorphic and partial melting events.

2. Geological Setting

The Carajás Province (Santos et al., 2000) is subdivided into the Rio Maria, at south,
and Carajás domains, at north (Vasquez and Rosa-Costa, 2008). The Rio Maria Domain
represents an exclusively Mesoarchean (2.98 to 2.86 Ga; Almeida et al., 2011) granite-
greenstone terrain composed of metavolcano-sedimentary sequences, TTG granitoids of low,
medium and high La/Yb ratios, sanukitoids and potassic leucogranites (Almeida et al., 2016,
2011). The basement of Carajás Domain is composed of middle Mesoarchean (3.0 Ga to 2.93
Ga) migmatitic orthogneiss, granulites and greenstone belt sequences crosscut by late
Mesoarchean granitoids (2.87 to 2.83 Ga; DOCEGEO, 1988; Feio et al., 2013; Gabriel and
Oliveira, 2013; Machado et al., 1991; Moreto, 2013; Moreto et al., 2011; Pidgeon et al., 2000;
Santos, 2016). The basement is covered by metavolcano-sedimentary units of the Itacaiúnas
Supergroup and Águas Claras Formation and crosscut by Neoarchean felsic and mafic suites
and Paleoproterozoic granites (Dall’Agnol et al., 2005; DOCEGEO, 1988; Feio et al., 2012;
Gibbs et al., 1986; Machado et al., 1991; Vasquez and Rosa-Costa, 2008; Wirth et al., 1986).
The limit between these domains is defined by an E–W geophysical lineament that follows
the Sapucaia greenstone belt sequences, at south of Água Azul do Norte city (Dall’Agnol et
al. 2000; Santos et al., 2000; Sousa et al., 2015; Vasquez and Rosa-Costa, 2008).
36

During the Archean, the basement was affected by two main deformation phases: (i)
the Mesoarchean (ca. 2.8 Ga) ductile system of linked, moderately and steepy dipping
sinistral strike-slip and thrust-dominated shear zones (Itacaiúnas Shear Belt) and (ii) the
Neoarchean (ca. 2.7 Ga) upper crustal, brittle-ductile to brittle deformation related to the
Carajás, Cinzento and Canaã dos Carajás strike-slip systems (Araújo et al., 1988; Holdsworth
and Pinheiro, 2000; Pinheiro et al., 2013). The Mesoarchean system introduces the E–W trend
ductile fabric seeing in the basement rocks (Araújo and Maia, 1991). The metamorphic rocks
of the basement rocks were initially grouped in the Xingu (Silva et al., 1974) and Pium
(Araújo and Maia, 1991) complexes. The first comprehended granulites, gneisses, migmatites
and granitoids, in addition to greenstone belts and mafic-ultramafic complexes (DOCEGEO,
1988; Silva et al., 1974). The Pium Complex included mafic granulites, charnockites, charno-
enderbites and enderbites (Araújo and Maia, 1991).
Vasquez and Rosa-Costa (2008) redefined the Xingu Complex (Pink unit in Figure 1)
restricting it to the Mesoarchean orthogneisses e migmatites with sparse amphibolite enclaves.
This unit represents the basement or the host rocks of the Mesoarchean and Neoarchean
greenstone belt and the Neoarchean granitoids of the Carajás Domain. Among those gneisses,
predominate rocks with tonalitic compositions with subordinated granodioritic and
trondhjemitic composition (Araújo and Maia, 1991; Vasquez and Rosa-Costa, 2008).
According to Araújo and Maia (1991), the Xingu Complex rocks underwent intense
migmatization with little generation of mobilized granitic melt. Avelar et al. (1999) obtained a
crystallization age of 2.974 ± 15 Ma (Pb–Pb in zircon) from a granodioritic orthogneiss of the
Xingu Complex in the São Felix do Xingu city (Figure 1). In the Curionópolis Village (Fig.
1), Machado et al. (1991) obtained an age of 2.859 ± 2 Ma (U–Pb in zircon) for an
undeformed leucosome, which were interpreted representative of the last migmatization event
recorded in the Carajás Province. Since then efforts were carried out on distinction of several
granitoids from the Xingu Complex orthogneiss (i.e Feio et al., 2013; Gabriel and Oliveira,
2013; Moreto et al., 2015a; Moreto et al., 2015b; Santos, 2016).
The Mesoarchean granitoids were subdivided in the following groups: (i) sodic calc-
alkaline granitoids, including Bacaba Tonalite, Canaã dos Carajás Granite, São Carlos
Tonalite and Campina Verde Tonalític Complex (3.00 to 2.85 Ga; Feio et al., 2013; Moreto et
al., 2011; Silva et al., 2014); (ii) TTG granitoids, including the Rio Verde Trondhjemite and
the Colorado Tonalite (ca. 2.93 to 2.87 Ga; Feio et al., 2013; Silva et al., 2014); (iii) high Mg
granitoids, including the Água Limpa and Água Azul granodiorites (ca. 2.87 Ga; Gabriel et
al., 2015); (iv) high Ba–Sr granitoids, represented by the Nova Canadá Leucogranite (2.87
37

Ga; Leite-Santos, 2016); and (v) the potassic calc-alkaline granitoids, including the Boa Sorte
Granite, Cruzadão Granite, Bom Jesus Granite and Serra Dourada Granite (ca. 2.89 to 2.83
Ga; Feio et al., 2013; Moreto et al., 2011; Rodrigues et al., 2014). The potassic calc-alkaline
granitoids, often related to crustal anatexis, represent the dominant granitoid variety between
the cited groups. Additionally, another TTG granitoids (> 2.8 Ga) and syn-tectonic (2.87–2.86
Ga) granitoids were recently individualized in the southwestern boundary of the Carajás
Domain (Água Azul do Norte region; Santos, 2016; Silva et al., 2018).
The Pium Complex was defined at the margins of the Pium and Cateté rivers (Figure
1) by Araújo and Maia (1991) as a set of orthoderived garnet-free granulitic rocks of mafic to
felsic composition. According to Araújo and Maia (1991), the peak granulite-facies
paragenesis in these rocks is represented by plagioclase + diopside + orthopyroxene. In
addition, the mafic rocks predominate in the area between the Pium and Parauapebas rivers,
whereas the felsic granulites occur in the Cateté River area (Araújo and Maia, 1991).
However, Ricci and Carvalho (2006) pointed out that the mafic rocks found around the Pium
River were diopside-bearing norites with typical igneous textures. Granulites were considered
to be restricted to the Cateté River area, in the vicinity of the Xicrim Native Village (Figure 1;
Ricci and Carvalho, 2006).
In this context, the occurrence of granulites in the Carajás Province remained
restricted to charnockitic to enderbitic gneisses with mafic granulite enclaves (Araújo and
Maia, 1991; Ricci and Carvalho, 2006). Based on the new evidences, Vasquez and Rosa-
Costa (2008) proposed the abandonment of the Pium Complex, grouping the charnockitic to
enderbitic gneisses in the Xicrim–Cateté Orthogranulite. The igneous mafic rocks were
named as Diopside Norite Pium, which was recently interpreted as a Neoarchean mafic
intrusion (2.74 Ga; Pb–Pb in zircon; Santos et al., 2013a).
Previously, Pidgeon et al. (2000) obtained ages from zircon cores (3,002 ±14 Ma) and
rims (2859 ±9 Ma) from an enderbitic gneiss sampled near the Paraúapebas River. These ages
were interpreted as related to crystallization and granulite facies metamorphism, respectively
(Pidgeon et al. 2000). The sample was collected in the northern portion of the study area,
immediately south of the rocks mapped as Bom Jesus Granite (Figure 1).
Ricci and Carvalho (2006) consider that the enderbitic gneiss dated by Pidgeon et al.
(2000) is part of the Xicrim-Cateté Orthogranulite, suggesting that it could represent a
xenolith within the Diopside Norite Pium. However, Delinardo da Silva et al. (2015)
characterized metatexitic orthopyroxene-diopside gneisses of tonalitic to granodioritic
composition in the area mapped as Bom Jesus Granite by Feio et al. (2013; Figure 1). The
38

protoliths of these rocks showed a TTG affinity with medium HREE content (Delinardo da
Silva et al., 2015) and high-strain migmatitic structures (i.e., stromatic, net, schollen,
schlieren; Sawyer, 2008). These rocks show different melt rates that were associated with
water influx and retrograde metamorphism (Delinardo da Silva et al. 2015). In addition, those
migmatitic rocks recorded a Mesoarchean metamorphic history, which has not yet been
systematically studied in the Carajás Province.
39

Figure 1. A) The distribution of the Carajás Province in the Pará estate, in the north of Brazil, at right, and its subdivision into the Carajás (CD) and Rio Maria Domain (RM), at left. The black
rectangle shows the map area. B) Geological map of the northern portion of the Carajás Province, modified from Costa et al. (2016).
40

3. Analytical Procedures
3.1. Field and Petrography

Three field work stages, totaling 35 days, were performed, with survey of geological-
structural sections and sampling. The geological map of the area integrates the new field work
data and the compiled data of Araújo and Maia (1991), Costa et al. (2016), Marangoanha
(2016), Leite-Santos (2016) and Vasquez and Rosa-Costa (2016). Detailed petrography under
transmitted and reflected light was carried out on 74 thin and thin-polished sections.
Additional mineralogical and textural characterization was made using a Scanning Electron
Microscope (SEM) coupled with and Energy Dispersive Spectroscopy (EDS) at the
University of Campinas (UNICAMP).

3.2. Mineral Chemistry and Geothermobarometry

The mineral chemistry analysis was carried out in the JEOL JXA-8530F of the
Electron Microprobe Laboratory of the University of São Paulo. The polish thin-sections of
seven samples were selected for mineral chemistry based on the previous petrographic
characterization. Two samples represent the Xingu Complex, one of the pargasite–biotite
tonalite gneiss (XS105P) and one of the amphibolite xenoliths (XS33). Five samples represent
the Xicrim–Cateté Orthogranulite, two samples of the metanorite (XS111B and SM42), two
samples of the mafic granulite (XA15B and SM41R) and one sample of the metatexitic
orthopyroxene-diopside granodiorite gneiss (SM11P). In the samples of the Xingu Complex,
feldspar, amphibole and biotite were analysed. In samples of the Xicrim-Cateté
Orthogranulite, the mineral chemistry of orthopyroxene, clinopyroxene, amphibole and biotite
was obtained.
The analytical conditions consisted of voltage acceleration of 15 kV, current of 20 nA
and electron beam of 5 µm of diameter, exepct for feldspar analysis, for which a beam with
10 µm diameter was used. The calibration standards used in the microprobe were: albite (Si;
feldspar), diopside (Si; pyroxene and amphibole), anorthite (Al), fayalite (Fe; Mn),
wollastonite (Ca), microcline (K; feldspar), orthoclase (K; pyroxene and amphibole),
strontianite (Sr), beinitoite (Ba), rutile (Ti), albite (Na), diopside (Mg), cromite (Cr), sodalite
(Cl), fluorapatite (F). The Table 1 shows the counting time for all elements during each
specific routine.
41

Table 1. Counting time during the analyses of the minerals using the electron microprobe.
Counting Time (s)
Element
Feldspar Amphibole Pyroxene Mica

Si 10 10 10 10
Al 20 15 15 15
Fe 10 10 10 10
Mn 40 40 40 20
Zn - - - 30
Cl - 10 - 10
Ca 10 20 10 10
K 10 10 20 10
Sr 40 - - -
Ti 10 10 10 10
Ba 30 - - 10
F - 10 - 10
Cr - 20 10 30
Na 10 10 10 10
Mg 10 10 10 10

The geothermobarometry was carried out using the chemical compositions of the
mineral pairs and mineral assemblages related to the prograde and retrograde metamorphism.
In the metabasites the following geothermometers and geobarometers were applied: (i)
orthopyroxene–clinopyroxene; (ii) clinopyroxene–hornblende; (iii) hornblende–plagioclase;
and (iv) hornblende–plagioclase–clinopyroxene–quartz. In the orthogneisses, the pair
hornblende–plagioclase were used as geothermometer and geobarometer (Blundy and
Holland, 1990; Molina et al., 2015). The estimatives were made in the Excell spreadsheets
and using the softwares Thermocalc (Holland and Powell, 1998) and TWQEEU (Berman,
1991). The Thermocalc calculations of average pressure and temperature use the internally-
consistent dataset of Holland and Powell (1998; 1990; 1985) and Powell and Holland (1993;
1985). The thermodynamic databases used in the geothermobarometric calculations with
TWQEEU include the solution models of Berman (1988) and Fuhrman and Lindsley (1988).

3.3. Geochronology

Nine samples were selected for geochronology based on petrographic characterization.


Two samples represent biotite- and hornblende-bearing orthogneisses (samples XS04P and
XS105P), five samples were from orthopyroxene-diopside orthogneisses (samples CMS22P,
SM46N, XA12A; XA15A1; TS23A), one sample is from a mafic granulite (XA15B3) and the
last one belongs to a leucosome of a metatexitic mafic granulite (XS112A). The zircon
42

concentrates were extracted from the samples using conventional gravimetric and magnetic
techniques. The mineral fractions were handpicked under a binocular microscope. All of these
procedures were carried out in the Isotope Geology Laboratory of the University of
Campinas.
Two methods were used in the analysis of the zircon grains: (1) U–Pb LA–ICP–MS in
zircon and (2) U–Pb SHRIMP IIe in zircon. The U–Pb LA–ICP–MS analysis were carried out
in the Isotope Geology Laboratory of the University of Campinas. The laser ablation
equipment uses the Photon Machine Excite 193 system, an ablation cell HelEx, coupled to an
ICP-MS Thermo Scientific Element XR, with the following characteristics: laser beam
diameter of 25µm, repetition rate of 10Hz and fluency of 4.74J/cm2. It acquires eight isotopes
for each sample (202,204,206,207,208Pb, 232Th, 235,238U). All the isotopes are measured
by the ion count, except the 232Th and 238U, measured by the previous system combined to
an analogue method. The analytical procedure in the LA-ICP-MS equipment uses the standard
zircon 91500 (Wiedenbeck et al., 1995) and peixe (Geherls, not published). The data
reduction was made in the software VizualAge (Petrus and Kamber, 2012) from the IOLITE
package (Paton et al., 2010). The method involves the white gas subtraction followed by the
downhole fractioning correction, compared to behave of the standard zircon 91500. Other
criteria was used to evaluate the quality of the data, like: (i) anomalous common Pb quantity,
related to 206Pb/204Pb ratios lower than 3000 and to high percentage of 206Pb in the total Pb
(f206>1), (ii) high U contents, data by element value higher than 500 µg/g.
The SHRIMP IIe analysis were carried out in the High Resolution Laboratory of
Geochronology (LGAR) of the Center of Geochronological Research from the University of
São Paulo. The handpicked fractions of zircon were sent to the LGAR where they were
mounted together with the standard zircon TEMORA. The mount was polish until quasi-
central sections of the zircon crystals were exhibited. After the gold metallization, the
polished mounts were exanimated in the Scanning Electron Microscope FEI-QUANTA 250
equipped with secondary electrons and cathode luminescence detectors. The operational
conditions in the cathode luminescence analysis were: emission current of 60 µA; acceleration
voltage of 15 kV, aperture beam of 7 µm; acquisition time of 200 µs and resolution of 1024 x
884 pixels. The mounts were then analyzed by the U–Pb method in the SHRIMP IIe
equipment, following the procedures of Willians (1998). The correction in the content of the
common Pb was made by the measured 204Pb content. The error of the 206Pb/238U was
inferior to 2% and the uranium abundances and U/Pb ratios were calibrated by the TEMORA
43

standard. The statistical analysis and age calculations were made in the softwere Isoplot 4.1
(Ludwig, 2003).

4. Geology of the high-grade rocks of the Goiaba Creek Area

The Goiaba Creek area comprises a south to north section following the banks of the
namesake river. At its floodplains the Meso to Neoarchean units of the Carajás Domain are
exposed in variable size outcrops. Additionaly, the evidence of both Meso and Neoarchean
deformation are recoginized in the area. The ductile fabric of the Itacaiúnas Shear Belt (ca.
2.8 Ga; Holdsworth and Pinheiro, 2000) is represented by moderate dipping E–W foliations in
the Xicrim–Cateté Orthogranulite and in the Xingu Complex rocks (Figure 2). In addition, it
defines the curved and braided thrust faults displayed in the Goiaba Creek map (Figure 2).
The tectonic contact among the Rio Maria and Carajás Domain and the Xingu Complex and
Xicrim–Cateté Orthogranulite are also marked by these shear zones.
The relationship among the Mesoarchean metamorphic rocks, granitoids and
greenstone belt sequences is controlled by this ductile fabric. Moreover, the E–W geophysical
lineament between the Rio Maria and Carajás Domain (Dall’Agnol et al., 2000) is a northern
vergence thrust shear zone where lenses of basement rocks, greenstone belts and granitoids
are imbricated together. Among the greenstone belt rocks, serpentinites and high-grade
peridotites occur in imbricated lenses (Sousa et al., 2015). Additionaly, the volumous and
kilometer-scale milky quartz veins are always associated with the thrust faults. The high-Mg
granitoids of the Água Limpa Granodiorite (ca. 2.87 Ga; Gabriel et al., 2015; Gabriel and
Oliveira, 2014) and the high Ba–Sr sodic leucogranites of the Nova Canadá Leucogranite (ca.
2.87 Ga; Leite-Santos, 2016) are intrusive in the Xingu Complex. They occur as granite-
sheets in the foliation of the Xingu Complex. Outcrop-scale enclaves of the Xingu Complex
are observed inside the Nova Canadá Leucogranite (Figure 2).
The Neoarchean deformation is inserted in the context of the sinistral transpression of
the development of the Canaã Shear Zone (Pinheiro et al., 2013). It is locally represented by
rotation and displacement of the previous foliations and migmatitic structures and regionally
associated with the reactivation of the braided and anastomosed major shear zones (Figure 2).
The hydrated charnockite-series plutons of the Planato Granite Suite (ca. 2.74 Ga; Feio
et al., 2012) and the tholeiitic intraplate gabbroic rocks of the Pium Diopside Norite (ca. 2.74
Ga; Santos et al., 2013a) occur in the northern portion of the Goiaba Creek Area. The latter
occur as autolith fragments inside the Planato Granite Suite. They are intrusive in the Xingu
44

Complex and Xicrim-Cateté Orthogranulite. Additionaly, the deformation in those plutons is


concentrated in their margins. The occurrence of Paleoproterozoic A-type magmatism is
restricted to the porphyritic rocks of the Rio Branco Granite (ca. 1.88 Ga; Santos et al.,
2013b) at the northeastern portion of the Goiaba Creek Area (Figure 2).
The characteristics of the Xicrim–Cateté Orthogranulite and the Xingu Complex are
detailed in the next sections.
45

Figure 2. Geological map of the Goiaba Creek area. The map results from the field and petrographic studies integrated with
the compiled data from the previous works of Dall’Agnol et al. (2005; 1), Feio et al. (2012; 2), Santos et al. (2013a; 3),
Sousa et al. (2015; 4), Leite-Santos (2016; 5) Feito et al. (2013; 6) Gabriel and Oliveira (2013; 7); and Almeida et al. (2011;
8), Araújo and Maia (1991), Marangoanha (2016), Costa et al. (2016).
46

4.1. Xicrim-Cateté Orthogranulite


4.1.1. Metatexitic Mafic Granulite

The metatexitic mafic granulites occur as boudins inside the metatexitic


orthopyroxene–diopside gneiss and, locally, as larger boulders (Figure 3A). The boudins are
disrupted and oriented in the E–W gneissic foliation of the metatexitic orthopyroxene–
diopside gneiss (Figure 3). Locally, an abrupt contact between the metatexitic orthopyroxene-
diopside gneiss and the metatexitic mafic granulite can be observed (Figure 3). The rocks are
composed of plagioclase (An35-68; 40%) diopside (8–20%), enstatite (10–45%); Ca-
amphibole (5–25%); biotite (2–3%); magnetite (1–10%), ilmenite (~5%), rutile (~1%), quartz
(1-2%). They are melanocratic, medium- to coarse-grained, dark grey in colour, foliated and,
locally, have massive granulitic texture.
The metatexitic mafic granulites show extremely variable textural fabric and mineral
assemblage. It is observed in different thin-sections of rocks of the same outcrop, or in the
same thin-section. Two main lithotypes compose the metatexitic mafic granulites of the
Xicrim-Cateté Orthogranulite: the (i) metanorite and the (ii) mafic granulite and diopside-
pargasite-plagioclase granulite. The migmatization affected all the lithotypes.
The metanorite is recognized distally to the E–W shear zones and occurs mainly as
larger boulders. They are composed of plagioclase (An53-68; ~40%), enstatite (45%),
diopside (5%), Cl-rich pargasite (~3%), ilmenite (~4%), biotite (>1%), rutile (>1%),
magnetite (>1%), and quartz (~1%). These rocks show a widespread coarse-grained, blasto-
subophitic texture (Figure 3D). However, the enstatite show undulatory extinction and the
plagioclase has deformation and wedge-shaped twins (Figure 3D). Moreover, the subgrain
rotation recrystallization and the grain boundary migration also affect the plagioclase and
enstatite crystals. It allows the local development of a granoblastic texture (Figure 3E). The
textural transformation is associated with (i) diopside formation and enstatite and plagioclase
recrystallization (Figure 3H) and (ii) enstatite and Ca-rich plagioclase replacement by an
assemblage with: Na-rich plagioclase + Cl-rich pargasite + quartz + magnetite (Figure 3E).
Nevertheless, at sites of preserved blasto-subophic texture, the plagioclase crystals show Ca-
rich core and Na-rich rim (Figure 3F). It commonly occurs where the pargasite crystals grows
in between the enstatite and plagioclase (Figure 3F). Rutile occurs as exolutions in the
enstatite crystals (Figure 3G). Magnetite replaces the rutile exolutions inside the enstatite
(Figure 3G). Magnetite also occurs as exsolutions within the ilmenite crystals.
47

The mafic granulite and diopside-pargasite-plagioclase granulite occur as boudins,


within the E–W shear zones, or as larger boulders, outside the E–W shear zones. The mafic
granulite is composed of plagioclase (An36-41; ~40%), diopside (~20%), enstatite (~10%),
pargasite (up to 15%), biotite (up to 10%), magnetite (~3%), quartz (~2%) and locally zircon
(<1%). The diopside–pargasite–plagioclase granulite is composed of plagioclase (An35-45;
~40%), diopside (25%), pargasite (25%), biotite (5%), magnetite (~3%), quartz (~2%), and
locally zircon (<1%). These rocks show a granoblastic texture. However, grain boundary
migration, bulging and subgrain rotation recrystallization may obliterate the polygonal
interface among plagioclase, enstatite, diopside, pargasite and biotite (Figure 3I). The
undulatory exctintion is common among enstatite, diopside and plagioclase. The latter also
show deformation and wedge-shaped twins. The peak paragenesis is represented by the
paragenesis plagioclase + diopside + F-rich pargasite in the diopside-pargasite-plagioclase
granulite and by plagioclase + diopside + enstatite + F-rich pargasite in the mafic granulite
(Figure 3I). Such paragenesis are associated with medium-grained rocks with granoblastic
interlobate texture. In these samples, pargasite forms large crystals with magnetite exsolutions
along the cleavage planes. The subidioblastic pargasite is surrounded by idioblastic to
subidioblastic crystals of diopside, enstatite and labradorite (Figure 3I).
48

Figure 3. Field and petrographic features os the metatexitic mafic granulite and diopside-pargasite-plagioclase granulite. A)
Meter scale boulders of the metatexitic mafic granulites. The migmatization is parallel to their foliation. B) Boudins of the
metatexitic mafic granulite inside the metatexitic orthopyroxene-diopside gneiss. The black dots and the milk quartz veins in
the gneissic foliation highlight the boudinage process; C) A closer look in the contact between the mafic and felsic granulites
suggests that the first may be intrusive. D) Blasto-subophitic texture in the metanorites. The plagioclase crystals show the
deformational and wedge-sheped twins and locally subgrain rotation recrystallization. E) The texture transformation during
the replacement of enstatite (En) and Ca-rich plagioclase by Na-rich plagioclase (Pl) + pargasite (Prg) + quartz (Qtz) +
magnetite (Mgt) which is related to the melt textures showed below; F) Plagioclase crystals with Ca-rich (red) nuclei and Na-
rich (blue) boundary in the SEM compositional maps. The Na-rich boundary occurs together with the hornblende (Hbl)
crystals in between plagioclase and enstatite. The enstatite may show diopside (Di) crystals inside it. G) Rutile (Rt)
exolutions inside the enstatite. The inclusions are replaced by magnetite. H) The paragenesis of the metanorites represented
by plagioclase + diopside + enstatite. The granoblastic texture is defined by interlobate contacts due to grain boundary
migration, bulging and subgrain rotation recrystallization; I) The SEM image shows the paragenesis of the mafic granulite
(plagioclase + diopside + enstatite + pargasite). The updated list of Whitney and Evans (2010) for the abbreviations of the
rock-forming minerals was used.
49

The migmatization shows distinct styles among the metatexitic mafic granulites
(Figure 3A). In the diopside-pargasite granulite and the mafic granulite it is related to
stromatic structures that locally evolve to schollen and “jaguar skin” diatexites (Figure 4A–E).
The link between the metatexites and diatexites is the net-structure metatexites. These net-like
features allow the individualization of granulitic rocks into smaller fragments, easily involved
by the increasing volume of melt. The stromatic structures also segregate a plagioclase-poor
residuum from a plagioclase-rich leucosome (Figure 4B). The leucosome is leucocratic,
white, medium-grained and has orthopyroxene, plagioclase and diopside megacrysts (Figure
4D). They are composed of plagioclase (60%), quartz (15%), hornblende (4%), orthopyroxene
(10%) and diopside (5%), biotite (5%), magnetite (<1%) and zircon (<1%). The subidioblastic
to locally xenomorphic orthopyroxene phenocrysts show undulatory extinction (Figure 4D–
F). The plagioclase xenomorphic phenocrysts also show undulatory extinction, but their
remarkable feature is the corrosion-like textures showed in the Figure 4G.
The migmatization is not widespread in the metanorites. The rocks locally show
stromatic and net-like structures. The melt-related textures are amoeboids and textureless
films around the plagioclase crystals and plagioclase-bearing veins in between plagioclase and
orthopyroxene crystals (Figure 4H-I). The microtextures observed in the metanorites are
associated with the Cl-rich pargasite crystals (Figure 4H-I). The leucosomes are composed of
plagioclase (~65%), pargasite (20%) and quartz (10%), with subordinated biotite (3%) and
magnetite (~2%).
50

Figure 4. Field and petrographic features related to the migmatization in the metatexitic mafic granulites. A) Characteristic
aspect of the boulders of metatexitic mafic granulite with the stromatic and net structures. B) Detail in the leucosome-
residumm interface highlighting the hornblende-rich melanosome; C) Shollen diatexites highlighting the assimilation of the
mafic granulite by the leucosomes. D and E) The characteristics of the orthpyroxene phenocyrsts in the “jguar skin” diatexite
(D) and in the leucosomes of the metatexite (E); F and G) orthopyroxene (Opx) and plagioclase (Pl) megacrysts in the
leucosomes. The orthopyroxene crystals show undulatory extinction and the plagioclase crystals show corrosion-like
textures; H) plagioclase (Pl) crystals in the metatexitic mafic granulites showing amoeboid film of melt. The image highlights
the relation between hornblende (Hbl) crystals and melt pockets; I) plagioclase-bearing veins that occur in the metatexitic
mafic granulites, especially in the metanorites. The hornblende crystals generally occur in the boundaries of the veins; I). The
updated list of Whitney and Evans (2010) for the abbreviations of the rock-forming minerals was used.

4.1.2. Metatexitic Orthopyroxene-Diopside Gneiss

The Xicrim-Cateté Orthogranulite encompasses metatexitic orthopyroxene-diopside


gneisses of tonalitic and, subordinate, granodioritic composition. The rocks are composed of
quartz (20–25%), plagioclase (An25-31; 45–60%), K-feldspar (5–10%), biotite (15%),
enstatite (5-10%), diopside (2–3%), hornblende (3–5%), apatite (2–3%), zircon (~1%),
51

ilmenite and magnetite (1–2%), sericite (1%), pistacite and clinozoisite (~1%), chlorite
(~1%), scapolite (1-2%), titanite (~1%).
The metatexitic orthopyroxene-diopside gneiss is leucocratic, medium-grained,
generally light grey with granoblastic interlobate, lepidoblastic and massive granulitic
textures. The massive granulitic texture is observed in the metatexitic orthopyroxene-diopside
gneiss in the lower strain zones (Figure 5A). In these rocks, the peak paragenesis of the
metatexitic orthopyroxene-diopside gneiss is well preserved (Figure 5B). It is represented by
plagioclase + K-feldspar + quartz + enstatite + diopside ± ilmenite ± hornblende. The crystals
of plagioclase, K-feldspar, quartz, enstatite and diopside show undulatory extinction, grain
boundary migration, subgrain rotation and bulging recrystallization (Figure 5B-C).
Plagioclase also has deformational twins with cuspate terminations or wedge-shaped twins in
the corner of the crystals (Figure 5C). Together with enstatite crystals, plagioclase locally
preserves growth crystal faces (Figure 5D). Some examples of a well-developed granoblastic
texture can be observed in the metatexitic orthopyroxene-diopside gneisses (Figure 5E).
Inside the higher strain zones, the metatexitic orthopyroxene-diopside gneiss develop a
gneissic foliation (Sn) that segregates domains of granoblastic interlobate and lepidoblastic
textures. The continuous gneissic foliation (Sn) shows E–W trend dipping to SSW to SSE
plunge (Figure 5F) and a downdip to oblique stretching lineation. The gneissic foliation (Sn)
has variable dip. In the northern portion of the Xicrim-Cateté Orthogranulite, the low-angle
foliation (up to 39º) is common. Otherwise, in the southern portion of the body, at the tectonic
contact with the Xingu Complex, the foliation has a higher angle (up to 68º). However, the
paragenesis related to Sn development is the same: biotite + quartz + apatite + magnetite ±
titanite (Figure 5G). The granoblastic interlobate domains are quartz-feldspathic and the
lepidoblastic domains are rich in biotite (Figure 5H–I). The reddish brown biotite crystals
define lepidoblastic and, locally, decussate textures (Figure 5H–I). Some larger biotite
crystals show bronzite and diopside inclusions.
52

Figure 5. Field and petrographic features of the metatexitic orthopyroxene-diopside gneiss. A) The massive granulitic texture
of the metatexitic orthopyroxene-diopside gneiss. The red circle shows the enstatite crystal surrounded by biotite; B) the peak
paragenesis of the metatexitic orthopyroxene-diopside gneiss. In the image the undulatory extinction, bulging and subgrain
recrystallization and grain boundary migration can be observed in the plagioclase (Pl), quartz (Qtz), K-feldspar (Kfs),
diopside (Di) and enstatite (En) crystals; C) the image highlight the characteristic deformational twins of the plagioclase and
its wedge-shape aspect; D) plagioclase and enstatite crystals with preserved growth crystalline faces; E) the sites of well-
developed granoblastic texture; F) The EW south-dipping gnessic foliation developed along the shear zones; G) the
retrograde paragenesis with biotite (Bt) + quartz + magnetite (Mgt) in the metatexitic orthopyroxene-diopside gneiss with
massive texture; H) The lepidoblastic and decussate texture of biotite (Bt) in the metatexitic orthopyroxene-diopside gneiss
with gneissic foliation. I) Biotite-rich domains with hornblende crystals replacing enstatite. The blue dots, shows the
granoblastic texture locally developed. The updated list of Whitney and Evans (2010) for the abbreviations of the rock-
forming minerals was used.

The partial melting textures are spatially related to the retrograde paragenesis. The
metatexitic orthopyroxene-diopside gneiss shows net and stromatic structures (Figure 6A-B).
The volume of melt increases towards the shear zones. The leucosomes are hololeucocratic,
medium- to coarse-grained and porphyritic (Figure 6A–B). They are composed of plagioclase
53

(~45%); K-feldspar (~20%), quartz (~25%), biotite (~3%), magnetite (~2%), apatite (~3%),
zircon (~2%). The plagioclase megacrysts defines porphyritic texture (Figure 6E–F). K-
feldspar shows mymerkitic texture in their boundaries (Figure 6E–F). In the residumm
(metatexitic orthopyroxene-diopside gneiss), the melt-related textures occur in the
granoblastic interlobate domains. The textures are represented by cuspate-shaped K-feldspar
and quartz films and melt pockets around the corner of the crystals (Figure 6C–D). The
plagioclase, K-feldspar and quartz crystals of the leucosomes show undulatory extinction,
bulging and subgrain rotation and recrystallization (Figure 6E–F).

Figure 6. Field and petrographic features of the melt-related features in the metatexitic orthopyroxene-diopside gneiss. A)
The stromatic structure in the metatexitic orthopyroxene-diopside gneiss with massive granulite texture and in (B) the rocks
with gneissic foliation. The leucosome has a coarse-grained texture; C) the melt-related texture showed by K-feldspar crystal
crosscutting a plagioclase crystal; D) the amoieboid form of K-feldspar, plagioclase and quartz crystals in the boundary
between the metatexitic orthopyroxene-diopside gneiss and the leucosomes; E) The coarse-grained leucosomes with
plagioclase megacrysts, K-felspar with mymerkites and post-crystallization deformation, associtated with bulging and
subgrain rotation recrystallization and grain boundary migration; F) the plagioclase crystals show growth crystal faces in the
leucosomes. The updated list of Whitney and Evans (2010) for the abbreviations of the rock-forming minerals was used.

A spaced high angle foliation (Sn+1) with NW–SE direction occurs in the metatexitic
orthopyroxene-diopside gneisses. The paragenesis with quartz + sericite + chlorite + epidote +
clinozoesite ± scapolite could be related to it. It selectively replaces enstatite, biotite and
plagioclase crystals. Such feature is related to brittle to brittle-ductile structures in the
gneisses.
54

4.2. Xingu Complex


4.2.1. Metatexitic pargasite-biotite and biotite gneisses and amphibolite

The Xingu Complex is composed of metatexitic pargasite-biotite and biotite gneisses


(Figure 7A). Both rocks show amphibolite boudins of variable size.
The metatexitic pargasite-biotite gneiss (Figure 7B) occurs in the northern portion of
the Xingu Complex, at south of the tectonic contact with the Xicrim–Cateté Orthogranulite
(see Figure 2). The rocks are composed of albite (~50%), quartz (~20%), K-feldspar (~5%),
pargasite (~10%), biotite (~7%), apatite (Ap; ~2%), pistacite, clinozoisite, allanite (~2%),
titanite (~1%), magnetite (2%), and zircon (~1%). The metatexitic biotite gneiss (Figure 7C)
predominates in the sountern portion of the Xingu Complex (see Figure 2). The rock is
composed of albite (~50%), quartz (~20%), K-feldspar (~5%), biotite (~15%), apatite (~2%),
pistacite, clinozoesite, allanite (~3%), titanite (~1%), magnetite (~3%), and zircon (~1%).
Both, metatexitic pargasite-biotite and biotite gneisses are orthoderived and show tonalitic
composition. In addition, they show similar textural and structural evolution. They are
medium-grained, gray, leucocratic and show nematoblastic, lepidoblastic, decussate,
granoblastic and porphyroclastic textures. The metatexitic pargasite-biotite and biotite
gneisses show a continuous E–W gneissic foliation (Sn) with NNW and NNE low-angle
plunge (Figure 7D). The foliation is associated with a downdip NNW stretching lineation and
asymetrical tight drag and sheath folds with SSE vergence (Figure 7E–F).
The metamorphic segregation, recorded by granoblastic interlobate quartz- and
feldspar-rich layers and nematoblastic and lepidoblastic pargasite- and biotite-rich bands,
occurs in association with the development of the continuous gneissic foliation (Figure 7G).
Inside the granoblastic layers, the albite, quartz and K-feldspar crystals show undulatory
extinction and interlobate outlines due to grain boundary migration (Figure 7H). They also
show bulging and subgrain rotation recrystallization (Figure 7H). Albite shows deformational
and wedge-shaped twins and poikiloblastic texture. The dark green pargasite crystals, inside
the nematoblastic layers, also show poikiloblastic texture (Figure 7B). They are replaced by
brown biotite and magnetite crystals. Biotite may show undulatory extinction. Idioblastic to
subidioblastic pistacite, allanite, clinozoisite, titanite and apatite replace biotite and pargasite
crystals in the pargasite- and biotite-rich layers (Figure 7B-C). In the lepidoblastic layers, the
brownish pleochroism of biotite crystals turns green due to their partial replacement by
chlorite, defining a pseudomorphic texture (Figure 7I). Outwards the high strain zones, the
biotite metatexitic gneiss shows decussate texture. A paragenesis with pistacite + apatite +
55

quartz ± clinozoisite ± allanite ± titanite ± muscovite replaces the biotite and pargasite.
Locally, symplectites of quartz, pistacite and biotite may be observed (Figure 7I).

Figure 7. Field and petrography features of the metatexitic hornblende-biotite and biotite gneisses of the Xingu Complex. A)
Outcrop of the metatexitic biotite gneiss with the lesser melt-fraction observed in the Goiaba Creek area; B) metatexitic
pargasite-biotite gneiss with coarse pargasite crystals with plagioclase inclusions (poikiloblastic texture) being replaced by
biotite. The image also shows the apatite + pistacite + quartz replacing both, pargasite and biotite. C) Metatexitic biotite
gneiss with the granoblastic and decussate textures. The apatite + pistacite + quartz paragenesis is also present over biotite
and plagioclase. D) An outcrop of the metatexitic biotite gneiss inside the Nova Canadá Leucogranite. The gneiss shows a
high melt-fraction and the segregation of the neosome into quartz-feldsphatic leucosomes and biotite-rich melanosome.
Additionaly, the low angle foliation is well-defined and the melt-related structures are parallel to it. E and F) Sheath and drag
folds observed in the metatexitic horneblende-biotite and biotite gneisses. The leucosomes highlight the structures. In the
case of the sheath folds (E) the streaching lineation, parallel to the axial planes of the folds, is well-defined in the
leucosomes. G) The biotite-rich lepidoblastic and quartz-feldsphatic domains that defines the gneissic foliation of these
rocks; H) the red arrows show the main recrystallization features in the granoblastic interlobate domains: bulging and
subgrain rotation recrystallization and grain boundary migration. I) pistacite + quartz symplectites over the biotite,
plagioclase, K-feldspar and quartz crystals. The updated list of Whitney and Evans (2010) for the abbreviations of the rock-
forming minerals was used.
56

The boudins of amphibolite are composed of magnesio-hornblende (~60%), albite


(~30%), biotite (~4%), magnetite (<1%), clinozoisite (~2%), titanite (~2%) and quartz (~1%).
They occur as elongated bodies in the E–W gneissic foliation (Figure 8A), or as xenoliths in
the Nova Canadá Leucogranite (Figure 8B). The boudins of amphibolite are melanocratic,
dark gray, medium-grained and show granoblastic, nematoblastic and decussate texture. The
paragenesis albite + magnesio-hornblende + magnetite represent the peak conditions of the
amphibolite (Figure 8C). The minerals show a granoblastic texture with interlobate outlines
due to grain boundary migration, bulging and subgrain recrystallization (Figure 8C-D).
Hornblende has inclusions of quartz and albite (Figure 8C; 8E). The albite crystals show
deformation and wedge-shaped twins. Idioblastic biotite crystals crosscut hornblende and
albite crystals (Figure 8E). A latter paragenesis with idioblastic to subidioblastic titanite +
clinozoisite + quartz also occurs in the amphibolite, replacing the hornblende and biotite
crystals (Figure 8F).

Figure 8. Field and petrographic features of the amphibolites observed within the metatexitic pargasite-biotite and biotite
gneisses of the Xingu Complex and in the Nova Canadá Leucogranite. A and B) The occurrence mode of the amphibolite
represented by elongated boudins in the foliation of the hornblende-biotite and biotite gneisses (A). Foliated and folded
amphibolite xenoliths inside the Nova Canadá Leucogranite (B). C and D) Micro-estrutural aspects of the amphibolites.,
represented by interlobate granoblastic texture, with evident grain boudary migration (C). Intra-crystalline deformation of
albite, showing deformational twins in the corner of the crystals. E and F) Paragenetic evolution of the amphibolites
evidenced by replacement of the magnesio-hornblende by brown biotite crystals followed by the replacement of both by
titanite and epidote.
57

The metatexitic pargasite-biotite and biotite gneisses show evidences of partial melting
(see Figure 7E-F). The segragation into quartz-feldspar leucosomes and biotite-rich
melanosomes defines the stromatic structures observed in the gneisses (see Figure 7E). These
structures are developed in the low-angle gneissic foliation (Sn). Locally, the anatetic
structures are folded, showing the SSE vergence (see Figure 7E–F). The leucosomes crosscut
the boudins of amphibolite, but no macroscopical or microscopical melt-related features were
observed in them (see Figure 8A). The melt-related microstructures occur in the quartz- and
feldspar-rich granoblastic domains of the pargasite-biotite and biotite metatexitic gneisses.
They are represented by cuspate terminations of K-feldspars crystals, among the outlines of
plagioclase and quartz, and thinny veinlets of K-feldspar and plagioclase and quartz
crosscutting plagioclase megacrysts (Figure 9A–B). The stromatic and net leucosomes are
hololeucocratic, white and medium- to coarse-grained. They are composed of plagioclase
(~45%), quartz (~25%), K-feldspar (~25%), biotite (3%), pistacite, clinozoesite, allanite,
apatite, and zircon (2%). The leucosome has granodioritic composition. It shows
porphyroclastic and granoblastic amoeboid textures. The K-feldspar porphyroclasts show
mesoperthitic texture (Figure 9C). Rounded quartz crystals occur as inclusions in the K-
feldspar. The grain boundary migration, bulging and subgrain recrystallization affected the
plagioclase, quartz and K-feldspar crystals and define their interlobate and amoeboid outlines
of them (Figure 9D). The crystals also show undulatory extinction. The plagioclase shows
deformation and wedge-shaped twins. The K-feldspar megacrysts also show tartame twins in
their corners (Figure 9D). Biotite crystals sparsely occur in the matrix. Pistacite, clinozoesite,
allanite, apatite and zircon, occur together with the biotite crystals.
The Nova Canadá Leucogranite has enclaves of the Xingu metatexitic biotite gneiss
and amphibolite (Figure 8B; Figure 9E). It also has biotite-rich elongated stripes (Figure 9F).
The leucosomes in the stromatic structures of the metatexitic biotite gneiss enclaves are
connected to the leucogranite. These kinds of structures define local diatexitic pockets with
schollen and schlieren structure (Figure 9E–F). They are common in northern boundary of the
Nova Canadá Leucogranite, at the contact with the Xingu Complex. Acording to Leite-Santos
(2016), the Nova Canadá Leucogranite has monzogranitic and granodioritic composition. In
addition, the author also shows that the leucogranite has K-feldspar porphyroclasts with
perthitic and poikilitic textures. The sheets of the Água Limpa Granodiorite emplaced in the
E–W gneissic foliation of the Xingu Complex rocks are also crosscut by the leucosomes
58

associated with the Xingu metatexitic pargasite-biotite and biotite gneisses. Locally, the
leucosome assimilates the feldspar megacrysts of the Água Limpa Granodiorite (Figure 9G).
A high-angle spaced mylonitic foliation (Sn+1) transposes, and locally, rotates the
previous structures, but has the same E–W strike (Figure 9H). Quartz crystals show subgrain
rotation recrystallization features associated to Sn+1 (Figure 9I).

Figure 9. Field and petrographic characteristic of the melt-related features of the Xingu Complex rocks. A and B) The most
typical melt-related microstructure of the metatexitic hornblende-biotite and biotite tonalite gneisses. A) elongate films with
cuspate shape growing from the K-feldspar (Kfs) crystals surrounded by quartz (Qtz) and albite (Pl); B) granitic veinlets with
K-feldspar, quartz and plagioclase; C and D) microtextures of K-feldspar crystals in the leucosomes, which include the
mesoperthitic texture (C) and the development of tartame twining (D) in the corners of the crystals; E) relationship between
the leucosomes of the Xingu metatexitic biotite tonalite gneiss and the Nova Canadá Leucogranite whose connection could
be observed in the field; F) Biotite clusters and the biotite stripes that define the schilieren structure of the Nova Canadá
Leucogranite; G) the relationship between the Xingu Complex rocks and the Água Limpa Granodiorite. The intrusive contact
is marked by the contemporaneous leucosome formation. In these cases, the leucosome assimilate the K-feldspar pheocrysts
from the Água Limpa Granodiorite. H and I) field and petrographic characteristic of the spaced high angle mylonitic
foliation tha transposes the previous structures. It rotates the previous structures (H) and is responsible for the
recrystallization of quartz crystals (I).
59

5. Mineral Chemistry and Geothermobarometry


5.1. Xicrim-Cateté Orthogranulite

In the metatexitic mafic granulites, metanorites and orthopyroxene-diopside gneisses


of the Xicrim-Cateté Orthogranulite, the microprobe analyses were carried out in the feldspar,
pyroxene and amphibole crystals. The feldspar composition (n = 121) varies widely in the
rocks of Xicrim-Cateté Orthogranulite. The feldspar crystals in the mafic granulite and in
metanorite samples are Ca-Al-rich labradorite and andesine, respectively, with low Si, Na and
K content (7.95 to 13.95 wt% of CaO; 26.31 to 31.46 wt% Al2O3; 53.38 to 58.58 wt% of
SiO2; 3.69 to 7.11 wt% of Na2O; 0.05 to 0.13 wt% of K2O; Table 2; Figure 10F).
In the metatexitic orthopyroxene-diopside granodiorite gneiss, the feldspar is mostly
represented by Na-rich oligoclase crystals with low Al and K content (7.92–8.45 wt% of
Na2O; 23.67–24.59 wt% Al2O3; 0.09–0.13 K2O; Table 2). In addition, core to rim increase
of CaO and Al2O3 and decrease of SiO2 and Na2O can be observed in feldspar from the
metatexitic mafic granulites and orthopyroxene-diopside gneiss. In the metanorites, the CaO
and Al2O3 contents decrease and the Na2O and SiO2 content increase betweem the core and
the rim of the feldspar crystals (Table 2). The Fe2O3t content usually increases from core to
rim in all samples of the Xicrim-Cateté Orthogranulite (Table 2).
The amphibole (n = 47) compositions were analysed in metatexitic mafic granulites
and in metanorites. According to the classification of Hawthorne et al. (2012), amphibole
from these rocks plot in the pargasite field (Figure 10D). The amphibole in the mafic granulite
and diopside-pargasite-plagioclase granulite is associated with feldspar and pyroxene in the
metamorphic peak paragenesis. In the metanorites, it is associated with the retrograde
metamorphism and partial melting, replacing orthopyroxene and diopside. These two
generations of amphiboles show a huge contrast with respect to F and Cl contents (Figure
10E). The mafic granulite and diopside-pargasite-plagioclase granulite have F-rich
amphiboles (up to 0.63 wt% of F; Table 3), and the metanorites have Cl- and K-rich
amphiboles (up to 1.78 wt% of Cl; up to 2.06 wt% of K2O; Table 3). In addition, the
amphiboles in the mafic granulite and diopside-pargasite-plagioclase granulite have higher
MgO and Na2O contents (12.82 to 13.57 wt% of MgO; 1.49 to 1.74 wt% of Na2O; Table 3).
The amphiboles in the metanorites have higher FeOt content (14.59 to 15.36 wt% of FeOt;
Table 3).
60

The Cr2O3 and TiO2 contents vary widely in rocks that have both pyroxenes, with
respect the ones that have just one. The amphiboles in the diopside-pargasite-plagioclase
granulite (SM41R; Table 3) have the higher Cr2O3 (up to 0.27 wt% of Cr2O3; Table 3) and
the lowest TiO2 (up to 1.07 wt% of TiO2; Table 3) contents compared to the mafic granulite
(XA15B). In samples of metanorites with only orthopyroxene (SM42), the highest TiO2 and
Cr2O3 contents are observed (Table 3). The core to rim compositional variation among the
amphibole crystals in the mafic granulite and diopside-pargasite-plagioclase granulite also
depends on the presence of one or both pyroxenes in the paragenesis. The MgO and K2O
contents increase and the Na2O content decreases in the amphiboles of the diopside-pargasite-
plagioclase granulite (SM41R; Table 3). However, in the amphiboles of the mafic granulite,
the MgO and K2O contents decrease and the Na2O content increases (XA15B; Table 3). The
FeOt content decreases from core to rim in the amphiboles of both samples. Nevertheless, the
FeOt content in the amphiboles of the diopside-pargasite-plagioclase granulite is higher than
the content of the amphiboles in the mafic granulite (Table 3). The core to rim compositional
variations among the amphiboles in the metanorites have two exceptions regarded to the
presence of one or both pyroxenes, the Na2O and TiO2 contents. They decrease in
amphiboles from sample with just orthopyroxene (SM42) and increase in the metanorite
sample with diopside and orthopyroxene (XS111B; Table 3). MgO and F contents increase
from core to rim in the amphiboles of both samples, whereas FeOt, K2O, Cr2O3 and Cl
decrease (Table 3).
The pyroxenes were classified as enstatite (n = 70) and diopside (n = 42). Among the
metanorites, the sample SM42 has just enstatite. According to the classification of Rietmeijer
(1983) showed in the Figure 10A, mostly of the enstatite crystals of the metatexitic mafic
granulite and orthopyroxene-diopside granodiorite gneiss fell in the field of the metamorphic
orthopyroxene. However, the core analyses of enstatite crystals from the metanorites plot in
the field of the igneous orthopyroxene (Figure 10A-B). As the diagram shows, the igneous
enstatite cores have higher Ca content (up to 0.115 apfu; Table 4; Figure 10B). The enstatite
crystals distinguish themselves by the MgO and MnO content (Figure 10C). The crystals in
the metatexitic orthopyroxene-diopside granodiorite gneiss have the highest MnO content (up
to 1.13 wt% of MnO; Table 4; Figure 10C). The enstatite crystals in the metanorites have the
highest MgO content (up to 24.22 wt% of MgO; Table 4), but the lowest MnO content (less
than 0.41 wt% MnO).
61

The enstatite in the mafic granulite (XA15B) shows an intermediate content of MnO
and the same range of MgO variation showed by the orthopyroxene-diopside granodiorite
gneiss (Table 4). In addition, the FeOt content of the enstatite crystals in the granulites is also
higher than those of the clinonestatite crystals of the metanorites (Table 4). Looking further in
the core to rim variation, the enstatite crystal shows some specific features: (i) Al2O3 and
FeOt increase is observed in all the crystals, not depending of the sample; (ii) the crystals in
the metanorites show an increase of TiO2, Cr2O3 and MgO and a decrease of CaO; and (iii)
the crystals in the orthopyroxene-diopside granodiorite gneiss and in the mafic granulite show
a decrease in the MgO content (Table 4).
Diopside displays similar response regarded to the MnO, FeOt and MgO content. In
the granulites (SM11P, SM41R and XA15B), the diopside shows the higher contents of MnO
(up to 0.50 wt%) and FeOt (up to 9.16 wt%). The diopside of the metanorites shows the
higher content of MgO (up to 14.95 wt%). The MgO gradually decreases among the diopside
in the granulites, whereas the diopside-pargasite-plagioclase granulite shows the lowest MgO
contents in diopside (Table 4). The Al2O3 content is higher in the diopside of the mafic rocks
(SM41R, XA15B and XS111B), but metanorite diopside has the highest Al2O3 contents (up
to 1.84 wt%; Table 4) when compared to that of the orthopyroxene-diopside granodiorite
gneiss. The diopside in the granulites (SM11P, SM41R and XA15B) has the highest contents
of Na2O (up to 0.50 wt%). The avaliation of the core to rim compositional variation of the
diopside crystals showed the following patterns among the granulites and the metanorite: (i)
decrease of FeOt, MnO, and Na2O from core to rim; and (ii) increase of CaO from core to rim
(Table 4). In addition, the diopside crystals of the mafic granulite (XA15B; Table 4) show an
decrease of MgO and an increase of Al2O3 from core to rim, which goes in the opposite
direction with respect the behave of these oxides in the other samples.

5.2. Xingu Complex

The feldspar crystals of both metatexitic pargasite-biotite tonalite gneiss (XS105P; n =


7) and amphibolite (XS33; n = 3) were classified as albite due to their high Na2O and lower
CaO contents (up to 8.94 wt% of Na2O and 0.07-0.09 wt% of CaO; Table 2). In addition, the
feldspar crystals of both rocks show high K2O contents (4.32-5.53 wt%; Table 2). The core to
rim compositional variations of the feldspar crystals runs in opposition for these rocks. In the
crystals of the metatexitic hornblende-biotite tonalite gneiss, SiO2, Fe2O3t, CaO and Na2O
decrese, and K2O and Al2O3 increase (Table 2). The feldspar crystals of the amphibolite
62

show an increase of SiO2, Fe2O3t and Na2O, and a decrease of Al2O3 and K2O from core to
rim (Table 2; Figure 10F).
Amphibole (n = 7) of the metatexitic hornblende-biotite tonalite gneiss plot in the
pargasite field in the classification diagram of Hawthorne et al. (2012; Figure 10D). The
amphibole crystals (n = 7) of the amphibolite has composition of magnesio-hornblende
(Figure 10D). Other differences in the chemical composition of the amphiboles in these rocks
are significant. Amphibole in the metatexitic pargasite-biotite tonalite gneiss has higher
Al2O3, TiO2, MnO, FeOt, Na2O, K2O and Cl (13.08–13.12 wt%; 0.54–0.58 wt%; 0.51–0.53
wt%; 22.95–23.01 wt%; 1.24–1.25 wt%; 1.71–1.74 wt%; 0.23–0.34 wt%; Table 3) contents.
Magnesio-hornblende from amphibolites shows higher contents of SiO2, Cr2O3, MgO and F
(49.08-49.45 wt%; 0.13-0.15 wt%; 15.13-15.26 wt%; 0.32-0.37 wt%; Table 3). The
compositional variation between core and rim of the amphiboles in both rocks shows some
similar patterns. TiO2, Cr2O3, CaO, Na2O, F and Cl increase and K2O decreases in the
amphiboles from the metatexitic pargasite-biotite tonalite gneiss and from the amphibolite
(Table 3). However, SiO2, Al2O3, MnO, FeOt and MgO diplay an opposite variations in the
amphibole crystals of both rocks. SiO2 and MgO increase in the amphibole crystals of the
metatexitic pargasite-biotite tonalite gneiss (Table 3). Al2O3, MnO and Fe increase in the
amphibole crystals of the amphibolite (Table 3).
63

Table 2. Mineral chemistry data of the feldspar crystals from Xicrim-Cateté Orthogranulite and Xingu Complex Rocks.

Unit Xicrim-Cateté Orthogranulite Xingu Complex

Opx-Di
Di-Prg-Pl Hbl-Bt Tonalite
Lithotype Granodiorite Mafic Granulite Metanorite Amphibolite
Granulite Gneiss
Gneiss

Sample SM11P SM11P SM41R SM41R XA15B XA15B SM42 SM42 XS111B XS111B XS105P XS105P XS33 XS33

Spot Pl1.1 Pl1.8 Pl1.1 Pl1.3 Pl2.1 Pl2.2 Pl1.1 Pl1.6 Pl1.1 Pl1.4 Pl4 Pl3 Pl1 Pl2

Area core rim core rim core rim core rim core rim core rim core rim

SiO2 61.70 60.20 58.58 58.36 57.12 56.22 51.26 54.58 53.38 54.27 61.49 61.11 61.10 62.35

Al2O3 23.67 24.59 26.31 26.36 27.09 27.28 31.46 28.60 29.77 28.45 24.20 24.32 24.39 23.46

Fe2O3t 0.04 0.10 0.11 0.19 0.16 0.27 0.10 0.42 0.21 0.28 0.22 0.04 0.08 0.11

MnO 0.00 0.03 0.00 0.01 0.00 0.00 0.00 0.02 0.01 0.02 0.00 0.01 0.02 0.00

CaO 4.96 6.28 8.02 7.95 8.78 9.43 13.95 10.65 11.86 10.63 0.09 0.08 0.07 0.07

K2O 0.13 0.09 0.13 0.10 0.11 0.10 0.05 0.13 0.11 0.07 5.38 5.53 5.53 4.32

SrO 0.04 0.02 0.08 0.03 0.03 0.05 0.03 0.01 0.06 0.06 0.10 0.06 0.05 0.06

TiO2 0.09 0.02 0.01 0.03 0.02 0.04 0.00 0.07 0.09 0.00 0.04 0.00 0.00 0.00

BaO 0.00 0.00 0.00 0.02 0.01 0.02 0.03 0.00 0.03 0.00 0.00 0.00 0.00 0.03

Na2O 8.45 7.92 6.99 7.11 6.58 6.12 3.69 5.24 4.88 5.32 8.36 8.24 8.28 8.94

MgO 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.07 0.00 0.05 0.00 0.00 0.00 0.00

Total 99.09 99.25 100.25 100.17 99.90 99.53 100.58 99.81 100.41 99.16 99.88 99.38 99.52 99.33
64

Unit Xicrim-Cateté Orthogranulite Xingu Complex

Opx-Di
Di-Prg-Pl Hbl-Bt Tonalite
Lithotype Granodiorite Mafic Granulite Metanorite Amphibolite
Granulite Gneiss
Gneiss

Sample SM11P SM11P SM41R SM41R XA15B XA15B SM42 SM42 XS111B XS111B XS105P XS105P XS33 XS33

Spot Pl1.1 Pl1.8 Pl1.1 Pl1.3 Pl2.1 Pl2.2 Pl1.1 Pl1.6 Pl1.1 Pl1.4 Pl4 Pl3 Pl1 Pl2

Area core rim core rim core rim core rim core rim core rim core rim

Number of ions on the basis of 8 O

Si 2.757 2.698 2.613 2.607 2.564 2.539 2.319 2.467 2.408 2.469 2.764 2.767 2.760 2.805

Al 1.246 1.299 1.383 1.388 1.433 1.452 1.677 1.523 1.583 1.525 1.296 1.284 1.299 1.244

Fe 0.001 0.003 0.004 0.007 0.005 0.009 0.004 0.014 0.007 0.010 0.001 0.007 0.003 0.004

Mn 0.000 0.001 0.000 0.000 0.000 0.000 0.000 0.001 0.000 0.001 0.000 0.000 0.001 0.000

Ca 0.237 0.302 0.383 0.380 0.422 0.456 0.676 0.516 0.573 0.518 0.004 0.005 0.004 0.003

K 0.007 0.005 0.008 0.006 0.006 0.006 0.003 0.008 0.006 0.004 0.319 0.309 0.319 0.248

Sr 0.001 0.001 0.002 0.001 0.001 0.001 0.001 0.000 0.002 0.002 0.001 0.003 0.001 0.002

Ti 0.003 0.001 0.000 0.001 0.001 0.001 0.000 0.002 0.003 0.000 0.000 0.001 0.000 0.000

Ba 0.000 0.000 0.000 0.000 0.000 0.000 0.001 0.000 0.001 0.000 0.000 0.000 0.000 0.000

Na 0.732 0.688 0.605 0.616 0.573 0.536 0.324 0.459 0.427 0.469 0.722 0.729 0.725 0.780

Mg 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.005 0.000 0.003 0.000 0.000 0.000 0.000
65

Unit Xicrim-Cateté Orthogranulite Xingu Complex

Opx-Di
Di-Prg-Pl Hbl-Bt Tonalite
Lithotype Granodiorite Mafic Granulite Metanorite Amphibolite
Granulite Gneiss
Gneiss

Sample SM11P SM11P SM41R SM41R XA15B XA15B SM42 SM42 XS111B XS111B XS105P XS105P XS33 XS33

Spot Pl1.1 Pl1.8 Pl1.1 Pl1.3 Pl2.1 Pl2.2 Pl1.1 Pl1.6 Pl1.1 Pl1.4 Pl4 Pl3 Pl1 Pl2

Area core rim core rim core rim core rim core rim core rim core rim

Xor 0.01 0.01 0.01 0.01 0.01 0.01 0.00 0.01 0.01 0.00 0.31 0.30 0.30 0.24

Xab 0.75 0.69 0.61 0.61 0.57 0.54 0.32 0.47 0.42 0.47 0.69 0.70 0.69 0.76

Xan 0.24 0.30 0.39 0.38 0.42 0.46 0.67 0.52 0.57 0.52 0.00 0.00 0.00 0.00
66

Table 3. Mineral chemistry data of the amphiboles from the Xicrim-Cateté Orthogranulite and Xingu Complex rocks.

Unit Xicrim-Cateté Orthogranulite Xingu Complex

Lithotype Di-Prg-Pl Mafic Granulite Metanorite Metatexitic Hbl-Bt Amphibolite


Granulite Tonalite Gneiss

Sample SM41R SM41R XA15B XA15B SM42 SM42 XS111B XS111B XS105P XS105P XS33 XS33

Spot Hbl1.1 Hbl1.4 Hbl1.1 Hbl1.5 Hbl1.1 Hbl1.3 Hbl1.2 Hbl1.3 Hbl2 Hbl1 Hbl1 Hbl3

Area core rim core rim core rim core rim core rim core rim

SiO2 43.09 43.71 43.44 43.48 40.31 41.05 40.90 41.47 39.17 39.22 49.45 49.08

TiO2 1.07 0.97 1.86 1.88 1.55 1.38 1.03 1.21 0.59 0.54 0.24 0.22

Al2O3 11.15 11.11 10.04 10.35 12.24 11.79 11.38 11.10 13.12 13.08 6.23 6.31

Cr2O3 0.27 0.17 0.01 0.02 0.31 0.19 0.09 0.07 0.02 0.00 0.15 0.13

MnO 0.22 0.20 0.18 0.19 0.07 0.08 0.06 0.06 0.53 0.51 0.30 0.35

FeOt 13.51 13.35 12.81 12.62 15.36 14.59 15.04 14.68 23.01 22.95 11.87 12.08

MgO 12.82 13.17 13.57 13.51 10.83 11.94 11.68 12.28 6.12 6.38 15.26 15.13

CaO 11.35 11.55 11.54 11.49 11.45 11.52 11.61 11.50 11.23 11.11 12.07 11.90

Na2O 1.66 1.49 1.67 1.74 1.42 1.25 1.15 1.22 1.25 1.24 0.98 0.95

K2O 1.40 1.46 1.31 1.29 1.98 1.95 2.06 1.87 1.71 1.74 0.51 0.58

F 0.54 0.54 0.63 0.63 0.21 0.23 0.19 0.23 0.25 0.16 0.37 0.32

Cl 0.00 0.00 0.07 0.05 1.67 1.29 1.78 1.61 0.34 0.23 0.04 0.03

H2O+ 1.77 1.78 1.72 1.72 1.45 1.56 1.44 1.48 1.73 1.80 1.89 1.92
67

Unit Xicrim-Cateté Orthogranulite Xingu Complex

Lithotype Di-Prg-Pl Mafic Granulite Metanorite Metatexitic Hbl-Bt Amphibolite


Granulite Tonalite Gneiss

Sample SM41R SM41R XA15B XA15B SM42 SM42 XS111B XS111B XS105P XS105P XS33 XS33

Spot Hbl1.1 Hbl1.4 Hbl1.1 Hbl1.5 Hbl1.1 Hbl1.3 Hbl1.2 Hbl1.3 Hbl2 Hbl1 Hbl1 Hbl3

Area core rim core rim core rim core rim core rim core rim

Total 98.97 99.64 98.86 98.86 98.70 98.87 98.39 98.83 99.46 99.48 99.46 99.17

Number of ions on the basis of 22 O

Si 6.420 6.451 6.470 6.477 6.177 6.220 6.262 6.287 6.097 6.090 7.172 7.146

Aliv 1.580 1.549 1.530 1.523 1.823 1.780 1.738 1.713 1.903 1.910 0.828 0.854

Ti 0.120 0.108 0.208 0.211 0.179 0.157 0.119 0.138 0.069 0.063 0.026 0.025

Alvi 0.378 0.383 0.232 0.294 0.388 0.325 0.316 0.270 0.504 0.484 0.237 0.229

Cr 0.032 0.020 0.002 0.002 0.037 0.022 0.011 0.009 0.002 0.000 0.017 0.015

Fe3+ 0.386 0.414 0.317 0.182 0.360 0.512 0.531 0.578 0.680 0.744 0.283 0.339

Fe2+ 1.298 1.234 1.279 1.391 1.608 1.338 1.395 1.283 2.316 2.236 1.157 1.132

Mg 2.847 2.898 3.013 3.000 2.474 2.697 2.666 2.775 1.420 1.477 3.300 3.284

Mn 0.028 0.025 0.023 0.024 0.010 0.010 0.007 0.008 0.069 0.067 0.037 0.043

Ca 1.812 1.826 1.842 1.834 1.880 1.870 1.905 1.868 1.873 1.848 1.876 1.856

Na 0.479 0.426 0.482 0.502 0.422 0.368 0.342 0.359 0.376 0.373 0.256 0.353
68

Unit Xicrim-Cateté Orthogranulite Xingu Complex

Lithotype Di-Prg-Pl Mafic Granulite Metanorite Metatexitic Hbl-Bt Amphibolite


Granulite Tonalite Gneiss

Sample SM41R SM41R XA15B XA15B SM42 SM42 XS111B XS111B XS105P XS105P XS33 XS33

Spot Hbl1.1 Hbl1.4 Hbl1.1 Hbl1.5 Hbl1.1 Hbl1.3 Hbl1.2 Hbl1.3 Hbl2 Hbl1 Hbl1 Hbl3

Area core rim core rim core rim core rim core rim core rim

K 0.266 0.275 0.249 0.245 0.387 0.377 0.402 0.362 0.340 0.345 0.093 0.107

OH 1.745 1.748 1.686 1.687 1.466 1.559 1.447 1.478 1.788 1.860 1.819 1.846

F 0.253 0.252 0.297 0.299 0.101 0.109 0.091 0.108 0.123 0.080 0.171 0.145

Cl 0.001 0.000 0.018 0.014 0.434 0.331 0.462 0.414 0.09 0.061 0.010 0.009

Xfe 0.372 0.363 0.346 0.344 0.443 0.407 0.419 0.401 0.678 0.669 0.304 0.309

Xmg 0.628 0.637 0.654 0.656 0.557 0.593 0.581 0.599 0.322 0.331 0.696 0.691

Xtr 0.037 0.031 0.041 0.033 0.015 0.018 0.016 0.017 0.004 0.003 0.192 0.143

Xprg 0.033 0.017 0.033 0.018 0.018 0.011 0.017 0.002 0.001 - 0.027 0.014

XFe-act - - - - - - - - 0.002 0.004 - -


69

Table 4. Mineral chemistry data of the pyroxene crystals from the Xicrim-Cateté Orthogranulite and Xingu Complex rocks.

Unit Xicrim-Cateté Orthogranulite

Lithotype Di-Prg-Pl Mafic Granulite Metanorite


Opx-Di Granodiorite Gneiss
Granulite

Sample SM11P SM11P SM11P SM11P SM41R SM41R XA15B XA15B XA15B XA15B SM42 SM42 XS111B XS111B XS111B XS111B

Spot Opx1.2 Opx1.3 Cpx1.1 Cpx1.4 Cpx1.1 Cpx1.2 Opx1.1 Opx1.7 Cpx1.1 Cpx1.7 Opx1.2 Opx1.1 Opx1.1 Opx1.3 Cpx1.6 Cpx1.5

Area core rim core rim core rim core rim core rim core rim core rim core rim

SiO2 53.11 52.91 53.71 53.54 52.78 53.24 52.85 52.39 52.89 52.51 52.69 53.37 53.35 53.07 52.89 53.38

Al2O3 0.45 0.49 1.09 0.84 1.43 1.29 0.78 0.90 1.45 1.59 0.96 1.03 0.89 0.91 1.84 1.38

FeOt 23.22 23.22 8.64 6.71 9.24 8.01 23.13 23.89 9.16 8.38 20.59 21.30 18.66 20.60 6.65 5.89

MnO 1.13 1.10 0.50 0.42 0.40 0.40 0.87 0.89 0.40 0.36 0.37 0.36 0.32 0.41 0.15 0.11

CaO 0.70 0.77 21.97 23.59 22.32 23.37 0.76 0.35 21.70 23.05 1.12 0.36 2.93 0.39 24.03 24.29

K2O 0.00 0.01 0.01 0.00 0.00 0.00 0.01 0.05 0.00 0.00 0.01 0.00 0.00 0.01 0.00 0.00

TiO2 0.09 0.11 0.05 0.10 0.06 0.03 0.06 0.05 0.15 0.20 0.07 0.08 0.24 0.91 0.18 0.17

Cr2O3 0.00 0.00 0.00 0.02 0.04 0.05 0.03 0.03 0.00 0.00 0.03 0.06 0.08 0.11 0.18 0.17

Na2O 0.02 0.01 0.50 0.44 0.46 0.43 0.00 0.00 0.39 0.37 0.00 0.01 0.03 0.02 0.39 0.34

MgO 21.71 21.61 14.47 14.86 13.62 13.80 21.70 21.37 14.55 14.14 23.43 24.18 23.76 24.22 14.41 14.95

Total 100.44 100.23 100.94 100.51 100.35 100.62 100.19 99.93 100.69 100.60 99.28 100.74 100.25 100.65 100.72 100.69

Number of ions on the basis of 6 O

Si 1.983 1.981 1.981 1.976 1.967 1.973 1.976 1.971 1.960 1.950 1.965 1.961 1.962 1.949 1.949 1.961
70

Unit Xicrim-Cateté Orthogranulite

Lithotype Di-Prg-Pl Mafic Granulite Metanorite


Opx-Di Granodiorite Gneiss
Granulite

Sample SM11P SM11P SM11P SM11P SM41R SM41R XA15B XA15B XA15B XA15B SM42 SM42 XS111B XS111B XS111B XS111B

Spot Opx1.2 Opx1.3 Cpx1.1 Cpx1.4 Cpx1.1 Cpx1.2 Opx1.1 Opx1.7 Cpx1.1 Cpx1.7 Opx1.2 Opx1.1 Opx1.1 Opx1.3 Cpx1.6 Cpx1.5

Area core rim core rim core rim core rim core rim core rim core rim core rim

Al 0.010 0.011 0.024 0.018 0.031 0.028 0.017 0.020 0.032 0.035 0.021 0.022 0.019 0.020 0.040 0.030

Fe 0.725 0.727 0.267 0.207 0.288 0.248 0.723 0.752 0.284 0.260 0.642 0.654 0.574 0.633 0.205 0.181

Mn 0.036 0.035 0.016 0.013 0.013 0.013 0.027 0.028 0.013 0.011 0.012 0.011 0.010 0.013 0.005 0.003

Ca 0.028 0.031 0.868 0.933 0.891 0.928 0.031 0.014 0.862 0.917 0.045 0.014 0.115 0.015 0.949 0.956

K 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.001 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000

Ti 0.003 0.003 0.001 0.003 0.002 0.001 0.002 0.002 0.004 0.006 0.002 0.002 0.007 0.025 0.005 0.005

Cr 0.000 0.000 0.000 0.000 0.001 0.001 0.000 0.000 0.000 0.000 0.000 0.001 0.001 0.002 0.003 0.002

Na 0.001 0.000 0.018 0.016 0.017 0.015 0.000 0.000 0.014 0.013 0.000 0.000 0.001 0.001 0.014 0.012

Mg 1.209 1.206 0.796 0.818 0.757 0.762 1.210 1.199 0.804 0.783 1.302 1.324 1.302 1.326 0.792 0.819

Xfe 0.375 0.376 0.251 0.202 0.276 0.246 0.374 0.385 0.261 0.250 0.330 0.331 0.306 0.323 0.181 0.206

XMg 0.625 0.624 0.749 0.798 0.724 0.754 0.626 0.615 0.739 0.750 0.670 0.669 0.694 0.677 0.819 0.794

Xen 0.370 0.370 - - - - 0.370 0.360 - - 0.420 0.440 0.420 0.430 - -


71

Unit Xicrim-Cateté Orthogranulite

Lithotype Di-Prg-Pl Mafic Granulite Metanorite


Opx-Di Granodiorite Gneiss
Granulite

Sample SM11P SM11P SM11P SM11P SM41R SM41R XA15B XA15B XA15B XA15B SM42 SM42 XS111B XS111B XS111B XS111B

Spot Opx1.2 Opx1.3 Cpx1.1 Cpx1.4 Cpx1.1 Cpx1.2 Opx1.1 Opx1.7 Cpx1.1 Cpx1.7 Opx1.2 Opx1.1 Opx1.1 Opx1.3 Cpx1.6 Cpx1.5

Area core rim core rim core rim core rim core rim core rim core rim core rim

Xfs 0.130 0.140 - - - - 0.130 0.150 - - 0.092 0.110 0.075 0.103 - -

Xdi - - 0.690 0.750 0.680 0.690 - - 0.730 0.720 - - - - 0.730 0.760

Xhed - - 0.230 0.210 0.220 0.240 - - 0.250 0.260 - - - - 0.210 0.190


72

Figure 10. Mineral chemistry diagrams for the metatexitic orthopyroxene-diopside granodiorite gneiss (SM11P), metatexitic
diopside-pargasite-plagioclase granulite (SM41R), metatexitic mafic granulite (XA15B), metanorites (SM42; XS111B),
metatexitic pargasite-biotite tonalite gneiss (XS105P) and the amphibolite (XS33). A) Rietmeijer (1983) diagram for
chemical differentiation of igneous and metamorphic orthopyroxene; B) SEM map for the ternary Mg–Fe–Ca composition
showing the variation in the 100Ca/(Fe+2+Mg+Ca) parameter from core to rim of the enstatite crystals in the metanorite; C)
Mn vs. Mg binary diagram distinguishing the enstatite composition of the metatexitic orthopyroxene-diopside granodiorite
gneiss (SM11P), mafic granulite (XA15B) and metanorites (SM42; XS111B); D) amphibole classification diagram of
Hawthorne et al. (2012); E) F/(F+Cl+OH) vs. Mg/(Fe+Mg) diagram showing the distinct halogen composition of the
amphiboles in the granulites (XA15B; SM41R), metanorites (SM42; XS111B), pargasite-biotite gneiss (XS105P) and
amphibolite (XS33); F) Al vs. Si diagram for the feldspar crystals allowing the distinction between the feldspar crystals in the
Xicrim-Cateté Orthogranulite and Xingu Complex.

5.3. Geothermobarometry

For the geothermobarometric calculations, the following equilibrium paragenesis were


considered: (i) plagioclase–diopside–F-rich pargasite–(orthopyroxene) related to the peak
granulite facies metamorphism; (ii) Na-rich plagioclase–Cl-rich pargasite–quartz reflecting
the retrograde metamorphism and partial melting of the granulite facies rocks; (iii) albite–
pargasite–quartz related to the amphibolite facies metamorphism in the metatexitic gneisses of
the Xingu Complex; and (iv) albite–Al-poor magnesio-hornblende–quartz associated with the
contact metamorphism in the Xingu amphibolite assimilated by the Nova Canadá
Leucogranite.
The P–T conditions of the prograde metamorphism recorded in the diopside–
pargasite–plagioclase granulite and in mafic granulite from the Chicrim-Cateté Orthogranulite
73

were estimated using the TWQEEU (Berman, 1991) and Thermocalc Average PT (Holland
and Powell, 1998). Four areas of one sample of the diopside-pargasite-plagioclase granulite
were chosen for the geothermobarometric analysis. The TWQEEU analyses of the association
plagioclase-diopside-F-rich pargasite returned the following set of independent reactions:

Prg + 5Qtz + 5Hd = Ab + An + 4Di + FeTr; (1)


Ab + Di + Ts = Prg + 3Qtz + An; (2)
2Ab + 15Hd + 5Ts = 2Prg + 3FeTr + 7Di + 8An; (3)
5Hd + 2Qtz + Ts = FeTr + 3Di + 2An; (4)
5Hd + 8Qtz + 2Prg = Ts + FeTr + 5Di +2Ab; (5)
2Ab + 5Hd + 2Ts = 3Prg + FeTr + 7Qtz + 5An. (6)

Graphical and thermodynamic analysis of the intersections among the set of


independent reactions returned the following temperature–pressure conditions for the four
areas: (i) 907 ±136.7 ºC and 10.0 ±1.28 kbar; (ii) 912 ±64.4 ºC and 13.3 ±1.57 kbar; (iii) 920
± 84.3ºC and 11.7 ±0.03 kbar and (iv) 975 ±66.8 ºC and 13.7 ±0.03 kbar (Table 5).
The Thermocalc Average PT analyses of the diopside-pargasite-plagioclase granulite
returned the following set of independent reactions:

CaTs + Qtz = An; (7)


4An + Di +Prg = Ab + 5CaTs +Tr; (8)
4An + 5Hd + Prg = Ab + 5CaTs +Tr (9)
4An + 5Hd + Prg = Ab + 4Di + 5CaTs +FeAct (10)
5Di + 8CaTs + FeAct + Gln = 6An + 5Hd + 2Prg (11)
5Hd + 2Prg + 8Qtz = 2Ab + 5Di + FeAct + Ts; (12)
20An + 5Hd + 5Prg = 5Ab + 25CaTs + 4Tr + FeAct (13)

The thermodynamic analysis using Thermocalc returned the following results: (i) 925
±440.0 ºC and 10.3 ±2.10kbar (sigfit 1.13); (ii) 1128 ±548.0 ºC and 8.7 ±2.20 kbar (sigfit
0.69); (iii) 1064 ±447.0 ºC and 10.8 ±2.00 kbar (sigfit 0.94) and (iv) 975 ±407.0 ºC and 11.0
±1.90 kbar (sigfit = 0.90; Table 5).
The microprobe analysis in the mafic granulite were carried out in two areas of the
same thin section. TWQEEU database using all the solution models for the association
plagioclase–diopside–F-rich pargasite–orthopyroxene does not return any conclusive results
74

from these analyses. Moreover, the solution models for the association plagioclase–diopside–
F-rich pargasite returned the following set of independent reactions:

FePrg + 5Qtz + Hd = Ab + An + FeTr; (14)


Ab + 4Hd + Ts = FePrg + 3Qtz + Di + An; (15)
2Ab + 23Hd + 5Ts = 2FePrg + 3Qtz + 3Di + An; (16)
5Hd + 2Qtz + Ts = FeTr + 3Di + 2An; (4)
3Di + 8Qtz + 2FePrg = Ts + FeTr + 3Hd + 2Ab; (17)
5FePrg + 23Qtz + 3Di = 5Ab + 2An + 4 FeTr +Ts; (18)
Prg +5Qtz + 5Hd = Ab + An + 4Di + FeTr; (1)
Ab + Di + Ts = Prg + 3Qtz + An; (2)
2Ab + 15Hd + 5Ts = 2Prg + 3FeTr + 7Di + 8An; (19)
5Hd + 8Qtz + 2Prg = Ts + FeTr + 5Di + 2Ab; (20)
3Ab + 5Hd + 4Ts = 3Prg + FeTr + 7Qtz + 5An. (21)

Graphical and thermodynamic results returned the following values of temperature and
pressure: (i) 1077 ±164.1 ºC and 9.8 ±2.17 kbar; (ii) 976 ±223.3 ºC and 9.8 ±2.59 kbar (Table
5).
Thermocalc database allowed the analysis of all the components of the association
plagioclase–diopside–F-rich pargasite–orthopyroxene. The average pressure–temperature
calculations returned the following set of independent reactions:

CaTs + Qtz = An; (7)


2CaTs + 2En + 2FeAct = 2An + 4Di + 5Fs + 2H2O; (22)
Di + Prg + 5Qtz = An + Ab + Tr; (23)
4Prg + 18Qtz = 4An + 4Ab + 3En + 2Tr + 2H2O; (24)
4An + 5Hd + Prg = Ab + 4Di + 5CaTs + FeAct; (25)
5Hd + 2Prg + 8Qtz = 2Ab + 5Di + FeAct + Ts (26)

The thermodynamic analysis returned the following average temperatures and


pressures: (i) 1070 ±72 ºC and 8.1 ±2.40 kbar (sigfit = 0.99); (ii) 1071 ±71 ºC and 8.4 ±2.40
kbar (sigfit = 1.01; Table 5).
The retrograde metamorphism and partial melting was easily recognized in the
metanorite due to its textural transformations. The mineral association related to the
retrograde metamorphism is represented by Na-rich plagioclase–Cl-rich pargasite–quartz in
75

both samples of the metanorites. The average pressure was obtained through the
𝑝𝑙𝑔

thermodynamic reaction: 𝑃(𝑘𝑏𝑎𝑟) = (8.3144𝑇(𝐾)𝑙𝑛𝐷𝐴𝑙𝑎𝑚𝑝 − 8.7𝑇(𝐾) + 23377𝑋𝐴𝑏 − 11302)/(−274),


𝑆𝑖

𝑝𝑙𝑔

where P is the pressure in kbar, T is the entrance temperature in Kelvin, 𝐷𝐴𝑙𝑎𝑚𝑝 is the molar
𝑆𝑖

plagioclase/amphibole Al–Si partition coefficient and XAb is the albite fraction in the
plagioclase (Molina et al., 2015). The average temperature were obtained through the
0.677𝑃−48.98+𝑌
following reaction: 𝑇(𝐾) = −0.0429−0.008314𝑙𝑛𝐾 where T is the temperature in Kelvin, Y is the (for
XAb <0.5; samples SM42 and XS111B; Table 2) equal −8.06 + 25.5(1 − 𝑋𝐴𝑏 )², K is equal
(𝑆𝑖−4)
((8−𝑆𝑖)) 𝑋𝐴𝑏 and XAb is the albite fraction in the plagioclase.

Amphibole and plagioclase crystals of the metanorites satisfy all the necessary
conditions defined by Blundy and Holland (1990) and Molina et al (2016) to use the
thermodynamic reactions. The calculation of the average temperatures and pressures were
carried out in two areas, giving the following results: (i) 893 ºC and 5.5 kbar (sample SM42;
Table 5); and (ii) 905 ºC and 4.4 kbar (sample XS111B; Table 5). The temperature standard
deviation (±38 ºC) consider the complete coupling of Al on the T1 site and of Na in the A site
of the amphibole (Blundy and Holland, 1990). Molina et al (2016) consider a precision of
±1.5 to ±2.5 kbar.
The rocks of Xingu Complex are representative of two distinct scenarios. The
metatexitic pargasite-biotite tonalite gneiss (XS105P) occurs in the north boundary of the unit,
closer to the tectonic contact with the Xicrim-Cateté Orthogranulite. The amphibolite occurs
as xenolith inside the Nova Canadá Leucoganite (in the outcrop showed in the Figure 8B).
The thermobarometric calculations were carried out using the formula of Blundy and Holland
(1990) and Molina et al (2016). The average pressure and temperature results of the
metatexitic pargasite-biotite tonalite gneiss place the albite–pargasite paragenesis observed in
this rock in the upper amphibolite facies (785 ºC and 8.8k bar; Table 5). The albite + Al-poor
magnesio–pargasite paragenesis displayed by the amphibolite indicates a hornfels facies
metamorphism (672 ºC and 0.6 kbar; Table 5).
76

Table 5. Summary of the geothermobarometric data of the Xicrim-Cateté Orthogranulite and Xingu Complex.

Temperature Pressure Method Reference


(ºC) (kbar)

Xicrim-Cateté Orthogranulite

Peak Granulite Facies Metamorphism

Di-Prg Mafic Granulite 907 – 975 10.0 – 13.7 TWQ (Berman, 1991)
(SM41R)

925 – 1128 8.7 – 11.3 Thermocalc (Holland and Powell, 1998)

Mafic Granulite 976 – 1077 9.8 TWQ (Berman, 1991)


(XA15B)

1070 – 1071 8.1 – 8.4 Thermocalc (Holland and Powell, 1998)

Retrograde Amphibolite Facies Metamorphism

Metanorite (SM42) 893 5.5 Al-Si in plagioclase- (Blundy and Holland, 1990;
amphibole pairs Molina et al., 2015)

Metanorite (XS111B) 905 4.4 Al-Si in plagioclase- (Blundy and Holland, 1990;
amphibole pairs Molina et al., 2015)

Xingu Complex

Peak Anfibolite Facies Metamorphism

Hbl-Bt Tonalite Gneiss 785 8.8 Al-Si in the pair (Blundy and Holland, 1990;
(XS105P) plagioclase-amphibole Molina et al., 2015)

Contact Metamorphism

Amphibolite (XS33) 671 0.6 Al-Si in the pair (Blundy and Holland, 1990;
plagioclase-amphibole Molina et al., 2015)

6. U-Pb Geochronology
6.1. Xicrim-Cateté Orthogranulite
6.1.1. Metatexitic Mafic Granulite

Zircon grains were collected from the mafic granulite with the equilibrium assemblage
plagioclase-diopside-pargasite (XA15B3) and from the leucosome of the schollen diatexite
with orthopyroxene, and subordinated diopside and plagioclase porphyroblasts (XS112A).
The schollen diatexite shows diopside-pargasite-plagioclase granulite fragments (see Figure
4B).
The mafic granulite (XA15B3) zircon grains are light brown with crystal length
ranging from 80 to 100 µm and aspect ratios (length/width) around 1:1. The CL and BSE
images show that the grains are almost rounded with interior sector zoning (Figure 11A). The
77

sample was analyzed by the U–Pb LA–ICP–MS method. The Th/U ratios range from 0.4 to
2.8 with a mean value of 2.1. The ten spot analyzes by U–Pb LA–ICP–MS in the sample
XA15B3 yielded a concordia age of 2,890 ±7 Ma (MSWD = 0.6; Figure 11D).
The zircon grains of the schollen diatexite leucosome (XS112A) were analysed by U–
Pb LA–ICP–MS and U–Pb SHRIMP IIe methods. The grains are light brown with a crystal
length ranging from 80 to 200 µm, and aspect ratios (length/width) from 1:1 to 3:1. The
cathodoluminescence (CL) and the back-scattering images (BSE) evidence prismatic grains
with rounded to euhedral boundaries (Figure 11B-C). The grains mounted for the U–Pb
SHRIMP IIe analyses show internal cores with oscillatory zoning and external high
luminescence rims with euhedral external shape and sector zoning (Figure 11B). These
features cannot be observed in the CL images of the grains mounted for the U–Pb LA–ICP–
MS analyses, which are more metamitic with polygonal and sector zoning (Figure 11C). In
addition, they show corrosion and absorption features and convoluted zoning (Figure 11C).
The Th/U ratios range from 0.02 to 0.64 with a mean value of 0.11 (XS112A-SHRIMP IIe)
and from 0.01 to 1.36 with a mean value of 0.53 (XS112A-LA-ICP-MS). The nine spot
analyses by U–Pb SHRIMP IIe of the internal core in the sample XS112A yielded a
discordant result with an upper intercept age of 2,932 ±20 [22] Ma (MSWD = 1.03; Figure
11E) and the five spot analyses by U–Pb LA–ICP–MS in the sample XS112A yielded a
discordant result with an upper intercept age of 2,873 ±53 Ma (MSWD = 4.1; Figure 11F).
78

Figure 11. A) CL images of the main zircon types of the mafic granulite (XA15B3) analysed by U–Pb LA–ICP–MS with
respective 207Pb/206Pb ages; B) CL images of the main zircon grain-types of the schollen diatexite leucosome (XS112A)
analyzed by U-Pb SHRIMP IIe with respective 207Pb/206Pb ages; C) CL images of the main zircon grain-types of the schollen
diatexite leucosomes (XS112A) analyzed by U–Pb LA–ICP–MS with respective 207Pb/206Pb ages; D, E and F) 206Pb/238U vs.
207
Pb/235U diagram of the mafic granulite and the schollen diatexite leucosome. D) Results of the analyses of zircon cores in
the mafic granulite analyzed by U–Pb LA–ICP–MS; E) Results of the analyses of zircon cores in the leucosome of the
schollen diatexite analyzed by U–Pb SHRIMP IIe; F) Results of the analyses of zircon cores in the schollen diatexite
leucosome analyzed by U–Pb LA–ICP–MS.
79

6.1.2. Metatexitic Orthopyroxene-Diopside Gneiss

Zircon grains were collected from five samples of the metatexitic orthopyroxene-
diopside gneiss. The sample selection includes rocks of tonalitic (samples XA15A1, SM46N,
XA12A) and granodioritic (samples CMS22P, TS23A) composition. In addition, the sampling
considered the degree of partial melting of the gneisses. The samples with less degree of
melting were chosen. Zircon is colourless to light brown with crystal lengths ranging from 60
to 220µm, and aspect ratios (length/width) from 1:1 to 3:1. The cathodoluminescence (CL)
and back-scattering images (BSE) images show that the zircon grains are mainly prismatic
with oscillatory zoning (Figure 12A-F). However, the zircon grains are variably envolved and
penetrated by recrystallization homogeneous rims (Figure 12A-F). The sample XA15A1
shows the more aggressive effects of the recrystallization (Figure 12A-B). It has rounded
zircons with dominant homogeneus high cathodoluminescent rims and tinny cores with
ocilatory zoning (Figure 12B). The sample SM46N shows recrystallization areas disconnected
with the external surface of the zircon grains (ghost textures; Corfu et al., 2003; Figure 12C).
The sample CMS22P shows zircon grains with high cathodoluminescent homogeneous cores
and mantled by low cathodoluminescent oscillatory zoning or homogeneous newly grown
zircon (Figure 12D). Additionaly, these zircon grains show high cathodoluminescent tinny
recrystallization rims (Figure 12D).
Th/U ratios of zircon from the metatexitic orthopyroxene-diopside gneisses range from
0.01 to 1.49. Variation is observed in each sample: from 0.37 to 1.13 with a mean value of
0.60 (metatexitic orthopyroxene-diopside gneiss of tonalitic composition; XA15A1-LA-ICP-
MS), 0.33 to 1.49 with a mean value of 0.51 (metatexitic orthopyroxene-diopside gneiss of
tonalitic composition; XA15A1-SHRIMP IIe-Core), 0.08 to 0.61 with a mean value of 0.29
(metatexitic orthopyroxene-diopside gneiss of tonalitic composition; SM46N-Core), 0.41 to
0.84 with a mean value of 0.42 (SM46N-Rim), 0.01 to 0.89 with a mean value of 0.39
(metatexitic orthopyroxene-diopside gneiss granodioritic composition; CMS22P), 0.13 to 0.67
with a mean value of 0.31 (metatexitic orthopyroxene-diopside gneiss granodioritic
composition; TS23A) and 0.03 to 0.72 with a mean value of 0.35 (metatexitic orthopyroxene-
diopside gneiss of tonalitic composition; XA12A). However, zircon rims from the XA15A1
sample have higher Th/U ratios (1.87 to 3.18 with a mean value of 2.93).
One sample of the metatexitic orthopyroxene-diopside gneisses of tonalitic
composition (XA15A1) was analyzed by the U–Pb LA–ICP–MS method and by U–Pb
SHRIMP IIe method. Six analyses of the zircon core by LA–ICP–MS yielded a concordia
80

age of 2,935 ±8 Ma (MSWD = 0.27; Figure 13A), whereas five analyses of zircon cores from
the same sample carried out by SHRIMP IIe yielded a discordant result with an upper
intercept age of 2,955 ±8 [11] Ma (MSWD = 1.3; Figure 13B). The zircon rims from the
XA15A1 sample were analyzed by SHRIMP IIe and yielded a discordant result with an upper
intercept age of 2,853 ±19 [23] Ma (MSWD = 0.64; Figure 13B) defined by five spots.
The other samples of orthopyroxene-diopside metatexitic gneisses of tonalitic and
granodioritic composition (SM46N, CMS22P, TS23A and XA12A) were analyzed using the
U-Pb SHRIMP IIe method. In the sample SM46N, nine grains yielded a discordant result with
an upper intercept age of 2,953 ±18 [20] Ma (MSWD = 4.4; Figure 13C). Three analyses of
the zircon rims in the sample SM46N indicated an age of 2,819 ±13 [15] Ma (MSWD = 22;
Figure 13C).
In the sample CMS22P, four grains yielded a concordia age of 3,066 ±7 Ma (MSWD
= 0.072; Figure 13D) and other six grains yielded a discordant result with an upper intercept
age of 2,954 ±71 Ma (MSWD = 22; Figure 13D). Two of the four older zircon grains have
low-U homogeneous cores.
Eight grains from the sample TS23A yielded an upper intercept age of 2,987 ±18 [19]
Ma (MSWD = 7; Figure 13E). In the sample XA12A, eight grains yielded a discordant result
with an upper intercept age of 2,979 ±31 [33] Ma (MSWD = 4,4; Figure 13F).
81

Figure 12. CL images of zircon from the metatexitic orthopyroxene-diopside gneisses of the Xicrim-Cateté Orthogranulite.
A) CL images of the XA15A1 sample analyzed by LA-ICP-MS. The arrows show the spots analyzed with the respective
207
Pb/206Pb ages. B) CL images of the XA15A1 sample analyzed by SHRIMP IIe. The arrows show the analyzed spots with
the respective 207Pb/206Pb ages; C) CL images of the SM46N sample. The white black circles show the place where the grains
were analyzed by the SHRIMP IIe method with the respective 207Pb/206Pb ages. D) CL images of the CMS22P sample. The
white circles show the place where the grains were analyzed by the SHRIMP IIe method with the respective 207Pb/206Pb ages.
E) CL images of the TS23A sample. The white circles show the place where the grains were analyzed by the SHRIMP IIe
method with the respective 207Pb/206Pb ages. F) CL images of the XA12A sample. The white circles show the place where the
82

grains were analyzed by the SHRIMP IIe method with the respective 207Pb/206Pb ages. The zircon grains in all the samples
share similar features, such as preserved oscillatory zoning and high luminescence rims. The thickness of this rims are
variable. In some cases it can grow all over the grain, which is the case of the first grain in the B.

Figure 13. 206Pb/238U vs. 207Pb/235U diagram for the metatexitic orthopyroxene-diopside gneisses of the Xicrim-Cateté
Ortogranulite. A) Results of the analyses of the zircon cores in the XA15A1 sample analyzed by LA-ICP-MS. B) Analyses of
zircon cores and rims in the sample XA15A1 (SHRIMP IIe). C) Analyses of zircon cores and rims in the sample SM46N
(SHRIMP IIe). D) Analyses of zircon cores in the CMS22P sample (SHRIMP IIe), which showed two age groups; E) Results
of the analyses of zircon cores in the TS23A sample (SHRIMP IIe); F) Results of the analyses of zircon cores in the XA12A
sample (SHRIMP IIe).
83

6.2. Xingu Complex

The zircon grains where collected from two samples of the metatexitic gneiss of the
Xingu Complex: (i) XS04P, which corresponds to biotite metatexitic gneiss of granodioritic
composition and (ii) XS105P, hornblende-biotite metatexitic gneiss of tonalitic composition.
Zircon is colourless to light pink, with crystal length ranging from 70 to 250 µm, and aspect
ratios (length/width) from 1:1 to 3:1. The CL and BSE images show that the zircon grains are
predominantly prismatic with local rounded terminations (Figure 14A-B). Zircon in both
samples displays prominent oscillatory zoning in the CL images, as well as an incipient
metamitic character, regarded to few fractures and inclusion incidence (Figure 14A-B). Th/U
ratios range from 0.16 to 0.76 with a mean value of 0.46 (XS04P) and from 0.22 to 0.54 with
a mean value of 0.36 (XS105P).
The metatexitic biotite granodiorite gneiss (XS04P) sample was analyzed by the U–Pb
LA–ICP–MS method and eleven spot analyses yielded a concordant age of 2,936 ±6 Ma
(MSDW = 2.5; Figure 15A). The metatexitc hornblende-biotite tonalite gneiss (XS105P)
sample was analyzed by U–Pb SHRIMP IIe and twelve spot analyses yielded a discordant
result with an upper intercept age of 2,928 ±15 [17] Ma (MSWD = 14; Figure 15B).

Figure 14. A) CL images of the main zircon types of the biotite metatexitic gneiss of the Xingu Complex (XS04P), showing
analyzed spots and corresponding 207Pb/206Pb ages. B) CL images of the most concordant zircon grains of the pargasite-
biotite metatexitic gneiss (XS105P) and their 207Pb/206Pb ages.
84

Figure 15. 206Pb/238U vs. 207Pb/235U diagrams for the metatexitic gneisses of the Xingu Complex. A) Concordia diagram for
the biotite metatexitic gneiss of the Xingu Complex (sample XS04P); B) Concordia diagram for the hornblende-biotite
metatexitic gneiss of the Xingu Complex (sample XS105P).

7. Discussion
7.1. Periods of crust formation and reworking during the Mesoarchean in the Carajás
Domain

The ages obtained for the metamorphic rocks of the Carajás Domain allow the
construction of an integrated view of the growth and reworking of the Mesoarchean basement.
Feio et al. (2013) distinguish through geochronology four major magmatic events defining the
crustal growth of the Carajás Domain: 3.05–3.00, 2.96–2.93, 2.87–2.83 and 2.75–2.73 Ga.
The ages obtained for the Xicrim-Cateté Orthogranulite and the Xingu Complex fall in the
three Mesoarchean intervals distinguished by Feio et al. (2013).
The zircon grains of the Xicrim-Cateté Orthogranulite show a complex evolution.
According to the criteria of Rubatto (2017), the features observed in the zircon grains of the
metatexitic orthopyroxene-diopside gneiss fall in two categories: (i) overgrowths or new
growth and (ii) replacement/recrystallization. The overgrowths are represented by the newly
oscillatory zoning high-U domains that sharply surround homogeneous low-U relicts
(Rubatto, 2017). Such feature was observed in the metatexitic orthopyroxene-diopside
granodiorite gneisses (sample CMS22P; Figure 12D). The concordant relic cores reveal
207Pb/206Pb ages of 3,100 ±16 Ma and 3,056 ±10 Ma, associated with a concordia age of
3,066 ±7 Ma (MSWD = 0.072). The concordant newly formed domains showed a
207Pb/206Pb age of 2,963 ±8 Ma, associated with an upper intercept age of 2,954 ±71 Ma
(MSWD = 22).
85

The results obtained from the oscillatory zoning domains in the zircon grains of the
metatexitic orthopyroxene-diopside gneisses, the schollen diatexite leucosome (core analyses;
Figure 11C-11E) and in the metatexitic hornblende-biotite and biotite gneisses from the
Xingu Complex indicate a range of crust formation between ca. 2.93 Ga and 2.99 Ga.
Pidgeon et al. (2000) obtained a corcordia age of 3,002 ± 14 Ma (U–Pb SHRIMP in
zircon) from oscillatory zoned cores of zircon crystals in enderbites related to the Xicrim-
Cateté Orthogranulite (see Vasquez and Rosa-Costa, 2008). Moreto et al. (2015; 2011)
obtained, mostly in oscillatory zoning domains of zircon crystals, a range of ages from the
Sequeirinho granite, Bacaba Tonalite and Pista felsic metavolcanic rocks in between ca. 3.08
and 2.97 Ga. These results indicate that the main period of crustal growth in the Carajás
Domain occur between ca. 3.08 and 2.93 Ga
The replaced/recrystallized domains are associated with low-U rims with no internal
zoning and sharp boundaries (Corfu et al., 2003; Rubatto, 2017). Such features, including the
soccerball texture (sample XA15A1; Figure 12B), are observed in zircon grains of the
Xicrim-Cateté Orthogranulite and Xingu Complex rocks. This feature is associated with in
situ dissolution-precipitation process in subsolidus conditions during high-T events and is
commonly aided by fluids (Corfu et al., 2003; Geisler et al., 2007; Harley et al., 2007;
Rubatto, 2017). The metamorphic rims yielded an upper intercept age of 2,853 ±19 Ma
(MSWD = 0.64; sample XA15A1; Figure 13B). The Th/U ratio in the metamorphic rims of
the zircon grains are always >0.1 (0.41 – 3.18 ppm). High Th/U ratios in metamorphic zircons
occur in rocks whose monazite and allanite are absent (Rubatto, 2017). Nevetheless, UHT
rocks (T > 900ºC) are a distinct case where Th-phases (ex. monazite and allanite) are not
stable and the metamorphic zircons show Th/U above 0.1 with values as high as 3 (Rubatto,
2017).
Pidgeon et al. (2000) obtained an age of 2,859 ±9 Ma from the metamorphic rims of
the zircon grains of the enderbites related to the Xicrim-Cateté Orthogranulite. The
recrystallized zircon rims of the enderbites were related to a granulite facies event. However,
the ages of recrystallized rims provide the history of cooling from the high temperatures due
to the interaction with aqueous fluids and melts (Geisler et al., 2007; Harley et al., 2007).
Among the partially melted rocks of the Xicrim-Cataté Orthogranulite and Xingu
Complex, the zircon grains of the leucosome carefully sampled from an outcrop of a schollen
diatexite (Figure 4A-C) show two main zircon populations: (i) one with euhedral shape,
oscillatory zoning cores and sector zoning rims whose cores yielded an upper intercept age of
86

2,932 ±20Ma (MSWD 1:03); and (ii) one with euhedral shape, sector zoning truncated by
irregular-curved high-CL intensity domains, convoluted zoning and metamitic zones that
yielded an upper intercept age of 2,873 ±53Ma (MSWD = 4.1). The core ages of the first
population of zircon grains fits in the context of the massive crust formation period (ca. 3.08-
2.93 Ga). The second population of zircon grains represent metamorphic zircon formed due to
dissolution-repreciptation re-equilibration in the presence of melt (Geisler et al., 2007; Harley
et al., 2007). Additionaly, Machado et al. (1991) obtained from pink euhedral zircon grains of
undeformed leucosomes in the northern portion of the Carajás Domain the age of 2,859 ±2Ma
(U–Pb TIMS in zircon).
The zircon grains of the mafic granulite (sample XA15B3; Figure 11A) showed
features typically found in metamorphic zircon associated with granulite and UHT
metamorphic rocks, such as internal planar banding, sector zoning and high Th/U ratios (0.4 –
2.8 ppm; Corfu et al., 2003; Harley et al., 2007; Rubatto, 2017). These zircon grains yielded a
concordia age of 2,890 ±7 Ma (MSWD = 0.60). These several results indicate that the interval
of 2.89–2.86 Ga was a period of intense reworking of the Middle Mesoarchean (3.08–2.93
Ga) crust due to the high-grade metamorphism and partial melting. At west of the Goiaba
Creek area, Silva et al. (2018) point out the relevance of crust contribution for the genesis of
the granitoids emplaced in the the southern boundary of the Carajás Domain at 2.87–2.86 Ga.
The authors highlight the shape of the bodies and their alignment to the main E–W regional
foliation. Such features are observed in the (i) Água Limpa Granodiotite (ca. 2.87 Ga; Gabriel
et al., 2015) hosted by the low-angle foliation of the Xingu Complex rocks and in the (ii)
Nova Canadá Leucogranite (2.87 Ga; Leite-Santos, 2016) whose field and petrographic
features resemble those of typical schlieren diatexite.
Additionaly, the data of several syn- to post-tectonic granitoids with important crustal
contribution emplaced all over the Carajás Domain in ca. 2.87–2.83 Ga period (Feio et al.,
2013; Gabriel and Oliveira, 2014; Leite-Santos, 2016; Moreto et al., 2011; Rodrigues et al.,
2014) also highlights the widespread reworking of the previous Middle Mesoarchean (3.08–
2.93 Ga) crust. This scenario of intense tectonic-metamorphic activity characterizes the
collisional cycle between the Carajás and Rio Maria domains, related to the amalgamation of
the Carajás Province in ca. 2.89–2.83 Ga. Silva et al. (2018) also point out this collisional
setting due to the characteristics of the syn-tectonic magmatism in the southern Carajás
Domain boundary in ca. 2.87–2.86 Ga. This conclusion is also consistent with the results
87

obtained by the metamorphic zircon and zircon rims of the Xicrim–Cataté Orthogranulite and
the Xingu Complex (Machado et al., 1991; Pidgeon et al., 2000; this work).

7.2. New constrains on metamorphic evolution of the Carajás Domain.

Araújo and Maia (1991) carried out the first insights of the metamorphic evolution of
the Carajás Domain. At that time, the authors concluded that the juxtaposition of the Xingu
Complex rocks and the rocks recently renamed as Xicrim-Cataté Orthogranulite occurred due
to a tectonic elevation in a low-angle ductile shear zone related to an imbricated tectonic
system. The metamophic peak at granulite facies was estimated at 700–800 ºC and pressures
around 4 and 6 kbar in the Xicrim-Cateté Orthogranulite, previously called as Pium Complex
(Araújo and Maia, 1991). This intermediate pressure–high temperature metamorphism was
related to the predominant assemblage with plagioclase + orthopyroxene + diopside in both
felsic and mafic rocks (Araújo and Maia, 1991).
However, F-rich pargasite demonstrated to play an important role in the
characterization of the metamorphic peak of the Xicrim-Cateté Orthogranulite rocks. Recent
reviews on the halogen cycle in the deeper Earth indicate that the sub-arc lithospheric mantle
represents an important fluorine source (Bénard et al., 2017; Beyer et al., 2012; Urann et al.,
2017).
The experimental results indicate that F-rich pargasite may be stable at pressures and
temperatures up to 35 kbar and 1300 ºC together with clinopyroxene and orthopyroxene
(Foley, 1991). In the ultra-high temperature granulites (T > 1100ºC) of the Napier Complex
(Tonagh Island, East Antartica), the high fluorine content of the pargasitic amphiboles holds
the thermal stability (Tsunogae et al., 2003). In the Chicrim-Cateté Orthogranulite, diopside-
pargasite-plagioclase granulite and mafic granulite show XF (0.10–0.15 a.p.f.u) inside the
range showed by the two pyroxene mafic granulite of the Napier Complex (0.08–0.48;
Tsunogae et al., 2003). The Ti (0.108–0.211 a.p.f.u) and Na2O content (1.49–1.74 wt%)
together with the higher XF in the pargasite can hold the stability of the mineral toward ultra-
high temperature conditions (Sajeev et al., 2009). Additionaly, the ultra-high temperature
granulites of Highland Complex (ca. 950 ºC; ~11kbar; Sri Lanka), have clinopyroxene,
orthopyroxene and plagioclase occurring as polygonal aggregates (locally symplectitic)
surrounding the paragasite porphyroblast. This characteristic texture occurs in the Xicrim-
Cateté Orthogranulite, especially in the diopside-pargasite-plagioclase granulite and in the
88

mafic granulite, where diopside + plagioclase or diopside + orthopyroxene + plagioclase


polygonal aggregates develop around the F-rich pargasites (Figure 3I).
The set of independent reactions show that the general process related to granulite
facies metamorphism in the mafic rocks of Xicrim-Cateté Orthogranulite is the breakup of the
pargasite to form an anhydrous assemblage. This process could be related to the general
equilibrium reaction: pargasite + quartz = clinopyroxene + orthopyroxene + plagioclase +
H2O (Tsunogae et al., 2003). Since quartz is not in excess in the mafic rocks, the equilibrium
assemblage should be pargasite–plagioclase–clinopiroxene–orthopyroxene. Additionally, in
dryer environments, the relicts of the previous textures tend to be preserved and the fast
exhumation could “frozen” the metamorphic evolution (Brown, 2009; Vernon, 2004). The
thermodynamic calculations indicate a temperature–pressure range for the diopside–
pargasite–plagioclase granulite between 907–1128 ºC and 8.7–13.7 kbar. The interval of T–P
conditions for the mafic granulite show slightly higher temperatures and lower pressures
(976–1077 ºC and 8.1–9.8 kbar).
The results for the mafic rocks indicate that the Xicrim-Cateté Orthogranulite reaches
conditions of the ultra-high-temperature granulite metamorphic facies series (Brown, 2009,
2006). It almost reaches the temperature–pressure condition of the high-pressure and high- to
ultra-high temperature granulites (> 900º C and 15kbar; O’Brien and Rötzler, 2003). The high
temperature may reflect the higher thermal gradients of the Precambrian, especially in the
Archean whose upper mantle temperatures should have been 160–175K higher than today’s
temperature (Abbott et al., 1994; Labrosse and Jaupart, 2007; Sizova et al., 2010). Moreover,
the mineral reactions in ultra-high temperature rocks indicate that the peak conditions are
followed by decompression, which suggest that these rocks form in the mid levels of a
thickened crust during continental collision (Clark et al., 2011).
The one-dimensional models indicate the following potential heat sources for the UHT
metamorphism: (i) elevated radioactive heat production in thickened crust; (ii) increased
mantle heat production towards back-arc basins; and (iii) mechanical heating in ductile shear
zones (Clark et al., 2011). These authors concluded that the most potential heat source is
related to the elevated concentration of elements that produce heat during radioactive decay in
a thickened crust (Andreoli et al., 2006; Chamberlain and Sonder, 1990; Clark et al., 2011). In
addition, the ultra-high temperature conditions are easily attained in terranes that have already
undergone an episode of partial melting, unless this also reduces the concentration of heat-
producing elements. Residual ultra-high temperature granulites can be enriched in U and Th,
89

while the crustal granites derived from them have lower contents of these elements than
expected (Villaseca et al., 2007). The age of the granulite-ultra-high-temperature
metamorphism in the Xicrim-Cateté Orthogranulite is defined by the metamorphic zircon
grains of the ultra-high temperature mafic granulite and the zircon rims in the metatexitic
orthpyroxene-diopside gneisses at ca. 2.89 – 2.85 Ga (some metamorphic zircon grains have
Th/U ratio up to 3.18 ppm).
The partial melting (ca. 2.86 Ga; Machado et al., 1991; this work) of the Xicrim-
Cateté Orthogranulite and Xingu Complex is characterized by the following microtextures
and outcrop features: (i) cuspate terminations of the K-feldspar grains in contact with
plagioclase and quartz (Figure 6C; Figure 9A); (ii) leucosomes veinlets crosscutting and
splitting apart phenocrysts (Figure 4B; Figure 9B;); (iii) corrosion amoeboid patches in the
plagioclase crystals (Figure 4E; Figure 6D); (iv) the evolution of the metatexitic mafic
granulite into the “jaguar skin” diatexite (Figure 4A-D); (v) stromatic structures with clear
individualization of leucosomes, melanosome and the gneissic residuum (Figure 7D); (vi)
connections between the metatexite gneiss leucosomes and the Nova Canadá Leucogranite
(Figure 9E); (vii) schilieren structures in the Nova Canadá Leucogranite (Figure 9F); (viii)
assimilation of K-feldspar phenocrysts of the Água Limpa Granodiorite by the metatexitic
gneiss leucosomes (Figure 9G); and (ix) textural transformations in the residumm during the
melting volume increase (Figure 6A-B). These textures are widely known in other migmatitic
terranes (e.g. Holness et al., 2011; Johnson et al., 2013; Pawley et al., 2015; Sawyer, 2010;
Vernon, 2011). These features are related to two processes: (i) dehydration melting; and (ii)
water-fluxed melting.
The dehydratation melt can be associated with the ultra-high temperature
metamorphism in the granulites (Clark et al., 2011; Sajeev et al., 2009; Tsunogae et al., 2003).
This relationship is observed in the transition between the metatexitic diopside-pargasite-
plagioclase granulite and the “jaguar skin” diatexite with orthopyroxene, and to some extent
diopside and plagioclase porphyroblasts showed in the Figures 4A-E. This transition may be
related to the following dehydration reactions:

Amph + Qtz = Opx + Cpx + Melt (Vielzeuf and Schmidt, 2001)


Prg + Qtz = Opx + Cpx + Pl + H2O (Tsunogae et al., 2003)

Such features have a restrict occurrence and were only being documented in the
diopside-pargasite-plagioclase granulite and in the mafic granulite. The orthopyroxene,
90

clinopyroxene and plagioclase phenocrysts represent peritectic phases in the dehydration


melting of these rocks (i.e, Johnson et al., 2013; Vernon, 2011). However, the plagioclase
reacts with the melt and develops the corrosion-like textures showed in the Figure 4G. Melt
loss can limit the partial melting during prograde metamorphism, giving more chance to the
rocks reach the ultra-high temperature metamorphism (Clark et al., 2011).
The other melt-related textures founded in the rocks of the Xicrim-Cateté
Orthogranulite and Xingu Complex, which are widespread in these rocks, are not related to
the breakdown of hydrated phases. In the tiny sites of melt, it is possible to see that only
plagioclase, quartz and K-feldspar participate of the reactions. The only exception is the Cl-
rich hornblende in the metanorites, which participate in the melting process of the metanorites
(Figure 4H-I). In the field, these melt-related features develop in the low angle E–W foliation.
In the same way, milky quartz veins are associated with the foliation. They also form
kilometer scale accumulations in the main E–W shear zones displayed in the geological map
of the Goiaba Creek Area (Figure 2). In addition, quartz, plagioclase and K-feldspar are only
nominally anhydrous minerals; they have water in their structures (Qtz can have 8 wt% of
H2O in its structure; Johnson, 2006).
Nevertheless, the evidences indicate that the migmatization of the Xicrim-Cateté
Orthogranulite and Xingu Complex are related to water influx melt reactions:

Qtz + Kfs + Pl + H2O = Melt (Stevens and Clemens, 1993)


Qtz + Kfs + Pl + H2O = Pl + Melt (Genier et al., 2008)

The dehydration melting of the mafic granulites could trigger these water influx
reactions leading to a chain effect and the widespread migmatization observed in these rocks.
The metanorites also show influence of added-water in the retrograde metamorphism
paragenesis and in the partial melting. The rocks show the opposite reaction that defines the
ultra-high temperture metamorphism and the dehydration melting in the diopside-pargasite-
plagioclase granulite and in the mafic granulite. Such process could be described by the
following reaction:

CaPl + Opx ± Cpx + H2O = Na-rich Pl + Cl-rich Prg + Melt

The Cl-rich pargasite and at some extent the Na-rich plagioclase, occurs as a peritectic
phase during the melting processes, as shown in the Figure 4H. According to the
thermobarometric data, the paragenesis Na-rich plagioclase + Cl-rich pargasite + quartz and
91

the melting is related to an almost isothermic decompression (904–893 ºC and 5.5–4.4 kbar;
Blundy and Holland, 1990; Molina et al., 2015). The symplectitic texture between biotite and
epidote (Figure 7I) can also be an evidence of this decompression in the middle and lower
crust.
The thermobarometric data of the Xingu Complex is consistent with field and
petrographic features observed in the samples of the metatexitic hornblende-biotite gneisses
(i.e grains boundary migration and development of deformational twin in the plagioclase;
Passchier and Trouw, 2005; Vernon, 2004, 1999). The Xingu Complex showed a clear
assemblage change in the south to north section. This is due to different thermal regimes,
which is lower in the south, related to metatexitic biotite gneisses, higher in the north, and
related to the metatexitic pargasite-biotite gneisses. The albite-pargasite equilibrium in the
metatexitic hornblende-biotite tonalite gneiss of the Xingu Complex was defined at 785 ºC
and 8.8 kbar (Blundy and Holland, 1990; Molina et al., 2015). The amphibolite sample
showed a hornfels facies metamorphism at 671 ºC and 0.6 kbar. The low pressure is related to
the anomalous low content of Al2O3 in the magnesio-hornblende (6.31-6.23 wt%), since the
geobarometer (and geothermometer) is Al-dependent (Molina et al., 2015). The Nova Canadá
shows the typical diatexite-like features (discussed above). In addition, the microtextures of
the leucosome in the metatexitic gneisses of the Xingu Complex are similar to those found in
the Nova Leucogranite. The amphibolite occurs as a resistive lithotype inside a portion of the
diatexitic Nova Canada Leucogranite with the schlieren structure. The record of regional
metamorphism in the amphibolite likely was strongly obliterated by the thermal
metamorphism associated with the partial melting of the Xingu Complex orthogneisses and
related magma generation.
The onset of the partial melting changes the structure and the strain rates in the crust
(Sawyer et al., 2011). The pre-melting strain field will control the initial pattern of the melt
flow that will further lead to heterogeneous deformation in the crust, as the melt volume
increases (Brown, 2005; Holness et al., 2011). The melt migration and accumulation in the
base of the upper crust can lead to the upper–lower crust detachment and the development of
large scale shear zones (Brown, 2005; Brown et al., 2011). Such process can enhance the fast
tectonic exhumation of the lower crust, the fast cooling of the migmatite source and
crystallization of the residual melt (Brown, 2005; Brown et al., 2011).
The exhumation model of the Carajás Domain, in an imbricated tectonic system
(Araújo and Maia, 1991) related to the development of the Itacaiúnas Shear Belt (Holdsworth
92

and Pinheiro, 2000) could be related to this process. The almost perfect preservation of the
metamorphic structures, textures and paragenesis related to the high-grade metamorphism,
especially in the mafic rocks, in which a prograde paragenesis could be yet characterized
(plagioclase + diopside ± orthopyroxene + pargasite), is a good evidence of such process. This
kind of preservation in high-grade and anatetic terrains is only possible due to a sudden
exhumation (Clark et al., 2011; Holness et al., 2011; Vernon, 2004).

8. Conclusion

• The Carajás Domain represents the oldest crustal segment of the Amazon Craton.
The period of predominant crustal growth occurred between ca. 3.08 Ga and 2.93
Ga. Between 2.89 and 2.83 Ga, the crust was submitted to an intense reworking
period associated with the syn-tectonic magmatism during the collision of the
Carajás and Rio Maria Domain.
• The metamorphic evolution during the Mesoarchean was related to the collisional
event. The rocks of the Xicrim-Cateté Orthogranulite were submitted to a ultra-
high temperature (UHT) granulite metamorphism, whereas the Xingu Complex
rocks reach the upper amphibolite facies memorphism. The dehydration melting
enhanced the high grade metamorphism, the widespread water-influx
migmatization of those rocks and, as consequence, the fast exhumation of the
lower crust along the detachment E–W shear zones.

Reference

Abbott, D., Burgess, L., Longhi, J., Smith, W.H.F., 1994. An empirical thermal history of the
Earth’s upper mantle. J. Geophys. Res. 99, 13835. doi:10.1029/94JB00112
Almeida, J. de A.C. de, Dall’Agnol, R., de Oliveira, M.A., Macambira, M.J.B., Pimentel,
M.M., Rämö, O.T., Guimarães, F.V., Leite, A.A. da S., 2011. Zircon geochronology,
geochemistry and origin of the TTG suites of the Rio Maria granite-greenstone terrane:
Implications for the growth of the Archean crust of the Carajás province, Brazil.
Precambrian Res. 187, 201–221. doi:10.1016/j.precamres.2011.03.004
Almeida, J. de A.C. de, Oliveira, V.E.S., Rocha, M.C., Rocha, K.P.P., 2016. Geologia,
Petrografia e Geoquímica dos Granitoides Arqueanos da Porção Sul do Domínio Rio
Maria, Província Carajás, in: SBG (Ed.), Anais Do 46o Congresso Brasileiro de
Geologia. SBG - Sociedade Brasileira de Geologia, Porto Alegre, RS.
93

Andreoli, M.A.G., Hart, R.J., Ashwal, L.D., Coetzee, H., 2006. Correlations between U, Th
content and metamorphic grade in the Western Namaqualand belt, South Africa, with
Implications for radioactive heating of the crust. J. Petrol. 47, 1095–1118.
doi:10.1093/petrology/egl004
Araújo, O.J.B. de, Maia, R.G.N., Jorge João, X. da S., Costa, J.B.S., 1988. Megaestruturação
Arqueana da Folha Serra dos Carajás, in: SBG (Ed.), Anais Do VII Congresso Latino-
Americano de Geologia. SBG, Belém, PA, pp. 324–338.
Araújo, O.J.B., Maia, R.G.N., 1991. Programa Levantamento Geológicos Básicos do Brasil.
Programa Grande Carajás. Folha SB.22-Z-A. Estado do Pará. Texto Explicativo.
Brasilia.
Avelar, V.G., Lafon, J.M., Correia Jr., F.C., Macambira, E.M.B., 1999. O magmatismo
arqueano da região de Tucumã - Província Mineral De Carajás: Novos resultados
geocronológicos. Rev. Bras. Geociências 29, 453–460.
Bénard, A., Koga, K.T., Shimizu, N., Kendrick, M.A., Ionov, D.A., Nebel, O., Arculus, R.J.,
2017. Chlorine and fluorine partition coefficients and abundances in sub-arc mantle
xenoliths (Kamchatka, Russia): Implications for melt generation and volatile recycling
processes in subduction zones. Geochim. Cosmochim. Acta 199, 324–350.
doi:10.1016/j.gca.2016.10.035
Berman, R.G. 1991. Thermobarometry using multi-equilibrium calculations: A new
technique, with petrological applications. Canadian Mineralogist, 29 (4): 833-855.
Berman, R.G., 1988. Internally-Consistent Thermodynamic Data for Minerals in the System
Na2O-K2O-CaO-MgO-FeO-Fe2O3-Al2O3-SiO2-TiO2-H2O-CO2. J. Petrol. 29, 445–
522. doi:10.1093/petrology/29.2.445
Beyer, C., Klemme, S., Wiedenbeck, M., Stracke, A., Vollmer, C., 2012. Fluorine in
nominally fluorine-free mantle minerals: Experimental partitioning of F between olivine,
orthopyroxene and silicate melts with implications for magmatic processes. Earth Planet.
Sci. Lett. 337–338, 1–9. doi:10.1016/j.epsl.2012.05.003
Blundy, J.D., Holland, T.J.B., 1990. Calcic amphibole equilibria and a new amphibole-
plagioclase geothermometer. Contrib. to Mineral. Petrol. 104, 208–224.
doi:10.1007/BF00306444
Brown, M., 2005. Synergistic effects of melting and deformation: an example from the
Variscan belt, western France. Geol. Soc. London, Spec. Publ. 243, 205–226.
doi:10.1144/GSL.SP.2005.243.01.15
94

Brown, M., 2006. Duality of thermal regimes is the distinctive characteristics of plate
tectonics since the Neoarchean. Geology 34, 961–964. doi:10.1130/G22853A.1
Brown, M., 2008. Characteristic thermal regimes of plate tectonics and their metamorphic
imprint throughout Earth history: When did Earth first adopt a plate tectonics mode of
behavior. Spec. Pap. 440 When Did Plate Tectonics Begin Planet Earth? 2440, 97–128.
doi:10.1130/2008.2440(05)
Brown, M., 2009. Metamorphic patterns in orogenic systems and the geological record. Geol.
Soc. London, Spec. Publ. 318, 37–74. doi:10.1144/SP318.2
Brown, M., Korhonen, F.J., Siddoway, C.S., 2011. Organizing Melt Flow through the Crust.
Elements 7, 261–266. doi:10.2113/gselements.7.4.261
Chamberlain, C.P., Sonder, L.J., 1990. Heat-producing elements and the thermal and baric
patterns of metamorphic belts. Science (80-. ). 250, 763–769.
doi:10.1126/science.250.4982.763
Clark, C., Fitzsimons, I.C.W., Healy, D., Harley, S.L., 2011. How does the continental crust
get really hot? Elements 7, 235–240. doi:10.2113/gselements.7.4.235
Corfu, F., Hanchar, J.M., Hoskin, P.W.O., Kinny, P.D., 2003. Atlas of Zircon Textures. Rev.
Mineral. Geochemistry 53, 469–500. doi:10.2113/0530469
Costa, U.A.P., Paula, R.R., Silva, D.P.B., Barbosa, J.P.O., Silva, C.M.G., Tavares, F.M.,
Oliveira, J.K.M., Justo, A.P., 2016. Programa geologia do Brasil-PGB. Mapa de
integração geológico-geofísica da ARIM Carajás. Escala 1:250.000 Estado do Pará.
Belém, CPRM.
Dall’Agnol, R., Teixeira, N.P., Rämö, O.T., Moura, C.A. V, Macambira, M.J.B., de Oliveira,
D.C., 2005. Petrogenesis of the Paleoproterozoic rapakivi A-type granites of the Archean
Carajás metallogenic province, Brazil. Lithos 80, 101–129.
doi:10.1016/j.lithos.2004.03.058
Delinardo da Silva, M.A., 2014. Metatexitos e diatexitos do Complexo Xingu na região de
Canaã dos Carajás: Implicações para a Evolução Mesoarqueana do Domínio Carajás.
University of Campinas (UNICAMP).
Delinardo da Silva, M.A., Monteiro, L.V.S., Moreto, C.P.N., Sousa, S.D. de, 2015.
Metamorfismo e geoquímica do Complexo Xingu na região de Canaã dos Carajás :
implicações para a evolução mesoarqueana do domínio Carajás , Província Carajás, in:
Gorayeb, P., Menguins, A. (Eds.), Contribuições a Geologia Da Amazônia. SBG-NO,
Belém, PA, pp. 279–298.
95

DOCEGEO, 1988. Revisão litoestratigráfica da Província Mineral de Carajás, in: CVRD,


SBG (Eds.), Anais Do 35o Congresso Brasileiro de Geologia. Sociedade Brasileira de
Geologia (SBG), Belém, PA, pp. 11–54.
Feio, G.R.L., Dall’Agnol, R., Dantas, E.L., Macambira, M.J.B., Gomes, A.C.B., Sardinha,
A.S., Oliveira, D.C., Santos, R.D., Santos, P.A., 2012. Geochemistry, geochronology,
and origin of the Neoarchean Planalto Granite suite, Carajás, Amazonian craton: A-type
or hydrated charnockitic granites? Lithos 151, 57–73. doi:10.1016/j.lithos.2012.02.020
Feio, G.R.L., Dall’Agnol, R., Dantas, E.L., Macambira, M.J.B., Santos, J.O.S., Althoff, F.J.,
Soares, J.E.B., 2013. Archean granitoid magmatism in the Canaã dos Carajás area:
Implications for crustal evolution of the Carajás province, Amazonian craton, Brazil.
Precambrian Res. 227, 157–185. doi:10.1016/j.precamres.2012.04.007
Foley, S., 1991. High-pressure stability of the fluor- and hydroxy-endmembers of pargasite
and K-richterite. Geochim. Cosmochim. Acta 55, 2689–2694. doi:10.1016/0016-
7037(91)90386-J
Fuhrman, M.L., Lindsley, D.H., 1988. Ternary-feldspar modeling and thermometry. Am.
Mineral. 73, 201–215.
Gabriel, E.O., Oliveira, D.C. de, 2013. Petrologia magnética dos granodioritos Água Azul e
Água Limpa, porção sul do Domínio Carajás - Pará. Geol. USP. Série Científica 13, 89–
110. doi:10.5327/Z1519-874X201300040005
Gabriel, E.O., Oliveira, D.C., 2014. Geologia, petrografia e geoquímica dos granitoides
arqueanos de alto magnésio da região de Água Azul do Norte, porção sul do Domínio
Carajás, Pará. Bol. do Mus. Para. Emílio Goeldi - Série Ciências Nat. 9, 533–564.
Gabriel, E.O., Oliveira, D.C., Galarza, M.A., Santos, M.S. 2015. Geocronologia e aspectos
estruturais dos sanukitoids Mesoarqueanos da área de Água Azul do Norte: Implicações
para a história evolutiva da porção sul do Domínio Carajás. In: Simp. de Geol. da
Amazônia, 14, Marabá (PA), CD-Rom.
Geisler, T., Schaltegger, U., Tomaschek, F., 2007. Re-equilibration of zircon in aqueous
fluids and melts. Elements 3, 43–50. doi:10.2113/gselements.3.1.43
Genier, F., Bussy, F., Epard, J.L., Baumgartner, L., 2008. Water-assisted migmatization of
metagraywackes in a Variscan shear zone, Aiguilles-Rouges massif, western Alps.
Lithos 102, 575–597. doi:10.1016/j.lithos.2007.07.024
Gibbs, A.K., Wirth, K.R., Hirata, W.K., Olszewski Jr, W.J., 1986. Age and composition of the
Grão Pará groups volcanics, Serra dos Carajás. Rev. Bras. Geociências 16, 201–211.
96

Harley, S.L., Kelly, N.M., Möller, a, 2007. Zircon behaviour and the thermal history of
mountain belts. Elements 3, 25–30; DOI: 10.2113/gselements.3.1.25.
doi:10.2113/gselements.3.1.25
Hawthorne, F.C., Oberti, R., Harlow, G.E., Maresch, W. V., Martin, R.F., Schumacher, J.C.,
Welch, M.D., 2012. Ima report: Nomenclature of the amphibole supergroup. Am.
Mineral. 97, 2031–2048. doi:10.2138/am.2012.4276
Holdsworth, R.E., Pinheiro, R.V.L., 2000. The anatomy of shallow-crustal transpressional
structures: Insights from the Archaean Carajas fault zone, Amazon, Brazil. J. Struct.
Geol. 22, 1105–1123. doi:10.1016/S0191-8141(00)00036-5
Holland, T.J.B., Powell, R., 1985. An internally consistent thermodynamic dataset with
uncertainties and correlations: 2. Data and results. J. Metamorph. Geol. 3, 343–370.
doi:10.1111/j.1525-1314.1985.tb00325.x
Holland, T.J.B., Powell, R., 1990. An enlarged and updated internally consistent
thermodynamic dataset with uncertainties and correlations: the system K2O–Na2O–
CaO–MgO–MnO–FeO–Fe2O3–Al2O3–TiO2–SiO2–C–H2–O2. J. Metamorph. Geol. 8,
89–124. doi:10.1111/j.1525-1314.1990.tb00458.x
Holland, T.J.B., Powell, R., 1998. An internally consistent thermodynamic data set for phases
of petrological interest. J. Metamorph. Geol. 16, 309–343. doi:10.1111/j.1525-
1314.1998.00140.x
Holness, M.B., Cesare, B., Sawyer, E.W., 2011. Melted Rocks under the Microscope:
Microstructures and Their Interpretation. Elements 7, 247–252.
doi:10.2113/gselements.7.4.247
Johnson, E.A., 2006. Water in Nominally Anhydrous Crustal Minerals: Speciation,
Concentration, and Geologic Significance. Rev. Mineral. Geochemistry 62, 117–154.
doi:10.2138/rmg.2006.62.6
Johnson, T.E., Fischer, S., White, R.W., 2013. Field and petrographic evidence for partial
melting of TTG gneisses from the central region of the mainland Lewisian complex, NW
Scotland. J. Geol. Soc. London. 170, 319–326. doi:10.1144/jgs2012-096
Labrosse, S., Jaupart, C., 2007. Thermal evolution of the Earth: Secular changes and
fluctuations of plate characteristics. Earth Planet. Sci. Lett. 260, 465–481.
doi:10.1016/j.epsl.2007.05.046
97

Leite, A.A.S., Dall’Agnol, R., Macambira, M.J.B., Althoff, F.J., 2004. Geologia e
geocronologia dos granitóides arqueanos da região de Xinguara (PA) e suas implicac¸
ões na evoluc¸ ão do terreno granito-greenstone de Rio Maria. Revista Brasileira de
Geociências 34, 447–458.
Leite-Santos, P.J., 2016. Petrologia e geocronologia das associações leucograníticas
arqueanas da região de Água Azul do Norte (PA): implicações para a evolução crustal da
Província Carajás. University of Pará (UFPA).
Leite-Santos, P.J., Oliveira, D.C. de, 2014. Trondhjemitos da área de Nova Canadá : novas
ocorrências de associações magmáticas tipo TTG no Domínio Carajás. Bol. Mus. Para.
Emílio Goeldi. Cienc. Nat 9, 635–659.
Ludwig, K.R., 2003. Mathematical-Statistical Treatment of Data and Errors for 230Th/U
Geochronology. Rev. Mineral. Geochemistry 52, 631–656. doi:10.2113/0520631
Machado, N., Lindenmayer, Z., Krogh, T.E., Lindenmayer, D., 1991. U-Pb geochronology of
Archean magmatism and basement reactivation in the Carajás area, Amazon shield,
Brazil. Precambrian Res. 49, 329–354. doi:10.1016/0301-9268(91)90040-H
Marangoanha, B., 2016. Petrologia de Granitoides e Implicações para a Evolução Crustal
Arqueana da Porção Oeste do Domínio Canaã dos Carajás, Província Carajás. University
of Pará (UFPA).
Molina, J.F., Moreno, J.A., Castro, A., Rodríguez, C., Fershtater, G.B., 2015. Calcic
amphibole thermobarometry in metamorphic and igneous rocks: New calibrations based
on plagioclase/amphibole Al-Si partitioning and amphibole/liquid Mg partitioning.
Lithos 232, 286–305. doi:10.1016/j.lithos.2015.06.027
Moreto, C.P.N., 2013. Geocronologia U-Pb e Re-Os aplicada à evolução metalógenética do
Cinturão Sul do Cobre da Província Mineral de Carajás. University of Campinas
(UNICAMP).
Moreto, C.P.N., Monteiro, L.V.S., Xavier, R.P., Amaral, W.S., dos Santos, T.J.S., Juliani, C.,
de Souza Filho, C.R., 2011. Mesoarchean (3.0 and 2.86 Ga) host rocks of the iron oxide–
Cu–Au Bacaba deposit, Carajás Mineral Province: U–Pb geochronology and
metallogenetic implications. Miner. Depos. 46, 789–811. doi:10.1007/s00126-011-0352-
9
98

Moreto, C.P.N., Monteiro, L.V.S., Xavier, R.P., Creaser, R.A., DuFrane, S.A., Melo, G.H.C.,
Delinardo da Silva, M.A., Tassinari, C.C.G., Sato, K., 2015. Timing of multiple
hydrothermal events in the iron oxide–copper–gold deposits of the Southern Copper
Belt, Carajás Province, Brazil. Miner. Depos. 50, 517–546. doi:10.1007/s00126-014-
0549-9
Moreto, C.P.N., Monteiro, L.V.S., Xavier, R.P., Creaser, R.A., DuFrane, S.A., Tassinari,
C.C.G., Sato, K., Kemp, A.I.S., Amaral, W.S., 2015. Neoarchean and Paleoproterozoic
Iron Oxide-Copper-Gold Events at the Sossego Deposit, Carajas Province, Brazil: Re-Os
and U-Pb Geochronological Evidence. Econ. Geol. 110, 809–835.
doi:10.2113/econgeo.110.3.809
Morfin, S., Sawyer, E.W., Bandyayera, D., 2013. Large volumes of anatectic melt retained in
granulite facies migmatites: An injection complex in northern Quebec. Lithos 168–169,
200–218. doi:10.1016/j.lithos.2013.02.007
O’Brien, P.J., Rötzler, J., 2003. High-pressure granulites: Formation, recovery of peak
conditions and implications for tectonics. J. Metamorph. Geol. 21, 3–20.
doi:10.1046/j.1525-1314.2003.00420.x
Passchier, C.W., Trouw, R. a J., 2005. Microtectonics. Springer-Verlag, Berlin/Heidelberg.
doi:10.1007/3-540-29359-0
Paton, C., Woodhead, J.D., Hellstrom, J.C., Hergt, J.M., Greig, A., Maas, R., 2010. Improved
laser ablation U-Pb zircon geochronology through robust downhole fractionation
correction. Geochemistry, Geophys. Geosystems 11, n/a-n/a.
doi:10.1029/2009GC002618
Pawley, M., Reid, A., Dutch, R., Preiss, W., 2015. Demystifying migmatites:
introduction for field-based geologist. Appl. Earth Sci. 124, 147–174.
doi:10.1179/1743275815Y.0000000014
Perchuk, L.L., Gerya, T. V., 2011. Formation and evolution of Precambrian granulite terranes:
A gravitational redistribution model. Geol. Soc. Am. Mem. 289–310.
doi:10.1130/2011.1207(15)
Petrus, J.A., Kamber, B.S., 2012. VizualAge: A Novel Approach to Laser Ablation ICP-MS
U-Pb Geochronology Data Reduction. Geostand. Geoanalytical Res. 36, 247–270.
doi:10.1111/j.1751-908X.2012.00158.x
99

Pidgeon, R.T., Macambira, M.J.B., Lafon, J.-M., 2000. Th–U–Pb isotopic systems and
internal structures of complex zircons from an enderbite from the Pium Complex,
Carajás Province, Brazil: evidence for the ages of granulite facies metamorphism and the
protolith of the enderbite. Chem. Geol. 159–171.
Pimentel, M. M., Machado, N. 1994. Geocronologia U-Pb dos terrenos granito-greenstone de
Rio Maria, Pará. Boletim de Resumos Expandidos do Congresso Brasileiro de Geologia
38 (2): 390-391
Pinheiro, R.V.L., Kadekaru, K., Soares, A.V., Freitas, C., Ferreira, S.N., Matos, F.M.V.,
2013. Carajás, Brazil - a short tectonic review, in: XIII Simpósio de Geologia Da
Amazônia. SBG-NO, Belém, PA, pp. 1086–1089.
Powell, R., Holland, T., 1993. The applicability of least squares in the extraction of
thermodynamic data from experimentally bracketed mineral equilibra. Am. Mineral. 78,
107–112.
Powell, R., Holland, T.J.B., 1985. An internally consistent thermodynamic dataset with
uncertainties and correlations: 1. Methods and a worked example. J. Metamorph. Geol.
3, 327–342. doi:10.1111/j.1525-1314.1985.tb00324.x
Ricci, P.S.F., Carvalho, M.A. 2006. Rocks of the Pium-Area, Carajás Block, Brazil – A deep
seated high-T gabbroic pluton (charnockitoid-like) with xenoliths of enderbitic gneisses
dated at 3002 Ma – the basement problem revisited. In: 8º Simp. Geol. Amaz., Manaus,
Proceedings [CD-ROM].
Rietmeijer, F.J.M., 1983. Chemical Distinction Between Igneous and Metamorphic
Orthopyroxenes Especially Those Coexisting with Ca-rich Clinopyroxenes: A Re-
Evaluation. Mineral. Mag. 47, 143–151. doi:10.1180/minmag.1983.047.343.04
Rodrigues, D.S., Oliveira, D.C., Macambira, M.J.B., 2014. Geologia, geoquímica e
geocronologia do Granito Mesoarqueano Boa Sorte, município de Água Azul do Norte,
Pará – Província Carajás. Bol. do Mus. Para. Emílio Goeldi - Série Ciências Nat. 9, 597–
633.
Rubatto, D., 2017. Zircon: The Metamorphic Mineral, Reviews in Mineralogy and
Geochemistry. doi:10.2138/rmg.2017.83.10
Sajeev, K., Osanai, Y., N, Y.K., Itaya, T., 2009. Stability of pargasite during ultrahigh-
temperature metamorphism: A consequence of titanium and REE partitioning? Am.
Mineral. 94, 535–545. doi:10.2138/am.2009.2815
100

Sajeev, K., Windley, B.F., Hegner, E., Komiya, T., 2013. High-temperature, high-pressure
granulites (retrogressed eclogites) in the central region of the Lewisian, NW Scotland:
Crustal-scale subduction in the Neoarchaean. Gondwana Res. 23, 526–538.
doi:10.1016/j.gr.2012.05.002
Santos, J.O.S., 2003. Geotectônica dos Escudos das Guianas e Brasil-Central, in: Bizzi, L.A.,
Schobbenhaus, C., Vidotti, R.M., Gonçalves, J.H. (Eds.), Geologia, Tectônica e
Recursos Minerais Do Brasil. CPRM - Serviço Geológico do Brasil, Brasilia, pp. 169–
195.
Santos, J.O.S., Hartmann, L.A., Gaudette, H.E., Groves, D.I., Mcnaughton, N.J., Fletcher,
I.R., 2000. A New Understanding of the Provinces of the Amazon Craton Based on
Integration of Field Mapping and U-Pb and Sm-Nd Geochronology. Gondwana Res. 3,
453–488.
Santos, M. da S., 2016. Granitoides TTG de Água Azul do norte: Implicações tectônicas para
a Província Carajás. University of Para (UFPA).
Santos, P.A. dos, Feio, G.R.L., Dall’Agnol, R., Costi, H.T., Lamarão, C.N., Galarza, M.A.,
2013b. Petrography, magnetic susceptibility and geochemistry of the Rio Branco
Granite, Carajás Province, southeast of Pará, Brazil. Brazilian J. Geol. 43, 2–15.
doi:10.5327/Z2317-48892013000100002
Santos, R.D., Galarza, M.A., Oliveira, D.C. de, 2013a. Geologia , geoquímica e
geocronologia do Diopsídio-Norito Pium, Província Carajás. Bol. Mus. Para. Emílio
Goeldi. Cienc. Nat. 8, 355–382.
Sawyer, E.W., 2008. Atlas of Migmatites. NRC Research Press and Mineralogical
Association of Canada. doi:10.1139/9780660197876
Sawyer, E.W., 2010. Migmatites formed by water-fluxed partial melting of a
leucogranodiorite protolith: Microstructures in the residual rocks and source of the fluid.
Lithos 116, 273–286. doi:10.1016/j.lithos.2009.07.003
Sawyer, E.W., Cesare, B., Brown, M., 2011. When the Continental Crust Melts. Elements 7,
229–234. doi:10.2113/gselements.7.4.229
Silva, A.C., Dall’Agnol, R., Guimarães, F.V., Oliveira, D.C., 2014. Geologia, petrografia e
geoquímica de Associações Tonalíticas e Trondhjemíticas Arqueanas de Vila Jussara,
Província Carajás, Pará. Bol. do Mus. Para. Emílio Goeldi - Série Ciências Nat. 9, 13–
45.
101

Silva, G.G. da, Lima, M.I.C. de, Andrade, A.R.F. de, Issler, R.S., Guimarães, G., 1974.
Geologia da Folha SB.22 Araguaia e Parte da Folha SC.22 Tocantins, in: Luz, A.A. da,
Almeida, A.L.S. de, Netto, O.B. (Eds.), Projeto Radam: Levantamento de Recursos
Naturais, Folha SB.22 Araguaia e Parte Da Folha SC.22 Tocantins, Volume 4. Rio de
Janeiro, pp. 1–177.
Silva, L.R., de Oliveira, D.C., dos Santos, M.N.S., 2018. Diversity, origin and tectonic
significance of the Mesoarchean granitoids of Ourilândia do Norte, Carajás province
(Brazil). J. South Am. Earth Sci. 82, 33–61. doi:10.1016/j.jsames.2017.12.004
Silva, M. L. T., Oliveira, D. C., Macambira, M. J. B. 2010. Geologia, petrografia e
geocronologia do magmatismo de alto K da região de Vila Jussara, Água Azul do Norte
– Província Mineral de Carajás. Congresso Brasileiro de Geologia 45: 1 CD-ROM
Sizova, E., Gerya, T., Brown, M., Perchuk, L.L., 2010. Subduction styles in the Precambrian:
Insight from numerical experiments. Lithos 116, 209–229.
doi:10.1016/j.lithos.2009.05.028
Sousa, S. D., Oliveira. D. C., Gabriel E. O., Macambira, M. J. B. 2010. Geologia, petrografia
e geocronologia das rochas granitoides do Complexo Xingu da porção a leste da cidade
de Água Azul do Norte (PA) – Província Mineral de Carajás. Congresso Brasileiro de
Geologia 45: 1 CD-ROM.
Sousa, S.D. de, Monteiro, L.V.S., Oliveira, D.C. de, Delinardo da Silva, M.A., Moreto,
C.P.N., Juliani, C., 2015. O Greenstone Belt Sapucaia na região de Água Azul do Norte,
Província Carajás: Contexto geológico e caracterização petrográfica e geoquímica, in:
Gorayeb, P., Menguins, A. (Eds.), Contribuições a Geologia Da Amazônia. SBG-NO,
Belem, Pará, pp. 317–338.
Stevens, G., Clemens, J.D., 1993. Fluid-absent melting and the roles of fluids in the
lithosphere: a slanted summary? Chem. Geol. 108, 1–17. doi:10.1016/0009-
2541(93)90314-9
Tsunogae, T., Osanai, Y., Owada, M., Toyoshima, T., Hokada, T., Crowe, W.A., 2003. High
fluorine pargasites in ultrahigh temperature granulites from Tonagh Island in the
Archean Napier Complex, East Antarctica. Lithos 70, 21–38. doi:10.1016/S0024-
4937(03)00087-2
Urann, B.M., Le Roux, V., Hammond, K., Marschall, H.R., Lee, C.T.A., Monteleone, B.D.,
2017. Fluorine and chlorine in mantle minerals and the halogen budget of the Earth’s
mantle. Contrib. to Mineral. Petrol. 172, 1–17. doi:10.1007/s00410-017-1368-7
102

Vasquez, M.L., Rosa-Costa, L.T., 2008. Geologia e Recursos Minerais do Estado do Pará :
Sistema de Informações Geográficas - SIG: texto explicativo dos mapas Geológico e
Tectônico e de Recursos Minerais do Estado do Pará. Belém.
Vernon, R.H., 1999. Quartz and feldspar microstructures in metamorphic rocks. Canedian
Mineral. 37, 513–524.
Vernon, R.H., 2004. A Practical Guide to Rock Microstructure. Cambridge University Press,
Cambridge. doi:10.1017/CBO9780511807206
Vernon, R.H., 2011. Microstructures of melt-bearing regional metamorphic rocks 1207, 1–11.
doi:10.1130/2011.1207(01)
Vielzeuf, D., Schmidt, M.W., 2001. Melting relations in hydrous systems revisited:
application to metapelites, metagreywackes and metabasalts. Contrib. to Mineral. Petrol.
141, 251–267. doi:10.1007/s004100100237
Villaseca, C., Orejana, D., Paterson, B.A., 2007. Zr-LREE rich minerals in residual
peraluminous granulites, another factor in the origin of low Zr-LREE granitic melts?
Lithos 96, 375–386. doi:10.1016/j.lithos.2006.11.002
Whitney, D.L., Evans, B.W., 2010. Abbreviations for names of rock-forming minerals. Am.
Mineral. 95, 185–187. doi:10.2138/am.2010.3371
WIEDENBECK, M., ALLÉ, P., CORFU, F., GRIFFIN, W.L., MEIER, M., OBERLI, F.,
QUADT, A. VON, RODDICK, J.C., SPIEGEL, W., 1995. THREE NATURAL
ZIRCON STANDARDS FOR U-TH-PB, LU-HF, TRACE ELEMENT AND REE
ANALYSES. Geostand. Geoanalytical Res. 19, 1–23. doi:10.1111/j.1751-
908X.1995.tb00147.x
Williams I.S., 1998, U-Th-Pb geochronology by ion microprobe. In: McKibben, M.A.
Shanks, W.C. and Ridley W.I. (eds) Applications of Microanalytical Techniques to
Understanding Mineralizing Processes. Reviews in Economic Geology 7: 1-35.
Wirth, K.R., Gibbs, A.K., Olszewski Jr., W.J., 1986. U-Pb ages of zircons from the Grão-Pará
Group and Serra dos Carajás Granites, Pará, Brazil. Rev. Bras. Geociências 16, 195–200.
103

Appedix
Supplementary Table 1. Summary of U-Pb LA-ICPMS zircon data from the Xicrim-Cateté Orthogranulite and Xingu Complex.

Isotopic Ratios Ages

206 207 207 206 207 207 206


Zircon Th U Pb/ Pb/ Pb/ ± Pb/ ± Pb/ Pb/ Pb/ %
Th/U ± ± ± ±
204 206 235 238 206 235 238
Grains (ppm) (ppm) Pb Pb U % U % Pb U U conc

Xicrim-Cateté Orthogranulite
Enderbitic gneiss (XA15A1)

ZR 3 0.54 90 165 72300000 0.2092 0.0025 16.70 2.095808 0.5758 1.684613 2881 28 2910 21 2922 40 101.4

ZR 34.2 0.37 45 122 50050000 0.2071 0.0030 16.09 2.734618 0.5660 1.766784 2859 40 2865 29 2891 41 101.1

ZR 9 0.74 140 190 85650000 0.2131 0.0025 17.10 1.812865 0.5730 1.483421 2930 25 2936 18 2917 34 99.6

ZR 24.1 1.13 97 86 42850 0.2099 0.0031 16.63 2.826218 0.5740 1.916376 2879 38 2904 29 2920 45 101.4

ZR 28 0.45 108 240 105000 0.2108 0.0024 17.44 1.548165 0.5905 1.30398 2907 25 2956 15 2988 31 102.8

ZR 32 0.39 57 145 62450000 0.2161 0.0028 17.39 1.955147 0.5845 1.556886 2919 28 2957 19 2968 37 101.7
Mafic granulite (XA15B3)

ZR 2.2 0.4 9 23 3815 0.1930 0.0061 15.93 3.26 0.588 3.40 2792 31 2851 30 2959 80 106.0

ZR 3.1 2.1 116 56 8870 0.2041 0.0051 16.39 2.38 0.569 2.81 2882 25 2900 22 2884 64 100.1

ZR 5.1 2.6 105 41 6415 0.2072 0.0052 16.52 2.36 0.561 2.67 2907 23 2893 22 2848 62 98.0

ZR 6.2 1.5 104 67 11075 0.2021 0.0041 16.54 2.06 0.579 2.42 2872 19 2900 20 2942 55 102.4
104

Isotopic Ratios Ages

206 207 207 206 207 207 206


Zircon Th U Pb/ Pb/ Pb/ ± Pb/ ± Pb/ Pb/ Pb/ %
Th/U ± ± ± ±
204 206 235 238 206 235 238
Grains (ppm) (ppm) Pb Pb U % U % Pb U U conc

ZR 9.2 2.8 148 52 8365 0.2052 0.0046 16.32 2.33 0.565 2.65 2907 22 2893 23 2870 60 98.7

ZR 11.1 0.7 59 80 12900 0.2035 0.0044 16.49 2.18 0.578 2.42 2877 22 2898 21 2911 58 101.2

ZR 3.2 1.9 80 43 6755 0.2027 0.0052 16.16 2.60 0.563 2.66 2888 25 2870 25 2860 62 99.0

ZR 3.3 2.2 137 61 9880 0.1960 0.0043 16.14 2.23 0.582 2.58 2827 21 2875 21 2934 58 103.8

ZR 5.2 2.1 78 38 5845 0.2096 0.0057 15.95 2.45 0.553 2.89 2920 25 2862 23 2819 64 96.5

ZR 9.1 2.8 148 53 8300 0.2043 0.0046 15.88 2.33 0.558 2.51 2882 21 2863 22 2834 57 98.3
Leucosome of a metatexitic mafic granulite (XS112A)

ZR 1.2 0.39 119.3 305 128500 0.2048 0.0019 16.08 1.49 0.5648 1.29 2860 20 2876 15 2882 30 100.8

ZR 7.2 1.36 384.3 282 119100000 0.2042 0.0019 16.00 1.31 0.5646 1.19 2844 17 2877 12 2884 27 101.4

ZR 9.2 1.09 210 191.8 81400000 0.2021 0.0018 15.95 1.44 0.5743 1.13 2825 20 2873 14 2925 26 103.5

ZR 12.1 0.69 195.6 283.4 132550000 0.2001 0.0015 16.25 1.17 0.5881 1.05 2825 15 2890 12 2979 25 105.5

ZR 13 0.04 26.72 616 175000 0.1970 0.0018 10.04 1.79 0.3699 1.46 2781 16 2435 17 2024 25 72.8

ZR 14 0.53 139.3 264 91500000 0.1937 0.0020 11.88 1.77 0.4431 1.40 2741 23 2589 17 2362 28 86.2

ZR 15.2 0.01 6.05 612 103300000 0.1621 0.0019 5.05 2.77 0.2287 2.01 2457 26 1818 23 1323 24 53.8

Xingu Complex
Granodioritic gneiss (XS04P)
105

Isotopic Ratios Ages

206 207 207 206 207 207 206


Zircon Th U Pb/ Pb/ Pb/ ± Pb/ ± Pb/ Pb/ Pb/ %
Th/U ± ± ± ±
204 206 235 238 206 235 238
Grains (ppm) (ppm) Pb Pb U % U % Pb U U conc

ZR 19 0.42 37 86 43400 0.2129 0.0028 17.15 2.16 0.5742 1.50 2923 35 2933 22 2919 35 99.9

ZR 21 0.70 87 124 62000000 0.2114 0.0024 16.90 2.07 0.5679 1.48 2924 29 2929 20 2893 35 98.9

ZR 3 0.16 40 253 134500 0.2142 0.0019 17.26 1.51 0.5842 1.22 2922 22 2945 15 2962 29 101.4

ZR 24 0.45 34 76 40200 0.2137 0.0027 17.09 2.69 0.5753 1.69 2927 35 2936 27 2927 39 100.0

ZR 25 0.59 72 121 59300000 0.2134 0.0026 16.67 1.98 0.5676 1.46 2910 29 2910 20 2894 34 99.5

ZR 27 0.41 47 114 56500000 0.2152 0.0023 17.39 2.01 0.5760 1.41 2937 28 2946 20 2930 33 99.8

ZR 29 0.46 45 98 48450 0.2104 0.0031 16.52 2.66 0.5720 1.75 2858 42 2902 27 2907 43 101.7

ZR 30 0.46 36 78 39950 0.2129 0.003 16.91 2.78 0.5680 1.74 2923 39 2911 29 2900 40 99.2

ZR 33 0.64 99 154 72550000 0.2136 0.003 17.22 1.92 0.5750 1.74 2928 29 2941 19 2920 41 99.7

ZR 38 0.72 106 148 71850000 0.2118 0.0022 16.65 1.62 0.5685 1.41 2901 23 2915 16 2904 32 100.1
106

Supplementary Table 2. Summary of U-Pb SHRIMP IIe zircon data from the Xicrim-Cateté Orthogranulite and Xingu Complex.

Isotopic Ratios Ages

204 207 207 206 207 206


Zircon Th U Pb/ Pb/ Pb/ ± Pb/ ± Pb/ Pb/
%
Th/U ± ± ±
206 206 235 238 206 238 disc
Grains (ppm) (ppm) Pb Pb U % U % Pb U

Xicrim-Cateté Orthogranulite
Enderbitic Gneiss (XA15A1) - Core

ZR 1.1 0.44 35 83 0.00048 0.2216 0.81 17.20 2.7 0.5790 2.6 2951 14 2943 60 0

ZR 3.2 0.33 47 147 0.00009 0.2161 0.32 16.20 2.7 0.5460 2.7 2944 5 2807 61 6

ZR 5.4 0.98 98 103 0.00014 0.2159 0.97 16.60 2.8 0.5610 2.6 2938 16 2869 60 3

ZR 4.1 0.51 63 126 0.00014 0.2151 0.72 16.10 2.6 0.5470 2.5 2932 12 2813 58 5

ZR 8.1 1.49 401 278 0.00036 0.1801 0.34 6.20 2.5 0.2580 2.5 2611 8 1477 33 48
Enderbitic Gneiss (XA15A1) – Rim

ZR 5.1 2.93 54 19 0.00040 0.2035 0.87 14.10 3.2 0.5140 3 2816 18 2672 65 6

ZR 5.2 3.08 75 25 0.00017 0.2014 0.9 14.50 3.1 0.5270 2.9 2821 17 2727 66 4

ZR 6.1 2.32 87 39 0.00044 0.2089 0.61 15.50 4.2 0.5530 4.2 2857 13 2839 96 1

ZR 5.3 3.18 85 28 0.00028 0.2091 0.81 16.30 3.1 0.5750 2.9 2874 16 2929 69 -2

ZR 3.1 1.87 29 16 0.00093 0.2181 0.93 16.70 3.4 0.5830 3.1 2884 24 2960 74 -3
Opdalitic Gneiss (SM46N) - Core

ZR 8.1 0.32 67 217 0.000013 0.2051 0.9159 16.10 1.6 .5697 1.3 2866 15 2906.5 30.0 -1

ZR 12.1 0.30 46 161 0.000093 0.2119 0.5499 16.10 1.5 .5539 1.3 2911 10 2841.2 30.6 2
107

Isotopic Ratios Ages

204 207 207 206 207 206


Zircon Th U Pb/ Pb/ Pb/ ± Pb/ ± Pb/ Pb/
%
Th/U ± ± ±
206 206 235 238 206 238 disc
Grains (ppm) (ppm) Pb Pb U % U % Pb U

ZR 6.1 0.55 303 574 0.000073 0.1965 0.5096 8.48 1.3 .3142 1.2 2790 9 1761.6 19.1 58

ZR 3.1 0.56 304 563 0.000040 0.2034 0.3251 11.16 1.3 .3987 1.2 2850 6 2163.1 22.5 32

ZR 9.1 0.24 82 355 0.000091 0.2071 0.5045 11.80 1.4 .4155 1.3 2874 9 2239.9 25.4 28

ZR 2.1 0.10 29 319 0.000029 0.2094 0.4374 14.44 1.5 .5010 1.4 2898 7 2618.1 30.8 11

ZR 11.1 0.16 48 307 0.000017 0.2156 0.4159 15.59 1.3 .5249 1.3 2947 7 2719.8 28.1 8

ZR 10.1 0.08 12 144 0.000116 0.2164 0.5694 15.65 1.5 .5276 1.3 2944 10 2731.4 29.8 8

ZR 5.1 0.61 111 188 0.000041 0.2178 0.4830 16.22 1.4 .5412 1.3 2961 8 2788.3 29.3 6

ZR 1.1 0.18 29 167 0.000033 0.2163 0.5232 16.92 1.4 .5683 1.3 2951 9 2900.9 31.0 2

ZR 7.1 0.29 35 125 0.000066 0.2185 0.5742 18.16 1.5 .6049 1.4 2964 10 3049.5 33.0 -3
Opdalitic gneiss (SM46N) - Rim

ZR 1.2 0.84 189 232 0.000026 0.2019 0.4566 15.95 1.6 .5738 1.5 2839 8 2923.5 35.6 -3

ZR 4.1 0.42 59 145 0.000079 0.1962 0.6119 13.97 1.5 .5190 1.3 2787 10 2694.8 29.6 3

ZR 9.2 0.41 58 149 0.000226 0.2039 0.7938 16.02 1.7 .5774 1.4 2836 15 2938.2 34.2 -3
Opdalitic gneiss (CMS22P) – Xenocrystic cores

ZR 10.1 0.68 151 230 0.000035 0.2303 0.9807 19.13 1.6 .6035 1.3 3052 16 3043.8 31.0 0

ZR 7.1 0.72 92 132 0.000055 0.2376 0.9642 19.83 1.9 .6069 1.6 3100 16 3057.5 38.8 1
108

Isotopic Ratios Ages

204 207 207 206 207 206


Zircon Th U Pb/ Pb/ Pb/ ± Pb/ ± Pb/ Pb/
%
Th/U ± ± ±
206 206 235 238 206 238 disc
Grains (ppm) (ppm) Pb Pb U % U % Pb U

ZR 12.1 0.89 166 193 0.000021 0.2326 0.6656 19.47 1.6 .6066 1.4 3072 11 3056.3 34.6 1

ZR 5.1 0.81 75 95 0.000054 0.2311 0.6342 19.62 1.6 .6173 1.4 3056 10 3099.0 35.2 -1
Opdalitic gneiss (CMS22P)

ZR 1.1 0.25 42 176 0.000023 0.2055 0.5624 15.64 1.4 .5528 1.3 2868 9 2836.6 30.5 1

ZR 11.1 0.11 43 414 0.000191 0.1933 0.6639 11.58 1.4 .4399 1.2 2750 12 2350.3 24.4 17

ZR 9.1 0.07 52 788 0.000108 0.1891 0.9460 7.73 1.7 .2986 1.4 2723 16 1684.2 20.6 62

ZR 2.1 0.01 9 754 0.000035 0.2121 0.2913 14.41 1.2 .4937 1.2 2919 5 2586.9 25.8 13

ZR 3.1 0.39 95 248 0.000030 0.2156 0.4478 16.43 1.4 .5536 1.3 2945 8 2840.1 30.2 4

ZR 8.1 0.03 10 362 0.000038 0.2180 0.3668 17.64 1.3 .5880 1.3 2963 6 2981.5 30.0 -1

ZR 6.1 0.66 125 196 0.000029 0.2172 0.4576 17.34 1.4 .5779 1.3 2963 8 2940.1 30.6 1
Enderbitic gneiss (TS23A)

ZR 9.1 0.25 78 319 0.0021 0.2422 0.7800 11.2 1.8 0.372 1.4 2962 19 2041 25 36

ZR 2.1 0.13 52 423 0.0005 0.1996 0.3000 10.1 1.5 0.379 1.4 2770 11 2074 24 29

ZR 10.1 0.18 61 344 0.0004 0.2047 0.2300 11.9 1.0 0.434 0.9 2822 8 2322 17 21

ZR 7.1 0.41 132 329 0.0002 0.2186 0.2100 14.6 0.9 0.489 0.9 2949 4 2568 18 16

ZR 1.1 0.31 88 292 0.0002 0.2208 0.2200 15.1 1.6 0.503 1.5 2967 5 2625 33 14
109

Isotopic Ratios Ages

204 207 207 206 207 206


Zircon Th U Pb/ Pb/ Pb/ ± Pb/ ± Pb/ Pb/
%
Th/U ± ± ±
206 206 235 238 206 238 disc
Grains (ppm) (ppm) Pb Pb U % U % Pb U

ZR 8.1 0.67 287 442 0.0000 0.2192 0.1700 15.6 0.9 0.516 0.9 2971 3 2684 19 12

ZR 6.1 0.39 89 235 0.0002 0.2193 0.2500 15.7 1.0 0.526 0.9 2960 5 2723 20 10

ZR 8.2 0.57 155 281 0.0001 0.2233 0.2000 16.5 0.9 0.539 0.9 2997 4 2778 20 9

ZR 4.1 0.28 72 266 0.0001 0.2208 0.3300 17.5 1.0 0.577 0.9 2979 5 2937 21 2

ZR 5.1 0.34 76 229 0.0001 0.2215 0.2200 17.6 1.3 0.578 1.3 2987 4 2941 31 2

ZR 3.1 0.30 84 288 0.0001 0.2203 0.3800 17.5 1.2 0.578 1.2 2979 6 2940 27 2
Enderbitic gneiss (XA12A)

ZR 9.1 0.03 17 559 0.00011 0.1802 0.9700 7.00 1.9 0.285 1.7 2642 16 1617 24 44

ZR 2.1 0.35 64 187 0.00008 0.2057 0.6400 14.80 1.6 0.526 1.5 2865 11 2723 33 6

ZR 1.2 0.30 30 105 0.00011 0.2150 0.3200 16.40 1.6 0.556 1.6 2934 6 2851 37 3

ZR 7.1 0.48 56 122 0.00014 0.2214 0.5800 17.10 1.2 0.563 1.1 2979 10 2880 24 4

ZR 4.1 0.30 18 61 0.00025 0.2246 0.7800 17.70 1.5 0.579 1.2 2993 13 2944 28 2

ZR 3.1 0.39 26 68 0.00027 0.2230 1.1400 17.60 1.7 0.579 1.2 2980 19 2946 28 1

ZR 5.1 0.36 29 83 0.00017 0.2172 0.3500 17.30 1.2 0.583 1.1 2945 7 2961 27 -1

ZR 6.1 0.72 103 147 0.00006 0.2184 0.2700 17.90 1.0 0.595 1.0 2964 5 3010 24 -2

ZR 1.1 0.33 21 67 0.00004 0.2218 0.4000 18.40 1.3 0.601 1.2 2997 7 3034 30 -2
110

Isotopic Ratios Ages

204 207 207 206 207 206


Zircon Th U Pb/ Pb/ Pb/ ± Pb/ ± Pb/ Pb/
%
Th/U ± ± ±
206 206 235 238 206 238 disc
Grains (ppm) (ppm) Pb Pb U % U % Pb U
Leucossome of a metatexitic mafic granulite (XS112A)

ZR 11.1 0.02 20 869 0.00005 0.1705 0.8300 6.3 3.5 0.267 3.4 2557 ±14 1525 ±46 45

ZR 8.1 0.46 235 531 0.00010 0.1705 0.2400 6.6 2.5 0.284 2.5 2550 ±5 1611 ±35 41

ZR 9.1 0.11 72 671 0.00004 0.1777 0.2100 7.4 2.5 0.304 2.5 2627 ±4 1709 ±37 40

ZR 4.1 0.05 26 501 0.00022 0.1865 0.7700 7.9 3.9 0.314 3.8 2687 ±14 1759 ±58 39

ZR 12.1 0.15 141 960 0.00003 0.1819 0.3800 8 2.5 0.319 2.5 2667 ±6 1785 ±39 38

ZR 5.1 0.1 47 500 0.00011 0.1893 0.2500 8.4 2.6 0.325 2.6 2724 ±5 1814 ±41 38

ZR 7.1 0.09 32 359 0.00007 0.2038 0.2200 13.3 2.5 0.475 2.5 2851 ±4 2506 ±51 15

ZR 10.1 0.64 258 419 0.00005 0.2121 0.4600 15.5 2.7 0.533 2.6 2917 ±7 2755 ±59 7

ZR 6.1 0.34 64 196 0.00003 0.2085 0.5100 16.8 2.8 0.584 2.8 2891 ±8 2965 ±65 -3

Xingu Complex

Tonalitic gneiss (XS105P)

ZR 10.1 0.46 180 401 0.00024 0.1896 0.2900 7.2 2.6 0.28 2.6 2713 6 1592 37 46

ZR 6.1 0.5 102 212 0.00027 0.2056 0.5700 11.6 2.6 0.417 2.5 2846 10 2248 47 25

ZR 1.1 0.31 148 488 0.00003 0.2043 0.2000 12.9 2.9 0.46 2.9 2858 3 2441 58 17

ZR 3.1 0.34 61 186 0.00010 0.2093 0.3000 13.5 2.5 0.469 2.5 2891 5 2479 52 17
111

Isotopic Ratios Ages

204 207 207 206 207 206


Zircon Th U Pb/ Pb/ Pb/ ± Pb/ ± Pb/ Pb/
%
Th/U ± ± ±
206 206 235 238 206 238 disc
Grains (ppm) (ppm) Pb Pb U % U % Pb U

ZR 12.1 0.25 70 284 0.00004 0.2115 0.2400 15.7 2.5 0.539 2.5 2914 4 2780 56 6

ZR 9.1 0.4 44 114 0.00013 0.2129 0.6100 16.8 2.6 0.578 2.6 2916 10 2939 60 -1

ZR 5.1 0.36 57 160 0.00020 0.2137 0.3300 15.2 2.5 0.522 2.5 2916 6 2709 56 9

ZR 11.1 0.43 80 191 0.00008 0.2127 0.3100 15.5 2.6 0.53 2.5 2920 5 2740 56 8

ZR 4.1 0.37 71 198 0.00008 0.2134 0.2700 14.8 2.5 0.507 2.5 2924 5 2644 54 12

ZR 8.1 0.36 51 146 0.00012 0.2146 0.3300 16.9 2.6 0.573 2.5 2929 6 2922 59 0

ZR 7.1 0.54 48 91 0.00020 0.2171 0.3900 16.9 2.6 0.57 2.6 2942 7 2908 60 1

ZR 2.1 0.22 60 279 0.00003 0.2158 0.2300 17.6 2.5 0.594 2.5 2947 4 3006 60 -3
112

Anexo 02
Artigo:
Mesoarchean arc-related system between 3.0 – 2.9 Ga in the Carajás Domain, Carajás
Province, PA

Delinardo da Silva et al.


113

Abstract

The arc-related system in the Carajás Domain was responsible for TTG magmatism
(ca. 3.06–2.93 Ga) associated with high-pressure slab melting and for volcanic arc mafic
magmatism related to melting of hydrated suprasubduction mantle. The TTG rocks: (i) are
silica-rich rocks (65.58 < wt% SiO2 < 72.90), (ii) have high Na2O contents (4.36 – 6.10
wt%), (iii) are poor in ferromagnesian elements (FeOt + MgO + MnO + TiO2 < 4wt.%), (iv)
have Nb and Ta negative anomalies, (v) have fractionated pattern due to low HREE
concentrations (0.10 < Yb ppm < 1.90; 5.48 < LaN/YbN ppm < 138.33), (vi) have low Y
contents (1.10 < Y ppm < 14.40) and (vii) high Sr/Y ratios (20.93–731.82 ppm). The mafic
magmatism is characterized by: (i) calc-alkaline affinity; (ii) metaluminous to slightly
peraluminous saturation index; (iii) Ti (3477 < Ti ppm < 7734), Zr (11 < Zr ppm < 29), Y (11
< Y ppm < 29) and V (114 < V ppm < 371) contents similar to those of volcanic arc basalts
(Pearce, 1996); and (iv) Nb, Ta and Ti negative anomalies. Additionally, these arc-related
rocks recorded in ca. 2.89 to 2.83 Ga the transition between the arc to the collisional setting
responsible by the amalgamation of the Rio Maria and Carajás domains. This collisional stage
is evidenced by the ultra-high temperature (UHT) metamorphism, partial melting and syn-
tectonic magmatism related to the melting of the metasomatized mantle and the lower crust.
The suture zone is marked by serpentinites and high-grade peridotites in the juxtaposition of
the Carajás and Rio Maria domains.

Keywords: TTG, Mesoarchean, Geochemistry, Carajás Province


114

1. Introduction

The Archean geological record is manly composed of distinct granitoid types (Batuk
Joshi et al., 2017; Laurent et al., 2014; Moyen and Martin, 2012), which have been classified
into four groups: (i) volumetrically-dominant and juvenile tonalites, trondhjemites and
granodiorites (TTG); (ii) Mg-, Fe- and K-rich, metaluminous (monzo) diorites and
granodiorites, referred to as sanukitoids s.l.; (iii) peraluminous and K-rich biotite- and two-
mica granites; and (iv) hybrid-K granites with mixed characteristics from the first three
groups (Laurent et al., 2014). Additionally, the geochemical data showed that the
volumetrically-dominant TTG magmatism is related to melting of mafic rocks under variably
pressure regimes, including rocks with eclogite paragenesis (omphacite + garnet; Moyen,
2011; Moyen and Martin, 2012; Moyen and Stevens, 2006). The sanukitoids s.l. derive from
hybridization between mantle peridotite and a component rich in incompatible elements,
whereas the peraluminous and K-rich biotite- and two-mica granites are associated with
melting of older crustal rocks (Laurent et al., 2014; Martin et al., 2009; Moyen and Martin,
2012).
The Archean granitoid record points to subduction-collisional settings (Laurent et al.,
2014; Moyen and Martin, 2012; Moyen and Stevens, 2006). Moreover, the occurrence of
paired high temperature (HT) – high pressure (HP) metamorphic belts and the petrological-
thermomechanical numerical models indicate that the modern plate-tectonic style would be
operating since the Mesoarchean-Neoarchean transition (ca. 3.2 to 2.5 Ga; Brown, 2009,
2006; Mints et al., 2015; Sizova et al., 2010). Those models also indicate that the slab
breakoff should be usual during the subduction events in the Archean, explaining the relative
rarity of evidences of ultra-high pressure (UHP) rocks (de Wit, 1998; Sizova et al., 2010; van
Hunen and van den Berg, 2008). However, the record of UHP rocks in Archean terrains has
become more frequent, which indicates the opposite (Mints et al., 2015; Sajeev et al., 2013).
On the other hand, the ultra-high temperature (UHT) metamorphism played an important role
in the subduction-collisional Archean environments (Brown, 2008, 2006; Sajeev et al., 2009).
Mostly related to the final collisional stage, the combination of high thermal gradients and fast
exhumation implies a tangential stress related to the horizontal plate motion (Brown, 2005;
Brown et al., 2011; Clark et al., 2011).
To build the whole scenario of the Carajás Domain, in the Carajás Province, it is
important to fill the knowledge gap about the magmatic system associated with the
emplacement of the protoliths of the Xicrim-Cateté Orthogranulite and Xingu Complex rocks.
115

These units record a Mesoarchean magmatic and UHT metamorphic evolution still to be
unravelled. This contribution presents new lithochemical data on the Xicrim-Cateté
Orthogranulite and Xingu Complex within a north to south section in the southern portion of
the Carajás Domain, along the Goiaba Creek banks. The sample selection include only the
rocks preserved from the partial melting aiming to the understanding of the behaviour of the
major, minor and trace elements in the protoliths of the high to medium grade metamorphic
rocks. The identification of the tectonic setting of the rocks of Xicrim-Cateté Orthogranulite
and Xingu Complex could contribute in the discussion about the global tectonic setting in the
Archean and crustal evolution (Batuk Joshi et al., 2017; Hamilton, 2011; Kerrich and Polat,
2006; Laurent et al., 2014; Moyen and Martin, 2012; Van Kranendonk, 2011).

2. Geological Setting

The Carajás Province (Santos et al., 2000) is subdivided into the Rio Maria, at south,
and Carajás domains, at north (Vasquez and Rosa-Costa, 2008). The limit between these
domains is defined by an E-W geophysical lineament that follows the trend of the greenstone
belt units of the Sapucaia Group, passing right below Água Azul do Norte city (Dall’Agnol et
al. 2000; Santos, 2003; Vasquez and Rosa-Costa, 2008; Fig. 1).
The Rio Maria Domain (ca. 3.0-2.8Ga; Almeida et al., 2011) has the classical elements
of an Archean granite-greenstone terrain: (i) metavolcano-sedimentary sequences; (ii) TTG
granitoids of low, medium and high La/Yb ratio; (iii) sanukitoids; (iv) and potassic
leucogranites (Almeida et al., 2011).
The Carajás Domain comprises a high-grade Archean terrain and supracrustal units. It
encompasses Middle Mesoarchean (3.08 Ga to 2.93 Ga; Delinardo da Silva, 2014; Feio et al.,
2013; Moreto et al., 2015a; 2015b; 2011; Pidgeon et al., 2000) migmatitic orthogneiss,
granulites and greenstone belt sequences; Late Mesoarchean granitoids (2.88 to 2.83 Ga; Feio
et al., 2013; Gabriel and Oliveira, 2014; Leite-Santos, 2016; Rodrigues et al., 2014; Silva et
al., 2014); Neoarchean metavolcano-sedimentary units and intrusive mafic-ultramafic, mafic
and felsic rocks (2.76–2.57Ga; DOCEGEO, 1988; Feio et al., 2012; Gibbs et al., 1986;
Machado et al., 1991; R. D. Santos et al., 2013; Wirth et al., 1986). In addition,
Paleoproterozoic A-type (ca. 1.88 Ga) granites are widespread throughout the Carajás
Province (Dall’Agnol et al., 2005; Santos et al., 2013a).
The Archean basement was affected by two main deformation phases, resulting in
continuous low-angle E–W-trending foliation, which was transposed by a spaced high-angle
116

E–W to WNW–ESE foliation associated with regional shear zones (Araújo et al., 1988;
Holdsworth and Pinheiro, 2000). The Mesoarchean basement rocks were initially grouped in
the Xingu (Silva et al., 1974) and Pium (Araújo and Maia, 1991) complexes. The Xingu
Complex encompasses granulites, gneisses, migmatites and granitoids, in addition to
greenstone belts and mafic-ultramafic complexes (Silva et al., 1974; DOCEGEO, 1988). The
Pium Complex is composed of mafic granulites, charnockites, charno-enderbites and
enderbites (Araújo and Maia, 1991).
In the recent geological mapping campaigns, the composition of the Xingu Complex
(Pink unit in Fig.1) was redefined (Vasquez and Rosa-Costa, 2008; Costa et al. 2016). The
authors considered that the unit comprises only tonalitic, and subordinately granodioritic,
orthogneisses and migmatites, with sparse amphibolite enclaves (Araújo and Maia, 1991;
Vasquez and Rosa-Costa, 2008; Costa et al., 2016). The rocks show strongly asymmetric and
anastomosed fabric defined by an imbricated tectonic regime (Araújo & Maia, 1991; Pinheiro
and Holdsworth, 2000; Vasquez et al., 2008). According to Araújo & Maia (1991) and
Delinardo da Silva et al. (in prep.), the Xingu Complex rocks are characterized by the intense
migmatization in a high strain deformational regime, leading to the development of stromatic
and net metatexites and schlieren and schollen diatexites. The protoliths of these migmatitic
gneisses have crystallization ages between ca. 2.97 Ga and 2.93 Ga (Avelar et al., 1999; Pb-
Pb in zircon; Delinardo da Silva et al., in prep.; U–Pb in zircon). Machado et al. (1991)
obtained an age of 2.859 ± 2 Ma (U-Pb in zircon) from an undeformed leucosome, which was
interpreted as the last migmatization event recorded by the Xingu Complex.
The several granitoids individualized from the Xingu Complex showed a diverse
magmatic affinity. The Table 1 shows the main types of magmatism and the related units.
117

Table 1. The Mesoarchean (ca. 3.0-2.8Ga) granitoids of Carajás Domain.


Unit Age Reference
Calc-Alkaline Sodic Granitoids

Campina Verde Tonalitic Complex 2.87 – 2.85 Ga (Feio et al., 2013)


Canaã dos Carajás Granite 2.96 Ga (Feio et al., 2013)
São Carlos Tonalite 2.93 Ga (Silva et al., 2014)
Bacaba Tonalite 3.0 Ga (Moreto et al., 2011)
TTG granitoids

Rio Verde Trondhjemite 2.93 – 2.87 Ga (Feio et al., 2013)


Colorado Tonalite 2.87 Ga (Silva et al., 2014)
High-Mg Granitoids

Água Limpa Granodiorite 2.87 Ga (Gabriel et al., 2015)


Água Azul Granodiorite 2.87 Ga (Gabriel et al., 2015)
High-Ba-Sr Granitoids

Nova Canadá Leucogranite 2.89 – 2.87 Ga (Leite-Santos, 2016)


Calc-Alkaline Potassic Granitoids

Serra Dourada Granite 2.86 – 2.83 Ga (Feio et al., 2013; Moreto et al., 2011)
Bom Jesus Granite 2.83 Ga (Feio et al., 2013)
Cruzadão Granite 2.85 Ga (Feio et al., 2013)
Boa Sorte Granite 2.89-2.85 Ga (Rodrigues et al., 2014)

After Ricci and Carvalho (2006) and Vasquez and Rosa-Costa (2008), the Pium
Complex was renamed and individualized into Xicrim-Cateté Orthogranulite (3.06-2.93 Ga;
U–Pb in zircon; Pidgeon et al., 2000; Delinardo da Silva et al., in prep) and Pium Diopside
Norite (2.74Ga; Pb-Pb in zircon; Santos et al., 2013). The Xicrim-Cateté Orthogranulite
encompasses metatexitic orthopyroxene-diopside gneisses of tonalitic to granitic composition
with enclaves of mafic granulites of variable size (Araújo and Maia, 1991; Ricci and
Carvalho, 2006; Vasquez and Rosa-Costa, 2008; Delinardo da Silva et al., 2015). The
granulite facies ultra-high temperature metamorphism and partial melting at ca. 2.89–2.85Ga
(Pidgeon et al., 2000; Delinardo da Silva et al., in prep.) enhanced the development of high
strain melt-related structures (i.e. stromatic, net, schollen, schlieren; Saywer, 2008) in such
rocks. On the other hand, the Pium Diopside Norite comprises norites, gabbronorites, quartzo-
gabbros and cumulatic rocks of tholeiitic subalkaline affinity (Santos et al., 2013b). The rocks
share geochemical affinity with intra-plate basalts and have Sm-Nd TDM model ages between
3.14 and 3.06 Ga and crystallization Pb–Pb zircon ages of 2,745 ±1 Ma and 2,744 ±1 Ma
(Santos et al., 2013b).
118

Fig. 1. Geological map of the Carajás Domain, modified from Costa et al. (2016). Include the geochronological data from
Machado et al. (1991), Moreto et al., (2011), Feio et al., (2013), Rodrigues et al. (2014), Sousa et al., (2010), Leite-Santos
(2016), Pimentel and Machado (1994), Silva et al., (2010) and Leite et al., (2004).
119

3. Methods
3.1. Field and Petrography
Three fieldwork campaings, totalizing 35 days, were performed, with survey of
geological-structural sections and sampling. The geological map of the area integrates the new
fieldwork data and the compiled data of Araújo and Maia (1991), Costa et al. (2016),
Marangoanha, (2016), Leite-Santos (2016) and Vasquez and Rosa-Costa (2016). Detailed
petrography under transmitted and reflected light was carried out on 74 thin and thin-polished
sections. Additional mineralogical and textural characterization was made using a Scanning
Electron Microscope (SEM) coupled with and Energy Dispersive Spectroscopy (EDS) at the
University of Campinas (UNICAMP).

3.2. Lithogeochemistry
The samples were collected along a south to north section in the Carajás Domain, in a
geographic distribution that allowed representative information of the group of metamorphic
lithotypes of the Xingu Complex and the Xicrim-Cateté Ortogranulite. The rocks are variably
migmatitic, so the sampling was carefully done in order to separate residuum and neosome.
Twenty seven samples were comminuted (< 75 µm) for chemical analysis using a jaw crusher
(Pulverisette 2, Fritsch, Germany) and a planetary and vibratory mill (Pulverisette 5 and 7,
Fritsch, Germany) with agate grinding devices. The major and minor elements were analysed
by the inductively coupled plasma emission spectrometry (ICP–ES) technique. The trace
elements were analysed by the inductively coupled plasma mass spectrometry (ICP–MS)
technique. Both analysis are care out in the Acme Labs and ALS Global Analytical, Canadá.
The data was treated in the Microsoft Excell 2010, using the reference papers to build plots
and templates.

4. Geology of the Goiaba Creek area

The Goiaba Creek area is inserted in the context of the Itacaiúnas Shear Belt (ca. 2.8
Ga; (Araújo et al., 1988; Holdsworth and Pinheiro, 2000). The moderate dipping E–W
foliations in the Xicrim-Cateté Orthogranulite and in the Xingu Complex are related to this
deformation phase. It controls the relationship among the Mesoarchean metamorphic rocks,
greenstone belts and granitoids in the Goiaba Creek area. The slices of greenstone belts occur
all over the area, often imbricated in the Xicrim-Cateté Orthogranulite and Xingu Complex
rocks. The major occurrences, however, are concentred in the Carajás and Rio Maria tectonic
120

contact, where serpentinites and high-grade peridotites were characterized (Sousa et al.,
2015). A generally south dipping shear zone defines this tectonic contact (Fig. 2). Kilometer-
scale milk quartz veins occur in these and in the other E–W shear zones. Another south
dipping shear zone marks the tectonic contact between the Xicrim-Cataté Orthogranulite and
the Xingu Complex (Fig. 2).
The Água Limpa Granodiorite (ca. 2.87 Ga; Gabriel et al., 2015) and the Nova Canadá
Leucogranite (ca. 2.87 Ga; Leite-Santos, 2016) are hosted in the moderate dipping E–W
foliation of the metamorphic rocks. In addition, the Água Limpa Granodiorite occurs as sheet
granite locally. The Neoarchean intrusions (Planalto Granite Suite and Pium Diopside Norite),
mostly occur in the northern part of the area. Their emplacement seems to be controlled by the
reactivation of the Mesoarchean structure in ca. 2.7 Ga (Holdsworth and Pinheiro, 2000).
These rocks show only marginal deformation in the field. In the northeast portion of the
Goiaba Creek area, some outcrops of the isotropic Paleoproterozoic A-type Rio Branco
Granite can be observed (ca. 1.88 Ga; Dall’Agnol et al., 2005; Santos et al., 2013a)
121

Fig. 2. Geological map of Goiaba Creek area. The map results from the field and petrographic studies integrated with the
compiled data from the previous works of Dall’Agnol et al. (2005; 1), Feio et al. (2012; 2), Santos et al. (2013; 3), Souza et
al. (2015; 4), Leite-Santos (2016; 5), Feio et al. (2013; 6) Gabriel and Oliveira (2013; 7); and Almeida et al. (2011; 8).
122

4.1. Xicrim-Cateté Orthogranulite

The Xicrim–Cateté Orthogranulite encompasses metatexitic orthopyroxene-diopside


gneisses of tonalitic to granodioritic composition with mafic granulite enclaves, which were
restricted to the vicinity of the Xicrim Native Village, according to Vasquez and Rosa-Costa
(2008). In this study these rocks were also identified in the northern portion of the Goiaba
Creek area, in which the major outcrops occur in the Princesa Farm (Fig. 2). In this area, the
tonalitic gneisses predominate over the granodioritic and granitic ones. The rocks are
leucocratic and generally light grey, medium-grained and composed of quartz (20–25%),
plagioclase (An25-31; 45–60%) and subordinate K-feldspar (5–10%), biotite (15%), enstatite
(5–10%), diopside (2–3%) Ca-amphibole (3–5%), apatite (2–3%), zircon (~1%), magnetite
(1–2%), sericite (~1%), pistacite and clinozoisite (~1%), chlorite (~1%), scapolite (1–2%),
ilmenite and magnetite (~1%) with grano-lepidoblastic, and subordinated, granoblastic texture
(Fig. 3a-b).
A massive granulitic texture is typical of less deformed granulites outwards the main
shear zones. These rocks show a peak metamorphic paragenesis defined as quartz +
plagioclase + orthopyroxene ± diopside ± Ca-amphibole (Fig. 3b-c). Otherwise, the foliated
texture is widespread, mainly associated with the replacement of orthopyroxene and diopside
by biotite and magnetite (Fig. 3c), especially near the main thrust shear zones (Fig. 2). The
grano-lepidoblastic texture defines the continuous gneissic foliation (Sn), which shows an E–
W trend dipping to SE to SSE (Fig. 3c). The foliation is associated with a downdip stretching
lineation. A spaced high angle foliation (Sn+1) with the same E–W trend can also be observed
in the field. Stromatic and net structures were developed parallel to the gneissic foliation (Sn)
and have hololeucocratic, medium- to coarse-grained phaneritic and equigranular leucosomes.
Orthopyroxene and diopside are variably replaced by lower grade assemblage with
hornblende, biotite, magnetite, pistacite and clinozoisite.
The metatexitic mafic granulites occur as boudins in the metatexitic orthopyroxene-
diopside gneisses or as boulders. They are composed of three lithotypes: (i) metanorite; (ii)
mafic granulite; and (iii) diopside-pargasite-plagioclase granulite (Fig. 3d-f). The rocks are
garnet-free, melanocratic, medium- to coarse-grained and dark grey. The metanorite is
composed of plagioclase (An53-68; 40%); diopside (~5%); enstatite (~45%); pargasite (~3%);
ilmenite (~4%); biotite (>1%); magnetite (>1%); quartz (~1%) and rutile (>1%). The
diopside-pargasite granulite is composed of plagioclase (An35-45; ~40%), diopside (~25%);
pargasite (~25%); biotite (~5%); magnetite (~3%); quartz (~2%) and zircon (<1%). The mafic
123

granulite is composed of plagioclase (An36-41; ~40%); diopside (~20%); enstatite (~10%);


pargasite (~15%); biotite (~10%); magnetite (~3%); quartz (~2%); and zircon (<1%). These
rocks are foliated close to main shear zones or massive with granoblastic and, locally, blasto-
subophic textures (Fig. 3f-i) in the distal areas. The diopside-pargasite-plagioclase granulite
and the mafic granulite with granoblastic texture have an equilibrium assemblage with
plagioclase–diopside–F-rich pargasite–clinoenstatite. The coarse-grained blasto-subophitic
texture is locally preserved in the metanorite (Fig. 3h). These rocks develop a Sn foliation that
accommodates the stromatic structures. Locally, dilatant zones form allowing the migration of
melt and the development of net structures. The areas of intense migmatization are associated
with shear zones and thicker milk quartz veins occurrences (Fig. 3i). The retrograde
paragenesis, associated with the melting event and the Sn foliation development, is
represented by Na-rich plagioclase + Cl-rich pargasite ± quartz. Biotite and magnetite are the
last mineral phases to appear in the retrograde metamorphism sequence.
124

Fig. 3. Field and petrographic features of the Xicrim-Cateté Orthogranulite. A) Coarse-grained metatexitic orthopyroxene-
diopside granodiorite gneiss with granulitic texture and anastomosed Sn foliation; B) Metatexitic orthopyroxene-diopside
tonalite gneiss with its typical coarse-grained granoblastic texture and the plagioclase (Pl) + quartz (Qtz) + K-feldspar (Kfs) +
orthopyroxene (Opx) assemblage, which locally include diopside and pargasite ; C) Retrograde metamorphism in the
granulitic gneisses: orthopyroxene + diopside  biotite + magnetite; D) The abrupt contact between the mafic granulites and
the granulitic gneisses; E) Meter scale fragments of mafic granulites highly deformed inside of a leucosome pocket
associated with an metatexitic orthopyroxene-diopside tonalite gneiss; F) Coarse-grained mafic granulite with local patch of
melting and melt mobilization evidences; G) Mafic granulites with coarse-grained granoblastic texture associated with the
assemblage with plagioclase + orthopyroxene + diopside; H) Metanorite with relic of blasto-subophic texture locally
preserved in the thin section; I) The stromatic and net structures typical of the metatexitic mafic granulites.
125

4.2. Xingu Complex

In the Goiaba Creek area, the Xingu Complex is represented by fine-grained grey
leucocratic rocks, composed of quartz (20–30%), albite (50–60%), K-feldspar (2–3%), biotite
(10–15%), with subordinated apatite (3–5%), epidote group minerals (allanite, clinozoisite
and pistacite; ~2%), titanite, magnetite, and zircon (1-2%). Specifically close to the granulite
facies zone, hornblende (pargasite) represent up to 5–10% in the rock. The Xingu Complex
rocks show commonly tonalitic modal composition (Fig. 4).
The rocks occur as large outcrops, which locally can reach 20 m2. They show an E-W
gneissic foliation (Sn) with NNW and NNE low-angle dip. The Sn represents a spaced
anastomosed foliation in the micro-scale, associated with a grano-lepidoblastic texture.
Asymmetrical tight drag and sheath folds with SSE vergence are related to the gneissic
foliation. The gneisses are variably melted, being classified as metatexites (Fig. 4a). The
stromatic and net structures of the metatexitic orthogneisses develop along the low angle
gneissic foliation (Sn). Locally, the anatetic structures are also folded (Fig 4a). Moreover, the
Sn foliation controlled the emplacement of the Água Limpa Granodiorite, which occurs as
sheets between the gneissic foliation (Fig. 4b).
The leucosomes of the metatexitic orthogneisses and the Água Limpa Granodiorite
mutually crosscut each other and some megacrysts of the second are assimilated by the first.
The relation between the Nova Canadá Leucogranite and the metatexitic orthogneisses of the
Xingu Complex is also marked by these sheets of the first interleaved in the low angle
gneissic foliation of the second. This is a common feature in the north boundary of the Nova
Canadá Leucogranite (see Fig. 2). Differently from the Água Limpa, the Nova Canadá
Leucogranite shows, schollen and schlieren structures which encompass fragments of
orthogneisses and biotite-rich melanocratic thin strips or clusters (Fig. 4c). The metatexitic
orthogneisses enclaves and the biotite-rich schlieren structures are oriented along the low
angle foliation (Sn). A continuous link between the stromatitc and net-structured leucosomes
in the fragments of the migmatitic orthogneisses and the surrounding Nova Canadá
Leucogranite can be observed in the outcrops (Fig. 4c). Under the microscope, dark-brown
biotite grows over hornblende, together with the development of cuspate terminations in K-
feldspar crystals and veinlets of quartz + plagioclase + K-feldspar (Fig. 4d-f). The main melt-
in zones are ultimately associated with milk quartz veins. The retrograde assemblage of these
rocks is represented by titanite–pistacite–allanite–clinozoisite–apatite (Fig. 4g). In some
126

samples, the symplectitic texture is defined by intergrowths of quartz, pistacite and biotite
(Fig. 4g)
The metatexitic orthogneisses show boudin-like enclaves of fine-grained amphibolites
composed of dark-green magnesio-hornblende (~60%), albite (30%), dark-brown biotite
(10%), quartz, pistacite and magnetite (~1%; Fig. 4h-i). Biotite essentially grows over
hornblende. The enclaves are oriented parallel to the low gneissic angle foliation. A high-
angle spaced mylonitic foliation (Sn+1) transposes, and locally, rotates the previous
structures, but has the same E–W strike.

Fig. 4. Field and petrographic features of the Xingu Complex. A) The intrusive contact between the High-Mg granodiorites
from Água Limpa Sanukitoid suite and the Xingu Complex (Xc), controlled by its low angle foliation; B) Metatexitic
orthogneiss from Xingu Complex, showing its common field characteristics: light grey, medium-grained paleosome (P),
white-pink coarse-grained leucosomes (L) and slender stripes of black medium-grained melanosome (M); C) Schollen
fragment of the Xingu Complex (Xc) involved by the Nova Canadá Leucogranite (Nc). The in-source leucosomes in the
127

metatexitic gneiss of Xingu Complex shows a gradual contact with the Nova Canadá Leucogranite (Nc); D) Typical
assemblage of the gneiss: hornblende (Hbl)+ plagioclase (Pl)+ K-feldspar (Kfs)+ Quartz, being replaced by biotite (Bt) 
epidote (Ep); E) Veins of leucosomes composed of quartz + plagioclase + K-feldspar crosscutting plagioclase megacrystals;
F) cuspate films of K-feldspar along the boundary of plagioclase + quartz + K-feldspar; G) symplectitic texture with epidote
(Ep), quartz and biotite (Bt) related to the retrograde metamorphism; H) the boudin-like occurrence of amphibolite enclaves
in the Xingu Complex orthogneiss; I) Amphibolite with a decussate texture and the paragenesis with hornblende (Hbl) +
plagioclase (Pl) which is replaced by biotite  epidote.

5. Geochemistry
5.1. The orthogneisses of the Xicrim-Cateté Orthogranulite and Xingu Complex

The behaviour of the major, minor and trace elements of metatexitic orthopyroxene-
diopside gneisses of the Xicrim-Cateté Orthogranulite and metatexitic hornblende-biotite and
biotite gneisses of the Xingu Complex were evaluated together. In some diagrams, the data of
the Xicrim-Cataté Orthogranulite and Xingu Complex were also compared with the available
data of several TTG rocks of the Rio Maria Domain, formed in the interval of ca. 3.0 and
2.93Ga (Almeida et al., 2011), and with the data of the Canaã dos Carajás Granite (2.95-2.92
Ga; Feio et al., 2013). The trace element data were plotted together with the average values of
the high, medium and low pressure sodic TTG of Moyen (2011).
The protoliths of the rocks of the Xicrim-Cateté Orthogranulite are mainly classified
as tonalites in the normative diagram of Barker (1979), showing a trondhjemitic tendency
(Fig. 5a-b). The exception is the sample SM11P, classified as granodiorite. The rocks are
magnesian, calcic to calc-alkalic (Fig. 6a-b), metaluminous to slightly peraluminous (0.85 <
A/CNK < 1.02; Fig. 6c), silica-rich (61.98< SiO2 <71.90 wt.%; Table 1) and have low
contents of ferro-magnesian oxides (FeOt+MgO+MnO+TiO2 < 4wt.%; an exception is the
sample SM39P with 9 wt.%). These rocks are K-depleted and Na-enriched (0.16 <
K2O/Na2O < 0.53 wt.%; Table 1).
Based on the Laurent et al. (2014) classification, the contents of Na2O, K2O, Al2O3,
CaO, FeOt, MgO, Sr and Ba in the granulitic rocks of the Xicrim-Cateté Orthogranulite are
similar to those of the TTG rocks (Fig. 6d). The average values of REE and trace elements of
Moyen (2011) were used to compare the granulitic gneisses with those of low, medium and
high pressure TTG. The Xicrim-Cateté Orthogranulite samples distribute in all the fields,
however the majority of the samples are similar to high pressure TTG due to their low
contents of Ti, Y, Nb, Rb, Th, U, Ce, Pr, Nd, Sm, Gd and HREE (Table 1). They have high
La/YbN (49.04 – 138.33 ppm) and Sr/Y (168.45 – 731.82 ppm) ratios, extremely positive Eu
anomalies (5.63< Eu/Eu* <1.61) and low contents of Rb, Th and U in comparison with the
medium- and low-pressure TTG samples (Table 1; Fig. 7a-b;). The positive anomalies of Sr
128

and Ti reflect the low concentration of Ce, Nd, Gd and Dy. The samples in the field of the
medium pressure TTG show slightly positive Eu anomalies (1.28< Eu/Eu* <1.13) and
moderate La/YbN ratio (37.59 – 40.07 ppm), whereas those in the field of the low pressure
TTG show negative Eu anomalies (0.81< Eu/Eu* <0.58) and low La/YbN ratios (27.94 –
11.46 ppm; Fig. 8; Table 1). All the samples have a well-marked Na/Ta negative anomaly.
The sample XA17A1 has an anomalous low concentration of Zr and Hf, showing a negative
anomaly in the multi-elementary diagram.
129

Table 2. Lithogeochemical data of the orthogneisses from Xicrim-Cateté Ortogranulite and Xingu Complex. The acronyms HP, MP and LP TTG correspond to the high, medium and low
pressure TTG of Moyen (2011). The normalization was carried out using the Sun and McDounough values of the C1 condrite 1. The Eu/Eu* ratio was calculated by the formulae: 𝑬𝒖𝑵 ÷
√(𝑺𝒎𝑵 ∗ 𝑮𝒅𝑵 )2
Xicrim-Cateté
Xingu Complex
Ortogranulite
Class HP TTG MP TTG LP TTG HP TTG MP TTG LP TTG
Samples XA15A1 XA17B1 SM46N SM54 SM44N SM11P XA12A XA41A CMS22P SM39P SM53N XA47A XS37 XS05P XS02 P XS04P XS105P TS09
SiO2 (wt%) 71.90 65.80 71.76 70.81 69.36 70.53 69.80 66.80 68.35 61.98 69.08 71.00 72.70 72.90 67.70 67.20 67.70 72.70
TiO2 0.29 0.12 0.29 0.22 0.30 0.23 0.30 0.54 0.46 0.45 0.22 0.13 0.15 0.19 0.41 0.45 0.39 0.13
Al2O3 15.80 20.30 15.60 16.23 16.11 15.00 16.15 17.35 15.81 15.91 15.95 17.30 15.60 15.55 15.85 15.75 16.15 15.70
FeOt 2.10 1.03 1.48 1.37 2.39 2.23 2.29 3.36 2.88 5.36 2.73 1.30 1.50 2.14 3.31 4.11 3.76 1.19
MnO 0.03 0.01 0.02 0.02 0.03 0.05 0.03 0.03 0.04 0.08 0.03 0.02 0.02 0.03 0.05 0.08 0.06 0.02
MgO 0.87 0.40 0.44 0.70 1.19 1.01 0.73 1.59 1.51 2.99 0.89 0.40 0.43 0.56 1.30 1.41 1.22 0.26
CaO 4.01 4.61 3.53 3.94 3.98 3.07 3.50 4.82 3.98 5.41 3.63 3.12 3.27 2.69 3.40 4.08 3.90 2.10
Na2O 4.44 6.10 4.53 4.76 4.79 4.36 4.88 5.20 4.58 4.66 5.07 5.94 4.98 5.38 4.49 4.56 4.58 5.28
K2O 0.95 1.31 1.38 0.98 0.89 2.31 1.37 0.84 0.98 1.09 1.06 1.12 0.94 0.99 1.57 1.17 1.44 2.36
P2O5 0.09 0.02 0.08 0.08 0.09 0.06 0.12 0.18 0.15 0.11 0.15 0.04 0.05 0.05 0.14 0.17 0.14 0.04
LOI 0.44 0.30 0.40 0.60 0.40 0.70 0.62 0.46 0.70 1.10 0.70 0.35 0.36 0.45 0.71 0.65 0.50 0.25
Total 101.29 100.31 99.74 99.84 99.82 99.78 100.18 101.67 99.76 99.78 99.81 101.00 100.24 101.27 99.46 100.16 100.37 100.39
P (ppm) 393 87 349 349 393 262 524 786 655 480 655 175 218 218 611 742 611 175
K 7886 10874 11455 8135 7388 19176 11372 6973 8135 9048 8799 9297 7803 8218 13033 9712 11954 19591
Rb 6.80 9.10 22.60 7.30 7.70 42.30 42.40 43.70 13.00 18.40 15.50 43.30 27.10 41.80 64.90 55.70 75.10 59.10
Sr 644 805 560 507 571 522 573 569 649 549 617 650 360 632 478 290 425 631
Cs 0.16 0.10 n.d n.d n.d n.d 0.19 0.32 0.10 0.20 n.d 1.31 0.52 1.08 1.83 1.32 1.28 0.85
Ba 587 867 1436 629 588 1092 520 544 642 382 603 462 154 258 791 306 474 1370
Pb 10.00 13.00 0.60 0.50 1.00 4.20 8.00 9.00 1.60 3.30 3.60 10.00 8.00 12.00 7.00 10.00 14.00 16.00
Sc 2.00 1.00 2.00 2.00 3.00 3.00 2.00 5.00 5.00 18.00 7.00 2.00 2.00 2.00 5.00 16.00 8.00 2.00
Ti 1738 719 1738 1319 1799 1379 1799 3237 2758 2698 1319 779 899 1139 2458 2698 2339 779
Ga 18.40 22.10 16.00 16.40 18.70 16.60 21.10 20.10 16.30 19.90 18.20 17.90 19.30 19.10 19.40 19.60 20.10 19.90
Y 1.60 1.10 1.60 1.30 1.30 3.10 2.40 4.50 5.20 12.80 9.20 2.00 2.00 3.50 6.30 14.00 12.20 14.40
130

Zr 121.00 5.00 74.60 111.80 141.00 62.00 108.00 209.00 254.60 80.30 75.90 24.00 106.00 164.00 151.00 169.00 214.00 86.00
Hf 3.40 n.d 1.80 2.60 3.50 1.60 3.10 5.80 7.30 2.00 1.90 0.80 3.20 4.60 3.90 4.70 5.70 3.10
Nb 2.00 2.40 2.10 1.40 2.10 1.70 3.30 3.60 3.00 4.30 2.00 2.50 2.50 3.60 9.70 10.70 5.10 2.10
Ta 0.10 0.20 n.d n.d n.d n.d 0.10 0.20 n.d 0.20 n.d 0.20 0.20 0.20 0.50 0.80 0.30 0.60
La 19.60 21.00 13.90 12.60 12.90 13.40 24.00 29.20 29.10 21.40 26.30 4.60 17.90 20.80 28.20 21.50 42.10 15.80
Ce 27.90 26.40 22.80 20.80 19.70 21.00 40.20 49.80 54.00 44.20 53.00 6.20 28.10 35.00 58.20 48.70 58.90 25.30
Pr 2.54 2.22 2.18 2.03 1.89 2.21 3.95 4.96 5.76 5.37 6.26 0.64 3.02 3.61 5.83 5.13 7.53 2.99
Nd 8.10 6.30 7.70 5.00 6.60 7.50 12.80 16.10 20.00 22.20 23.60 2.40 9.80 12.60 19.10 19.70 25.10 9.80
Sm 0.94 0.59 0.73 0.74 0.70 1.23 1.56 2.24 2.55 3.78 3.88 0.46 1.22 1.65 2.70 3.83 4.06 1.82
Eu 0.79 0.84 0.73 0.57 0.62 0.72 0.62 0.80 0.81 0.91 0.64 0.46 0.61 0.37 0.84 1.24 1.10 0.45
Gd 0.53 0.36 0.67 0.47 0.61 0.90 0.91 1.66 1.93 3.16 3.03 0.44 0.99 1.03 1.98 3.84 3.57 1.69
Tb 0.07 0.02 0.07 0.06 0.07 0.11 0.10 0.17 0.21 0.41 0.39 0.05 0.10 0.12 0.23 0.46 0.47 0.21
Dy 0.30 0.24 0.32 0.23 0.35 0.67 0.49 0.96 1.12 2.54 1.95 0.34 0.50 0.63 1.34 2.59 2.29 1.17
Ho 0.06 0.03 0.06 0.05 0.05 0.10 0.09 0.16 0.21 0.46 0.32 0.04 0.06 0.10 0.21 0.50 0.41 0.26
Er 0.15 0.09 0.18 0.16 0.12 0.27 0.26 0.37 0.51 1.29 0.80 0.19 0.20 0.27 0.43 1.27 1.11 0.92
Tm 0.03 0.02 0.02 0.02 0.02 0.04 0.04 0.08 0.09 0.19 0.11 0.05 0.04 0.05 0.07 0.22 0.15 0.21
Yb 0.24 0.10 0.16 0.13 0.16 0.18 0.18 0.48 0.51 1.23 0.62 0.21 0.20 0.28 0.43 1.53 1.00 1.90
Lu 0.02 n.d 0.02 0.01 0.01 0.04 0.03 0.05 0.09 0.16 0.07 0.02 0.02 0.03 0.07 0.18 0.15 0.29
Th 0.16 0.15 n.d n.d 1.20 0.50 3.64 3.53 3.70 0.90 6.10 0.89 3.88 12.50 7.21 8.39 3.29 4.02
U 0.08 n.d n.d n.d n.d n.d 0.20 0.44 0.20 n.d n.d 0.22 0.34 0.57 0.33 1.37 0.51 0.38
V 40.00 16.00 25.00 20.00 28.00 44.00 42.00 65.00 48.00 102.00 30.00 20.00 16.00 21.00 55.00 52.00 51.00 12.00
Cr 20.00 20.00 164.21 34.21 47.89 191.58 30.00 40.00 41.05 218.95 150.53 20.00 30.00 20.00 40.00 30.00 30.00 10.00
Co 6.00 4.00 3.30 4.40 6.70 6.90 5.00 18.00 8.10 16.70 5.40 2.00 3.00 5.00 9.00 10.00 8.00 1.00
Ni 5.00 2.00 4.70 7.90 10.40 19.50 4.00 25.00 18.10 38.90 9.50 1.00 1.00 1.00 10.00 10.00 8.00 1.00
Cu 11.00 4.00 7.10 7.10 21.50 4.70 18.00 107.00 31.50 25.30 16.10 4.00 1.00 3.00 14.00 30.00 4.00 1.00
Zn 33.00 17.00 14.00 10.00 15.00 26.00 38.00 43.00 29.00 44.00 22.00 29.00 29.00 49.00 62.00 77.00 66.00 28.00
K2O/Na2O 0.21 0.21 0.30 0.21 0.19 0.53 0.28 0.16 0.21 0.23 0.21 0.19 0.19 0.18 0.35 0.26 0.31 0.45
mg# 0.43 0.41 0.35 0.48 0.47 0.45 0.36 0.46 0.48 0.50 0.37 0.35 0.34 0.32 0.41 0.38 0.37 0.28
A/CNK 1.01 1.02 1.02 1.01 1.00 0.98 1.02 0.95 1.00 0.85 0.99 1.04 1.03 1.05 1.04 0.97 1.00 1.04
131

Sr/Y 402.50 731.82 350.19 390.23 439.46 168.45 238.75 126.44 124.83 42.88 67.02 325.00 180.00 180.57 75.87 20.93 34.84 43.82
(La/Yb)N1 53.80 138.33 57.23 63.85 53.11 49.04 87.83 40.07 37.59 11.46 27.94 14.43 58.96 48.93 43.20 9.26 27.73 5.48
2
Eu/Eu* 3.46 5.63 3.22 2.98 2.93 2.11 1.61 1.28 1.13 0.81 0.58 3.16 1.71 0.88 1.12 1.00 0.89 0.79
Lu/Hf 0.01 n.d 0.01 0.00 0.00 0.03 0.01 0.01 0.01 0.08 0.04 0.03 0.01 0.01 0.02 0.04 0.03 0.09
Nb/Ta 20.00 12.00 n.d n.d n.d n.d 33.00 18.00 n.d 21.50 n.d 12.50 12.50 18.00 19.40 13.38 17.00 3.50
132

Fig. 5. A) Normative classification of the gneisses of the Xicrim-Cateté Orthogranulite and Xingu Complex in the diagram of
Barker (1979); B) K-Na-Ca triangle comparing the evolution of the samples from Xicrim-Cateté Orthogranulite and Xingu
Complex with the modern calc-alkaline magmas (black CA trend).

The protoliths of the Xingu Complex orthogneisses are distributed in the fields of
tonalites and trondhjemites in the normative diagram of Barker (1979) and the rocks have in
general a trondhjemitic affinity (Fig. 5a-b). The orthogneisses are magnesian, calcic to calc-
alkalic (Fig. 6a-b), slightly peraluminous (0.97 < A/CNK < 1.05; Fig. 6c), silica-rich (67.20 <
SiO2 < 72.90 wt.%; Table 1) with low contents of ferro-magnesian oxides (2 <
FeOt+MgO+MnO+TiO2 < 6 wt.%). The rocks are also K-poor and Na-rich rocks (0.2 <
K2O/Na2O <4 wt.%; Table 1).
In the diagram of Laurent et al. (2014), the orthogneisses of Xingu Complex are
placed in the TTG field (Fig. 6d). The average values of high, medium and low pressure TTG
of Moyen (2011) were also used to individualize the samples of the Xigu Complex
orthogneisses. The rocks in the field of the high-pressure TTG follow almost the same pattern
of the granulitic gneisses of Xicrim-Cateté Orthogranulite, as indicated by their low contents
of Ti, Y, Ce, Pr, MREE and HREE and high Sr/Y ratio (325 – 180 ppm) in comparison with
the rocks classified as medium- and low-pressure TTG (Table 1; Fig. 7c-d;Fig. 8). The
La/YbN (58.96 – 5.48 ppm) and the Eu/Eu* (3.16 – 0.79 ppm) ratios do not have a regular
pattern of distribution, as observed for the Xicrim-Cateté Orthogranulite (Table 1). The
sample XA47A shows the highest Eu anomaly (Eu/Eu* = 3.16 ppm) followed by the lowest
La/YbN ratio (14.43 ppm) in relation to the rocks in the field of high-pressure TTG (Fig. 7c-
d). All the samples commonly have Na/Ta ratios and Ti anomalies that decrease and increase,
respectively, from high- to low-pressure TTG (Fig. 7c-d). The sample XA47A has a Th/U
133

negative anomaly, whereas the sample TS09 shows a positive Ba anomaly and concave
HREE pattern in the REE diagram (Fig. 7c-d).

Fig. 6. A-B) Granite classification of Frost et al. (2001) applied to the orthogneisses of Xicrim-Cateté Orthogranulite e Xingu
Complex; C) Schand (1943) diagram for the aluminium saturation index; D) Individualization of Archean granitoids
compiled in a single diagram, according to Laurent et al. (2014).
134

Fig. 7. Spider plots of the Xicrim-Cateté Orthogranulite and Xingu Complex gneisses. The average values of high, medium
and low pressure TTG from Moyen (2011) were plotted together with the sample data.
135

Fig. 8. Diagrams of Moyen (2011) for HP (High Pressure), MP (Medium Pressure) and LP (Low Pressure) TTG.

5.2. The mafic rocks of the Xicrim-Cateté Orthogranulite and Xingu Complex

The mafic granulites of the Xicrim-Cateté Orthogranulite were sampled in the


southern and northern portion of this unit, around the intrusive Neoarchean Planalto Granite
Suite and the Pium Diopside Norite (Fig. 2). Most of the samples were collected in large
outcrops, in portions relatively isolated of the surrounding granulitic gneisses (XA41B;
XS112B; SM41R; XA15B3; SM51R; SM36R). Other samples were extracted from meter
scale enclaves (SM11A; XA12A; see Fig. 3). The amphibolites could only be found as meter
scale enclaves (XA45B; XS33A see Fig. 4) inside the migmatitic orthogneisses of the Xingu
Complex and the Nova Canadá Leucogranite. The samples are also compared to the mafic
granulites of the Lewisian Complex, northwest Scotland (data from Weaver and Tarney,
1981).
136

At first, the approach of Cann (1970) and Pearce (1996) were used to verify the
influence of metamorphism or metasomatism through the analysis of the relatively immobile
elements. Table 2 shows the results of correlation matrices among these immobile elements,
allowing the identification of least contaminated mafic granulites. Through this previous
analysis, the use Zr, Y, Ti, Nb, Th and V appear to be reasonably trustworthy in the following
diagrams for this set of relatively uncontaminated samples. The plots of Fig. 9 exemplify the
positive correlation of Zr and Eu, V and Ti in the least contaminated protoliths of the mafic
granulites. On the other hand, two samples of mafic granulites of the Xicrim-Cateté
Orthogranulite and two samples of amphibolite enclaves in the Xingu Complex orthogneisses
show evidences of contamination or metasomatism. These samples tend to approximate to the
average values of the lower and upper crust (Rudnick and Gao, 2014).

Fig. 9. Plots of trace element vs. Zr, exemplifying the good correlation of the considered least contaminated mafic granulites.
Average values of the lower and upper crust of Rudnick and Gao (2014) are also plotted for comparison.

The mafic granulites of the Xicrim-Cateté Orthogranulite have a narrow range of SiO2
contents (48.30 to 53.60 wt.%), whereas the amphibolites of the Xingu Complex reach higher
silica contents (58.10–59.50 wt.%). K2O, Rb, Sr, Zr, Hf, LREE, and Th contents are low in
the uncontaminated protoliths of granulites when compared to the contaminated samples of
granulites and amphibolites (Table 3). All samples, including mafic granulites and
amphibolites, have low content of Ti, Y, Nb, Ta, HREE, V, Co, Ni (Table 3). Based on its
magmatic affinity and aluminium saturation index, the uncontaminated and contaminated
protoliths of the mafic granulites were considered to be part of a metaluminous, calc-alkaline
series (0.53< A/CNK <0.62; Fig. 10). The protoliths of the amphibolites were also calc-
alkaline, however their slightly peraluminous tendency (0.75< A/CNK <1.04; Fig. 10) may be
due to contamination.
137

∑𝑵 ̅ )×(𝒚𝒊 −𝒚
𝒊=𝟏(𝒙𝒊 −𝒙 ̅)
Table 3. Correlation matrices for the least contaminated samples of the mafic granulites. The correlation is carried out using the Pearson correlation coefficient: 𝐜𝐨𝐬 𝜶 =
𝟐 𝟐
√∑𝑵 ̅) ×√∑𝑵
𝒊=𝟏(𝒙𝒊 −𝒙 𝒊=𝟏(𝒚𝒊 −𝒚
̅)

Sc Ti V Cr Ga Y Zr Nb Hf Ta Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Tm Yb Lu Th
Sc 1.00
Ti 0.50 1.00
V 0.76 0.94 1.00
Cr -0.57 -0.70 -0.78 1.00
Ga 0.26 0.80 0.68 -0.19 1.00
Y 0.62 0.98 0.98 -0.76 0.77 1.00
Zr 0.57 0.99 0.96 -0.70 0.79 0.99 1.00
Nb -0.12 0.68 0.44 -0.06 0.92 0.60 0.63 1.00
Hf 0.71 0.94 0.97 -0.62 0.82 0.97 0.96 0.59 1.00
Ta -0.50 -0.32 -0.46 0.89 0.23 -0.41 -0.34 0.36 -0.29 1.00
Ce -0.08 0.49 0.36 -0.74 0.05 0.44 0.46 0.19 0.23 -0.67 1.00
Pr 0.11 0.78 0.63 -0.77 0.43 0.72 0.75 0.50 0.56 -0.55 0.92 1.00
Nd 0.34 0.86 0.79 -0.90 0.51 0.86 0.83 0.48 0.72 -0.66 0.80 0.92 1.00
Sm 0.50 0.95 0.91 -0.84 0.66 0.95 0.95 0.55 0.87 -0.57 0.67 0.88 0.95 1.00
Eu 0.54 0.98 0.93 -0.66 0.85 0.98 0.99 0.70 0.96 -0.31 0.43 0.74 0.82 0.95 1.00
Gd 0.66 0.93 0.97 -0.82 0.69 0.98 0.93 0.49 0.93 -0.53 0.43 0.67 0.88 0.93 0.92 1.00
Tb 0.59 0.98 0.97 -0.75 0.78 1.00 0.98 0.62 0.96 -0.40 0.44 0.72 0.87 0.95 0.97 0.98 1.00
Dy 0.58 0.99 0.96 -0.70 0.82 0.99 0.99 0.66 0.97 -0.35 0.42 0.73 0.84 0.95 0.99 0.96 0.99 1.00
Ho 0.60 0.97 0.97 -0.81 0.72 0.99 0.97 0.56 0.94 -0.49 0.51 0.76 0.91 0.97 0.96 0.99 0.99 0.98 1.00
Er 0.67 0.98 0.99 -0.75 0.76 1.00 0.98 0.56 0.98 -0.41 0.40 0.69 0.83 0.94 0.98 0.97 0.99 0.99 0.99 1.00
Tm 0.71 0.93 0.98 -0.75 0.75 0.98 0.94 0.52 0.97 -0.45 0.32 0.61 0.81 0.91 0.94 0.99 0.98 0.97 0.97 0.98 1.00
Yb 0.71 0.95 0.99 -0.73 0.76 0.99 0.96 0.54 0.99 -0.42 0.33 0.63 0.80 0.92 0.96 0.98 0.98 0.98 0.97 0.99 0.99 1.00
Lu 0.68 0.97 0.99 -0.74 0.77 0.99 0.98 0.56 0.99 -0.42 0.37 0.67 0.82 0.94 0.97 0.98 0.99 0.99 0.98 1.00 0.99 1.00 1.00
Th -0.49 -0.73 -0.71 0.12 -0.83 -0.70 -0.75 -0.65 -0.81 -0.28 0.18 -0.20 -0.28 -0.50 -0.74 -0.60 -0.70 -0.73 -0.63 -0.72 -0.68 -0.72 -0.72 1.00
138

The ternary diagram of Pearce and Cann (1973) based on the relatively immobile
elements (Ti, Y, Zr) shows increase of Zr content towards the values of the lower and upper
crust in the samples of the contaminated protoliths of mafic rocks from both, Xicrim-Cateté
Orthogranulite and Xingu Complex (Fig. 11a). This is expressed by the magma-crust
interaction vector of Pearce (1996). In the binary diagram of Winchester and Floyd (1977),
based on the Zr/Ti and Nb/Y ratios, both ratios increase towards the average composition of
the lower and upper crust from the uncontaminated protoliths to the contaminated ones,
mostly guided by the increase of Nb as well Zr (Fig. 11b). The Fig. 11a also shows that the
least contaminated protoliths of the mafic granulites fall in the MORB + VAB field. In the
Fig. 11b, the variation of the Nb/Y ratios for the uncontaminated protoliths along the garnet
lherzolite vector of Pearce (1996) can also be observed and related to the relative increase of
Nb and decrease of Y contents (Table 3).
In the multi-elementary diagram, the contaminated protoliths of Xicrim-Cateté
Orthogranulite and Xingu Complex units display the same pattern of the average data of the
lower and upper crust, especially the samples SM11A and XA45B, which have enrichment of
Th, LREE, Zr, when compared to the uncontaminated protoliths (Fig. 12). The contaminated
protoliths also show expressive Na/Ta and Ti anomalies (Fig. 12).
In the diagrams of Pearce (2014), the data from the mafic granulites of the Xicrim-
Cateté Orthogranulite and amphibolites of the Xingu Complex display a trajectory towards
the field of the magmatic arc (Fig. 13a). The trajectory is expressed by Th and Nb enrichment
from values of depleted and enriched MORB sources following the subduction zone vector
(Fig. 13a). Even the least contaminated protoliths of the mafic granulites follow the
subduction zone vector (SZ; Fig. 13a). The samples display a major Th enrichment, with
respect Nb (Fig. 13a). The contaminated protoliths of the mafic granulites and the
amphibolites are distributed in the continental arc field, surrounding the lower and upper crust
average values (Fig. 13a). The uncontaminated samples follow the path to the island arc field.
The Fig. 13b shows the fields statistically defined by Pearce (1996) for the volcanic
arc basalts (VAB) and the within-plate basalts (WPB), where the hot colours represent the
high percentage of data cluster. In this scenario, the uncontaminated protoliths of the mafic
granulites consistently plot in the volcanic arc basalt field. The least contaminated protoliths
of the mafic granulites have higher V/(Ti/1000) ratios (48 – 66; Table 3) and the samples fall
in the island arc tholeiites (IAT) field in the Fig. 13c, being related to the slab proximal melts.
The contaminated protoliths of both mafic granulites and amphibolites plot in the MORB field
139

in the V vs. Ti diagram (Fig. 13c). This is due to the Ti relative enrichment towards the lower
and upper crust average values that these rocks showed in the previous diagrams.
In the Fig. 14, the least contaminated protoliths of the mafic granulites are compared
with different types of basaltic rocks. The multi-elementary diagram highlights the negative
Nb/Ta anomalies, the low concentration of Zr, Hf, Sm and Ti and negative Ho/Y anomalies of
the protoliths of the mafic granulites (Fig. 14). It also shows that most of the samples display
a pattern very similar to the continental arc basalts (CAB; Fig. 14). The sample XA41B
display a pattern that is more similar to the volcanic-rifted margin basalts (VRMB; Fig. 14).
140

Table 4. Lithogeochemical data of the Xicrim-Cateté Orthogranulite mafic granulites and Xingu Complex amphibolites.
UGP: Uncontaminated Granulite Protoliths; CGP: Contaminated Granulite Protoliths; CAP: Contaminated Amphibolite
Protoliths.
Xicrim-Cateté Ortogranulite Xingu Complex
UGP CGP CAP
Samples XA41B XS112B SM41R XA15B3 SM51R SM36R XA12B SM11A XA45B XS33A
SiO2 (wt%) 48.70 48.30 51.18 47.60 49.07 48.70 53.60 48.85 58.10 59.50
Al2O3 14.35 14.45 16.19 14.65 14.44 15.18 20.40 14.86 16.50 17.30
Fe2O3t 13.60 13.05 8.90 12.15 11.27 12.15 7.53 11.78 8.75 6.87
CaO 9.83 10.80 10.46 11.65 11.86 10.04 6.82 8.80 7.48 3.36
MgO 6.72 7.50 6.59 7.52 7.42 7.58 2.48 8.05 3.64 3.18
Na2O 3.24 2.97 3.81 2.74 3.02 3.34 5.62 3.45 4.61 4.56
K2O 0.48 0.91 0.92 0.79 0.82 0.93 1.22 1.60 0.89 2.77
TiO2 1.29 0.82 0.49 0.64 0.74 0.84 0.87 0.90 0.74 0.58
MnO 0.19 0.20 0.14 0.18 0.19 0.18 0.09 0.20 0.16 0.12
P2O5 0.10 0.05 0.05 0.04 0.07 0.07 0.25 0.30 0.18 0.22
LOI 0.03 0.33 1.00 0.47 0.80 0.70 0.42 0.80 0.61 0.90
Total 98.59 99.45 99.79 98.51 99.77 99.74 99.44 99.69 101.77 99.47
P (ppm) 436 218 218 175 305 305 1091 1309 786 960
K 3985 7554 7637 6558 6807 7720 10128 13282 7388 22995
Rb 2.2 7.6 9.1 6.8 5.6 7.3 30.6 62.6 16.1 165.5
Sr 134 120 206 122 161 269 862 347 441 517
Cs 0.07 0.06 n.d 0.63 n.d n.d 0.13 0.40 0.29 4.43
Ba 82 64 170 144 131 248 405 460 283 541
Pb 2.0 n.d 0.8 4.0 0.6 3.1 4.0 2.8 8.0 14.0
Sc 43 47 34 40 39 35 12 31 17 12
Ti 7734 4916 2938 3837 4436 5036 5216 5395 4436 3477
Ga 19.6 16.1 16.2 14.5 14.4 16.0 23.8 20.8 18.4 25.9
Y 29.1 20.6 11.7 16.9 16.7 19.0 11.4 23.7 19.4 12.9
Zr 70 49 29 36 42 46 220 98 114 123
Hf 2.2 1.7 1.0 1.2 1.2 1.3 4.9 2.3 3.2 3.7
Nb 3.9 2.0 2.8 1.8 1.9 2.9 3.4 4.1 5.9 7.5
Ta 0.3 0.2 0.5 0.2 n.d 0.2 0.3 0.1 0.5 0.5
La 4.8 4.2 3.8 4.6 5.9 7.0 10.8 40.3 33.0 11.2
Ce 13.0 10.4 7.5 11.6 13.0 16.7 27.1 86.4 52.6 32.3
Pr 2.09 1.46 0.95 1.35 1.63 2.22 3.77 10.67 7.83 3.22
Nd 10.3 7.2 4.1 7.6 6.7 9.4 15.2 43.7 29.8 11.8
Sm 3.18 2.30 1.07 1.83 1.85 2.55 3.37 7.59 5.28 3.01
Eu 1.22 0.85 0.51 0.59 0.65 0.82 1.11 1.73 1.31 0.51
Gd 4.90 3.46 1.58 3.16 2.48 3.06 2.97 6.09 4.22 2.89
Tb 0.84 0.55 0.30 0.47 0.44 0.52 0.4 0.84 0.57 0.45
Dy 5.74 3.84 2.07 2.83 2.90 3.58 2.47 4.87 3.42 2.49
Ho 1.13 0.78 0.41 0.69 0.64 0.76 0.42 0.95 0.66 0.48
Er 3.25 2.38 1.26 1.83 1.85 2.05 1.19 2.48 1.78 1.21
Tm 0.52 0.38 0.18 0.31 0.25 0.3 0.18 0.34 0.26 0.18
Yb 3.22 2.45 1.29 1.89 1.75 1.98 1.26 2.07 1.57 1.14
141

Xicrim-Cateté Ortogranulite Xingu Complex


UGP CGP CAP
Samples XA41B XS112B SM41R XA15B3 SM51R SM36R XA12B SM11A XA45B XS33A
Lu 0.50 0.37 0.19 0.28 0.27 0.31 0.18 0.36 0.27 0.15
Th 0.21 0.37 0.40 0.50 0.40 0.50 0.77 7.70 3.91 3.34
U 0.15 0.19 0.40 0.20 n.d 0.20 0.84 1.30 1.66 0.83
V 371 305 188 255 259 255 114 229 141 122
Cr 230 290 602 280 335 267 30 356 170 30
Co 49 52 37 53 45 47 18 43 30 21
Ni 115 148 59.9 190 97 88 20 67 94 35
Cu 101 58 8 30 88 58 44 46 1 9
Zn 105 96 25 93 31 41 87 59 85 177
mg# 0.49 0.53 0.59 0.55 0.57 0.55 0.39 0.58 0.45 0.48
A/CNK 0.60 0.57 0.62 0.55 0.53 0.61 0.89 0.63 0.75 1.04
Zr/Ti 0.010 0.009 0.009 0.010 0.009 0.010 0.042 0.018 0.026 0.035
Nb/Y 0.239 0.153 0.134 0.114 0.107 0.097 0.298 0.173 0.304 0.581
Th/Yb 0.31 0.25 0.07 0.23 0.26 0.15 0.61 3.72 2.49 2.93
Nb/Yb 2.17 1.46 1.21 1.09 0.95 0.82 2.70 1.98 3.76 6.58
V/(Ti/1000) 64 51 48 58 66 62 22 42 32 35

Fig. 10. The diagrams for the classification of magmatic affinity (left; AFM; Irvine and Baragar, 1971) and aluminium
saturation index (right; Schand, 1943) for the rocks of Xicrim-Cateté Orthogranulite and Xingu Complex.
142

Fig. 11. Diagrams of Pearce and Cann (1973; A) and Winchester and Floyd (1977; B) modified by Pearce (1996). VAB:
Volcanic Arc Basalts; MORB: Mid-Ridge Oceanic Basalt; WPB: Within Plate Basalts; Fractional Crystallization Vectors:
plagioclase+olivine+augite (POA);plagioclase+olivine+augite+magnetite (POAM); plagioclase+olivine+augite+
hornblende+magnetite (POAHM). Petrogenetic vectors: Spinel Lherzolite (sp. lherz), Garnet Lherzolite (gt. lherz). MM:
MORB mantle source.

Fig. 12. Spider plots of samples from Xicrim-Cateté Orthogranulites (mafic granulites) and Xingu Complex amphibolites.
The primitive mantle normalization values come from McDonough and Sun (1995).
143

Fig. 13. Plots of tectonic setting of Pearce (1996) and Pearce (2014) for the mafic rocks of Xicrim-Cateté Orthogranulite and
Xingu Complex rocks. OIB: ocean island basalt; E-MORB: enriched middle ocean ridge basalt; N-MORB: normal middle
ocean ridge basalt; WPB: within-plate basalt; VAB: volcanic arc basalt; IAT: island arc tholeiite; BABB: back arc basin
basalt; FAB: front arc basalt.

Fig. 14. Spider plots of the uncontaminated protoliths of the mafic granulites compared to ocean island basalts (OIB),
continental arc basalts (CAB), volcanic rifted margin basalts (VRMB), normal middle ocean ridge basalts (N-MORB) and
enriched middle ocean ridge basalts (E-MORB). OIB and MORB values are from Sun and McDonough (1989), primitive
mantle values from the pyrolite model of McDonough and Sun (1995), CAB values are from Tamura et al. (2014), VRMB
values come of Hooper et al. (1999).
144

6. Discussion
6.1. The fingerprints of the Mesoarchean arc magmatism

The available lithochemical data of the granitoids characterized in the Mesoarchean


(ca. 3.0-2.8 Ga) basement of the Carajás Domain indicate that the magmatism has the
following characteristics: (i) Na-rich and TTG-like; (ii) high Mg; and (iii) high Ba–Sr and;
(iv) K-rich. It covers all the spectrum of Archean granitoid magmatism (Batuk Joshi et al.,
2017; Laurent et al., 2014). The protoliths of the metamorphic gneisses of Xicrim-Cateté
Orthogranulite (ca. 3.06–2.93 Ga) and Xingu Complex (ca. 2.97–2.93 Ga) reveal the
following geochemical characteristics: (i) silica-rich rocks (65.58 < wt% SiO2 < 72.90), (ii)
with high Na2O contents (4.36 – 6.10 wt%), (iii) poor in ferromagnesian elements
(FeOt+MgO+MnO+TiO2 < 4wt.%), (iv) have Nb and Ta negative anomalies, (v) fractionated
pattern due to low HREE concentrations (0.10 < Yb ppm < 1.90; 5.48 < LaN/YbN ppm <
138.33), (vi) low Y contents (1.10 < Y ppm < 14.40) and (vii) high Sr/Y ratios (20.93 –
731.82 ppm). According to these geochemical characteristics, the protoliths of these gneisses
should be classified as TTG rocks (Laurent et al., 2014; Moyen and Martin, 2012).
Additionally, the samples could be subdivided into the high, medium and low pressure groups
of TTG settled down by Moyen (2011). However, there is no clear relationship between the
classification of the protoliths as high, medium and low pressure TTG and the crystallization
ages of the protoliths, indicating that melts generated at different depths were emplaced
contemporaneously.
According to Vielzeuf and Schmidt (2001), the genesis of the sodic rocks are related
to the breakdown of plagioclase + mafic hydrated silicates (amphibole ± epidote). Despite the
fact that the high, medium and low pressure TTG share the same source (hydrated
amphibolites), a closer examination reveal the specific characteristic of the source and
residual assemblage (Moyen, 2011). The data suggest that the main difference among the
samples rest in the concentration of Y and HREE elements, which is extremely low in the
samples classified as high pressure TTG in comparison with those of the medium and low
pressure ones. The low and intermediate contents of Y and HREE are due to melting in the
presence of garnet (Moyen and Stevens, 2006; Rapp et al., 1991). Additionally, the
combination of clinopyroxene (omphacite) and garnet and the absence of plagioclase in the
residuum enhance the Y and HREE retention in the source (high partition coefficient; see
Rollinson, 1993 and references there in). This may explain the high La/YbN and Sr/Y ratios
and the extremely positive Eu anomalies in the samples in the field of the high-pressure TTG
145

samples, which are not observed in the samples in the field of medium- and low-pressure
TTG. The Eu negative anomalies and the concave HREE pattern in the REE diagram showed
by the low-pressure TTG would require the presence of plagioclase + amphibole in the
residuum (Foley et al., 2002; Laurie and Stevens, 2012; Moyen and Stevens, 2006; Rollinson,
1993; Vielzeuf and Schmidt, 2001). Ti, Nb and Ta depletion in the samples could be related to
the presence of rutile in the residuum, (stable above 15kbar; Moyen and Stevens, 2006; Foley
et al., 2002), at the high to low pressure sites of melting.
The high, medium and low pressure TTG are related to different geothermal gradients:
(i) ca. 10 ºC/Km for the high pressure TTG; (ii) 12–20º C/Km for the medium pressure TTG;
and (iii) 20–30 ºC/Km low pressure TTG (Moyen, 2011). The cold geothermal of the high
pressure TTG was related to hot subduction zones, favourably related to young slabs, related
to adakite (TTG-like) magmatism nowadays (Martin et al., 2009; Moyen, 2011). The stability
of the lithosphere in the Mesoarchean to Neoarchean period, due to the higher upper mantle
geothermal gradients, would favour the short-lived subduction events related to frequent slab-
breakoff (Moyen and van Hunen, 2012; Sizova et al., 2010). The low pressure TTG record,
which is related to higher geothermal gradients, could be related to the collapse of a thickened
crust or to plume magmatism at the base of an oceanic plateau (Moyen, 2011; Moyen and van
Hunen, 2012). The geochemical data of the Sapucaia Group, the greenstone belt sequence
between the Carajás and Rio Maria Domain, shows anthophyllite-chlorite-tremolite schist
with OIB affinity (Sousa et al., 2015). In addition, the ratios of the HFS elements of the
anthophyllite-chlorite-tremolite schist suggest that the plume magmatism would have
contribution of slab-recycled material (Sousa et al., 2015). The required geothermal gradients
associated with the medium pressure TTG, whose source is a poor-plagioclase garnet-
amphibolite, may be related to the early stages of the continent collision (Moyen, 2011 and
references there in). However, it could also be related to crustal thickening followed by
orogenic collapse and crustal delamination (Kisters et al., 2003; Moyen, 2011).
The mafic rocks of the Xicrim-Cateté Orthogranulite and Xingu Complex share the
following geochemical characteristics: (i) calc-alkaline affinity; (ii) metaluminous to slightly
peraluminous saturation index; (iii) low Ti (3477 < Ti ppm < 7734), Zr (11 < Zr ppm < 29), Y
(11 < Y ppm < 29) and V (114 < V ppm < 371); and (iv) Nb, Ta and Ti negative anomalies.
The enrichment of Th, Zr, and LREE in the rocks considered to be “contaminated” is
ambiguous. It could be influenced by the coeval TTG magmatism and metasomatism of the
depleted mantle source (Laurent et al., 2014; Moyen and van Hunen, 2012) or it could be due
146

to metasomatism of the mafic rocks during the partial melt event (e.g. Heaman et al., 2002).
During the high grade metamorphism, such elements would be mobilized from the mafic
granulites (James et al., 1987), which explain the low contents of Th, U and LREE in the
uncontaminated rocks distally sampled from the partially melted zones. However, the
“contaminated” mafic rocks were collected inside these zones; some of them were involved in
larger melt pockets (i.e. Fig. 3E). The behave of Th, U and LREE may be due to a buffering
effect during the partial melting of the surrounding gneisses, as the mafic rocks act like
resister lithotypes, leading to a conservative behaviour (Pearce and Peate, 1995).
Moreover, the Ti, Y, Nb, Ta, Zr contents of the least-altered mafic rocks of Xicrim-
Cateté Orthogranulite and Xingu Complex indicate a depleted mantle source (Pearce and
Cann, 1973; Pearce, 1996; Pearce and Peate, 1995). Additionally, the HREE content and the
Nb/Y indicate the partial melting of the garnet lherzolite source (Pearce, 1996). The Nb, Ta
and Ti anomalies also reflect the presence of rutile in the source, related to pressures over 15
kbar (Foley et al., 2002). The Th/Nb proxy for the mafic rocks of the Xicrim-Cateté and
Xingu Complex, evolve in a diagonal line, which indicate the modification of both Th and Nb
contents (Pearce, 2014). This feature indicate the contamination of the mantle melts by
assimilation of the continental crust or interaction with slab-melts (Pearce, 2014; Pearce and
Peate, 1995). V/Ti proxy (Fig. 13C) indicate a time trend of the subduction influence in the
arc magmatism due to V enrichment in the suprasubduction melts (Pearce, 2014). The mafic
rocks of Xicrim-Cateté Orthogranulite and Xingu Complex represent the periods of: (i)
subduction initiation related to low V/Ti ratios and consequently low percentage of water in
the mantle; and (ii) subduction-influenced setting related to intermediate V/Ti ratio and
consequently higher hydration of the mantle. The modification of the mantle source by
subduction-derived fluids and melts is regarded to the increase of water, making the melting
process more oxidising and increasing the proportion of V in the generated liquids (Pearce,
2014; Pearce and Peate, 1995). This scenario implies the unambiguously participation of the
coeval slab-derived TTG melts in the mafic magmatism during the stage of the arc setting.
The syn- to late-collisional setting, the magmatism would have fingerprints of the
crust and the lithospheric mantle, metasomatized by the previous TTG magmatism (Laurent et
al., 2014; Martin et al., 2009). In a short time span after the waning of the TTG magmatism,
or even coeval to it, the syn- to late-tectonic magmatism related to sanukitoid, peraluminous
K-rich magmatism and hybrid granitoids occur. In the Carajás Domain, the magmatism
related to the collisional period shows the same pattern. It is marked by the emplacement of
147

the high Mg rocks of sanukitoids affinity (Água Limpa Granodiorite and Água Azul
Granodiorite; ca. 2.87 Ga; Gabriel et al., 2015) that indicate the new metasomatic character of
the lithospheric mantle. This is coeval to the emplacement of the high Ba-Sr granitoids of
Nova Canadá Leucogranite (ca. 2.87 Ga), interpreted as a result of fractional crystallization of
sanukitoid magmas and mixture with sodic contemporaneous magmas (Leite-Santos, 2016).
Together with that, the volumetrically dominant Late Mesoarchean K-rich granitoids (ca.
2.87–2.83Ga) occur and are related to crustal anatexis at variable crustal depths (Feio et al.,
2013; Leite-Santos, 2016; Silva et al., 2018).
The scenario of diverse magmatism is contemporaneous to the high-grade
metamorphic event associated to the collision of the Carajás and Rio Maria domains at ca.
2.89–2.85 Ga (Delinardo da Silva et al., in prep.; Machado et al., 1991; Pidgeon et al., 2000).
During this event, the Carajás Domian crust underwent ultra-high temperature metamorphism
and local dehydration melting that trigger the widespread water fluxed melting of the Xicrim-
Cateté Orthogranulite and Xingu Complex rocks (Delinardo da Silva et al., in prep.). The
dehydration melting of mafic source (ca. 2.87 Ga; Delinardo da Silva et al., in prep.) in the
lower crust would generate sodic magmas and leave granulite residuum (Pl + Amph = Opx +
Cpx + Melt; Vielzeuf and Schmidt, 2001). Such process could be related to the younger sodic
magmatism (ca. 2.87–2.85 Ga; Rio Verde Trondhjemite, Campina Verde Tonalitic Complex
and Colorado Tonalite; Feio et al., 2013; Silva et al., 2014). In addition, the widespread water-
fluxed melting enhanced by the high thermal activity and intense fluid circulation (e.g
Sawyer, 2010), related to the reaction Qtz+Pl+Kfs+H2O = Melt (Stevens and Clemens,
1993), could play an important role in the genesis of the other granitoids that have a crustal
component.
The evolution in the Rio Maria Domain is also related to an active arc system
(Almeida et al., 2016, 2011). The older rocks (ca. 2.98–2.92 Ga), represented by low pressure
TTG, are associated with the melting of the thickened oceanic crust during a hot subduction
event (ca. 10kbar; Almeida et al., 2016, 2011). The medium and high pressure TTG were
associated with the slab melting (10–15 kbar and >15 kbar, respectively (Almeida et al., 2016,
2011). The authors suggest that part of the melts reacted with the mantle wedge during this
process, leaving behind a metasomatized source. The melting of this metasomatized source
produced the sanukitoid magmas emplaced at ca. 2.87 Ga (Almeida et al., 2016, 2011). The
heat generated by those magmas were related to the melting of the base of the crust,
producing the new TTG pulse at ca. 2.86 Ga (Almeida et al., 2016, 2011)
148

The characteristic of the Mesoarchean magmatic evolution in the Rio Maria Domain is
very similar to those observed in the Carajás Domain. Both terrains show evidences of an
evolution of an arc-related setting (3.0-2.9Ga) to a collisional setting (ca. 2.86 Ga). However,
in the Carajás Domain, the collisional setting is not also regarded to the syn-tectonic
magmatism, but to the granulite ultra-high temperature magmatism and dehydration melting
in the lower crust (Delinardo da Silva et al., in prep.). Such features is not described in the Rio
Maria Domain (Almeida et al., 2011). This indicates that despite the similar style of
evolution, the terrains evolved separately. In addition, the occurrence of serpentinites and
peridotites with granulite facies metamorphism (Sousa et al., 2015) right in the limit between
the terrains is also an evidence of the amalgamation process.
According to the geological evidences and thermomechanical models, the transition to
the modern plate tectonic regime occurred during the Mesoarchean-Neoarchean period (ca.
3.2-2.5Ga; Brown, 2008; Sizova et al., 2010). However, the plates were not so strong due to
the characteristic high thermal regime of that period (i.e. mantle temperatures 175-160 K
higher than today’s temperature; Sizova et al., 2010). The duration of the slabs should not be
long enough, and several subduction processes would occur (de Wit, 1998; Moyen and van
Hunen, 2012; Sizova et al., 2010; van Hunen and van den Berg, 2008). However, the recent
findings of Mesoarchean UHP in the HT-HP paired metamorphic belts indicate that some
slabs resisted (Sajeev et al., 2013 and references there in). The Carajás Domain magmatic and
metamorphic evolution implies a whole scenario related to one-sided asymmetrical
subduction due to the dominant occurrence of high pressure TTG and mafic magmatism with
arc signature. Moreover, the episodic contributions of plumes were suggested by little fraction
of low pressure TTG and by the OIB affinity of some rocks of the Sapucaia greenstone belt
(Sousa et al., 2015). Nevertheless, the search for high-pressure rocks will certainly improve
the tectonic model of the Carajás Domain.

7. Conclusions

• The protoliths of the orthogneisses of granulite and amphibolite facies of the


Xicrim-Cateté Orthogranulite and Xingu Complex, respectively, are related to an
early stage of subduction-related arc magmatic system developed between ca. 3.06
and 2.93 Ga. In this system, slab melting of hydrated amphibolites (hornblende +
plagioclase) and eclogites (omphacite + garnet) in the presence of rutile and
suprasubduction melting of a hydrated garnet lherzolite predominate.
149

• The Carajás–Rio Maria collision stage (ca. 2.89–2.83 Ga) was associated with a
combination of crustal and mantle derived melts. The mantle melts shows the
influence of the previous metasomatism in the stage one, related to the sanukitoid,
high Ba–Sr magmatism. This stage was coeval to the high-temperature
metamorphism and widespread partial melting. Such processes reflect the
volumetrically dominant occurrence of crustal derived granitoids in the Carajás
Domain.

References

Almeida, J. de A.C. de, Dall’Agnol, R., de Oliveira, M.A., Macambira, M.J.B., Pimentel,
M.M., Rämö, O.T., Guimarães, F.V., Leite, A.A. da S., 2011. Zircon geochronology,
geochemistry and origin of the TTG suites of the Rio Maria granite-greenstone terrane:
Implications for the growth of the Archean crust of the Carajás province, Brazil.
Precambrian Res. 187, 201–221. doi:10.1016/j.precamres.2011.03.004
Almeida, J. de A.C. de, Oliveira, V.E.S., Rocha, M.C., Rocha, K.P.P., 2016. Geologia,
Petrografia e Geoquímica dos Granitoides Arqueanos da Porção Sul do Domínio Rio
Maria, Província Carajás, in: SBG (Ed.), Anais Do 46o Congresso Brasileiro de
Geologia. SBG - Sociedade Brasileira de Geologia, Porto Alegre, RS.
Araújo, O.J.B. de, Maia, R.G.N., Jorge João, X. da S., Costa, J.B.S., 1988. Megaestruturação
Arqueana da Folha Serra dos Carajás, in: SBG (Ed.), Anais Do VII Congresso Latino-
Americano de Geologia. SBG, Belém, PA, pp. 324–338.
Araújo, O.J.B., Maia, R.G.N., 1991. Programa Levantamento Geológicos Básicos do Brasil.
Programa Grande Carajás. Folha SB.22-Z-A. Estado do Pará. Texto Explicativo.
Brasilia.
Barker, F., 1979. Trondhjemite: Definition, Environment and Hypotheses of Origin, in:
Developments in Petrology. ELSEVIER SCIENTIFIC PUBLISHING COMPANY, pp.
1–12. doi:10.1016/B978-0-444-41765-7.50006-X
Batuk Joshi, K., Bhattacharjee, J., Rai, G., Halla, J., Ahmad, T., Kurhila, M., Heilimo, E.,
Choudhary, A.K., 2017. The diversification of granitoids and plate tectonic implications
at the Archaean–Proterozoic boundary in the Bundelkhand Craton, Central India. Geol.
Soc. London, Spec. Publ. 449, 123–157. doi:10.1144/SP449.8
150

Brown, M., 2005. Synergistic effects of melting and deformation: an example from the
Variscan belt, western France. Geol. Soc. London, Spec. Publ. 243, 205–226.
doi:10.1144/GSL.SP.2005.243.01.15
Brown, M., 2006. Duality of thermal regimes is the distinctive characteristics of plate
tectonics since the Neoarchean. Geology 34, 961–964. doi:10.1130/G22853A.1
Brown, M., 2008. Characteristic thermal regimes of plate tectonics and their metamorphic
imprint throughout Earth history: When did Earth first adopt a plate tectonics mode of
behavior. Spec. Pap. 440 When Did Plate Tectonics Begin Planet Earth? 2440, 97–128.
doi:10.1130/2008.2440(05)
Brown, M., 2009. Metamorphic patterns in orogenic systems and the geological record. Geol.
Soc. London, Spec. Publ. 318, 37–74. doi:10.1144/SP318.2
Brown, M., Korhonen, F.J., Siddoway, C.S., 2011. Organizing Melt Flow through the Crust.
Elements 7, 261–266. doi:10.2113/gselements.7.4.261
Cann, J.R., 1970. Rb, Sr, Y, Zr and Nb in some ocean floor basalts. Epsl 10, 7–11.
Clark, C., Fitzsimons, I.C.W., Healy, D., Harley, S.L., 2011. How does the continental crust
get really hot? Elements 7, 235–240. doi:10.2113/gselements.7.4.235
Costa, U.A.P., Paula, R.R., Silva, D.P.B., Barbosa, J.P.O., Silva, C.M.G., Tavares, F.M.,
Oliveira, J.K.M., Justo, A.P., 2016. Programa geologia do Brasil-PGB. Mapa de
integração geológico-geofísica da ARIM Carajás. Escala 1:250.000 Estado do Pará.
Belém, CPRM.
Dall’Agnol, R., Teixeira, N.P., Rämö, O.T., Moura, C.A. V, Macambira, M.J.B., de Oliveira,
D.C., 2005. Petrogenesis of the Paleoproterozoic rapakivi A-type granites of the Archean
Carajás metallogenic province, Brazil. Lithos 80, 101–129.
doi:10.1016/j.lithos.2004.03.058
de Wit, M.J., 1998. On Archean granites, greenstones, cratons and tectonics: does the
evidence demand a verdict? Precambrian Res. 91, 181–226.
Delinardo da Silva, M.A., 2014. Metatexitos e diatexitos do Complexo Xingu na região de
Canaã dos Carajás: Implicações para a Evolução Mesoarqueana do Domínio Carajás.
University of Campinas (UNICAMP).
151

Delinardo da Silva, M.A., Monteiro, L.V.S., Moreto, C.P.N., Sousa, S.D. de, 2015.
Metamorfismo e geoquímica do Complexo Xingu na região de Canaã dos Carajás :
implicações para a evolução mesoarqueana do domínio Carajás , Província Carajás, in:
Gorayeb, P., Menguins, A. (Eds.), Contribuições a Geologia Da Amazônia. SBG-NO,
Belém, PA, pp. 279–298.
DOCEGEO, 1988. Revisão litoestratigráfica da Província Mineral de Carajás, in: CVRD,
SBG (Eds.), Anais Do 35o Congresso Brasileiro de Geologia. Sociedade Brasileira de
Geologia (SBG), Belém, PA, pp. 11–54.
Feio, G.R.L., Dall’Agnol, R., Dantas, E.L., Macambira, M.J.B., Gomes, A.C.B., Sardinha,
A.S., Oliveira, D.C., Santos, R.D., Santos, P.A., 2012. Geochemistry, geochronology,
and origin of the Neoarchean Planalto Granite suite, Carajás, Amazonian craton: A-type
or hydrated charnockitic granites? Lithos 151, 57–73. doi:10.1016/j.lithos.2012.02.020
Feio, G.R.L., Dall’Agnol, R., Dantas, E.L., Macambira, M.J.B., Santos, J.O.S., Althoff, F.J.,
Soares, J.E.B., 2013. Archean granitoid magmatism in the Canaã dos Carajás area:
Implications for crustal evolution of the Carajás province, Amazonian craton, Brazil.
Precambrian Res. 227, 157–185. doi:10.1016/j.precamres.2012.04.007
Foley, S.F., Tiepolo, M., Vannucci, R., 2002. Growth of early continental crust controlled by
melting of amphibolite in subduction zones. Nature 417, 837–840.
doi:10.1038/nature00792.1.
Frost, B.R., Barnes, C.G., Collins, W.J., Arculus, R.J., Ellis, D.J., Frost, C.D., 2001. A
Geochemical Classification for Granitic Rocks. J. Petrol. 42, 2033–2048.
doi:10.1093/petrology/42.11.2033
Gabriel, E.O., Oliveira, D.C., 2014. Geologia, petrografia e geoquímica dos granitoides
arqueanos de alto magnésio da região de Água Azul do Norte, porção sul do Domínio
Carajás, Pará. Bol. do Mus. Para. Emílio Goeldi - Série Ciências Nat. 9, 533–564.
Gabriel, E.O., Oliveira, D.C., Galarza, M.A., Santos, M.S. 2015. Geocronologia e aspectos
estruturais dos sanukitoids Mesoarqueanos da área de Água Azul do Norte: Implicações
para a história evolutiva da porção sul do Domínio Carajás. In: Simp. de Geol. da
Amazônia, 14, Marabá (PA), CD-Rom.
Gibbs, A.K., Wirth, K.R., Hirata, W.K., Olszewski Jr, W.J., 1986. Age and composition of the
Grão Pará groups volcanics, Serra dos Carajás. Rev. Bras. Geociências 16, 201–211.
Hamilton, W.B., 2011. Plate tectonics began in Neoproterozoic time, and plumes from deep
mantle have never operated. Lithos 123, 1–20. doi:10.1016/j.lithos.2010.12.007
152

Heaman, L.M., Creaser, R.A., Cookenboo, H.O., 2002. Extreme enrichment of high field
strength elements in Jericho eclogite xenoliths: A cryptic record of Paleoproterozoic
subduction, partial melting, and metasomatism beneath the Slave craton, Canada.
Geology 30, 507–510. doi:10.1130/0091-7613(2002)030<0507:EEOHFS>2.0.CO;2
Holdsworth, R.E., Pinheiro, R.V.L., 2000. The anatomy of shallow-crustal transpressional
structures: Insights from the Archaean Carajas fault zone, Amazon, Brazil. J. Struct.
Geol. 22, 1105–1123. doi:10.1016/S0191-8141(00)00036-5
Hooper, P., Rehacek, J., Morris, G., 1999. 10 . DATA REPORT : MAJOR AND TRACE
ELEMENT COMPOSITION , STRONTIUM , NEODYMIUM , AND OXYGEN
ISOTOPE RATIOS , AND MINERAL COMPOSITIONS OF SAMPLES 1, in: Larsen,
H.-C., Duncan, R.A., Allan, J.F., Brooks, K. (Eds.), Proceedings of the Ocean Drilling
Program, Scientific Results. pp. 113–117.
Irvine, T.N., Baragar, W.R.A., 1971. A Guide to the Chemical Classification of the Common
Volcanic Rocks. Can. J. Earth Sci. 8, 523–548. doi:10.1139/e71-055
James, S.D., Pearce, J.A., Oliver, R.A., 1987. The Geochemistry of the Lower Proterozoic
Willyama Complex Volcanics, Broken Hill Block, New South Wales. Geol. Soc.
London, Spec. Publ. 33, 395–408. doi:10.1144/GSL.SP.1987.033.01.27
Kerrich, R., Polat, A., 2006. Archean greenstone-tonalite duality: Thermochemical mantle
convection models or plate tectonics in the early Earth global dynamics? Tectonophysics
415, 141–165. doi:10.1016/j.tecto.2005.12.004
Kisters, A.F.M., Stevens, G., Dziggel, A., Armstrong, R.A., 2003. Extensional detachment
faulting and core-complex formation in the southern Barberton granite–green- stone
terrain, South Africa: evidence for a 3.2 Ga orogenic collapse. Precambrian Research
127, 355–378.
Laurent, O., Martin, H., Moyen, J.F., Doucelance, R., 2014. The diversity and evolution of
late-Archean granitoids: Evidence for the onset of “modern-style” plate tectonics
between 3.0 and 2.5Ga. Lithos 205, 208–235. doi:10.1016/j.lithos.2014.06.012
Laurie, A., Stevens, G., 2012. Water-present eclogite melting to produce Earth’s early felsic
crust. Chem. Geol. 314–317, 83–95. doi:10.1016/j.chemgeo.2012.05.001
Leite, A.A.S., Dall’Agnol, R., Macambira, M.J.B., Althoff, F.J., 2004. Geologia e
geocronologia dos granitóides arqueanos da região de Xinguara (PA) e suas implicac¸
ões na evoluc¸ ão do terreno granito-greenstone de Rio Maria. Revista Brasileira de
Geociências 34, 447–458.
153

Leite-Santos, P.J., 2016. Petrologia e geocronologia das associações leucograníticas


arqueanas da região de Água Azul do Norte (PA): implicações para a evolução crustal da
Província Carajás. University of Pará (UFPA).
Machado, N., Lindenmayer, Z., Krogh, T.E., Lindenmayer, D., 1991. U-Pb geochronology of
Archean magmatism and basement reactivation in the Carajás area, Amazon shield,
Brazil. Precambrian Res. 49, 329–354. doi:10.1016/0301-9268(91)90040-H
Marangoanha, B., 2016. Petrologia de Granitoides e Implicações para a Evolução Crustal
Arqueana da Porção Oeste do Domínio Canaã dos Carajás, Província Carajás. University
of Pará (UFPA).
Martin, H., Moyen, J.-F., Rapp, R., 2009. The sanukitoid series: magmatism at the Archaean–
Proterozoic transition. Earth Environ. Sci. Trans. R. Soc. Edinburgh 100, 15–33.
doi:10.1017/S1755691009016120
McDonough, W.F., Sun, S. -s., 1995. The composition of the Earth. Chem. Geol. 120, 223–
253. doi:10.1016/0009-2541(94)00140-4
Mints, M. V., Dokukina, K.A., Konilov, A.N., Kaulina, T. V., Belousova, E.A., Dokukin,
P.A., Natapov, L.M., Van, K. V., 2015. 2. Mesoarchean Kola-Karelia continent, in: East
European Craton: Early Precambrian History and 3D Models of Deep Crustal Structure.
Geological Society of America, pp. 15–88. doi:10.1130/2015.2510(02)
Moreto, C.P.N., Monteiro, L.V.S., Xavier, R.P., Amaral, W.S., dos Santos, T.J.S., Juliani, C.,
de Souza Filho, C.R., 2011. Mesoarchean (3.0 and 2.86 Ga) host rocks of the iron oxide–
Cu–Au Bacaba deposit, Carajás Mineral Province: U–Pb geochronology and
metallogenetic implications. Miner. Depos. 46, 789–811. doi:10.1007/s00126-011-0352-
9
Moreto, C.P.N., Monteiro, L.V.S., Xavier, R.P., Creaser, R.A., DuFrane, S.A., Melo, G.H.C.,
Delinardo da Silva, M.A., Tassinari, C.C.G., Sato, K., 2015. Timing of multiple
hydrothermal events in the iron oxide–copper–gold deposits of the Southern Copper
Belt, Carajás Province, Brazil. Miner. Depos. 50, 517–546. doi:10.1007/s00126-014-
0549-9
Moreto, C.P.N., Monteiro, L.V.S., Xavier, R.P., Creaser, R.A., DuFrane, S.A., Tassinari,
C.C.G., Sato, K., Kemp, A.I.S., Amaral, W.S., 2015. Neoarchean and Paleoproterozoic
Iron Oxide-Copper-Gold Events at the Sossego Deposit, Carajas Province, Brazil: Re-Os
and U-Pb Geochronological Evidence. Econ. Geol. 110, 809–835.
doi:10.2113/econgeo.110.3.809
154

Moyen, J.-F., 2011. The composite Archaean grey gneisses: Petrological significance, and
evidence for a non-unique tectonic setting for Archaean crustal growth. Lithos 123, 21–
36. doi:10.1016/j.lithos.2010.09.015
Moyen, J.-F., Martin, H., 2012. Forty years of TTG research. Lithos 148, 312–336.
doi:10.1016/j.lithos.2012.06.010
Moyen, J.-F., Stevens, G., 2006. Experimental constraints on TTG petrogenesis: Implications
for Archean geodynamics, in: Archean Geodynamics and Environments. pp. 149–175.
doi:10.1029/164GM11
Moyen, J.F., van Hunen, J., 2012. Short-term episodicity of Archaean plate tectonics.
Geology 40, 451–454. doi:10.1130/G322894.1
Pearce, J. a., Cann, J.R., 1973. Tectonic setting of basic volcanic rocks determined using trace
element analyses. Earth Planet. Sci. Lett. 19, 290–300. doi:10.1016/0012-
821X(73)90129-5
Pearce, J.A., 1996. A User’s Guide to Basalt Discrimination Diagrams. Geol. Assoc. Canada,
Short Course Notes.
Pearce, J.A., 2014. Immobile Element Fingerprinting of Ophiolites. Elements 10, 101–108.
doi:10.2113/gselements.10.2.101
Pearce, J.A., Peate, D.W., 1995. Tectonic implications of Volcanic Arc Magmas. Annu. Rev.
Earth Planet. Sci. 23, 251–285. doi:DOI: 10.1146/annurev.ea.23.050195.001343
Pidgeon, R.T., Macambira, M.J.B., Lafon, J.-M., 2000. Th–U–Pb isotopic systems and
internal structures of complex zircons from an enderbite from the Pium Complex,
Carajás Province, Brazil: evidence for the ages of granulite facies metamorphism and the
protolith of the enderbite. Chem. Geol. 159–171.
Pimentel, M. M., Machado, N. 1994. Geocronologia U-Pb dos terrenos granito-greenstone de
Rio Maria, Pará. Boletim de Resumos Expandidos do Congresso Brasileiro de Geologia
38 (2): 390-391
Rapp, R.P., Watson, E.B., Miller, C.F., 1991. Partial melting of amphibolite/eclogite and the
origin of Archean trondhjemites and tonalites. Precambrian Res. 51, 1–25.
Ricci, P.S.F., Carvalho, M.A. 2006. Rocks of the Pium-Area, Carajás Block, Brazil – A deep
seated high-T gabbroic pluton (charnockitoid-like) with xenoliths of enderbitic gneisses
dated at 3002 Ma – the basement problem revisited. In: 8º Simp. Geol. Amaz., Manaus,
Proceedings [CD-ROM].
155

Rodrigues, D.S., Oliveira, D.C., Macambira, M.J.B., 2014. Geologia, geoquímica e


geocronologia do Granito Mesoarqueano Boa Sorte, município de Água Azul do Norte,
Pará – Província Carajás. Bol. do Mus. Para. Emílio Goeldi - Série Ciências Nat. 9, 597–
633.
Rollinson, H.R., 1993. Using Geochemical Data: Evaluation, Presentation, Interpretation, 1st
ed. Pearson Education Limited, Harlow.
Rudnick, R.L., Gao, S., 2014. Composition of the Continental Crust, in: Treatise on
Geochemistry. Elsevier, pp. 1–51. doi:10.1016/B978-0-08-095975-7.00301-6
Sajeev, K., Osanai, Y., N, Y.K., Itaya, T., 2009. Stability of pargasite during ultrahigh-
temperature metamorphism: A consequence of titanium and REE partitioning? Am.
Mineral. 94, 535–545. doi:10.2138/am.2009.2815
Sajeev, K., Windley, B.F., Hegner, E., Komiya, T., 2013. High-temperature, high-pressure
granulites (retrogressed eclogites) in the central region of the Lewisian, NW Scotland:
Crustal-scale subduction in the Neoarchaean. Gondwana Res. 23, 526–538.
doi:10.1016/j.gr.2012.05.002
Santos, J.O.S., Hartmann, L.A., Gaudette, H.E., Groves, D.I., Mcnaughton, N.J., Fletcher,
I.R., 2000. A New Understanding of the Provinces of the Amazon Craton Based on
Integration of Field Mapping and U-Pb and Sm-Nd Geochronology. Gondwana Res. 3,
453–488.
Santos, P.A. dos, Feio, G.R.L., Dall’Agnol, R., Costi, H.T., Lamarão, C.N., Galarza, M.A.,
2013. Petrography, magnetic susceptibility and geochemistry of the Rio Branco Granite,
Carajás Province, southeast of Pará, Brazil. Brazilian J. Geol. 43, 2–15.
doi:10.5327/Z2317-48892013000100002
Santos, R.D., Galarza, M.A., Oliveira, D.C. de, 2013. Geologia , geoquímica e geocronologia
do Diopsídio-Norito Pium, Província Carajás. Bol. Mus. Para. Emílio Goeldi. Cienc.
Nat. 8, 355–382.
Sawyer, E.W., 2010. Migmatites formed by water-fluxed partial melting of a
leucogranodiorite protolith: Microstructures in the residual rocks and source of the fluid.
Lithos 116, 273–286. doi:10.1016/j.lithos.2009.07.003
Schand, S.J. (1943). Eruptive Rocks: their genesis, composition, classification and their
relation to ore deposits. 2nd edition, Hafner Publishing Co., New York
156

Silva, A.C., Dall’Agnol, R., Guimarães, F.V., Oliveira, D.C., 2014. Geologia, petrografia e
geoquímica de Associações Tonalíticas e Trondhjemíticas Arqueanas de Vila Jussara,
Província Carajás, Pará. Bol. do Mus. Para. Emílio Goeldi - Série Ciências Nat. 9, 13–
45.
Silva, G.G. da, Lima, M.I.C. de, Andrade, A.R.F. de, Issler, R.S., Guimarães, G., 1974.
Geologia da Folha SB.22 Araguaia e Parte da Folha SC.22 Tocantins, in: Luz, A.A. da,
Almeida, A.L.S. de, Netto, O.B. (Eds.), Projeto Radam: Levantamento de Recursos
Naturais, Folha SB.22 Araguaia e Parte Da Folha SC.22 Tocantins, Volume 4. Rio de
Janeiro, pp. 1–177.
Silva, L.R., de Oliveira, D.C., dos Santos, M.N.S., 2018. Diversity, origin and tectonic
significance of the Mesoarchean granitoids of Ourilândia do Norte, Carajás province
(Brazil). J. South Am. Earth Sci. 82, 33–61. doi:10.1016/j.jsames.2017.12.004
Silva, M. L. T., Oliveira, D. C., Macambira, M. J. B. 2010. Geologia, petrografia e
geocronologia do magmatismo de alto K da região de Vila Jussara, Água Azul do Norte
– Província Mineral de Carajás. Congresso Brasileiro de Geologia 45: 1 CD-ROM
Sizova, E., Gerya, T., Brown, M., Perchuk, L.L., 2010. Subduction styles in the Precambrian:
Insight from numerical experiments. Lithos 116, 209–229.
doi:10.1016/j.lithos.2009.05.028
Sousa, S. D., Oliveira. D. C., Gabriel E. O., Macambira, M. J. B. 2010. Geologia, petrografia
e geocronologia das rochas granitoides do Complexo Xingu da porção a leste da cidade
de Água Azul do Norte (PA) – Província Mineral de Carajás. Congresso Brasileiro de
Geologia 45: 1 CD-ROM.
Sousa, S.D. de, Monteiro, L.V.S., Oliveira, D.C. de, Delinardo da Silva, M.A., Moreto,
C.P.N., Juliani, C., 2015. O Greenstone Belt Sapucaia na região de Água Azul do Norte,
Província Carajás: Contexto geológico e caracterização petrográfica e geoquímica, in:
Gorayeb, P., Menguins, A. (Eds.), Contribuições a Geologia Da Amazônia. SBG-NO,
Belem, Pará, pp. 317–338.
Stevens, G., Clemens, J.D., 1993. Fluid-absent melting and the roles of fluids in the
lithosphere: a slanted summary? Chem. Geol. 108, 1–17. doi:10.1016/0009-
2541(93)90314-9
Sun, S. -s., McDonough, W.F., 1989. Chemical and isotopic systematics of oceanic basalts:
implications for mantle composition and processes. Geol. Soc. London, Spec. Publ. 42,
313–345. doi:10.1144/GSL.SP.1989.042.01.19
157

Tamura, Y., Ishizuka, O., Stern, R.J., Nichols, A.R.L., Kawabata, H., Hirahara, Y., Chang, Q.,
Miyazaki, T., Kimura, J.-I., Embley, R.W., Tatsumi, Y., 2014. Mission Immiscible:
Distinct Subduction Components Generate Two Primary Magmas at Pagan Volcano,
Mariana Arc. J. Petrol. 55, 63–101. doi:10.1093/petrology/egt061
van Hunen, J., van den Berg, A.P., 2008. Plate tectonics on the early Earth: Limitations
imposed by strength and buoyancy of subducted lithosphere. Lithos 103, 217–235.
doi:10.1016/j.lithos.2007.09.016
Van Kranendonk, M.J., 2011. Two types of Archean continental crust: Plume and plate
tectonics on early Earth. Am. J. Sci. 310, 1187–1209. doi:10.2475/10.2010.01
Vasquez, M.L., Rosa-Costa, L.T., 2008. Geologia e Recursos Minerais do Estado do Pará :
Sistema de Informações Geográficas - SIG: texto explicativo dos mapas Geológico e
Tectônico e de Recursos Minerais do Estado do Pará. Belém.
Vielzeuf, D., Schmidt, M.W., 2001. Melting relations in hydrous systems revisited:
application to metapelites, metagreywackes and metabasalts. Contrib. to Mineral. Petrol.
141, 251–267. doi:10.1007/s004100100237
Winchester, J.A., Floyd, P.A., 1977. Geochemical discrimination of different magma series
and their differentiation products using immobile elements. Chem. Geol. 20, 325–343.
doi:10.1016/0009-2541(77)90057-2
Wirth, K.R., Gibbs, A.K., Olszewski Jr., W.J., 1986. U-Pb ages of zircons from the Grão-Pará
Group and Serra dos Carajás Granites, Pará, Brazil. Rev. Bras. Geociências 16, 195–200.
158

6. CONSIDERAÇÕES FINAIS

a) As rochas metamórficas do embasamento do Domínio Carajás são


representadas por ortognaisses de afinidade TTG com enclaves de metabasitos.
Os protólitos destas rochas cristalizaram entre ca. 3,06–2,93 Ga. Os dados
geoquímicos apontam para um contexto de arco magmático, no qual a fusão da
crosta ocênica em subducção desempenhou papel mais importante na geração
das rochas. Minoritariamente, algumas rochas com baixa razão LaN/YbN
sugerem contribuição de crosta máfica inferior. Os dados das rocha máficas
reforçam a sugestão de um magmatismo cálcio-alcalino típico de arco
magmático.
b) Os ortognaisses TTG e metabasitos das unidades Ortogranulito Xicrim-Cateté
e Complexo Xingu foram afetados por um metamorfismo de alto grau durante
a colagem dos blocos Carajás e Rio Maria entre ca. 2,89 e ca. 2,85 Ga. Nesse
contexto a base da crosta registrou metamorfismo de ultra-alta temperatura,
associado a reações de fusão relacionadas com a quebra de anfibólios
pargasíticos. Esse processo, em combinação com a colocação de granitoides de
derivação mantélica (sanukitoides; ca. 2,87 Ga) catalizou reações de fusão
assistida por fluidos que tornam a migmatização expressiva na região.
c) A migmatização modificou o comportamento reológico da crosta do Domínio
Carajás, levando a uma deformação heterogênea associada ao incremento do
volume de fundido. A migração e acumulação do fundido na base da crosta
superior levou ao destacamento da base da crosta e desenvolvimento das
grandes zonas de cisalhamento E–W do Cinturão Itacaiúnas (ca. 2,8 Ga). Isso
permitiu a rápida exumação do terreno e o “congelamento” das estruturas
formadas durante o metamorfismo progressivo de alto grau com fusão parcial.
159

REFERÊNCIAS BIBLIOGRÁFICAS

Abbott, D., Burgess, L., Longhi, J., Smith, W.H.F., 1994. An empirical thermal history of the
Earth’s upper mantle. J. Geophys. Res. 99, 13835. doi:10.1029/94JB00112
Almeida, J. de A.C. de, Dall’Agnol, R., de Oliveira, M.A., Macambira, M.J.B., Pimentel,
M.M., Rämö, O.T., Guimarães, F.V., Leite, A.A. da S., 2011. Zircon geochronology,
geochemistry and origin of the TTG suites of the Rio Maria granite-greenstone terrane:
Implications for the growth of the Archean crust of the Carajás province, Brazil.
Precambrian Res. 187, 201–221. doi:10.1016/j.precamres.2011.03.004
Almeida, J. de A.C. de, Oliveira, V.E.S., Rocha, M.C., Rocha, K.P.P., 2016. Geologia,
Petrografia e Geoquímica dos Granitoides Arqueanos da Porção Sul do Domínio Rio
Maria, Província Carajás, in: SBG (Ed.), Anais Do 46o Congresso Brasileiro de
Geologia. SBG - Sociedade Brasileira de Geologia, Porto Alegre, RS.
Amaldev, T., Santosh, M., Tang, L., Baiju, K.R., Tsunogae, T., Satyanarayanan, M., 2016.
Mesoarchean convergent margin processes and crustal evolution: Petrologic,
geochemical and zircon U–Pb and Lu–Hf data from the Mercara Suture Zone, southern
India. Gondwana Res. 37, 182–204. doi:10.1016/j.gr.2016.05.017
Araújo, O.J.B., Maia, R.G.N., 1991. Programa Levantamento Geológicos Básicos do Brasil.
Programa Grande Carajás. Folha SB.22-Z-A. Estado do Pará. Texto Explicativo.
Brasilia.
Avelar, V.G., Lafon, J.M., Correia Jr., F.C., Macambira, E.M.B., 1999. O magmatismo
arqueano da região de Tucumã - Província Mineral De Carajás: Novos resultados
geocronológicos. Rev. Bras. Geociências 29, 453–460.
Batuk Joshi, K., Bhattacharjee, J., Rai, G., Halla, J., Ahmad, T., Kurhila, M., Heilimo, E.,
Choudhary, A.K., 2017. The diversification of granitoids and plate tectonic implications
at the Archaean–Proterozoic boundary in the Bundelkhand Craton, Central India. Geol.
Soc. London, Spec. Publ. 449, 123–157. doi:10.1144/SP449.8
Berman, R.G. 1991. Thermobarometry using multi-equilibrium calculations: A new
technique, with petrological applications. Canadian Mineralogist, 29 (4): 833-855.
Berman, R.G. 1991. Thermobarometry using multi-equilibrium calculations: A new
technique, with petrological applications. Canadian Mineralogist, 29 (4): 833-855.
Blundy, J.D., Holland, T.J.B., 1990. Calcic amphibole equilibria and a new amphibole-
plagioclase geothermometer. Contrib. to Mineral. Petrol. 104, 208–224.
doi:10.1007/BF00306444
160

Brown, M., 2006. Duality of thermal regimes is the distinctive characteristics of plate
tectonics since the Neoarchean. Geology 34, 961–964. doi:10.1130/G22853A.1
Brown, M., 2009. Metamorphic patterns in orogenic systems and the geological record. Geol.
Soc. London, Spec. Publ. 318, 37–74. doi:10.1144/SP318.2
Costa, U.A.P., Paula, R.R., Silva, D.P.B., Barbosa, J.P.O., Silva, C.M.G., Tavares, F.M.,
Oliveira, J.K.M., Justo, A.P., 2016. Programa geologia do Brasil-PGB. Mapa de
integração geológico-geofísica da ARIM Carajás. Escala 1:250.000 Estado do Pará.
Belém, CPRM.
Dall’Agnol, R., Teixeira, N.P., Rämö, O.T., Moura, C.A. V, Macambira, M.J.B., de Oliveira,
D.C., 2005. Petrogenesis of the Paleoproterozoic rapakivi A-type granites of the Archean
Carajás metallogenic province, Brazil. Lithos 80, 101–129.
doi:10.1016/j.lithos.2004.03.058
DOCEGEO, 1988. Revisão litoestratigráfica da Província Mineral de Carajás, in: CVRD,
SBG (Eds.), Anais Do 35o Congresso Brasileiro de Geologia. Sociedade Brasileira de
Geologia (SBG), Belém, PA, pp. 11–54.
Feio, G.R.L., Dall’Agnol, R., Dantas, E.L., Macambira, M.J.B., Gomes, A.C.B., Sardinha,
A.S., Oliveira, D.C., Santos, R.D., Santos, P.A., 2012. Geochemistry, geochronology,
and origin of the Neoarchean Planalto Granite suite, Carajás, Amazonian craton: A-type
or hydrated charnockitic granites? Lithos 151, 57–73. doi:10.1016/j.lithos.2012.02.020
Feio, G.R.L., Dall’Agnol, R., Dantas, E.L., Macambira, M.J.B., Santos, J.O.S., Althoff, F.J.,
Soares, J.E.B., 2013. Archean granitoid magmatism in the Canaã dos Carajás area:
Implications for crustal evolution of the Carajás province, Amazonian craton, Brazil.
Precambrian Res. 227, 157–185. doi:10.1016/j.precamres.2012.04.007
Gabriel, E.O., Oliveira, D.C. de, 2013. Petrologia magnética dos granodioritos Água Azul e
Água Limpa, porção sul do Domínio Carajás - Pará. Geol. USP. Série Científica 13, 89–
110. doi:10.5327/Z1519-874X201300040005
Gabriel, E.O., Oliveira, D.C., Galarza, M.A., Santos, M.S. 2015. Geocronologia e aspectos
estruturais dos sanukitoids Mesoarqueanos da área de Água Azul do Norte: Implicações
para a história evolutiva da porção sul do Domínio Carajás. In: Simp. de Geol. da
Amazônia, 14, Marabá (PA), CD-Rom.
Gerya, T. V., Stern, R.J., Baes, M., Sobolev, S. V., Whattam, S.A., 2015. Plate tectonics on
the Earth triggered by plume-induced subduction initiation. Nature 527, 221–225.
doi:10.1038/nature15752
161

Gibbs, A.K., Wirth, K.R., Hirata, W.K., Olszewski Jr, W.J., 1986. Age and composition of the
Grão Pará groups volcanics, Serra dos Carajás. Rev. Bras. Geociências 16, 201–211.
Guitreau, M., Mukasa, S.B., Loudin, L., Krishnan, S., 2017. New constraints on the early
formation of the Western Dharwar Craton (India) from igneous zircon U-Pb and Lu-Hf
isotopes. Precambrian Res. 302, 33–49. doi:10.1016/j.precamres.2017.09.016
Hamilton, W.B., 2011. Plate tectonics began in Neoproterozoic time, and plumes from deep
mantle have never operated. Lithos 123, 1–20. doi:10.1016/j.lithos.2010.12.007
Holland, T.J.B., Powell, R., 1998. An internally consistent thermodynamic data set for phases
of petrological interest. J. Metamorph. Geol. 16, 309–343. doi:10.1111/j.1525-
1314.1998.00140.x
Kerrich, R., Polat, A., 2006. Archean greenstone-tonalite duality: Thermochemical mantle
convection models or plate tectonics in the early Earth global dynamics? Tectonophysics
415, 141–165. doi:10.1016/j.tecto.2005.12.004
Labrosse, S., Jaupart, C., 2007. Thermal evolution of the Earth: Secular changes and
fluctuations of plate characteristics. Earth Planet. Sci. Lett. 260, 465–481.
doi:10.1016/j.epsl.2007.05.046
LaFlamme, C., McFarlane, C.R.M., Corrigan, D., 2014. U-Pb, Lu-Hf and REE in zircon from
3.2 to 2.6 Ga Archean gneisses of the Repulse Bay block, Melville Peninsula, Nunavut.
Precambrian Res. 252, 223–239. doi:10.1016/j.precamres.2014.07.012
Laurent, O., Martin, H., Moyen, J.F., Doucelance, R., 2014. The diversity and evolution of
late-Archean granitoids: Evidence for the onset of “modern-style” plate tectonics
between 3.0 and 2.5Ga. Lithos 205, 208–235. doi:10.1016/j.lithos.2014.06.012
Laurent, O., Zeh, A., 2015. A linear Hf isotope-age array despite different granitoid sources
and complex Archean geodynamics: Example from the Pietersburg block (South Africa).
Earth Planet. Sci. Lett. 430, 326–338. doi:10.1016/j.epsl.2015.08.028
Leite, A.A.S., Dall’Agnol, R., Macambira, M.J.B., Althoff, F.J., 2004. Geologia e
geocronologia dos granitóides arqueanos da região de Xinguara (PA) e suas implicac¸
ões na evoluc¸ ão do terreno granito-greenstone de Rio Maria. Revista Brasileira de
Geociências 34, 447–458.
Leite-Santos, P.J., 2016. Petrologia e geocronologia das associações leucograníticas
arqueanas da região de Água Azul do Norte (PA): implicações para a evolução crustal da
Província Carajás. University of Pará (UFPA).
162

Ludwig, K.R., 2003. Mathematical-Statistical Treatment of Data and Errors for 230Th/U
Geochronology. Rev. Mineral. Geochemistry 52, 631–656. doi:10.2113/0520631
Machado, N., Lindenmayer, Z., Krogh, T.E., Lindenmayer, D., 1991. U-Pb geochronology of
Archean magmatism and basement reactivation in the Carajás area, Amazon shield,
Brazil. Precambrian Res. 49, 329–354. doi:10.1016/0301-9268(91)90040-H
Martin, H., Moyen, J.-F., Rapp, R., 2009. The sanukitoid series: magmatism at the Archaean–
Proterozoic transition. Earth Environ. Sci. Trans. R. Soc. Edinburgh 100, 15–33.
doi:10.1017/S1755691009016120
Molina, J.F., Moreno, J.A., Castro, A., Rodríguez, C., Fershtater, G.B., 2015. Calcic
amphibole thermobarometry in metamorphic and igneous rocks: New calibrations based
on plagioclase/amphibole Al-Si partitioning and amphibole/liquid Mg partitioning.
Lithos 232, 286–305. doi:10.1016/j.lithos.2015.06.027
Moreto, C.P.N., 2013. Geocronologia U-Pb e Re-Os aplicada à evolução metalógenética do
Cinturão Sul do Cobre da Província Mineral de Carajás. University of Campinas
(UNICAMP).
Moreto, C.P.N., Monteiro, L.V.S., Xavier, R.P., Amaral, W.S., dos Santos, T.J.S., Juliani, C.,
de Souza Filho, C.R., 2011. Mesoarchean (3.0 and 2.86 Ga) host rocks of the iron oxide–
Cu–Au Bacaba deposit, Carajás Mineral Province: U–Pb geochronology and
metallogenetic implications. Miner. Depos. 46, 789–811. doi:10.1007/s00126-011-0352-
9
Moyen, J.-F., 2011. The composite Archaean grey gneisses: Petrological significance, and
evidence for a non-unique tectonic setting for Archaean crustal growth. Lithos 123, 21–
36. doi:10.1016/j.lithos.2010.09.015
Moyen, J.-F., Martin, H., 2012. Forty years of TTG research. Lithos 148, 312–336.
doi:10.1016/j.lithos.2012.06.010
Moyen, J.-F., Stevens, G., 2006. Experimental constraints on TTG petrogenesis: Implications
for Archean geodynamics, in: Archean Geodynamics and Environments. pp. 149–175.
doi:10.1029/164GM11
O’Brien, P.J., Rötzler, J., 2003. High-pressure granulites: Formation, recovery of peak
conditions and implications for tectonics. J. Metamorph. Geol. 21, 3–20.
doi:10.1046/j.1525-1314.2003.00420.x
163

Oriolo, S., Oyhantçabal, P., Basei, M.A.S., Wemmer, K., Siegesmund, S., 2016. The Nico
Pérez Terrane (Uruguay): From Archean crustal growth and connections with the Congo
Craton to late Neoproterozoic accretion to the Río de la Plata Craton. Precambrian Res.
280, 147–160. doi:10.1016/j.precamres.2016.04.014
Paton, C., Woodhead, J.D., Hellstrom, J.C., Hergt, J.M., Greig, A., Maas, R., 2010. Improved
laser ablation U-Pb zircon geochronology through robust downhole fractionation
correction. Geochemistry, Geophys. Geosystems 11, n/a-n/a.
doi:10.1029/2009GC002618
Perchuk, L.L., Gerya, T. V., 2011. Formation and evolution of Precambrian granulite
terranes: A gravitational redistribution model. Geol. Soc. Am. Mem. 289–310.
doi:10.1130/2011.1207(15)
Petrus, J.A., Kamber, B.S., 2012. VizualAge: A Novel Approach to Laser Ablation ICP-MS
U-Pb Geochronology Data Reduction. Geostand. Geoanalytical Res. 36, 247–270.
doi:10.1111/j.1751-908X.2012.00158.x
Pidgeon, R.T., Macambira, M.J.B., Lafon, J.-M., 2000. Th–U–Pb isotopic systems and
internal structures of complex zircons from an enderbite from the Pium Complex,
Carajás Province, Brazil: evidence for the ages of granulite facies metamorphism and the
protolith of the enderbite. Chem. Geol. 159–171.
Pimentel, M. M., Machado, N. 1994. Geocronologia U-Pb dos terrenos granito-greenstone de
Rio Maria, Pará. Boletim de Resumos Expandidos do Congresso Brasileiro de Geologia
38 (2): 390-391
Pinheiro, R.V.L., Holdsworth, R.E., 2000. Evolução tectonoestratigráfica dos sistemas
transcorrentes Carajás e Cinzento, Cinturão Itacaiúnas, na borda leste do Cráton
Amazônico, Pará. Rev. Bras. Geociências 30, 597–606.
Ricci, P.S.F., Carvalho, M.A. 2006. Rocks of the Pium-Area, Carajás Block, Brazil – A deep
seated high-T gabbroic pluton (charnockitoid-like) with xenoliths of enderbitic gneisses
dated at 3002 Ma – the basement problem revisited. In: 8º Simp. Geol. Amaz., Manaus,
Proceedings [CD-ROM].
Rodrigues, D.S., Oliveira, D.C., Macambira, M.J.B., 2014. Geologia, geoquímica e
geocronologia do Granito Mesoarqueano Boa Sorte, município de Água Azul do Norte,
Pará – Província Carajás. Bol. do Mus. Para. Emílio Goeldi - Série Ciências Nat. 9, 597–
633.
164

Rollinson, H.R., 1993. Using Geochemical Data: Evaluation, Presentation, Interpretation, 1st
ed. Pearson Education Limited, Harlow.
Sajeev, K., Osanai, Y., N, Y.K., Itaya, T., 2009. Stability of pargasite during ultrahigh-
temperature metamorphism: A consequence of titanium and REE partitioning? Am.
Mineral. 94, 535–545. doi:10.2138/am.2009.2815
Sajeev, K., Windley, B.F., Hegner, E., Komiya, T., 2013. High-temperature, high-pressure
granulites (retrogressed eclogites) in the central region of the Lewisian, NW Scotland:
Crustal-scale subduction in the Neoarchaean. Gondwana Res. 23, 526–538.
doi:10.1016/j.gr.2012.05.002
Santos, J.O.S., 2003. Geotectônica dos Escudos das Guianas e Brasil-Central, in: Bizzi, L.A.,
Schobbenhaus, C., Vidotti, R.M., Gonçalves, J.H. (Eds.), Geologia, Tectônica e
Recursos Minerais Do Brasil. CPRM - Serviço Geológico do Brasil, Brasilia, pp. 169–
195.
Santos, J.O.S., Hartmann, L.A., Gaudette, H.E., Groves, D.I., Mcnaughton, N.J., Fletcher,
I.R., 2000. A New Understanding of the Provinces of the Amazon Craton Based on
Integration of Field Mapping and U-Pb and Sm-Nd Geochronology. Gondwana Res. 3,
453–488.
Santos, P.A. dos, Feio, G.R.L., Dall’Agnol, R., Costi, H.T., Lamarão, C.N., Galarza, M.A.,
2013. Petrography, magnetic susceptibility and geochemistry of the Rio Branco Granite,
Carajás Province, southeast of Pará, Brazil. Brazilian J. Geol. 43, 2–15.
doi:10.5327/Z2317-48892013000100002
Santos, R.D., Galarza, M.A., Oliveira, D.C. de, 2013. Geologia , geoquímica e geocronologia
do Diopsídio-Norito Pium, Província Carajás. Bol. Mus. Para. Emílio Goeldi. Cienc.
Nat. 8, 355–382.
Silva, A.C., Dall’Agnol, R., Guimarães, F.V., Oliveira, D.C., 2014. Geologia, petrografia e
geoquímica de Associações Tonalíticas e Trondhjemíticas Arqueanas de Vila Jussara,
Província Carajás, Pará. Bol. do Mus. Para. Emílio Goeldi - Série Ciências Nat. 9, 13–
45.
Silva, G.G. da, Lima, M.I.C. de, Andrade, A.R.F. de, Issler, R.S., Guimarães, G., 1974.
Geologia da Folha SB.22 Araguaia e Parte da Folha SC.22 Tocantins, in: Luz, A.A. da,
Almeida, A.L.S. de, Netto, O.B. (Eds.), Projeto Radam: Levantamento de Recursos
Naturais, Folha SB.22 Araguaia e Parte Da Folha SC.22 Tocantins, Volume 4. Rio de
Janeiro, pp. 1–177.
165

Silva, L.R., de Oliveira, D.C., dos Santos, M.N.S., 2018. Diversity, origin and tectonic
significance of the Mesoarchean granitoids of Ourilândia do Norte, Carajás province
(Brazil). J. South Am. Earth Sci. 82, 33–61. doi:10.1016/j.jsames.2017.12.004
Silva, M. L. T., Oliveira, D. C., Macambira, M. J. B. 2010. Geologia, petrografia e
geocronologia do magmatismo de alto K da região de Vila Jussara, Água Azul do Norte
– Província Mineral de Carajás. Congresso Brasileiro de Geologia 45: 1 CD-ROM
Sizova, E., Gerya, T., Brown, M., Perchuk, L.L., 2010. Subduction styles in the Precambrian:
Insight from numerical experiments. Lithos 116, 209–229.
doi:10.1016/j.lithos.2009.05.028
Sousa, S. D., Oliveira. D. C., Gabriel E. O., Macambira, M. J. B. 2010. Geologia, petrografia
e geocronologia das rochas granitoides do Complexo Xingu da porção a leste da cidade
de Água Azul do Norte (PA) – Província Mineral de Carajás. Congresso Brasileiro de
Geologia 45: 1 CD-ROM.
Sousa, S.D. de, Monteiro, L.V.S., Oliveira, D.C. de, Delinardo da Silva, M.A., Moreto,
C.P.N., Juliani, C., 2015. O Greenstone Belt Sapucaia na região de Água Azul do Norte,
Província Carajás: Contexto geológico e caracterização petrográfica e geoquímica, in:
Gorayeb, P., Menguins, A. (Eds.), Contribuições a Geologia Da Amazônia. SBG-NO,
Belem, Pará, pp. 317–338.
van Hunen, J., van den Berg, A.P., 2008. Plate tectonics on the early Earth: Limitations
imposed by strength and buoyancy of subducted lithosphere. Lithos 103, 217–235.
doi:10.1016/j.lithos.2007.09.016
Vasquez, M.L., Rosa-Costa, L.T., 2008. Geologia e Recursos Minerais do Estado do Pará :
Sistema de Informações Geográficas - SIG: texto explicativo dos mapas Geológico e
Tectônico e de Recursos Minerais do Estado do Pará. Belém.
WIEDENBECK, M., ALLÉ, P., CORFU, F., GRIFFIN, W.L., MEIER, M., OBERLI, F.,
QUADT, A. VON, RODDICK, J.C., SPIEGEL, W., 1995. THREE NATURAL
ZIRCON STANDARDS FOR U-TH-PB, LU-HF, TRACE ELEMENT AND REE
ANALYSES. Geostand. Geoanalytical Res. 19, 1–23. doi:10.1111/j.1751-
908X.1995.tb00147.x
Williams I.S., 1998, U-Th-Pb geochronology by ion microprobe. In: McKibben, M.A.
Shanks, W.C. and Ridley W.I. (eds) Applications of Microanalytical Techniques to
Understanding Mineralizing Processes. Reviews in Economic Geology 7: 1-35.
166

Williams I.S., 1998, U-Th-Pb geochronology by ion microprobe. In: McKibben, M.A.
Shanks, W.C. and Ridley W.I. (eds) Applications of Microanalytical Techniques to
Understanding Mineralizing Processes. Reviews in Economic Geology 7: 1-35.
Wirth, K.R., Gibbs, A.K., Olszewski Jr., W.J., 1986. U-Pb ages of zircons from the Grão-Pará
Group and Serra dos Carajás Granites, Pará, Brazil. Rev. Bras. Geociências 16, 195–200.

Você também pode gostar