Acessibilidade / Reportar erro

Amphibole crystallization conditions as record of interaction between ultrapotassic enclaves and monzonitic magmas in the Glória Norte Stock, South of Borborema Province

Abstract

The Glória Norte Stock (GNS) is made up of predominantly porphyritic biotite- amphibole-bearing quartz monzonite, enclosing a large number of microgranular mafic enclaves (MME). The GNS MME are fine-grained rocks with rounded and ellipsoid shapes that show sharp and gradational contacts with the host rocks, suggesting they coexisted with the host monzonite as magmas. The studied amphibole crystals of the two rock types are calcic and correspond to pargasite, edenite, and magnesium-hornblende. Their compositions are influenced by substitutions involving Al3+, Na+, Fe2+, Si4+ and Mg2+ ions, reflecting the decrease in temperature and the increase in oxygen fugacity. They present low Ti and Al contents, and varied Si content (6.2-7.7 apuf). The mg# values range from 0.48-0.84. Pressures for the crystallization of the MME amphiboles vary between 2.6 and 7.8 kbar and the temperatures of solidus and liquidus were estimated between 600-659 and 887-908ºC, respectively. The early MME amphibole crystals with lower Si and Mg content were formed under high pressure (7.8 kbar) and temperature (908 °C). The presence of amphiboles with 862ºC and 5 kbar in MME and GNS reflects that the interaction between these magmas occurred at 18 km of depth.

KEYWORDS:
mineral chemistry; amphibole thermobarometry; Sergipano Orogenic System

INTRODUCTION

Amphiboles constitute a complex group of rock-forming minerals, which present large compositional variation and occur in several mineral associations. Their structures and chemical compositions are affected by changes in conditions during or after their crystallization (Hawthorne 1983Hawthorne F.C. 1983. The crystal chemistry of the amphiboles. Canadian Mineralogist, 21(2):173-480.). Amphiboles occur in igneous rocks, from acidic to ultrabasic, and in medium- to high-temperature metamorphic rocks (Deer et al. 1997Deer W.A., Howie R.A., Zussman J. 1997. Rock-Forming Minerals. 2B. Double-Chain Silicates. London, Geological Society of London, 764 p., Martin 2007Martin R.F. 2007. Amphiboles in the igneous environment. Reviews in Mineralogy and Geochemistry, 67(1):323-358. https://doi.org/10.2138/rmg.2007.67.9
https://doi.org/https://doi.org/10.2138/...
). Additionally, the occurrence of amphibole in mafic-ultramafic enclaves associated with granites and basalts indicates that it is stable in the mantle (e.g., Green and Ringwood 1963Green D.H., Ringwood A.E. 1963. Mineral assemblages in a model mantle composition. Journal of Geophysical Research, 68(3):937-945. https://doi.org/10.1029/JZ068i003p00937
https://doi.org/https://doi.org/10.1029/...
, Krawczynski et al. 2012Krawczynski M.J., Grove T.L., Behrens H. 2012. Amphibole stability in primitive arc magmas: effects of temperature, H2O content, and oxygen fugacity. Contributions to Mineralogy and Petrology, 164:317-339. https://doi.org/10.1007/s00410-012-0740-x
https://doi.org/https://doi.org/10.1007/...
).

Granites make up around 25% of the outcrop area of the Macururé Domain (MD), in the Sergipano Orogenic System (SOS), (Davison and Santos 1989Davison I., Santos R.A. 1989. Tectonic evolution of the Sergipano fold belt, NE Brazil, during the Brasiliano orogeny. Precambrian Research, 45(4):319-342. https://doi.org/10.1016/0301-9268(89)90068-5
https://doi.org/https://doi.org/10.1016/...
). Conceição et al. (2016Conceição J.A., Rosa M.L.S., Conceição H. 2016. Sienogranitos leucocráticos do Domínio Macururé, Sistema Orogênicos Sergipano, Nordeste do Brasil: Stock Glória Sul. Brazilian Journal of Geology, 46(1):63-77. https://doi.org/10.1590/2317-4889201620150044
https://doi.org/https://doi.org/10.1590/...
) grouped the MD intrusions into four distinct groups:

  • High-K calc-alkaline granodiorites;

  • Shoshonitic monzonites;

  • Leucocratic syenogranites;

  • Rapakivi-textured calc-alkaline granites.

Microgranular mafic enclaves (MME) are common in the SOS granites and some of them show high amphibole contents (Sial et al. 1998Sial A.N., Ferreira V.P., Fallick A.E., Cruz M.J.M. 1998. Amphibole-rich clots in calc-alkalic granitoids in the Borborema province, northeastern Brazil. Journal of South American Earth Sciences, 11(5):457-471. https://doi.org/10.1016/S0895-9811(98)00034-0
https://doi.org/https://doi.org/10.1016/...
). The genesis of amphibole enclaves in the MD has been ascribed to the mixing between felsic and mafic magmas (Sial et al. 1998Sial A.N., Ferreira V.P., Fallick A.E., Cruz M.J.M. 1998. Amphibole-rich clots in calc-alkalic granitoids in the Borborema province, northeastern Brazil. Journal of South American Earth Sciences, 11(5):457-471. https://doi.org/10.1016/S0895-9811(98)00034-0
https://doi.org/https://doi.org/10.1016/...
, Conceição et al. 2016Conceição J.A., Rosa M.L.S., Conceição H. 2016. Sienogranitos leucocráticos do Domínio Macururé, Sistema Orogênicos Sergipano, Nordeste do Brasil: Stock Glória Sul. Brazilian Journal of Geology, 46(1):63-77. https://doi.org/10.1590/2317-4889201620150044
https://doi.org/https://doi.org/10.1590/...
, Lisboa et al. 2019Lisboa V.A.C., Conceição H., Rosa M.L.S., Fernandes D.M. 2019. The onset of post-collisional magmatism in the Macururé Domain, Sergipano Orogenic System: The Glória Norte Stock. Journal of South American Earth Sciences, 89:173-188. https://doi.org/10.1016/j.jsames.2018.11.005
https://doi.org/https://doi.org/10.1016/...
). Interaction between magmas with distinct compositions can generate total or partial mixing, and partial corrosion textures in early formed crystals (Barbarin and Didier 1992Barbarin B., Didier J. 1992. Genesis and evolution of mafic microgranular enclaves through various types of interaction between coexisting felsic and mafic magmas. Earth and Environmental Science Transactions of the Royal Society of Edinburgh, 83(1-2):145-153. https://doi.org/10.1017/S0263593300007835
https://doi.org/https://doi.org/10.1017/...
). The minerals formed during mixing may have their chemical composition modified or rebalanced due to reactions with the hybrid magma (Şahin et al. 2010Şahin S.Y., Orgun Y., Gungor Y., Goker A.F., Gultekin A.H., Karacik Z. 2010. Mineral and Whole-rock Geochemistry of the Kestanbol Granitoid (Ezine-Çanakkale) and its Mafic Microgranular Enclaves in Northwestern Anatolia: Evidence of Felsic and Mafic Interaction. Turkish Journal of Earth Sciences, 19:101-122. https://doi.org/10.3906/yer-0809-3
https://doi.org/https://doi.org/10.3906/...
).

This work presents and discusses the result of the chemical analysis obtained on amphibole crystals of microgranular mafic enclaves of the Glória Norte Stock (GNS), their textural relations and crystallization conditions.

GEOLOGICAL CONTEXT

The SOS defines the N-NE limit of the São Francisco Craton (SFC) (Fig. 1). The formation of the SOS is attributed to the neoproterozoic collision between the SFC and the Pernambuco-Alagoas Superterraine (PEAL; Davison and Santos 1989Davison I., Santos R.A. 1989. Tectonic evolution of the Sergipano fold belt, NE Brazil, during the Brasiliano orogeny. Precambrian Research, 45(4):319-342. https://doi.org/10.1016/0301-9268(89)90068-5
https://doi.org/https://doi.org/10.1016/...
). According to Oliveira et al. (2010Oliveira E.P., Windley B.F., Araujo M.N.C. 2010. The Neoproterozoic Sergipano orogenic belt, NE Brazil: a complete plate tectonic cycle in western Gondwana. Precambrian Research, 181(1-4):64-84. https://doi.org/10.1016/j.precamres.2010.05.014
https://doi.org/https://doi.org/10.1016/...
), the convergence between SFC and PEAL led to the deformation of passive margins and granitic plutons emplacement in the central-northern portion of the SOS, during 628-621 and 590-570 Ma. These granites have calc-alkaline and shoshonitic signatures (Silva Filho et al. 1992Silva Filho A.F., Guimarães I.P., Silva M.R.R. 1992. Utilização de granitóides na definição de domínios tectônicos; aplicação no lado oriental do Sistema de Dobramento Sergipano. In: Congresso Brasileiro de Geologia, 37., 1992. Boletim de Resumos Expandidos., Conceição et al. 2016Conceição J.A., Rosa M.L.S., Conceição H. 2016. Sienogranitos leucocráticos do Domínio Macururé, Sistema Orogênicos Sergipano, Nordeste do Brasil: Stock Glória Sul. Brazilian Journal of Geology, 46(1):63-77. https://doi.org/10.1590/2317-4889201620150044
https://doi.org/https://doi.org/10.1590/...
, Lisboa et al. 2019Lisboa V.A.C., Conceição H., Rosa M.L.S., Fernandes D.M. 2019. The onset of post-collisional magmatism in the Macururé Domain, Sergipano Orogenic System: The Glória Norte Stock. Journal of South American Earth Sciences, 89:173-188. https://doi.org/10.1016/j.jsames.2018.11.005
https://doi.org/https://doi.org/10.1016/...
).

Figure 1.
(A) Geological compartmentation of the Sergipano Orogenic System (Pinho Neto et al. 2019Pinho Neto M.A., Rosa M.L.S., Conceição H. 2019. Petrologia do Batólito Sítios Novos, Sistema Orogênico Sergipano, Província Borborema, NE do Brasil. Geologia USP. Série Científica, 19(2):135-150. https://doi.org/10.11606/issn.2316-9095.v19-152469
https://doi.org/https://doi.org/10.11606...
). (B) Geological sketch map of the Macururé Domain proposed by Conceição et al. 2016Conceição J.A., Rosa M.L.S., Conceição H. 2016. Sienogranitos leucocráticos do Domínio Macururé, Sistema Orogênicos Sergipano, Nordeste do Brasil: Stock Glória Sul. Brazilian Journal of Geology, 46(1):63-77. https://doi.org/10.1590/2317-4889201620150044
https://doi.org/https://doi.org/10.1590/...
. (C) Geological scheme of the Glória Norte Stock (Lisboa et al. 2019Lisboa V.A.C., Conceição H., Rosa M.L.S., Fernandes D.M. 2019. The onset of post-collisional magmatism in the Macururé Domain, Sergipano Orogenic System: The Glória Norte Stock. Journal of South American Earth Sciences, 89:173-188. https://doi.org/10.1016/j.jsames.2018.11.005
https://doi.org/https://doi.org/10.1016/...
).

The GNS, which was crystallized at 588 ± 5 Ma (zircon U-PbSHRIMP) is the main representative of the MD post-orogenic shoshonitic magmatism (Lisboa et al. 2019Lisboa V.A.C., Conceição H., Rosa M.L.S., Fernandes D.M. 2019. The onset of post-collisional magmatism in the Macururé Domain, Sergipano Orogenic System: The Glória Norte Stock. Journal of South American Earth Sciences, 89:173-188. https://doi.org/10.1016/j.jsames.2018.11.005
https://doi.org/https://doi.org/10.1016/...
). This stock has an area of 45 km2 (Fig. 1) and is essentially composed of porphyritic quartz-monzonite. Biotite, amphibole, and diopside are the mafic minerals present in these rocks. In quartz-monzonites, the enclaves and feldspar phenocrysts mark a magmatic flow foliation. The GNS contains abundant mafic enclaves, comprising three different types: microgranular mafic, lamprophyric, and cumulatic ultramafic.

The lamprophyric enclaves occur in the midwestern portion of the GNS. They are dark colored with varied shape and their length is up to 1.3 meters, with sharp contacts with quartz-monzonite. The diopside biotite cumulate enclaves (Fontes et al. 2018Fontes M.P., Conceição H.C., Rosa M.L.S., Lisboa V.A.C. 2018. Minettes do Stock monzonítico Glória Norte: Evidência de magmatismo ultrapotássico pós-orogênico, com assinatura de subducção, no Sistema Orogênico Sergipano. Geologia USP - Série Científica, 18(1):1-16. http://dx.doi.org/10.11606/issn.2316-9095.v18-133599
https://doi.org/http://dx.doi.org/10.116...
) are medium-grained rocks with olive green color and rounded shape.

MME are abundant in the GNS (Fig. 2A). They have different sizes and shapes (mostly rounded and ellipsoidal), often with the longest axis orientated by the magmatic flow. The contacts of these enclaves with the host quartz monzonitic rocks are gradational or sharp and their cores are more coarse-grained than the border (Fig. 2B). The decreasing grain size toward the edge of the MME is interpreted as resulting from the rapid cooling of the mafic magma droplets, due to temperature contrast with the felsic magma. Color gradation forms darker to lighter MME types, reflecting the mafic mineral contents. This variation is interpreted with different degrees of mixing between magmas. In the felsic magma, the interaction is less pervasive, limited to the contact (a few centimeters) with the mafic magma.

Figure 2.
Field aspects of the studied enclaves. (A) Dark-colored MME with ellipsoidal shape. Note the finer grain size as compared to the host porphyritic monzonite. The white arrow marks the feldspar xenocrysts. (B) Microgranular mafic enclave correspond to alkali feldspar syenite and is composed of biotite and amphibole with a finer-grained border (with darker coloration) as well as feldspar xenocrysts (white arrow).

Lisboa et al. (2019Lisboa V.A.C., Conceição H., Rosa M.L.S., Fernandes D.M. 2019. The onset of post-collisional magmatism in the Macururé Domain, Sergipano Orogenic System: The Glória Norte Stock. Journal of South American Earth Sciences, 89:173-188. https://doi.org/10.1016/j.jsames.2018.11.005
https://doi.org/https://doi.org/10.1016/...
) characterized the MME as ultrapotassic alkali feldspar syenite and monzonite rocks (44 < %SiO2 < 47, MgO > 3%, K2O > 3%, and K2O/Na2O > 2). According to these authors, the MME of GNS rocks correspond to intermediate magma, while the composition of the most primitive microgranular mafic enclaves correspond to mafic magmas. The low Ta, Nb, and Ti contents, which characterize orogenic magmas, are characteristics of both GSN and MME samples.

MATERIALS AND METHODS

The investigation was executed in polished thin sections of representative samples from the host quartz monzonite and mafic enclaves. The petrographic study was carried out with the aid of a transmitted and reflected light petrographic microscope from the Microscopy Laboratory at the Geoscience Laboratories, Universidade Federal de Sergipe (CLGeo-UFS).

The chemical compositions of the amphibole crystals were obtained from analysis through wavelength dispersion (WDS) and dispersive energy (EDS). The analyses with WDS were performed with the JEOL JXA-8230 electronic probe from the Microanalysis Laboratory, Universidade Federal do Pará (LABMEV-UFPA). Constant accelerating voltage of 15kV, electron beam current of 20 nA, electron beam diameter of 10 µm, with counting time for the analysis of major and minor elements of 20 and 40s, respectively. The following calibration standards were used for WDS analysis: fluorite (F), sodalite (Na2O), diopside (MgO), anortite (Al2O3), orthoclase (SiO2, K2O), wollastonite (CaO), celestine (SrO), magnetite (FeO), rhodonite (MnO), barite (BaO), rutile (TiO2), sodalite (Cl), vanadinite (V2O3), Cr2O3 (Cr2O3), and NiO (NiO).

EDS analysis was performed with an Oxford Instrument® X-Act model EDS attached to a Tescan® Scanning Electron Microscope (SEM) Vega 3 LMU model at the Laboratory of Microanalysis, Universidade Federal de Sergipe (CLGeo-UFS). Potential acceleration of 20 kV and current intensity of 17 nA, generating a beam with a diameter of 400 nm; analysis time was 30 s. The CLGeo’s EDS-SEM is calibrated with copper energy and the reliability of the analyses is assessed from the analysis of ASTIMEX international mineral standards (Tab. 1). The analytical uncertainties, given by the difference between the certified values and those obtained with the EDS, were less than 2%wt for the elements with content greater than 10%wt in weight and errors between 4 and 19%wt for elements with content below 5% in weight.

Table 1.
Comparison between the composition of ASTIMEX reference standards and the results obtained through the EDS, adapted from Leandro (2018Leandro M.V.S. 2018. Mineraloquímica de Rochas do Batólito Sienítico Itabuna, Sul do Estado da Bahia. Monografia, Departamento de Geologia, Universidade Federal de Sergipe, 67 p.).

RESULTS

Petrographic aspects

The MME studied correspond to alkali feldspar syenites and monzonites that exhibit fine- to medium-grained or equigranular porphyritic texture. The essential mineralogy of these enclaves is: amphibole, biotite, plagioclase (andesine and oligoclase), alkali feldspar and quartz (Fig. 3). Accessory minerals are diopside, epidote, allanite, titanite, apatite, zircon, ilmenite, magnetite, pyrite, and chalcopyrite.

Figure 3.
Representative texture of the microgranular mafic enclaves of the Glória Norte Stock. It is possible to observe a fine- to medium-grained granular rock.

Green amphibole is the dominant mafic phase in the MME (up to 53% volume). Most crystals are subhedral with a length of up to 1.5 mm. Pleochroism ranges between olive green (Y’), yellowish green (Z’) to light yellowish green (X’). The grain limits with most minerals are rectilinear and irregular to interpenetrating with biotite and diopside. It shows inclusions of anhedral quartz, subhedral biotite, as well as apatite and ilmenite euhedral grains. Green amphibole is replaced by colorless amphibole at the borders.

The amphibole crystals are well distributed in the MME. They occur as grains up to 0.5 mm, as subhedral to anhedral crystals in mafic agglomerates, as well as euhedral inclusions in alkali feldspar and, more rarely, in plagioclase.

Mafic minerals such as amphibole, diopside, titanite, ilmenite, and magnetite, occur as agglomerates with up to 3.0 mm in length. Diopside is anhedral and rimmed by dark green amphibole (Fig. 4A). The limits between diopside and amphibole grains are diffuse and the diopside cleavage planes contain subhedral grains of ilmenite and magnetite. The association of amphibole with biotite is more common. In this case, biotite occurs as subhedral crystals with size up to 0.8 mm, arranged between amphibole crystals (Fig. 4B) or along their cleavage planes.

Figure 4.
Photomicrography of the GNS microgranular mafic enclaves. (A) Contact relationships between amphibole and diopside grains. Also note the presence of subhedral amphibole and stubby apatite crystals included in alkali feldspar; (B) Subhedral biotite crystals between amphibole crystals.

Lisboa et al. (2019Lisboa V.A.C., Conceição H., Rosa M.L.S., Fernandes D.M. 2019. The onset of post-collisional magmatism in the Macururé Domain, Sergipano Orogenic System: The Glória Norte Stock. Journal of South American Earth Sciences, 89:173-188. https://doi.org/10.1016/j.jsames.2018.11.005
https://doi.org/https://doi.org/10.1016/...
) described abundant crystals of stubby apatite and mafic mineral agglomerates, composed by amphibole, biotite, and diopside, in the GNS porphyritic quartz monzonite. Baxter and Feely (2002Baxter S., Feely M. 2002. Magma mixing and mingling textures in granitoids: examples from the Galway Granite, Connemara, Ireland. Mineralogy and Petrology, 76:63-74. https://doi.org/10.1007/s007100200032
https://doi.org/https://doi.org/10.1007/...
) propose that agglomerates of this nature are formed from a reaction between pyroxene and magma, thus leading to crystallization of amphibole. The occurrence of these agglomerates in the studied rocks suggests that they originated from mafic magma, which was responsible for the formation of MME, and their presence in the GNS enclaves and monzonites is interpreted as due to the interaction between felsic and mafic magmas, or could be only due to equilibrium crystallization of the quartz monzonite magma.

Mineral chemistry

Nomenclature and chemical composition

Chemical classification of the amphiboles and structural formula were based on 23 oxygens, assuming the general formula A0-1B2C5T8O22(OH, F, Cl), whose site occupation complied with the recommendations of Leake et al. (1997Leake B.E., Woolley A.R., Arps C.E., Birch W.D., Gilbert M.C., Grice J.D., Hawthorne F.C., Kato A., Kisch H.J., Krivovichev V.G., Linthout K., Laird J., Mandarino J. 1997. Report. Nomenclature of amphiboles: report of the subcommittee on amphiboles of the international mineralogical association commission on new minerals and mineral names. Mineralogical Magazine, 61:295-321.). In total, 60 amphibole crystals were analyzed. Representative chemical analyses of these crystals are presented in Table 2.

Table 2.
Representative chemical analyses of the studied amphiboles. The Fe2+/Fe3+ partition was made using the Stout’s method (1972Stout J.H. 1972. Phase petrology and mineral chemistry of coexisting amphiboles from Telemark, Norway. Journal of Petrology, 13(1):99-145. https://doi.org/10.1093/petrology/13.1.99
https://doi.org/https://doi.org/10.1093/...
). H2O* was calculated by stoichiometry.

The MME amphibole crystals are calcic, with CaB content between 1.5-2.0 apuf (atoms per unit formula), and can be divided into two groups (Figs. 5A and 5B):

  • (Na + K)A > 0.5 apuf (pargasite and edenita);

  • (Na + K)A < 0.5 apuf (magnesium-hornblende and actinolite).

Figure 5.
Classification diagrams for the studied calcic amphiboles, after Leake et al. (1997Leake B.E., Woolley A.R., Arps C.E., Birch W.D., Gilbert M.C., Grice J.D., Hawthorne F.C., Kato A., Kisch H.J., Krivovichev V.G., Linthout K., Laird J., Mandarino J. 1997. Report. Nomenclature of amphiboles: report of the subcommittee on amphiboles of the international mineralogical association commission on new minerals and mineral names. Mineralogical Magazine, 61:295-321.). (1) The composition of the crystal core corresponds to the circle, and (2) the rim corresponds to the diamond. (3) The grey area represents amphiboles from the Glória Norte Stock.

These amphiboles show mg# [Mg2+/(Mg2+ + Fe2+)] ratios ranging from 0.48-0.84. TiO2 (0.0-2.0%), Al2O3 (3.9-12.3%), FeO (5.0-16.0%), MgO (9.0-15.4%), MnO (0.0-0.4%), CaO (10.0-12.4%), Na2O (0.5-1.9%), and K2O (0.2-3.8%). The intermediate contents of SiO2 and low TiO2 and alkali (Na2O) contents are similar to those of amphiboles found in rocks formed in orogenic environments (Coltorti et al. 2007Coltorti M., Bonadiman C., Faccini B., Grégoire M., O'Reilly S.Y., Powell W. 2007. Amphiboles from suprasubduction and intraplate lithospheric mantle. Lithos, 99(1-2):68-84. https://doi.org/10.1016/j.lithos.2007.05.009
https://doi.org/https://doi.org/10.1016/...
, Martin 2007Martin R.F. 2007. Amphiboles in the igneous environment. Reviews in Mineralogy and Geochemistry, 67(1):323-358. https://doi.org/10.2138/rmg.2007.67.9
https://doi.org/https://doi.org/10.2138/...
).

In some amphibole crystals, the increase in Si and Mg associated with the depletion in IVAl from nucleus to border, form actinolite (Fig. 5B). The existence of such zoning may be a consequence of the increased oxygen fugacity in the magma, or it may also result from the reaction between magmatic amphibole and late fluids (Kawakatsu and Yamaguchi 1987Kawakatsu K., Yamaguchi Y. 1987. Successive zoning of amphiboles during progressive oxidation in the Daito-Yokota granitic complex, San-in belt, southwest Japan. Geochimica et Cosmochimica Acta, 51(3):535-540. https://doi.org/10.1016/0016-7037(87)90067-6
https://doi.org/https://doi.org/10.1016/...
, Martin 2007Martin R.F. 2007. Amphiboles in the igneous environment. Reviews in Mineralogy and Geochemistry, 67(1):323-358. https://doi.org/10.2138/rmg.2007.67.9
https://doi.org/https://doi.org/10.2138/...
). Similar zoning is also described as evidence of mixing between magmas by Sato et al. (1999Sato H., Nakada S., Fujii T., Nakamura M., Suzuki-Kamata K. 1999. Groundmass pargasite in the 1991-1995 dacite of Unzen volcano: phase stability experiments and volcanological implications. Journal of Volcanology and Geothermal Research, 89(1):197-212. https://doi.org/10.1016/S0377-0273(98)00132-2
https://doi.org/https://doi.org/10.1016/...
).

Calcium amphiboles are also important mafic minerals in the GNS monzonites, where they reach up to 14% of the volume (Lisboa et al. 2019Lisboa V.A.C., Conceição H., Rosa M.L.S., Fernandes D.M. 2019. The onset of post-collisional magmatism in the Macururé Domain, Sergipano Orogenic System: The Glória Norte Stock. Journal of South American Earth Sciences, 89:173-188. https://doi.org/10.1016/j.jsames.2018.11.005
https://doi.org/https://doi.org/10.1016/...
). Analyzes from representative samples of the GNS revealed that amphibole crystals from this stock correspond to edenite (Fig. 5A) and to magnesium-hornblende (Fig. 5B). There is an overlap among samples of the GSN and MME.

Leake (1971Leake B.E. 1971. On aluminous and edenitic hornblendes. Mineralogical Magazine, 38(296):389-407. https://doi.org/10.1180/minmag.1971.038.296.01
https://doi.org/https://doi.org/10.1180/...
, 1978Leake B.E. 1978. Nomenclature of amphiboles. The Canadian Mineralogist, 16(4):501-520.) used the relationship between Si and (Na + Ca + K), to distinguish magmatic from post-magmatic amphiboles. Using these criteria, it is clear that some compositions in the MME amphiboles evolve from the magmatic to the post-magmatic field (Fig. 6). Crystals with higher Si content (> 7.1 apuf) are also enriched in Mg, depleted in Na + Ca + K (< 2.4 apuf) and present similar compositional variation to that observed by Yamaguchi (1985Yamaguchi Y. 1985. Hornblende-cummingtonite and hornblende-aclinolite intergowths from the Koyama calc-alkaline intrusion, Susa, southwest Japan. American Mineralogist, 70(9-10):980-986.) for late-crystallized crystals, which had their chemical compositions partially modified due to reactions with magma, or formed under subsolidus conditions in the presence of fluids (Chivas 1982Chivas A.R. 1982. Gechemical evidence for magmatic fluids in porphyry copper mineralization. Part I. Mafic silicates from the Koloula Igneous Complex. Contributions to Mineralogy and Petrology, 78:389-403. https://doi.org/10.1007/BF00375201
https://doi.org/https://doi.org/10.1007/...
, Hendry et al. 1985Hendry D.A.F., Chivas A.R., Long J.V.P., Reed S.J.B. 1985. Chemical differences between minerals from mineralizing and barren intrusions from some North American porphyry copper deposits. Contributions to Mineralogy and Petrology, 89:317-329. https://doi.org/10.1007/BF00381554
https://doi.org/https://doi.org/10.1007/...
).

Figure 6.
Si versus Na + Ca + K diagram of Leake (1971Leake B.E. 1971. On aluminous and edenitic hornblendes. Mineralogical Magazine, 38(296):389-407. https://doi.org/10.1180/minmag.1971.038.296.01
https://doi.org/https://doi.org/10.1180/...
) applied to the studied MME amphibole. The curve separates the compositions of magmatic from post-magmatic amphiboles.

Substitutions

Spear (1981Spear F.S. 1981. An experimental study of hornblende stability and compositional variability in amphibolite. American Journal of Science, 281:697-734. https://doi.org/10.2475/ajs.281.6.697
https://doi.org/https://doi.org/10.2475/...
) listed the most common amphibole substitutions. The compositional variations in the MME studied crystals can be described in terms of pargasite, tschermakite, and Ti-tschermakite substitutions, in addition to the simple replacement of Fe2+ by Mg (Spear 1981Spear F.S. 1981. An experimental study of hornblende stability and compositional variability in amphibolite. American Journal of Science, 281:697-734. https://doi.org/10.2475/ajs.281.6.697
https://doi.org/https://doi.org/10.2475/...
).

The diagram in Figure 7A shows a good positive correlation between (NaA + KA) and IVAl for MME amphiboles. The relationship shown in this diagram is close to the 2:1 ratio, implying the predominance of pargasite substitution, indicating that the partial occupation of site A by alkalis requires the replacement of Si for IVAl at the tetrahedral site. The highest vacancy rates at site [A] are observed for the crystals with higher Mg and Si contents. The IVAl versus VIAl ratio (Fig. 7B) demonstrates the importance of edenite and pargasite substitutions in the evolution of the studied MME amphibole compositions. The inverse behavior of Fe and Mg suggests important participation in the simple substitution Fe+2 ↔ Mg, and also shows the good negative correlation existent between these elements (Fig. 7C).

Figure 7.
Chemical substitutions in the amphiboles, microgranular mafic enclaves of the Glória Norte Stock. (A) IVAl and the occupancy rate of Site “A”; (B) IVAl and VIAl; (C) Fe+2 and Mg. Substitutions are named according to Spear (1981Spear F.S. 1981. An experimental study of hornblende stability and compositional variability in amphibolite. American Journal of Science, 281:697-734. https://doi.org/10.2475/ajs.281.6.697
https://doi.org/https://doi.org/10.2475/...
).

Conditions of crystallization

Several studies used amphibole composition to infer pressure (e.g., Schmidt 1992Schmidt M.W. 1992. Amphibole composition in tonalite as a function of pressure: an experimental calibration of the Al-in-hornblende barometer. Contributions to Mineralogy and Petrology , 110:304-310. https://doi.org/10.1007/BF00310745
https://doi.org/https://doi.org/10.1007/...
, Mutch et al. 2016Mutch E.J.F., Blundy J.D., Tattitch B.C., Cooper F.J., Brooker R.A. 2016. An experimental study of amphibole stability in low-pressure granitic magmas and a revised Al-in-hornblende geobarometer. Contributions to Mineralogy and Petrology, 171:85. https://doi.org/10.1007/s00410-016-1298-9
https://doi.org/https://doi.org/10.1007/...
), temperature (e.g., Holland and Blundy 1994Holland T., Blundy J. 1994. Non-ideal interactions in calcic amphiboles and their bearing on amphibole-plagioclase thermometry. Contributions to Mineralogy and Petrology, 116:433-447. https://doi.org/10.1007/BF00310910
https://doi.org/https://doi.org/10.1007/...
, Putirka 2016Putirka K. 2016. Amphibole thermometers and barometers for igneous systems and some implications for eruption mechanisms of felsic magmas at arc volcanoes. American Mineralogist, 101(4):841-858. https://doi.org/10.2138/am-2016-5506
https://doi.org/https://doi.org/10.2138/...
), and oxygen fugacity (e.g., Anderson and Smith 1995Anderson J.L., Smith D.R. 1995. The effects of temperature and fO2 on the Al-in-hornblende barometer. American Mineralogist, 80(5-6):549-559. https://doi.org/10.2138/am-1995-5-614
https://doi.org/https://doi.org/10.2138/...
) magma conditions of crystallization. In calcium amphiboles, with increased pressure and temperature, K, Na, Ti, and Al contents are generally increased, although Al is more sensitive to changes in pressure (Hammarstron and Zen 1986Hammarstron J.M., Zen E. 1986. Aluminium in hornblende: an empirical igneous geobarometer. American Mineralogist, 71:1297-1313., Féménias et al. 2006Féménias O., Mercier J.C.C., Nkono C., Diot H., Berza T., Tatu M., Demaiffe D. 2006. Calcic amphibole growth and compositions in calc-alkaline magmas: Evidence from the Motru Dike Swarm (Southern Carpathians, Romania). American Mineralogist, 91(1):73-81. https://doi.org/10.2138/am.2006.1869
https://doi.org/https://doi.org/10.2138/...
, Nandedkar et al. 2014Nandedkar R.H., Ulmer P., Müntener O. 2014. Fractional crystallization of primitive hydrous arc magmas: an experimental study at 0.7 GPa. Contributions to Mineralogy and Petrology, 167:1015. https://doi.org/10.1007/s00410-014-1015-5
https://doi.org/https://doi.org/10.1007/...
).

The main calcium amphibole geobarometers are based on the existence of a linear relationship between totalAl and crystallization pressure (Anderson 1996Anderson J.L. 1996. Status of thermobarometry in granitic batholiths. Earth and Environmental Science Transactions of the Royal Society of Edinburgh, 87(1-2):125-138. https://doi.org/10.1017/S0263593300006544
https://doi.org/https://doi.org/10.1017/...
). The identification of the amphibole in association with biotite, titanite, magnetite (and/or ilmenite), quartz, alkaline feldspar, and plagioclase in the studied MME allowed the use of these geobarometers. Pressure values calculated for the studied amphiboles by different algorithms are presented in Table 3.

Table 3.
Geobarometry data obtained for the studied MME amphiboles. P1 (Hammarstron and Zen 1986Hammarstron J.M., Zen E. 1986. Aluminium in hornblende: an empirical igneous geobarometer. American Mineralogist, 71:1297-1313.); P2 (Hollister et al. 1987Hollister L.S., Grissom G.C., Peters E.K., Stowell H.H., Sisson V.B. 1987. Confirmation of the empirical correlation of Al in hornblende with pressure of solidification of calc-alkaline plutons. American Mineralogist, 72(3-4):231-239.); P3 (Johnson and Rutherford 1989Johnson M.C., Rutherford M.J. 1989. Experimental calibration of the aluminum-in-hornblende geobarometer with application to Long Valley caldera (California) volcanic rocks. Geology, 17(9):837-841. https://doi.org/10.1130/0091-7613(1989)017%3C0837:ECOTAI%3E2.3.CO;2
https://doi.org/https://doi.org/10.1130/...
); P4 (Schmidt 1992Schmidt M.W. 1992. Amphibole composition in tonalite as a function of pressure: an experimental calibration of the Al-in-hornblende barometer. Contributions to Mineralogy and Petrology , 110:304-310. https://doi.org/10.1007/BF00310745
https://doi.org/https://doi.org/10.1007/...
); P5 (Anderson and Smith 1995Anderson J.L., Smith D.R. 1995. The effects of temperature and fO2 on the Al-in-hornblende barometer. American Mineralogist, 80(5-6):549-559. https://doi.org/10.2138/am-1995-5-614
https://doi.org/https://doi.org/10.2138/...
); P6 (Mutch et al. 2016Mutch E.J.F., Blundy J.D., Tattitch B.C., Cooper F.J., Brooker R.A. 2016. An experimental study of amphibole stability in low-pressure granitic magmas and a revised Al-in-hornblende geobarometer. Contributions to Mineralogy and Petrology, 171:85. https://doi.org/10.1007/s00410-016-1298-9
https://doi.org/https://doi.org/10.1007/...
).

The Altotal content in the analyzed crystals ranges from 1.42 to 2.22 apuf. Considering all calibrations, the pressures (P) obtained to magmatic amphiboles range from 2.6 to 7.8 kbar, corresponding to high- and low-pressure amphiboles (Fig. 8A), according to Hammarstron and Zen (1986Hammarstron J.M., Zen E. 1986. Aluminium in hornblende: an empirical igneous geobarometer. American Mineralogist, 71:1297-1313.). The identification of both high- and low-pressure amphiboles in orogenesis-associated magmas is described in several papers (e.g., Samaniego et al. 2010Samaniego P., Robin C., Chazot G., Bourdon E., Cotten J. 2010. Evolving metasomatic agent in the Northern Andean subduction zone, deduced from magma composition of the long-lived Pichincha volcanic complex (Ecuador). Contributions to Mineralogy and Petrology , 160(2):239-260. https://doi.org/10.1007/s00410-009-0475-5
https://doi.org/https://doi.org/10.1007/...
, Ribeiro et al. 2016Ribeiro J.M., Maury R.C., Grégoire M. 2016. Are Adakites slab melts or high-pressure fractionated mantle melts? Journal of Petrology, 57(5):839-862. https://doi.org/10.1093/petrology/egw023
https://doi.org/https://doi.org/10.1093/...
).

Figure 8.
Diagrams showing the variation of chemical parameters and the conditions of pressure and oxygen fugacity for the crystallization of calcium amphiboles. (A) IVAl versus Altotal diagram with high and low pressure, amphibole fields defined by Hammarstron and Zen (1986Hammarstron J.M., Zen E. 1986. Aluminium in hornblende: an empirical igneous geobarometer. American Mineralogist, 71:1297-1313.). (B) Fe/(Fe + Mg) versus IVAl diagram, showing fields with different conditions of oxygen fugacity (ƒO2), according to Anderson and Smith (1995Anderson J.L., Smith D.R. 1995. The effects of temperature and fO2 on the Al-in-hornblende barometer. American Mineralogist, 80(5-6):549-559. https://doi.org/10.2138/am-1995-5-614
https://doi.org/https://doi.org/10.2138/...
). The composition of grain cores corresponds to the circle; the grey area represents compositions of amphibole from monzonites of the Glória Norte Stock.

As demonstrated by Anderson and Smith (1995Anderson J.L., Smith D.R. 1995. The effects of temperature and fO2 on the Al-in-hornblende barometer. American Mineralogist, 80(5-6):549-559. https://doi.org/10.2138/am-1995-5-614
https://doi.org/https://doi.org/10.2138/...
), the increase of Altotal may also be related to the increase in temperature. In this case, these authors recommended the use of barometers that consider the effect of temperature. When applying the calibration proposed by Anderson and Smith (1995Anderson J.L., Smith D.R. 1995. The effects of temperature and fO2 on the Al-in-hornblende barometer. American Mineralogist, 80(5-6):549-559. https://doi.org/10.2138/am-1995-5-614
https://doi.org/https://doi.org/10.2138/...
) in the studied MME amphiboles, it is seen that the pressure values range from 3.8 to 6.6 kbar, similarly or very close to those obtained using other calibrations (Tab. 3).

The MME solidus temperature estimates were obtained from the amphibole-plagioclase pair thermometry, which is based on the amount of IVAl in the amphibole coexisting with plagioclase in silica-saturated rocks. The results can be seen in Table 4.

Table 4.
Temperature data obtained for the studied MME amphiboles.

The crystallization temperature for the studied MME amphiboles uses the calibration made by Holland and Blundy (1994Holland T., Blundy J. 1994. Non-ideal interactions in calcic amphiboles and their bearing on amphibole-plagioclase thermometry. Contributions to Mineralogy and Petrology, 116:433-447. https://doi.org/10.1007/BF00310910
https://doi.org/https://doi.org/10.1007/...
) for the edenite-richterite reaction. The temperatures were calculated based on the pressures obtained by the Schmidt (1992Schmidt M.W. 1992. Amphibole composition in tonalite as a function of pressure: an experimental calibration of the Al-in-hornblende barometer. Contributions to Mineralogy and Petrology , 110:304-310. https://doi.org/10.1007/BF00310745
https://doi.org/https://doi.org/10.1007/...
) and Anderson and Smith (1995Anderson J.L., Smith D.R. 1995. The effects of temperature and fO2 on the Al-in-hornblende barometer. American Mineralogist, 80(5-6):549-559. https://doi.org/10.2138/am-1995-5-614
https://doi.org/https://doi.org/10.2138/...
) barometers. The temperature values obtained from this geothermometer are situated between 600 and 659ºC (Tab. 4), below the temperature of the granite solidus, which is around 700ºC (Piwinskii 1968Piwinskii A.J. 1968. Experimental Studies of Igneous Rock Series Central Sierra Nevada Batholith, California. The Journal of Geology, 76(5):548-570. https://doi.org/10.1086/627359
https://doi.org/https://doi.org/10.1086/...
, McDowell and Willye 1971McDowell S.D., Willye P.J. 1971. Experimental studies of igneous rock series: the Kungnat complex of southwest Greenland. Journal of Geology, 79(2):173-194. https://doi.org/10.1086/627607
https://doi.org/https://doi.org/10.1086/...
).

The composition of the cores of the MME pargasite crystals was used to estimate the temperature using the empirical geothermometer of Ridolfi et al. (2010Ridolfi F., Renzulli A., Puerini M. 2010. Stability and chemical equilibrium of amphibole in calc-alkaline magmas: an overview, new thermobarometric formulations and application to subduction-related volcanoes. Contributions to Mineralogy and Petrology , 160:45-66. https://doi.org/10.1007/s00410-009-0465-7
https://doi.org/https://doi.org/10.1007/...
). The temperatures obtained range from 887 to 908ºC (Tab. 4).

The fO2 has a strong chemical control over major mafic phases in igneous rocks (Anderson and Smith 1995Anderson J.L., Smith D.R. 1995. The effects of temperature and fO2 on the Al-in-hornblende barometer. American Mineralogist, 80(5-6):549-559. https://doi.org/10.2138/am-1995-5-614
https://doi.org/https://doi.org/10.2138/...
). The studied amphiboles show Fe/(Fe + Mg) ratios below 0.6. When these ratios are correlated to the IVAl content of the MME and monzonite amphibole, they indicate crystallization at high fO2 (Fig. 8B). The presence of the magnetite + titanite paragenesis in the MME reinforces the hypothesis of crystallization under oxidizing conditions (Wones 1989Wones D.R. 1989. Significance of the assemblage titanite+ magnetite+ quartz in granitic rocks. American Mineralogist, 74(7-8):744-749.).

The mean pressure values calculated for GNS monzonite amphiboles are approximately 5 kbar, indicating crystallization at around 18 km depth. Using the Holland and Blundy (1994Holland T., Blundy J. 1994. Non-ideal interactions in calcic amphiboles and their bearing on amphibole-plagioclase thermometry. Contributions to Mineralogy and Petrology, 116:433-447. https://doi.org/10.1007/BF00310910
https://doi.org/https://doi.org/10.1007/...
) and Ridolfi et al. (2010Ridolfi F., Renzulli A., Puerini M. 2010. Stability and chemical equilibrium of amphibole in calc-alkaline magmas: an overview, new thermobarometric formulations and application to subduction-related volcanoes. Contributions to Mineralogy and Petrology , 160:45-66. https://doi.org/10.1007/s00410-009-0465-7
https://doi.org/https://doi.org/10.1007/...
) geothermometers, the temperature of amphibole crystallization values obtained range between 615-747 and 772-862ºC, respectively.

DISCUSSION

Amphiboles are very useful for the investigation of mafic magma genesis, especially the ultrapotassic ones, in which amphibole is an early crystallizing phase (Foley 1992Foley S.F. 1992. Petrological characterization of the source components of potassic magmas: geochemical and experimental constraints. Lithos, 28(3-6):187-204. https://doi.org/10.1016/0024-4937(92)90006-K
https://doi.org/https://doi.org/10.1016/...
, Conceição and Green 2004Conceição R.V., Green D.H. 2004. Derivation of potassic (shoshonitic) magmas by decompression melting of phlogopite+pargasite lherzolite. Lithos, 72(3-4):209-229. https://doi.org/10.1016/j.lithos.2003.09.003
https://doi.org/https://doi.org/10.1016/...
, Förster et al. 2017Förster M.W., Prelević D., Schmück H.R., Buhre S., Veter M., Mertz-Kraus R., Foley S.F., Jacob D.E. 2017. Melting and dynamic metasomatism of mixed harzburgite+ glimmerite mantle source: Implications for the genesis of orogenic potassic magmas. Chemical Geology, 455:182-191. http://dx.doi.org/10.1016/j.chemgeo.2016.08.037
https://doi.org/http://dx.doi.org/10.101...
). Compositional variation in amphibole is strongly affected by changes in pressure, temperature, oxidation conditions, magma composition, and water activity (Scaillet and Evans 1999Scaillet B., Evans B.W. 1999. The 15 June 1991 eruption of Mount Pinatubo. I. Phase equilibria and pre-eruption P-T-fO2-fH2O conditions of the dacite magma. Journal of Petrology, 40(3):381-411. https://doi.org/10.1093/petroj/40.3.381
https://doi.org/https://doi.org/10.1093/...
, Dalpé and Baker 2000Dalpé C., Baker D.R. 2000. Experimental investigation of large-ion-lithophile-element-, high-field-strength-element and rare-earth-element-partitioning between calcic amphibole and basaltic melt: the effects of pressure and oxygen fugacity. Contributions to Mineralogy and Petrology, 140:233-250. https://doi.org/10.1007/s004100000181
https://doi.org/https://doi.org/10.1007/...
, Rooney et al. 2011Rooney T.O., Franceschi P., Hall C.M. 2011. Water-saturated magmas in the Panama Canal region: a precursor to adakite-like magma generation? Contributions to Mineralogy and Petrology , 161:373-388. https://doi.org/10.1007/s00410-010-0537-8
https://doi.org/https://doi.org/10.1007/...
).

Experimental studies performed on amphiboles which crystallized in equilibrium with primitive basaltic magma have shown high mg# values (0.83-0.75; Grove et al. 2003Grove T.L., Elkins-Tanton L.T., Parman S.W., Chatterjee N., Muntener O., Gaetani G.A. 2003. Fractional crystallization and mantle-melting controls on calc-alkaline differentiation trends. Contributions to Mineralogy and Petrology, 145(5):515-533. https://doi.org/10.1007/s00410-003-0448-z
https://doi.org/https://doi.org/10.1007/...
) as compared to amphiboles equilibrated with more evolved magmas, where lower values are observed (mg# ≤ 0.65; Alonso-Perez et al. 2009Alonso-Perez R., Müntener O., Ulmer P. 2009. Igneous garnet and amphibole fractionation in the roots of island arcs: experimental constraints on andesitic liquids. Contributions to Mineralogy and Petrology, 157:541-558. https://doi.org/10.1007/s00410-008-0351-8
https://doi.org/https://doi.org/10.1007/...
). The amphiboles of the MME show mg# ranging between 0.48 and 0.84; these values correspond to those observed in primitive basaltic magmas and magmatic reequilibrium. This suggests that this mineral phase is a product of the early crystallization of these magmas (Hidalgo and Rooney 2010Hidalgo P.J., Rooney T.O. 2010. Crystal fractionation processes at Baru volcano from the deep to shallow crust. Geochemistry, Geophysics, Geosystems, 11(12):1-29. https://doi.org/10.1029/2010GC003262
https://doi.org/https://doi.org/10.1029/...
, Rooney et al. 2011Rooney T.O., Franceschi P., Hall C.M. 2011. Water-saturated magmas in the Panama Canal region: a precursor to adakite-like magma generation? Contributions to Mineralogy and Petrology , 161:373-388. https://doi.org/10.1007/s00410-010-0537-8
https://doi.org/https://doi.org/10.1007/...
). The lower mg# values (below 0.65) of the studied MME amphiboles may be related to the physicochemical changes in the magmatic chamber (i.e., P, T, fO2) or crystallization in equilibrium with more evolved magmas (Ribeiro et al. 2016Ribeiro J.M., Maury R.C., Grégoire M. 2016. Are Adakites slab melts or high-pressure fractionated mantle melts? Journal of Petrology, 57(5):839-862. https://doi.org/10.1093/petrology/egw023
https://doi.org/https://doi.org/10.1093/...
).

The equation Mg#parental magma = 1/(1 + KDFe/Mg × (FeT/Mg)amphibole), KDFe/Mg = 0.38 by Alonso-Perez et al. (2009Alonso-Perez R., Müntener O., Ulmer P. 2009. Igneous garnet and amphibole fractionation in the roots of island arcs: experimental constraints on andesitic liquids. Contributions to Mineralogy and Petrology, 157:541-558. https://doi.org/10.1007/s00410-008-0351-8
https://doi.org/https://doi.org/10.1007/...
) was used to estimate the degree of differentiation of the mafic magma of the studied MME amphiboles. Primitive magmas, formed in the mantle, have Mg# > 0.70, while the values of Mg# in evolved magma do not exceed 0.60 (Grove et al. 2003Grove T.L., Elkins-Tanton L.T., Parman S.W., Chatterjee N., Muntener O., Gaetani G.A. 2003. Fractional crystallization and mantle-melting controls on calc-alkaline differentiation trends. Contributions to Mineralogy and Petrology, 145(5):515-533. https://doi.org/10.1007/s00410-003-0448-z
https://doi.org/https://doi.org/10.1007/...
, 2012Grove T.L., Till C.B., Krawczynski M.J. 2012. The role of H2O in subduction zone magmatism. Annual Review of Earth and Planetary Sciences, 40:413-439. https://doi.org/10.1146/annurev-earth-042711-105310
https://doi.org/https://doi.org/10.1146/...
). Magmas in equilibrium with the studied amphiboles of the mafic enclaves have Mg# ranging from 0.71 to 0.93 (mean 0.78), indicating that these amphiboles represent early crystallized phases in mantle-derived magmas.

The compositions of the studied MME amphibole crystals, in particular the Al content, indicate that there are two groups of amphibole, respectively, crystallized at high- and low-pressures (Fig. 8A).

The highest pressure values correspond to lower Si and Mg content, and higher Al values, and should represent the earlier crystals crystallized at lower crustal levels, at ~28 km (considering 1kbar = 3.7 km of continental crust; Tulloch and Challis 2000Tulloch A.J., Challis G.A. 2000. Emplacement depths of Paleozoic-Mesozoic plutons from western New Zealand estimated by hornblende-AI geobarometry. New Zealand Journal of Geology and Geophysics, 43(4):555-567. https://doi.org/10.1080/00288306.2000.9514908
https://doi.org/https://doi.org/10.1080/...
). The lowest pressure values were obtained in Si- rich and Al-poor crystals, and may represent a second generation of amphiboles or crystals reequilibrated under shallower crustal conditions.

In the Macururé Domain plutons, the interaction between mafic and felsic magmas has been pointed out by several authors as the process responsible for generating the mafic enclaves in these massifs (e.g., Conceição et al. 2016Conceição J.A., Rosa M.L.S., Conceição H. 2016. Sienogranitos leucocráticos do Domínio Macururé, Sistema Orogênicos Sergipano, Nordeste do Brasil: Stock Glória Sul. Brazilian Journal of Geology, 46(1):63-77. https://doi.org/10.1590/2317-4889201620150044
https://doi.org/https://doi.org/10.1590/...
, Fontes et al. 2018Fontes M.P., Conceição H.C., Rosa M.L.S., Lisboa V.A.C. 2018. Minettes do Stock monzonítico Glória Norte: Evidência de magmatismo ultrapotássico pós-orogênico, com assinatura de subducção, no Sistema Orogênico Sergipano. Geologia USP - Série Científica, 18(1):1-16. http://dx.doi.org/10.11606/issn.2316-9095.v18-133599
https://doi.org/http://dx.doi.org/10.116...
, Lisboa et al. 2019Lisboa V.A.C., Conceição H., Rosa M.L.S., Fernandes D.M. 2019. The onset of post-collisional magmatism in the Macururé Domain, Sergipano Orogenic System: The Glória Norte Stock. Journal of South American Earth Sciences, 89:173-188. https://doi.org/10.1016/j.jsames.2018.11.005
https://doi.org/https://doi.org/10.1016/...
). The following evidence, coming from our data, points out that the interaction between mafic and felsic magmas is physically possible: textural evidence (e.g., mafic mineral agglomerates); the compositional similarity between MME and the GNS amphiboles; similar crystallization temperature and liquidus temperature of the MME amphiboles and the GNS monzonite.

Yamaguchi et al. (2003Yamaguchi Y., Wada H., Ohta Y., Harayama S. 2003. Amphibole zoning, a record of progressive oxidation during crystallization of mafic microgranular enclaves in the Kurobegawa Granitic Pluton. Journal of Mineralogical and Petrological Sciences, 98(4):151-155. https://doi.org/10.2465/jmps.98.151
https://doi.org/https://doi.org/10.2465/...
) stated that the amphiboles formed at an early stage in mafic magma continue to crystallize after the interaction between magmas. The pre-interaction conditions can be obtained from early amphiboles and the reequilibrium conditions can be obtained from those phases which crystallized after the interaction. Our results suggest that the parental magmas of the MME were stored in deeper (~28 km) pressures, and the mixing with monzonitic magma was shallower (~18 km) at crustal conditions. The interaction between these magmas occurred at temperatures ranging from 772 to 862ºC as indicated by Ridolfi et al. (2010Ridolfi F., Renzulli A., Puerini M. 2010. Stability and chemical equilibrium of amphibole in calc-alkaline magmas: an overview, new thermobarometric formulations and application to subduction-related volcanoes. Contributions to Mineralogy and Petrology , 160:45-66. https://doi.org/10.1007/s00410-009-0465-7
https://doi.org/https://doi.org/10.1007/...
), considered by us as the best geothermometers for involving different chemical elements in the calculation.

CONCLUSION

The amphiboles of the microgranular mafic enclaves of the Glória Norte Stock are calcic and correspond to pargasite, edenite, magnesium-hornblende, and actinolite. They have varying Si contents (6.2-7.7 apuf) and mg# ratio between 0.48 and 0.84. The main important substitutions involve Al, Si, Na, Fe, and Mg (edenite and pargasite types) and reflect a decrease in temperature and an increase of fO2. The magmatic amphiboles of the GNS are edenite and magnesium-hornblende. They crystallize at 772 and 862ºC and the mean pressure values are in the order of 5 kbar, indicating crystallization at around 18 km depth.

The identification of similar chemical compositions between amphiboles from the MME and host GSN, as well as the estimates of similar pressure and temperature obtained and the compositional similarity between amphibole from the MME and GNS monzonites all suggest chemical reequilibration between the magmas which originated the monzonites and the enclaves. The interaction between these magmas occurred under ± 5 kbar pressure (~18 km depth), with temperature ranging between 772 and 862ºC.

ACKNOWLEDGMENTS

The first author thanks the IFPB, Campus Picuí, for granting the removal for completion of the Doctorate in Geology at UFBA. The authors thank CNPq processes: 140117/2015-6 (VACL Doctorate grant), 311008/2017-8 (CNPq-PQ), 310391/2017-2 (CNPq-PQ), 405387/2016-4 (CNPq-Universal 2016). We are grateful to the anonymous reviewers for the comments and suggestions that have improved considerably the original manuscript.

REFERENCES

  • Alonso-Perez R., Müntener O., Ulmer P. 2009. Igneous garnet and amphibole fractionation in the roots of island arcs: experimental constraints on andesitic liquids. Contributions to Mineralogy and Petrology, 157:541-558. https://doi.org/10.1007/s00410-008-0351-8
    » https://doi.org/https://doi.org/10.1007/s00410-008-0351-8
  • Anderson J.L. 1996. Status of thermobarometry in granitic batholiths. Earth and Environmental Science Transactions of the Royal Society of Edinburgh, 87(1-2):125-138. https://doi.org/10.1017/S0263593300006544
    » https://doi.org/https://doi.org/10.1017/S0263593300006544
  • Anderson J.L., Smith D.R. 1995. The effects of temperature and fO2 on the Al-in-hornblende barometer. American Mineralogist, 80(5-6):549-559. https://doi.org/10.2138/am-1995-5-614
    » https://doi.org/https://doi.org/10.2138/am-1995-5-614
  • Barbarin B., Didier J. 1992. Genesis and evolution of mafic microgranular enclaves through various types of interaction between coexisting felsic and mafic magmas. Earth and Environmental Science Transactions of the Royal Society of Edinburgh, 83(1-2):145-153. https://doi.org/10.1017/S0263593300007835
    » https://doi.org/https://doi.org/10.1017/S0263593300007835
  • Baxter S., Feely M. 2002. Magma mixing and mingling textures in granitoids: examples from the Galway Granite, Connemara, Ireland. Mineralogy and Petrology, 76:63-74. https://doi.org/10.1007/s007100200032
    » https://doi.org/https://doi.org/10.1007/s007100200032
  • Chivas A.R. 1982. Gechemical evidence for magmatic fluids in porphyry copper mineralization. Part I. Mafic silicates from the Koloula Igneous Complex. Contributions to Mineralogy and Petrology, 78:389-403. https://doi.org/10.1007/BF00375201
    » https://doi.org/https://doi.org/10.1007/BF00375201
  • Coltorti M., Bonadiman C., Faccini B., Grégoire M., O'Reilly S.Y., Powell W. 2007. Amphiboles from suprasubduction and intraplate lithospheric mantle. Lithos, 99(1-2):68-84. https://doi.org/10.1016/j.lithos.2007.05.009
    » https://doi.org/https://doi.org/10.1016/j.lithos.2007.05.009
  • Conceição J.A., Rosa M.L.S., Conceição H. 2016. Sienogranitos leucocráticos do Domínio Macururé, Sistema Orogênicos Sergipano, Nordeste do Brasil: Stock Glória Sul. Brazilian Journal of Geology, 46(1):63-77. https://doi.org/10.1590/2317-4889201620150044
    » https://doi.org/https://doi.org/10.1590/2317-4889201620150044
  • Conceição R.V., Green D.H. 2004. Derivation of potassic (shoshonitic) magmas by decompression melting of phlogopite+pargasite lherzolite. Lithos, 72(3-4):209-229. https://doi.org/10.1016/j.lithos.2003.09.003
    » https://doi.org/https://doi.org/10.1016/j.lithos.2003.09.003
  • Dalpé C., Baker D.R. 2000. Experimental investigation of large-ion-lithophile-element-, high-field-strength-element and rare-earth-element-partitioning between calcic amphibole and basaltic melt: the effects of pressure and oxygen fugacity. Contributions to Mineralogy and Petrology, 140:233-250. https://doi.org/10.1007/s004100000181
    » https://doi.org/https://doi.org/10.1007/s004100000181
  • Davison I., Santos R.A. 1989. Tectonic evolution of the Sergipano fold belt, NE Brazil, during the Brasiliano orogeny. Precambrian Research, 45(4):319-342. https://doi.org/10.1016/0301-9268(89)90068-5
    » https://doi.org/https://doi.org/10.1016/0301-9268(89)90068-5
  • Deer W.A., Howie R.A., Zussman J. 1997. Rock-Forming Minerals 2B. Double-Chain Silicates. London, Geological Society of London, 764 p.
  • Féménias O., Mercier J.C.C., Nkono C., Diot H., Berza T., Tatu M., Demaiffe D. 2006. Calcic amphibole growth and compositions in calc-alkaline magmas: Evidence from the Motru Dike Swarm (Southern Carpathians, Romania). American Mineralogist, 91(1):73-81. https://doi.org/10.2138/am.2006.1869
    » https://doi.org/https://doi.org/10.2138/am.2006.1869
  • Foley S.F. 1992. Petrological characterization of the source components of potassic magmas: geochemical and experimental constraints. Lithos, 28(3-6):187-204. https://doi.org/10.1016/0024-4937(92)90006-K
    » https://doi.org/https://doi.org/10.1016/0024-4937(92)90006-K
  • Fontes M.P., Conceição H.C., Rosa M.L.S., Lisboa V.A.C. 2018. Minettes do Stock monzonítico Glória Norte: Evidência de magmatismo ultrapotássico pós-orogênico, com assinatura de subducção, no Sistema Orogênico Sergipano. Geologia USP - Série Científica, 18(1):1-16. http://dx.doi.org/10.11606/issn.2316-9095.v18-133599
    » https://doi.org/http://dx.doi.org/10.11606/issn.2316-9095.v18-133599
  • Förster M.W., Prelević D., Schmück H.R., Buhre S., Veter M., Mertz-Kraus R., Foley S.F., Jacob D.E. 2017. Melting and dynamic metasomatism of mixed harzburgite+ glimmerite mantle source: Implications for the genesis of orogenic potassic magmas. Chemical Geology, 455:182-191. http://dx.doi.org/10.1016/j.chemgeo.2016.08.037
    » https://doi.org/http://dx.doi.org/10.1016/j.chemgeo.2016.08.037
  • Green D.H., Ringwood A.E. 1963. Mineral assemblages in a model mantle composition. Journal of Geophysical Research, 68(3):937-945. https://doi.org/10.1029/JZ068i003p00937
    » https://doi.org/https://doi.org/10.1029/JZ068i003p00937
  • Grove T.L., Elkins-Tanton L.T., Parman S.W., Chatterjee N., Muntener O., Gaetani G.A. 2003. Fractional crystallization and mantle-melting controls on calc-alkaline differentiation trends. Contributions to Mineralogy and Petrology, 145(5):515-533. https://doi.org/10.1007/s00410-003-0448-z
    » https://doi.org/https://doi.org/10.1007/s00410-003-0448-z
  • Grove T.L., Till C.B., Krawczynski M.J. 2012. The role of H2O in subduction zone magmatism. Annual Review of Earth and Planetary Sciences, 40:413-439. https://doi.org/10.1146/annurev-earth-042711-105310
    » https://doi.org/https://doi.org/10.1146/annurev-earth-042711-105310
  • Hammarstron J.M., Zen E. 1986. Aluminium in hornblende: an empirical igneous geobarometer. American Mineralogist, 71:1297-1313.
  • Hawthorne F.C. 1983. The crystal chemistry of the amphiboles. Canadian Mineralogist, 21(2):173-480.
  • Hendry D.A.F., Chivas A.R., Long J.V.P., Reed S.J.B. 1985. Chemical differences between minerals from mineralizing and barren intrusions from some North American porphyry copper deposits. Contributions to Mineralogy and Petrology, 89:317-329. https://doi.org/10.1007/BF00381554
    » https://doi.org/https://doi.org/10.1007/BF00381554
  • Hidalgo P.J., Rooney T.O. 2010. Crystal fractionation processes at Baru volcano from the deep to shallow crust. Geochemistry, Geophysics, Geosystems, 11(12):1-29. https://doi.org/10.1029/2010GC003262
    » https://doi.org/https://doi.org/10.1029/2010GC003262
  • Holland T., Blundy J. 1994. Non-ideal interactions in calcic amphiboles and their bearing on amphibole-plagioclase thermometry. Contributions to Mineralogy and Petrology, 116:433-447. https://doi.org/10.1007/BF00310910
    » https://doi.org/https://doi.org/10.1007/BF00310910
  • Hollister L.S., Grissom G.C., Peters E.K., Stowell H.H., Sisson V.B. 1987. Confirmation of the empirical correlation of Al in hornblende with pressure of solidification of calc-alkaline plutons. American Mineralogist, 72(3-4):231-239.
  • Johnson M.C., Rutherford M.J. 1989. Experimental calibration of the aluminum-in-hornblende geobarometer with application to Long Valley caldera (California) volcanic rocks. Geology, 17(9):837-841. https://doi.org/10.1130/0091-7613(1989)017%3C0837:ECOTAI%3E2.3.CO;2
    » https://doi.org/https://doi.org/10.1130/0091-7613(1989)017%3C0837:ECOTAI%3E2.3.CO;2
  • Kawakatsu K., Yamaguchi Y. 1987. Successive zoning of amphiboles during progressive oxidation in the Daito-Yokota granitic complex, San-in belt, southwest Japan. Geochimica et Cosmochimica Acta, 51(3):535-540. https://doi.org/10.1016/0016-7037(87)90067-6
    » https://doi.org/https://doi.org/10.1016/0016-7037(87)90067-6
  • Krawczynski M.J., Grove T.L., Behrens H. 2012. Amphibole stability in primitive arc magmas: effects of temperature, H2O content, and oxygen fugacity. Contributions to Mineralogy and Petrology, 164:317-339. https://doi.org/10.1007/s00410-012-0740-x
    » https://doi.org/https://doi.org/10.1007/s00410-012-0740-x
  • Leake B.E. 1971. On aluminous and edenitic hornblendes. Mineralogical Magazine, 38(296):389-407. https://doi.org/10.1180/minmag.1971.038.296.01
    » https://doi.org/https://doi.org/10.1180/minmag.1971.038.296.01
  • Leake B.E. 1978. Nomenclature of amphiboles. The Canadian Mineralogist, 16(4):501-520.
  • Leake B.E., Woolley A.R., Arps C.E., Birch W.D., Gilbert M.C., Grice J.D., Hawthorne F.C., Kato A., Kisch H.J., Krivovichev V.G., Linthout K., Laird J., Mandarino J. 1997. Report. Nomenclature of amphiboles: report of the subcommittee on amphiboles of the international mineralogical association commission on new minerals and mineral names. Mineralogical Magazine, 61:295-321.
  • Leandro M.V.S. 2018. Mineraloquímica de Rochas do Batólito Sienítico Itabuna, Sul do Estado da Bahia Monografia, Departamento de Geologia, Universidade Federal de Sergipe, 67 p.
  • Lisboa V.A.C., Conceição H., Rosa M.L.S., Fernandes D.M. 2019. The onset of post-collisional magmatism in the Macururé Domain, Sergipano Orogenic System: The Glória Norte Stock. Journal of South American Earth Sciences, 89:173-188. https://doi.org/10.1016/j.jsames.2018.11.005
    » https://doi.org/https://doi.org/10.1016/j.jsames.2018.11.005
  • Martin R.F. 2007. Amphiboles in the igneous environment. Reviews in Mineralogy and Geochemistry, 67(1):323-358. https://doi.org/10.2138/rmg.2007.67.9
    » https://doi.org/https://doi.org/10.2138/rmg.2007.67.9
  • McDowell S.D., Willye P.J. 1971. Experimental studies of igneous rock series: the Kungnat complex of southwest Greenland. Journal of Geology, 79(2):173-194. https://doi.org/10.1086/627607
    » https://doi.org/https://doi.org/10.1086/627607
  • Mutch E.J.F., Blundy J.D., Tattitch B.C., Cooper F.J., Brooker R.A. 2016. An experimental study of amphibole stability in low-pressure granitic magmas and a revised Al-in-hornblende geobarometer. Contributions to Mineralogy and Petrology, 171:85. https://doi.org/10.1007/s00410-016-1298-9
    » https://doi.org/https://doi.org/10.1007/s00410-016-1298-9
  • Nandedkar R.H., Ulmer P., Müntener O. 2014. Fractional crystallization of primitive hydrous arc magmas: an experimental study at 0.7 GPa. Contributions to Mineralogy and Petrology, 167:1015. https://doi.org/10.1007/s00410-014-1015-5
    » https://doi.org/https://doi.org/10.1007/s00410-014-1015-5
  • Oliveira E.P., Windley B.F., Araujo M.N.C. 2010. The Neoproterozoic Sergipano orogenic belt, NE Brazil: a complete plate tectonic cycle in western Gondwana. Precambrian Research, 181(1-4):64-84. https://doi.org/10.1016/j.precamres.2010.05.014
    » https://doi.org/https://doi.org/10.1016/j.precamres.2010.05.014
  • Pinho Neto M.A., Rosa M.L.S., Conceição H. 2019. Petrologia do Batólito Sítios Novos, Sistema Orogênico Sergipano, Província Borborema, NE do Brasil. Geologia USP. Série Científica, 19(2):135-150. https://doi.org/10.11606/issn.2316-9095.v19-152469
    » https://doi.org/https://doi.org/10.11606/issn.2316-9095.v19-152469
  • Piwinskii A.J. 1968. Experimental Studies of Igneous Rock Series Central Sierra Nevada Batholith, California. The Journal of Geology, 76(5):548-570. https://doi.org/10.1086/627359
    » https://doi.org/https://doi.org/10.1086/627359
  • Putirka K. 2016. Amphibole thermometers and barometers for igneous systems and some implications for eruption mechanisms of felsic magmas at arc volcanoes. American Mineralogist, 101(4):841-858. https://doi.org/10.2138/am-2016-5506
    » https://doi.org/https://doi.org/10.2138/am-2016-5506
  • Ribeiro J.M., Maury R.C., Grégoire M. 2016. Are Adakites slab melts or high-pressure fractionated mantle melts? Journal of Petrology, 57(5):839-862. https://doi.org/10.1093/petrology/egw023
    » https://doi.org/https://doi.org/10.1093/petrology/egw023
  • Ridolfi F., Renzulli A., Puerini M. 2010. Stability and chemical equilibrium of amphibole in calc-alkaline magmas: an overview, new thermobarometric formulations and application to subduction-related volcanoes. Contributions to Mineralogy and Petrology , 160:45-66. https://doi.org/10.1007/s00410-009-0465-7
    » https://doi.org/https://doi.org/10.1007/s00410-009-0465-7
  • Rooney T.O., Franceschi P., Hall C.M. 2011. Water-saturated magmas in the Panama Canal region: a precursor to adakite-like magma generation? Contributions to Mineralogy and Petrology , 161:373-388. https://doi.org/10.1007/s00410-010-0537-8
    » https://doi.org/https://doi.org/10.1007/s00410-010-0537-8
  • Şahin S.Y., Orgun Y., Gungor Y., Goker A.F., Gultekin A.H., Karacik Z. 2010. Mineral and Whole-rock Geochemistry of the Kestanbol Granitoid (Ezine-Çanakkale) and its Mafic Microgranular Enclaves in Northwestern Anatolia: Evidence of Felsic and Mafic Interaction. Turkish Journal of Earth Sciences, 19:101-122. https://doi.org/10.3906/yer-0809-3
    » https://doi.org/https://doi.org/10.3906/yer-0809-3
  • Samaniego P., Robin C., Chazot G., Bourdon E., Cotten J. 2010. Evolving metasomatic agent in the Northern Andean subduction zone, deduced from magma composition of the long-lived Pichincha volcanic complex (Ecuador). Contributions to Mineralogy and Petrology , 160(2):239-260. https://doi.org/10.1007/s00410-009-0475-5
    » https://doi.org/https://doi.org/10.1007/s00410-009-0475-5
  • Sato H., Nakada S., Fujii T., Nakamura M., Suzuki-Kamata K. 1999. Groundmass pargasite in the 1991-1995 dacite of Unzen volcano: phase stability experiments and volcanological implications. Journal of Volcanology and Geothermal Research, 89(1):197-212. https://doi.org/10.1016/S0377-0273(98)00132-2
    » https://doi.org/https://doi.org/10.1016/S0377-0273(98)00132-2
  • Scaillet B., Evans B.W. 1999. The 15 June 1991 eruption of Mount Pinatubo. I. Phase equilibria and pre-eruption P-T-fO2-fH2O conditions of the dacite magma. Journal of Petrology, 40(3):381-411. https://doi.org/10.1093/petroj/40.3.381
    » https://doi.org/https://doi.org/10.1093/petroj/40.3.381
  • Schmidt M.W. 1992. Amphibole composition in tonalite as a function of pressure: an experimental calibration of the Al-in-hornblende barometer. Contributions to Mineralogy and Petrology , 110:304-310. https://doi.org/10.1007/BF00310745
    » https://doi.org/https://doi.org/10.1007/BF00310745
  • Sial A.N., Ferreira V.P., Fallick A.E., Cruz M.J.M. 1998. Amphibole-rich clots in calc-alkalic granitoids in the Borborema province, northeastern Brazil. Journal of South American Earth Sciences, 11(5):457-471. https://doi.org/10.1016/S0895-9811(98)00034-0
    » https://doi.org/https://doi.org/10.1016/S0895-9811(98)00034-0
  • Silva Filho A.F., Guimarães I.P., Silva M.R.R. 1992. Utilização de granitóides na definição de domínios tectônicos; aplicação no lado oriental do Sistema de Dobramento Sergipano. In: Congresso Brasileiro de Geologia, 37., 1992. Boletim de Resumos Expandidos
  • Spear F.S. 1981. An experimental study of hornblende stability and compositional variability in amphibolite. American Journal of Science, 281:697-734. https://doi.org/10.2475/ajs.281.6.697
    » https://doi.org/https://doi.org/10.2475/ajs.281.6.697
  • Stout J.H. 1972. Phase petrology and mineral chemistry of coexisting amphiboles from Telemark, Norway. Journal of Petrology, 13(1):99-145. https://doi.org/10.1093/petrology/13.1.99
    » https://doi.org/https://doi.org/10.1093/petrology/13.1.99
  • Tulloch A.J., Challis G.A. 2000. Emplacement depths of Paleozoic-Mesozoic plutons from western New Zealand estimated by hornblende-AI geobarometry. New Zealand Journal of Geology and Geophysics, 43(4):555-567. https://doi.org/10.1080/00288306.2000.9514908
    » https://doi.org/https://doi.org/10.1080/00288306.2000.9514908
  • Wones D.R. 1989. Significance of the assemblage titanite+ magnetite+ quartz in granitic rocks. American Mineralogist, 74(7-8):744-749.
  • Yamaguchi Y. 1985. Hornblende-cummingtonite and hornblende-aclinolite intergowths from the Koyama calc-alkaline intrusion, Susa, southwest Japan. American Mineralogist, 70(9-10):980-986.
  • Yamaguchi Y., Wada H., Ohta Y., Harayama S. 2003. Amphibole zoning, a record of progressive oxidation during crystallization of mafic microgranular enclaves in the Kurobegawa Granitic Pluton. Journal of Mineralogical and Petrological Sciences, 98(4):151-155. https://doi.org/10.2465/jmps.98.151
    » https://doi.org/https://doi.org/10.2465/jmps.98.151

ARTICLE INFORMATION

  • 1
    Manuscript ID: 20190101.

Publication Dates

  • Publication in this collection
    07 Sept 2020
  • Date of issue
    2020

History

  • Received
    28 Sept 2019
  • Accepted
    13 July 2020
Sociedade Brasileira de Geologia R. do Lago, 562 - Cidade Universitária, 05466-040 São Paulo SP Brasil, Tel.: (55 11) 3459-5940 - São Paulo - SP - Brazil
E-mail: sbgeol@uol.com.br