<<

https://doi.org/10.1130/G46923.1

Manuscript received 10 June 2019 Revised manuscript received 28 August 2019 Manuscript accepted 3 September 2019

© 2019 The Authors. Gold Open Access: This paper is published under the terms of the CC-BY license. Published online 15 October 2019

Serra Sul diamictite of the Carajás Basin (Brazil): A glaciation on the Amazonian craton Raphael Araújo1,2* and Afonso Nogueira2 1Management of Geology and Mineral Resources, Geological Survey of Brazil, Av. Perimetral 3645, Belém 66095-904, Brazil 2Institute of Geoscience, Federal University of Pará, R. Augusto Corrêa 01, Belém 66075-110, Brazil

ABSTRACT we attempted to resolve this issue by producing­ This paper reports the discovery of glacial deposits of likely age (2.58– sedimentological and stratigraphic evidence that 2.06 Ga) in South America (Carajás Basin, Brazil), thereby expanding the potential reach helps to identify the origin of these . of Paleoproterozoic glaciations to the Amazonian craton for the first time. Glacially derived The results provide important evidence of the diamictites are stacked within a hitherto unrecognized ∼600-m-thick siliciclastic succession, paleoenvironmental conditions on the Amazo- here named the Serra Sul Formation. Well-preserved textures, with evidence of glaciotec- nian craton in the early Paleoproterozoic. tonism and ice rafting, indicate deposition in a coastal subglacial to glacial-fed submarine fan , in which the immediately underlying units ( and volcanic MATERIALS AND METHODS rock) were the main source and bedrock. The Serra Sul diamictite may be correlated with A sedimentological and stratigraphic study any of the known Paleoproterozoic glaciations, or with none of them. based on detailed descriptions of drill cores from throughout the Serra Sul Formation was INTRODUCTION mented as a set of ca. 2.75 Ga BIF and mafic undertaken. To cover as much of the lateral Paleoproterozoic glaciations (ca. 2.45–2.22 to felsic volcanic rock (Trendall et al., 1998) stratigraphic information as possible, detailed Ga; Rasmussen et al., 2013) are recorded in al- ­affected by Cu-Au mineralization at ca. 2.57 Ga measured sections were taken throughout an most every continent, but they have not yet been (­Tallarico et al., 2005). This was partially due to ∼600-m-thick succession from drill cores lo- reported in South America. These glaciations, the lack of an accurate stratigraphic framework cated in three different areas of the Carajás Ba- also called the Event (Tang for the Carajás Basin that has caused several sin (a, b, and c in Fig. 1; see Items DR3, DR4, and Chen, 2013), occurred at the time of the units to be neglected, misinterpreted, or even and DR5). For each of these areas, a simplified (GOE), and it is suggest- overestimated. sedimentary log was created and organized in ed that they were triggered by the superconti- The initial investigation of structures and the north-south section toward the basin dep- nent breakups at the end of the (Aspler textures from drill cores revealed that the strata ocenter (Fig. 2). Small samples (up to 10 cm and Chiarenzelli, 1998; Bleeker, 2003; Strand, of the Serra Sul succession are inverted. This is long) were collected from the drill cores, and 2012). While some studies have argued that they generally observed on the borders of the Carajás polished thin sections were made in order to ver- may represent an early snowball episode Basin that were deformed by regional folds asso- ify the microstructures and compositions of the (e.g., Evans et al., 1997; Hoffman, 2013), oth- ciated with the latest tectonic event that affected diamictite matrices and clasts. Scanning electron ers have suggested that there may have been the area, the Transamazonian cycle at ca. 2.0 microscope (SEM) analyses were performed to a series of more typical glaciations occurring Ga (Machado et al., 1991). In fact, the breccia investigate the microtextures on quartz grain at different times and in different places (e.g., intervals occur within a siliciclastic succession surfaces. For this analysis, small pieces of core Young, 2014). hitherto unrecognized in the basin, hereafter (n = 2) were gradually broken and quartz grains In the Carajás Basin, Amazonian craton formally named the Serra Sul Formation. This (n = 100) were randomly selected, then fixed in (Brazil), intervals of matrix-supported breccia formation unconformably overlies the Neoar- an electrically conductive carbon double-stick were previously reported as a particular rock chean units and unconformably underlies the tape, coated with gold, and imaged using sec- of the Serra Sul succession, and considered a siliciclastic deposits of the Águas Claras For- ondary electrons at the Geological Survey of magnetite-rich breccia of the low- mation of ca. 2.06 Ga (Mougeot et al., 1996), Brazil (Belém). ermost banded iron formation (BIF) unit (Ca- suggesting Paleoproterozoic (2.58–2.06 Ga) age bral et al., 2013). Strangely, these rocks were constraints (Fig. 1). RESULTS never described in the Neoarchean succession In this study, based on a detailed descrip- Foliated to Massive Diamictite Facies (Grão-Pará Group), which is generally docu- tion of well-preserved delicate primary textures Association throughout the Serra Sul Formation (see Items The association of foliated to mas- *E-mail: [email protected] DR1 and DR2 in the GSA Data Repository1), sive diamictite facies comprises diamictite

1GSA Data Repository item 2019397, complete set of core-based sedimentary logs, borehole locations, table of lithofacies, and supplementary figures, is available online at http://www.geosociety.org/datarepository/2019/, or on request from [email protected].

CITATION: Araújo, R., and Nogueira, A., 2019, Serra Sul diamictite of the Carajás Basin (Brazil): A Paleoproterozoic glaciation on the Amazonian craton: Geology, v. 47, p. 1166–1170, https://doi.org/10.1130/G46923.1

1166 www.gsapubs.org | Volume 47 | Number 12 | GEOLOGY | Geological Society of America

Downloaded from http://pubs.geoscienceworld.org/gsa/geology/article-pdf/47/12/1166/4870248/1166.pdf by guest on 27 September 2021 ­interlayered with massive diamictite and black Amazonian x x 50°W Stratigraphy Ages (Ma) craton x x N shale beds indicates an intrastratal deformation x x Carajás Basin 6 Carajás faulx t Dike swarms 1880.2 ± 1.5 at the time of deposition, and the coeval occur- x Carajás x x x 1 rence of deformed and preserved clasts in the Carajás granite 1880 ± 2 Basin same interval suggests a highly heterogeneous Brazil x Gorotire 5 c ALM-FD037 2011 ± 25 deformational setting (e.g., Menzies, 2012; Bus- ALM-FD083 6°S Formation x ALM-FD099 field and Le Heron, 2013, 2018). The observa- b 4 A43W-DH03 x Águas Claras 2060 ± 13 tion that fractures are restricted to the borders x ALM-FD52 GT41-FD02 Formation Drill hole GT41-F16 of the clasts and rarely cutting them, are prefer- GT41-F13 locations x a x SSD-FD672 GT41-F12 Serra Sul entially oriented perpendicular to foliation, and Synform SSD-FD927 Serra Sul area Formation are filled with quartz and dolomite that replaced Major faults x x 3 x x x 2575 ± 12 Igarapé Bahia v v v v v the original matrix minerals indicates that these Dike swarms x 20 km v v v 2 Formation v v v v v 2743 ± 11 structures were formed during sedimentation Granite Carajás v v Formation and subsequently infilled by secondary minerals.

Volcanic rock Granite Gabbro Group v v v v v Volcano-sedi v v v 2 Parauapebas v v v v v Grão-Par á v v v 2757 ± 7 Similar to the interpretation of Cabral et al. mentary units Iron formation Mn-bearing mudrock Formation vv vv v (2013), our observations indicate that the perva- xx Basement and Cu-Au mineralization metamorphic x x x 1 x x Migmatite Mud-rich diamictite Xingu Complex sive deformation fabric present in some diamic- terrains x x 2859 ± 2 Mudstone Iron-rich diamictite and correlates x x x tite intervals is derived from synsedimentary deformation, certainly linked to subglacial Figure 1. Simplified geological map of Carajás Basin (Brazil) highlighting study areas a( , b, mechanisms, rather than tectonic deformations. and c) and location of drill holes (black dots), and stratigraphic framework of basin including The observation in drill cores that the succession exact stratigraphic position of Serra Sul Formation and main geochronological data available (1—Machado et al., 1991; 2—Trendall et al., 1998; 3—Tallarico et al., 2005; 4—Mougeot et al., is only tilted and that no shear bands are present 1996; 5—Pereira et al., 2009; 6—Teixeira et al., 2018). strengthens this hypothesis, as does the fact that no evidence of metamorphism was recorded in ­interbedded with thin beds of black shale (as Conglomerate--Rhythmite- the thin section. Ultimately, the occurrence of much as 3 m thick). There are uninterrupted in- Diamictite Facies Association well-preserved delicate microtextures such as tervals of diamictite, 150 m thick, in the strati- Conglomerate and sandstone interbedded steps, striations, and fractures on the surfaces of graphic ­section (Fig. 2, log a; see Item DR1). multiple times with rhythmite and diamictite quartz sand grains confirms that the Serra Sul The diamictite has a foliated matrix that tends compose the final facies association observed diamictite is unmetamorphosed or was meta- toward a massive aspect. The BIF and mafic in this study (Fig. 2, log b; see Item DR2). morphosed to a very low grade (cf. Machado to felsic volcanic-subvolcanic rock clasts vary Normally graded to massive conglomerate and et al., 1991; Pinheiro and Holdsworth, 1997). in size from pebbles to boulders, and vary in sandstone are stacked into 0.30–30-m-thick fin- shape from angular to well rounded. Faceted, ing-upward cycles. In the stratigraphic section, Data Interpretation and Model flattened, rotated, boudinaged, sheared, and they reach uninterrupted intervals of as much as The occurrence of foliated to massive lens-shaped clasts are widespread (Figs. 3A 150 m thick. In sandstone, the grains are mainly diamictite with faceted, fractured, flattened, and 3B; see Item DR6). Fractures occur pref- composed of iron chert. Conglomerate clasts are sheared, boudinaged, and rotated clasts with erentially along the borders of clasts and rarely angular to well-rounded fragments of BIF, iron pressure shadows suggests that these rocks were cross-cut them; they are even found perpendic- chert, and volcanic rock ranging from pebble formed in subglacial environments and recorded ular to the foliation and filled with quartz and to cobble size, with cross-laminae and planar primary ice-contact sedimentation (i.e., tillite), dolomite. In some clasts, the inflection of the laminae weakly developed. Clast- to matrix- where high strain rates led to progressive de- foliation resulted in the development of asym- supported conglomerate has a sandy to granular formation (Menzies et al., 2006; Busfield and metrical and symmetrical pressure shadows. matrix (Fig. 3E; see Item DR8). Rhythmite and Le Heron, 2013, 2018; Le Heron et al., 2017). The fine-grained and dark-colored matrix is diamictite occur interbedded with one another. The quartz- and magnetite-rich matrices of these composed of micrometer-scale fragments (up The massive diamictite reaches as much as 10 m diamictites suggest that they were derived from to 100 μm) of quartz and magnetite, and the thick and consists of a mud-rich matrix in which a rock flour produced by intense abrasion of the alignment of these two minerals defines the pebble- to cobble-sized and angular to well- bedrock. The observation that volcanic rock and foliation (Fig. 3C). Isolated bullet-shaped and rounded BIF, iron chert, and volcanic rock clasts BIF clasts penetrate, puncture, and deform the boulder-sized clasts occur in black shale beds, occur scattered. Isolated pebble-sized clasts underlying laminae of mudstone suggests that sandwiched by diamictite intervals (Fig. 3D). occur with high dip angles and deformed un- these clasts are dropstones. The occurrence of derlying laminae (Fig. 3F), and coarse-grained beds composed of coarse-grained clast mounds Rhythmite Facies Association clasts are mound forming in concave-down beds surrounded by fine-grained sediments further Another facies association found within the (Fig. 3G). Faceted clasts, including elongated suggests that these beds are dumpstone (i.e., Serra Sul Formation is composed of only rhyth- pentagonal ones, also occur scattered within the rock formed from the melt-out of debris con- mite that constitutes a monotonous succession muddy matrix (Figs. 3H and 3I; see Item DR9). tained in floating dirt-laden icebergs; Thomas as much as 300 m thick at the base of the strati- Fractures, steps, and grooves occur on quartz and Connell, 1985). graphic section (Fig. 2, logs b and c; see Items grain surfaces (Fig. 3J; see Item DR10). The large number of faceted clasts embedded DR2 and DR7). The rhythmite is composed of in mudstone beds, some of which have elon- intercalated normally graded, fine-grained sand- DISCUSSION gated pentagonal shapes and high dip angles, stone and mudstone stacked into fining-upward Diamictite Textures: Tectonic versus indicates that they were dropped over a muddy cycles as much as 10 cm thick; the lithological Sedimentary substrate and associated with rapid burial by fine contact is generally marked by soft-sediment The observation that foliated diamictite sediments without subsequent reworking. Simi- deformation structures and rip-up clasts. is restricted to certain beds and that it occurs larly, the fairly common occurrence of faceted

Geological Society of America | GEOLOGY | Volume 47 | Number 12 | www.gsapubs.org 1167

Downloaded from http://pubs.geoscienceworld.org/gsa/geology/article-pdf/47/12/1166/4870248/1166.pdf by guest on 27 September 2021 b clasts sandwiched by mudstone indicates a hy- South North drodynamically incongruent sedimentary setting (Le Heron, 2015; Le Heron et al., 2017). The grooves observed on the surfaces of the quartz Facies Associations 600 Unconformity grains are straight, parallel, and V-shaped, sug- Foliated to massive diamictite gesting that they are striations produced by glacial (Whalley and Krinsley, 1974; Conglomerate-sandstone- rhythmite-diamictite Mahaney et al., 1996; Immonen, 2013). On the other hand, the rhythmic centimeter-scale Rhythmite intercalation between fine-grained sandstone 3G a with normal gradation and mudstone suggests 500 Águas Claras deposition by low-density turbidity currents, Formation and the meter-scale intercalations of massive ~20 km to normally graded conglomerate and sandstone 100 suggests deposition by the types of high-density 3F, 3I turbidity currents associated with submarine fan 3A, 3A', 3B, 3C systems (Pickering et al., 2015). A glacially influenced coastal to marine system is envisaged as the depositional envi- 400 3E Thickness (m) 3D ronment that accommodated the sediments of 0 the Serra Sul Formation (Fig. 4). This system was developed over and supplied by bedrock vvvv Serra Sul v vvvv mainly composed of Neoarchean BIF and vol- v v vv Formation v v v v v canic rock belonging to the Grão-Pará Group. This hypothesis is strongly supported by the Lithologies observation that quartz and magnetite compose Rhythmite 300 the matrices of subglacial diamictite, and by the fact that BIF, iron chert, and volcanic rock are Black shale 3H, 3J the main constituents of the clasts. The surface vvvv v v vvv Volcanic rock c that separates the Serra Sul deposits from the Iron formation Neoarchean strata represents a great unconfor- mity that may be used for correlation inside and Massive diamictite outside the basin. Ice overloading on the con- Foliated diamictite tinent, combined with the lateral expansion of Conglomerate 200 glaciers, promoted high rates of deformation of 300 the bedrock. This was likely strengthened by the Sandstone presence of a previously formed cover that greatly favored the glaciotectonic deforma- Structures tion and had a composition similar to that of Meter-scale fining- the bedrock. Short-lived marine transgressions upward cycle ~50 km into the continent associated with the retreat of Centimeter-scale glaciers may explain the occurrence of black fining-upward cycle 100 200 shale beds repeatedly interbedded with subgla- Cross-lamination cial diamictite at the base of coastal deposits. Cross-stratification In the marine setting, submarine fan systems were washed and supplied by glacial meltwater, Mud rip-up clasts which promoted the resedimentation of glacio-

Soft-sediment Thickness (m) genic debris in deep waters. The simultaneous deformation release of ice-rafted debris (IRD) from icebergs Volcanic rock clasts 0 in distal waters and deposition of fine-grained v vvvv vvvv v vvv v 100 particles generated massive diamictite beds with Iron formation clasts vvvv Unconformity dropstones and dump structures. The alternation v v vv Dropstone v v v v v of sedimentation by submarine fan systems and ice rafting explains the multiple intercalations of Dumpstone Grão-Pará Group IRD-bearing diamictite and sandy and/or con- glomeratic facies that is observed throughout Depocenter of the basin Thickness (m) the succession (Fig. 2, log b). Although gravity- 0 flow deposits constitute a greater part of the suc- v v v vv cession, IRD-bearing intervals of glaciogenic Figure 2. Sedimentary logs in Serra Sul Formation for three vvvv v vvv v distinct areas (a, b, and c) of Carajás Basin, Brazil (see Fig. 1 vvvv diamictite are sharply distinct. The interbed- v vvvv for location). Numbers and letters to right correspond to photos vvvv v v v vv ding of these deposits testifies to subaqueous in Figure 3. vvvv sedimentation within the basin. The similar

1168 www.gsapubs.org | Volume 47 | Number 12 | GEOLOGY | Geological Society of America

Downloaded from http://pubs.geoscienceworld.org/gsa/geology/article-pdf/47/12/1166/4870248/1166.pdf by guest on 27 September 2021 A B Figure 3. Images of Serra Sul Formation (Carajás Basin, Brazil). (A,A′) 360° panoramic image and line interpretation of core sample representative of foliated diamictite exhibit- ing pressure shadow (Ps) and faceted (Fc), 1 mm sheared (Sh), rotated (Rt), fractured (Ft), flat- tened (Fl), and boudinaged (Bt) clasts of C iron formation, iron chert, and volcanic to Mag subvolcanic rock. Background color in A′ Py represents facies association of foliated to 2 cm Qtz massive diamictite. (B,C) Thin section images of foliated diamictite showing angular clasts 200 µm of chert with asymmetrical pressure shadow (B; plane polarized light, 10 ) rotated in a A′ Rt Sh × Fc Fl E matrix containing quartz (Qtz), magnetite Bt (Mag), and pyrite (Py). (D) Bullet-shaped clast Ps Ft of porphyritic rhyolite isolated in black shale bed sandwiched by diamictite. (E) Matrix-sup- Ft Sh ported conglomerate with pebble-sized iron Fc 1 cm formation clast scattered in granular matrix. Fl H (F) Well-rounded clast of porphyritic rhyo- Fl Ps Ps lite displaying disruption and puncturing of Fl Ft underlying laminae beneath clast. (G) Thin Ps Sh beds of coarse-grained mounds of chert and volcanic rock clasts surrounded by mudstone. D F (H) Faceted and elongated pentagonal clasts with high-angle dips randomly embedded in muddy diamictite. (I) Thin section image of diamictite (plane polarized light, 10×) showing very angular chert clasts suspended in poorly sorted and mud-rich matrix. (J) Striations on the surface of quartz grain from diamictite 1 cm 1 cm (scanning electron microscope image). G I

­composition of clasts in the subglacial (Fig. 2, log a) and debris-flow diamictite (Fig. 2, logs 200 µm b and c) suggests a genetic correlation between J these environments and that the debris-flow deposits are glaciogenic (e.g., Le Heron et al., 2017). The thick beds of rhythmite at the base 2 cm 1 cm 100 µm of the succession (Fig. 2, logs b and c) suggest a prolonged sedimentation by low-density tur- bidity currents in the deepest parts of the basin, although the absence of IRD-bearing intervals in these deposits does not necessarily indicate Coastal diamictite Marine diamictite ice-free conditions (e.g., Le Heron, 2015). (quartz- and magnetite-rich matrix) (mud-rich matrix) Ultimately, the depositional system hypoth- esized for the Serra Sul Formation is similar to those interpreted for Paleoproterozoic glacia- a Meltwater outwash tions worldwide (e.g., Miall, 1985; Chen et al., Currents and waves IRD 2019). The Serra Sul succession represents the b first large-scale siliciclastic sedimentation after the predominantly chemical sedimentation that Erosion and Regolith characterized the Neoarchean of the Carajás Ba- Black shale v c glaciotectonism v sin. Compared to the Sturtian v v v Glacial-fed submarine v glaciation, in which iron formation was depos- v v v v v fan deposits ited simultaneously with diamictite in sub–ice v v v Bedrock (Neoarchean iron v Rhythmite shelf settings (Lechte and Wallace, 2016), the formation & volcanic rock) v (distal turbidites) deposition of diamictite in the Serra Sul system was completely decoupled from the deposition Figure 4. Model showing sedimentary mechanisms underlying diamictite-bearing succession of the BIF. Although the expressive rift-related of Serra Sul Formation (Carajás Basin, Brazil). Subglacial diamictite from coastal environment volcanism of the Carajás Basin represented by is characterized by quartz- and magnetite-rich matrix resulting from erosion and glaciotecto- the Grão-Pará Group (Olszewski et al., 1989; nism on bedrock, whereas marine diamictite results from sedimentation of ice-rafted debris Martins et al., 2017) may be related to the break- (IRD) and is characterized by mud-rich matrix. Glacially fed submarine fan deposits rework glaciogenic sediments and are interbedded with IRD-bearing diamictite. Neoarchean banded up of the southern supercontinent at the end of iron formation (BIF) and volcanic rock units were main bedrock and source of sediments. the Archean, and may have also been one of the Positions of sedimentary logs (a, b, and c; see Fig. 1 for location) are indicated. triggers of Serra Sul glaciation, this volcanic

Geological Society of America | GEOLOGY | Volume 47 | Number 12 | www.gsapubs.org 1169

Downloaded from http://pubs.geoscienceworld.org/gsa/geology/article-pdf/47/12/1166/4870248/1166.pdf by guest on 27 September 2021 event predates by nearly 200 m.y. (2.75–2.58 Hoffman, P.F., 2013, The Great Oxidation and a Side- Pereira, R.M.P., Rosiere, C.A., Santos, J.O.S., Lobato, Ga) the onset of this cooling episode. rian : MIF-S based correlation of L.M., Silva, R.C.F., and McNaughton, N.J., 2009, Paleoproterozoic glacial epochs: Chemical Geol- Unidade Caninana: Sequência clástica Paleopro- ogy, v. 362, p. 143–156, https://doi​.org/10.1016/​ terozóica revelada por datação U-Pb em zircões CONCLUSIONS j.chemgeo.2013.04.018. detríticos da Província Mineral Carajás: Abstract A core-based sedimentological and strati- Immonen, N., 2013, Surface microtextures of ice- presented at XI Simpósio de Geologia da Amazô- graphic investigation in the Carajás Basin rafted quartz grains revealing glacial ice in the nia, Manaus, Brazil, 2–5 August. (Brazil) revealed the occurrence of likely Side- Arctic: Palaeogeography, Palaeoclima- Pickering, K.T., Corregidor, J., and Clark, J.D., 2015, tology, Palaeoecology, v. 374, p. 293–302, https:// Architecture and stacking patterns of lower-slope rian–Rhyacian (2.58–2.06 Ga) glacial strata doi​.org/10.1016/​j.palaeo.2013.02.003. and proximal basin-floor channelized submarine within a succession hitherto unrecognized, Lechte, M., and Wallace, M., 2016, Sub–ice shelf fans, Middle Ainsa System, Spanish Pyr- newly named the Serra Sul Formation. We ironstone deposition during the Neoproterozoic enees: An integrated outcrop–subsurface study: suggest that Amazonia underwent glaciation : Geology, v. 44, p. 891–894, Earth-Science Reviews, v. 144, p. 47–81, https:// in the early Paleoproterozoic, which may or https://doi​.org/10.1130/G38495.1. doi​.org/10.1016/​j.earscirev.2014.11.017. Le Heron, D.P., 2015, The significance of ice-raft- Pinheiro, R.V.L., and Holdsworth, R.E., 1997, Reacti- may not be correlated with any of the known ed debris in Sturtian glacial successions: Sedi- vation of Archaean strike-slip systems, Ama- Paleoproterozoic glaciations. The Serra Sul mentary Geology, v. 322, p. 19–33, https://doi​ zon region, Brazil: Journal of the Geological So- diamictite represents the first reported occur- .org/10.1016/​j.sedgeo.2015.04.001. ciety, v. 154, p. 99–103, https://doi​.org/10.1144/ rence of Paleoproterozoic glacial deposits in Le Heron, D.P., Tofaif, S., Vandyk, T., and Ali, D.O., gsjgs.154.1.0099. 2017, A diamictite dichotomy: Glacial conveyor Rasmussen, B., Bekker, A., and Fletcher, I.R., 2013, South America, expanding the potential reach belts and olistostromes in the Neoproterozoic of Correlation of Paleoproterozoic glaciations based of glaciations of this time period to the Ama- Death Valley, California, USA: Geology, v. 45, on U-Pb zircon ages for tuff beds in the Transvaal zonian craton for the first time. p. 31–34, https://doi​.org/10.1130/G38460.1. and Huronian Supergroups: Earth and Planetary Machado, N., Lindenmayer, Z., Krogh, T.E., and Science Letters, v. 382, p. 173–180, https://doi​ Lindenmayer, D., 1991, U-Pb of .org/10.1016/​j.epsl.2013.08.037. ACKNOWLEDGMENTS Archean magmatism and basement reactivation Strand, K., 2012, Global and continental-scale gla- We thank Sérgio Huhn, Fernando Matos, and Luiz in the Carajás area, Amazon shield, Brazil: Pre- ciations on the earth: Marine and Costa of Vale S.A. (Parauapebas, Brazil) for making Research, v. 49, p. 329–354, https://doi​ Petroleum Geology, v. 33, p. 69–79, https://doi​ the drill cores available for the research; the Geo- .org/10.1016/0301-9268(91)90040-H. .org/10.1016/​j.marpetgeo.2012.01.011. logical Survey of Brazil (Belém, Brazil) and the Mahaney, W.C., Claridge, G., and Campbell, I., 1996, Tallarico, F.H.B., Figueiredo, B.R., Groves, D.I., PROPESP/UFPA for supporting this work; Werner Microtextures on quartz grains in from Ant- Kositcin, N., McNaughton, N.J., Fletcher, I.R., Truckenbrodt for his constructive comments on an arctica: Palaeogeography, Palaeoclimatology, and Rego, J.L., 2005, Geology and SHRIMP earlier version of the manuscript; and Roberto Araújo Palaeoecology, v. 121, p. 89–103, https://doi​ U-Pb geochronology of the Igarapé Bahia depos- Filho and Alexandre Ribeiro for fieldwork assistance. .org/10.1016/0031-0182(95)00069-0. it, Carajás copper-gold belt, Brazil: An Archean We are also very grateful to Daniel Le Heron, Paul Martins, P.L.G., Toledo, C.L.B., Silva, A.M., Chemale, (2.57 Ga) example of iron-oxide Cu-Au-(U-REE) Hoffman, Kari Strand, and an anonymous reviewer F., Jr., Santos, J.O.S., and Assis, L.M., 2017, Neo- mineralization: Economic Geology and the Bulle- for reviewing our manuscript; and to James Schmitt archean magmatism in the southeastern Amazoni- tin of the Society of Economic Geologists, v. 100, for his editorial work. an Craton, Brazil: Petrography, geochemistry and p. 7–28, https://doi​.org/10.2113/100.1.0007. tectonic significance of basalts from the Carajás Tang, H., and Chen, Y., 2013, Global glaciations REFERENCES CITED Basin: Precambrian Research, v. 302, p. 340–357, and atmospheric change at ca. 2.3 Ga: Geo- Aspler, L.B., and Chiarenzelli, J.R., 1998, Two Neo- https://doi.org/10.1016/​ j.precamres.2017.10.013​ . science Frontiers, v. 4, p. 583–596, https://doi​ archean supercontinents? Evidence from the Menzies, J., 2012, Strain pathways, internal ar- .org/10.1016/​j.gsf.2013.02.003. Paleoproterozoic: Sedimentary Geology, v. 120, chitecture and microstructures—Perspectives Teixeira, W., Hamilton, M.A., Girardi, V.A.V., Faleiros, p. 75–104, https://doi​.org/10.1016/S0037- on a general kinematic model—A ‘blueprint’ F.M., and Ernst, R.E., 2018, U-Pb baddeleyite ages 0738(98)00028-1. for till development: Science Re- of key dyke swarms in the Amazonian Craton (Cara- Bleeker, W., 2003, The late Archean record: A puzzle views, v. 50, p. 105–124, https://doi.org/10.1016/​ ​ jás/Rio Maria and Rio Apa areas): Tectonic impli- in ca. 35 pieces: Lithos, v. 71, p. 99–134, https:// j.quascirev.2012.07.012. cations for events at 1880, 1110 Ma, 535 Ma and doi​.org/10.1016/​j.lithos.2003.07.003. Menzies, J., van der Meer, J.J.M., and Rose, J., 2006, 200 Ma: Precambrian Research, v. 329, p. 138–155, Busfield, M.E., and Le Heron, D.P., 2013, Glaci- Till—As a glacial “tectomict”, its internal ar- https://doi.org/10.1016/​ j.precamres.2018.02.008​ . tectonic deformation in the Chuos Formation chitecture, and the development of a “typing” Thomas, G.S.P., and Connell, R.J., 1985, Iceberg drop, of northern­ Namibia: Implications for Neopro- method for till differentiation: Geomorphol- dump, and grounding structures from terozoic ice dynamics: Proceedings of the Ge- ogy, v. 75, p. 172–200, https://doi​.org/10.1016/​ glacio-lacustrine sediments, Scotland: Journal ologists’ Association, v. 124, p. 778–789, https:// j.geomorph.2004.02.017. of Sedimentary Petrology, v. 55, p. 243–249, doi​.org/10.1016/​j.pgeola.2012.10.005. Miall, A.D., 1985, Sedimentation on an early Pro- https://doi​.org/10.1306/212F8689-2B24-11D7- Busfield, M.E., and Le Heron, D.P., 2018, Snow- terozoic continental margin under glacial influ- 8648000102C1865D. ball Earth under the microscope: Journal of ence: The Gowganda Formation (Huronian), El- Trendall, A.F., Basei, M.A.S., Laeter, J.R., and Nelson, ­Sedimentary Research, v. 88, p. 659–677, https:// liot Lake area, Ontario, Canada: Sedimentology, D.R., 1998, SHRIMP zircon U-Pb constraints doi​.org/10.2110/jsr.2018.34. v. 32, p. 763–788, https://doi.org/10.1111/​ j.1365-​ on the age of the Carajás Formation, Grão Pará Cabral, A.R., et al., 2013, Trace-element and multi- 3091.1985.tb00733.x. Group, Amazon Craton: Journal of South Ameri- isotope geochemistry of Late-Archean black Mougeot, R., Respaut, J.P., Briqueu, L., Ledru, P., can Earth Sciences, v. 11, p. 265–277, https://doi​ shales in the Carajás iron-ore district, Brazil: Milesi, J.P., Lerouge, C., Marcoux, E., Huhn, .org/10.1016/S0895-9811(98)00015-7. Chemical Geology, v. 362, p. 91–104, https:// S.B., and Macambira, M.J.B., 1996, Isotope Whalley, W.B., and Krinsley, D.H., 1974, A scan- doi​.org/10.1016/​j.chemgeo.2013.08.041. geochemistry constraints for Cu, Au mineraliza- ning electron microscope study of surface tex- Chen, Y., Chen, W., Li, Q., Santosh, M., and Li, tions and of the Carajás Province (Pará, tures of quartz grains from glacial environments: J., 2019, Discovery of the Huronian Glacia- Brazil), in Proceedings, Congresso Brasileiro de Sedimentology, v. 21, p. 87–105, https://doi​ tion Event in China: Evidence from glacigenic Geologia 39, Salvador: Abstracts: Salvador, So- .org/10.1111/​j.1365-3091.1974.tb01783.x. diamictites in the Hutuo Group in Wutai Shan: ciedade Brasileira de Geologia, v. 7, p. 321–324. Young, G.M., 2014, Contradictory correlations of Precambrian Research, v. 320, p. 1–12, https:// Olszewski, W.J., Wirth, K.R., Gibbs, A.K., and Gaudette, Paleoproterozoic glacial deposits: Local, re- doi​.org/10.1016/​j.precamres.2018.10.009. H.E., 1989, The age, origin, and tectonics of the gional or global controls?: Precambrian Re- Evans, D.A., Beukes, N.J., and Kirschvink, J.L., Grão Pará Group and associated rocks, Serra dos search, v. 247, p. 33–44, https://doi.org/10.1016/​ ​ 1997, Low-latitude glaciation in the Paleopro- Carajás, Brazil: Archean continental volcanism and j.precamres.2014.03.023. terozoic : , v. 386, p. 262–266, https:// rifting: Precambrian Research, v. 42, p. 229–254, doi​.org/10.1038/386262a0. https://doi​.org/10.1016/0301-9268(89)90013-2. Printed in USA

1170 www.gsapubs.org | Volume 47 | Number 12 | GEOLOGY | Geological Society of America

Downloaded from http://pubs.geoscienceworld.org/gsa/geology/article-pdf/47/12/1166/4870248/1166.pdf by guest on 27 September 2021