You are on page 1of 21

Precambrian Research 200203 (2012) 3858

Contents lists available at SciVerse ScienceDirect

Precambrian Research
journal homepage: www.elsevier.com/locate/precamres

Review

Marinoan glaciation in east central Brazil


Fabrcio de Andrade Caxito a,b, , Galen P. Halverson a , Alexandre Uhlein b , Ross Stevenson c ,
Tatiana Goncalves Dias b , Gabriel J. Uhlein b
a
b
c

Department of Earth and Planetary Sciences/GEOTOP, McGill University, Montral, Qubec H3A 2A7, Canada
Instituto de Geocincias, Universidade Federal de Minas Gerais, Av. Antnio Carlos 6627, CEP 31270-901, Belo Horizonte, MG, Brazil
Dpartement des Sciences de la Terre et de lAtmosphre///GEOTOP, Universit du Qubec Montral, P.O. Box 8888, Montral, Qubec H3C 3P8, Canada

a r t i c l e

i n f o

Article history:
Received 24 September 2011
Received in revised form 12 January 2012
Accepted 17 January 2012
Available online 24 January 2012
Keywords:
Neoproterozoic
Isotope stratigraphy
Marinoan glaciation
So Francisco craton
Bambu Group
Cap carbonate

a b s t r a c t
Remnants of a Neoproterozoic glaciation in east central Brazil are represented by thin diamictite layers (Jequita Formation and correlative units), locally overlying striated pavements on the So Francisco
craton. The diamictites are covered by the Sete Lagoas Formation of the basal Bambu Group, which is
generally accepted to be a typical cap carbonate sequence. Although most authors have preferred a midCryogenian (post-Sturtian) age for it, based mainly on Pb-Pb whole rock data, the Sete Lagoas Formation
bears lithostratigraphic and isotopic characteristics that are identical to early Ediacaran cap carbonates
worldwide, including a basal thin (010 m) pale and inty cap dolostone, preserving a drop in 13 C values from around 3.2 to 4.5 with associated 18 O around 5, and crystal-fan facies interpreted
as aragonite pseudomorphs. Ediacaran zircons have been recovered from the middle of the Sete Lagoas
Formation, constraining the deposition of its upper half to be younger than 610 Ma (Rodrigues, 2008).
Although there is an unconformity below the point where the zircons were collected, it is short-lived, as
suggested by the identical, typically Ediacaran 87 Sr/86 Sr values above and below (0.70740.7076). Carbonate clasts from the Jequita Formation and correlative diamictite-bearing units in the fold belts that
surround the So Francisco craton (Canabravinha and Serra do Catuni formations) display similar ranges
in 13 C (6.7 to +2.6), suggesting the erosion of a pre-glacial carbonate platform with negative 13 C
values (i.e. the Islay and/or Trezona anomalies). The cratonic Carrancas Formation, on the other hand,
yielded pale dolostone clasts with 13 C in a small range between 4.2 and 3.4, and 18 O values around
6.5. These clasts could be derived from the cap dolostone unit itself, in which case the Carrancas Formation would represent resedimented basal Sete Lagoas Formation and imply that sections of the Sete
Lagoas Formation sitting atop the Carrancas Formation are incomplete. The base-truncated sections have
confused previous attempts to correlate the Sete Lagoas Formation with other cap carbonate successions. In light of the available lithostratigraphic, isotopic and U-Pb zircon data, we propose that the Sete
Lagoas Formation represents a basal Ediacaran cap carbonate sequence (635610 Ma) deposited after
the Marinoan glaciation in east central Brazil.
2012 Elsevier B.V. All rights reserved.

Contents
1.
2.

3.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Lithofacies of the Neoproterozoic glaciation in east central Brazil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.
Bambu Group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.
Jequita Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.
Carrancas Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.
Sete Lagoas Formation: the cap carbonate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.
Fold belts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
New data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.
Materials and methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Corresponding author at: Instituto de Geocincias, Universidade Federal de Minas Gerais, Av. Antnio Carlos 6627, CEP 31270-901, Belo Horizonte, MG, Brazil.
Tel.: +55 31 34095444; fax: +55 31 34095410.
E-mail address: boni@ufmg.br (F.d.A. Caxito).
0301-9268/$ see front matter 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.precamres.2012.01.005

39
39
40
41
42
43
45
45
45

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858

39

Sampling and analytical procedures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .


Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.1.
Correntina section . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.2.
Carbonate clast data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.3.
Interpretation of the carbonate clast data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Discussion and correlation of the Sete Lagoas Formation with other cap carbonate successions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

45
46
46
47
50
52
54
55
55

3.2.
3.3.

4.
5.

1. Introduction
The Neoproterozoic Era (1000542 Ma) was a time of extreme
climatic variation as recorded in sedimentary rocks of this age
across the globe (Gaucher et al., 2009). Especially intriguing is an
association of glacial deposits with warm climate carbonate platforms that is a trademark feature of the Neoproterozoic record
(Fairchild, 1993; Hoffman, 2011). This climatic dichotomy has
led to a number of controversial explanations, including extreme
icehouse-greenhouse uctuations (snowball Earth glaciations),
which are unprecedented in the Phanerozoic record (Kirschvink,
1992; Hoffman et al., 1998; Hoffman and Schrag, 2002).
One of the predictable consequences of an abrupt shift from
icehouse to greenhouse conditions is the worldwide distribution of lithologically and isotopically similar strata (Kirschvink,
1992; Hoffman and Schrag, 2002). Despite local differences in
the sedimentation regime (clastic, mixed or carbonate), these
events are represented by glacial deposits of diverse nature, overlain by distinctive cap carbonate sequences corresponding to
the mid-Cryogenian (Sturtian) and end-Cryogenian (Marinoan)
glaciations (Kennedy et al., 1998; Hoffman and Schrag, 2002;
Halverson et al., 2005). Unfortunately, Neoproterozoic successions are largely devoid of useful biostratigraphic markers, and
the general scarcity of directly datable material (i.e., interbedded
felsic volcanic units) leads to difculties in assigning sequences
to a particular glacial event. The chemostratigraphy of carbonate
sequences has been invoked as a promising alternative tool for
regional and global correlation, as well as for the assignment of relative ages (Knoll and Walter, 1992; Jacobsen and Kaufman, 1999;
Melezhik et al., 2001; Halverson et al., 2010). The isotopic signatures of 13 C and 87 Sr/86 Sr have proven especially useful in this
respect, as they provide proxies for seawater composition at the
time of deposition and vary considerably throughout the Neoproterozoic.
In east central Brazil, a number of Neoproterozoic glacial
deposits have long been recognized (Moraes and Guimares, 1931;
Isotta et al., 1969; Pug and Schll, 1975; Hettich, 1977, Hettich
and Karfunkel, 1978; Rocha-Campos and Hasui, 1981; Gravenor
and Monteiro, 1983; Karfunkel and Hoppe, 1988; Rocha-Campos
et al., 1996; Uhlein et al., 1999, 2004; Karfunkel et al., 2002). The
climatic dichotomy here is represented by the Jequita Formation
diamictites and overlying Sete Lagoas Formation carbonates at the
base of the Bambu Group, which is the Neoproterozoic cover of the
So Francisco craton in Minas Gerais (MG), Bahia (BA), Gois (GO)
and Tocantins (TO) states (Figs. 1 and 2). Although there is currently
a consensus that the basal Sete Lagoas Formation is a typical postglacial cap carbonate, the age of the glacial event is the subject of
much debate (e.g. Misi and Veizer, 1998; Babinski et al., 2007, 2012;
Vieira et al., 2007; Sial et al., 2009; Misi et al., 2011). In this paper,
we reassess the chronostratigraphic position of the Neoproterozoic
glaciation documented in east central Brazil using available stratigraphic, sedimentary, geochronological and isotopic data from the
Bambu Group. To this data set, we add our own new stratigraphic
and C, O and Sr isotopic data from a previously unpublished section

of the Sete Lagoas Formation in the central portion of the basin and
from carbonate clasts from four separate diamictite occurrences.
The combination of new and published data overwhelmingly favors
an Ediacaran age for the basal Bambu Group, implying an end
Cryogenian age for the Jequita Formation glacial sediments.
2. Lithofacies of the Neoproterozoic glaciation in east
central Brazil
The So Francisco craton (Fig. 1; Almeida, 1977; Alkmim, 2004)
is a geotectonic unit characterized by an Archean to Paleoproterozoic basement covered by predominantly clastic Mesoproterozoic
units (Espinhaco Supergroup, Parano and Canastra groups) and
a Neoproterozoic mixed carbonate-siliciclastic sequence (Bambu
Group). It is surrounded by the late Neoproterozoic (Brasiliano

Fig. 1. The So Francisco craton and surrounding fold belts. The dashed rectangle
indicates the location of the map in Fig. 2. Modied from Alkmim (2004).

40

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858

Fig. 2. Simplied geologic map of the So Francisco Basin, showing the main outcrop areas of the Sete Lagoas Formation. The schematic lithostratigraphic and isotopic proles
are simplied after: (a) Vieira et al. (2007), Babinski et al. (2007); (b) Kuchenbecker (2011); (c) Alvarenga et al. (2007); and (d) Powis et al. (2001) in Misi et al. (2007). The
square near Correntina (BA) indicates the approximate location of Fig. 6.

Cycle, 630500 Ma) Aracua, Braslia, Rio Preto, Riacho do Pontal and Sergipano fold belts (Fig. 1; Trompette, 1994; Sial et al.,
2009), which contains deformed and metamorphosed diamictitebearing rift to passive margin sequences that were deposited on
the margins of the craton (Figs. 3 and 4).
2.1. Bambu Group
The Bambu Group covers over 300,000 km2 of the So Francisco
craton in east central Brazil (Figs. 1 and 2; Costa and Branco, 1961;
Braun, 1968; Dardenne, 1978a, 1979; Sial et al., 2009). The classically dened stratigraphy of the Bambu Group (Dardenne, 1978a)

includes, from bottom to top: the Jequita/Carrancas (diamictite,


sandstone, rhythmite), Sete Lagoas (carbonates), Serra de Santa
Helena (siltstone), Lagoa do Jacar (oolitic, dark and fetid limestone), Serra da Saudade (siltstone, arkose and green sandstone)
and Trs Marias (uvial to platformal sandstones) formations. The
whole of the Bambu Group is typically interpreted to have been
deposited in a foreland basin, with a N-S foredeep parallel to the
Braslia Orogen on the western margin of the craton (Chang et al.,
1988; Castro and Dardenne, 2000; Dardenne, 2000; Martins-Neto
and Alkmim, 2001; Alkmim and Martins-Neto, 2001; Martins-Neto
et al., 2001). Recently, some authors have suggested that there is a
major unconformity in the middle of the Sete Lagoas Formation,

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858

41

Fig. 3. Comparative stratigraphy of the diamictite-bearing sequences in fold belts rimming the So Francisco craton. The ages on the rectangles represent estimates of the
Brasiliano deformation and peak metamorphism for each fold belt. Although the sequences are generally metamorphosed to greenschist facies, sedimentary names of the
rocks have been kept for simplication, as most sedimentary features can still be recognized. The west and east sectors of the Aracua Fold Belt are separated by a basement
block (Porteirinha block); the Vaza-Barris Group is thrusted southwards upon the Miaba Group on the Sergipano Fold Belt. From: (a) Pedrosa-Soares et al., 2000, 2011;
Queiroga et al., 2007; Martins et al., 2008; Babinski et al., 2012; (b) Pereira et al., 1994; Valeriano et al., 2004; Rodrigues et al., 2010; Amorim Dias, 2011; (c) Egydio-Silva
et al., 1989; Caxito et al., 2011; (d) Oliveira et al., 2005, 2006, 2010; Sial et al., 2010.

which might divide the Bambu Group into a lower, glaciallyrelated sequence, and an upper, foreland basin sequence (Martins
and Lemos, 2007; Zaln and Romeiro-Silva, 2007).
On the craton margins, the Bambu Group was affected
by the Brasiliano Orogeny and cratonic-directed deformation
of the surrounding Aracua, Braslia and Rio Preto fold belts
(Figs. 1 and 4; Dardenne, 1978b; Chang et al., 1988; Chemale
et al., 1993; Dardenne, 2000; Alkmim and Martins-Neto, 2001).
Thin-skin deformation fronts on the Bambu Group reach up
to 200 km in width before passing laterally into horizontal
cratonic cover (Fig. 4). The timing of emplacement of the
external nappes of the Braslia fold belt, which were thrust
upon the cratonic cover on the southern So Francisco craton, provides a minimum depositional age for the Bambu
Group at around 567 Ma (K-Ar muscovite age; Valeriano et al.,
2000).
2.2. Jequita Formation
The Jequita Formation outcrops on the borders of the Cabral
and gua Fria ranges in north central Minas Gerais, near the
boundary between the Aracua fold belt and the So Francisco craton (Figs. 2 and 4; Isotta et al., 1969; Viveiros and Walde, 1976;
Uhlein et al., 1999, 2011; Cukrov et al., 2005). These ranges are
the topographic expression of east-west open anticlinal folding
with a north-plunging axis, with younger strata towards the anks
(Fig. 4). The Jequita Formation consists mainly of massive, clastpoor diamictite with rare and thin (cm to dm) intercalations of
sandstone and rhythmite, reaching up to 100 m thick. The diamictite contains clasts of carbonates, gneiss, quartzite, granite and
quartz, up to boulder size, oating in a pelitic matrix (Fig. 5c).

On the northeastern ank of the gua Fria range, 10 km southeast of Jequita, Isotta et al. (1969) described beautifully preserved
striated pavements in quartzites underlying a diamictite of the
Jequita Formation (Figs. 4 and 5a and b). East-west trending striae
vary from ne, V-shaped thin scratches to U-shaped grooves up
to 20 cm wide 5 cm deep, bearing crescent-shaped cracks that
consistently indicate a roughly eastward ice ow direction (Fig. 5a
and b). Individual grooves reach up to 18 meters in length with no
widening or shallowing at the ends (Isotta et al., 1969). This spectacular striated pavement is the best evidence for Neoproterozoic
glaciation in east central Brazil (e.g. Uhlein et al., 1999 and references therein). Karfunkel and Hoppe (1988) also recovered striated
and faceted pebbles from a diamictite layer 4 km north of Jequita.
Although Isotta et al. (1969) originally interpreted the striae to
be formed over an indurated quartzite pavement, Rocha-Campos
et al. (1996) re-interpreted the glacial abrasion marks as having
been formed over a soft-sediment substratum, in the uctuating grounding zone of a marine ice sheet, based on evidence that
includes internal striae in the U-shaped groove walls that are covered by slumped plow ridges, clasts partially embedded in the
basement quartzites, and striae and ripple marks occupying the
same bedding plane. These observations imply that the underlying sandstones were still unlithied during glaciation, and hence
belong to the basal Jequita Formation and not to the Mesoproterozoic Espinhaco Supergroup, as conventionally interpreted (Uhlein
et al., 2011).
Until recently, the Jequita Formation diamictites were interpreted as tillites resting directly above the grooved and striated
pavement (Fig. 5d, Gravenor and Monteiro, 1983; Karfunkel and
Hoppe, 1988; Karfunkel et al., 2002). However, Uhlein et al. (1999,
2011), Cukrov et al. (2005) and Chaves et al. (2010) noted the

42

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858

Fig. 4. Schematic cross-sections across the craton-fold belt transition zone in the northwestern (a) and eastern (b) margins of the So Francisco craton. See Fig. 2 for location
of the sections. The arrows indicate sites where carbonate clasts were collected. (b) Modied from Uhlein et al. (1999).

scarcity of outwash facies deposits around Jequita, and interpreted


the monotonous, massive clast-poor diamictites as glacio-marine
deposits. Sandstone lenses within the diamictites, previously interpreted as eskers (Karfunkel and Hoppe, 1988) or as ice-thrust
features (Gravenor and Monteiro, 1983) were re-interpreted as
submarine channel lls. We could not nd inter-tillite pavements,
as suggested by Gravenor and Monteiro (1983) and Karfunkel et al.
(2002), in the Jequita area. Rather, the pavement is developed on
coarse-grained quartzites with partially embedded clasts. The presence of striae and ripple marks on the same bedding plane (Fig. 5b)
further demonstrates that the pavement was not developed on
a thin tillite layer, but instead on a shallow marine sand body,
on top of which large clasts transported by the overriding glacier
were embedded during soft-sediment deformation (Rocha-Campos
et al., 1996).
Sparse diamictite occurrences bordering the Braslia Fold belt
from Cristalina to Campos Belos, GO (Fig. 2), as well as very thin
(02 m) diamictite layers found below the Sete Lagoas Formation
in boreholes on the So Francisco basin, have also been assigned as
equivalents to the Jequita Formation, although in its type-area (e.g.
gua Fria and Cabral ranges), the Jequita Formation is not overlain
by the Sete Lagoas cap carbonate, but instead, by the Serra de Santa
Helena Formation siltstone (Fig. 4b).
The age of deposition of the Jequita Formation is roughly constrained to between 880 and 575 Ma based on the U-Pb age of
the youngest detrital zircons from the diamictite and the age of
deformational and metamorphic peak on the Aracua fold belt,
which caused the anticlinal folding on the Cabral and gua Fria

ranges (Buchwaldt et al., 1999; Uhlein et al., 1999; Rodrigues, 2008;


Pedrosa-Soares et al., 2011).
2.3. Carrancas Formation
The Carrancas Formation is composed of thin and patchy paraconglomerate and rhythmite lenses directly overlying the cratonic
basement north of Belo Horizonte and south of Sete Lagoas in
south central Minas Gerais (Fig. 2; Costa and Branco, 1961; Tuller
et al., 2009). Its classic outcrop is a meter thick paraconglomerate layer overlain by Sete Lagoas Formation limestone at km 30
of the MG-424 road (Figs. 2 and 5f). The paraconglomerate is composed of pebble- to boulder-sized gneiss, dolostone, granite, quartz,
and limestone clasts oating in a microsparite matrix with sandsized grains of dolomite, quartz, plagioclase, microcline, biotite and
granite/gneiss lithoclasts (Vieira et al., 2007; Tuller et al., 2009).
Interpretation of the Carrancas Formation depositional environment is hampered by its limited spatial distribution and sparse
outcrops. The highly calcitic matrix and the occurrence of limestone clasts near the contact with the Sete Lagoas Formation,
which are identical to the overlying carbonates, suggest the local
reworking of the basal carbonate platform, along with its basement
(Fig. 5f).
Although the Jequita and Carrancas formations are generally
regarded as equivalent, most authors agree that there is no direct
evidence for glacial deposition of the Carrancas Formation in its
type-area (Vieira et al., 2007; Tuller et al., 2009). The detrital zircon U-Pb age spectra of the Carrancas Formation is distinct from

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858

43

Fig. 5. (a and b) The Serra da gua Fria striated pavement. (a) deep and wide U-shaped furrows with internal crescent-shaped cracks showing eastward ice ow direction
(arrow). Note large quartzite cobble partially embedded on the striated pavement near hammers head. (b) Striae and ripple marks on the same bedding plane. (c) Jequita
Formation diamictite, at an abondoned quarry on the km 66 of the BR-365 road, 36 km northeast of Jequita. Note stromatolite clast in lower left corner. (d) Serra do Catuni
Formation metadiamictite at Couto de Magalhes (MG). (e) Canabravinha Formation metadiamictite at Canabravinha stream, near Monte Alegre dos Cardosos (BA). Note
angular carbonate clast in upper right corner. (f) Contact between the Carrancas paraconglomerate (below) and the Sete Lagoas limestone (above), at km 30 of the MG-424
road between Belo Horizonte and Vespasiano (MG). Hammer is 30 cm long; diameter of coin is 2.5 cm; bottle on d is 30 cm long.

that of the Jequita Formation, with a range in ages from 1.4 to


3.1 Ga (Rodrigues, 2008), which might reect erosion of only local
sources.
Other sparse diamictite and rhythmite occurrences fringing the
southern margin of the So Francisco craton and covered by carbonates of the Sete Lagoas Formation are typically interpreted as
equivalents to the Carrancas Formation, e.g. in the Onca do Pitangui and Moema (MG) regions (Fig. 2; Romano and Knauer, 2003;

Romano, 2007; Rocha-Campos et al., 2007, 2011; Ribeiro et al.,


2008; Sial et al., 2009).
2.4. Sete Lagoas Formation: the cap carbonate
The Sete Lagoas Formation consists of up to 200 m of
limestone, dolomite, and subordinate marly siltstones, sitting
atop ArcheanPaleoproterozoic granite-gneiss basement, or, more

44

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858

locally, in sharp contact atop thin diamictite layers, generally correlated to the Jequita Formation (Fig. 2). Recent studies have
suggested that an unconformity in the middle of the Sete Lagoas
Formation, marked by a facies, isotopic and seismic break, divides
it into two shallowing-upwards cycles (Martins and Lemos, 2007;
Zaln and Romeiro-Silva, 2007; Vieira et al., 2007).
The base of the lower Sete Lagoas Formation is marked by
a thin (010 m), lenticular, pink or pale inty dolostone layer
(Misi et al., 2007; Sial et al., 2009). It generally consists of a
microcrystalline massive or nely laminated dololutite, sometimes
containing peloids (Alvarenga et al., 2007; Lima, 2011). Locally,
reworked equivalents, containing detrital quartz, carbonate and
lithic clasts, have been described (Kuchenbecker, 2011). Recent
studies have shown that this basal dolostone yields an upwards
decreasing 13 C pattern from around 3.2 to 4.5, with associated 18 O between 4.5 and 6.5 (Fig. 2b and c; Alvarenga
et al., 2007; Kuchenbecker, 2011). Above this thin dolostone, a
100 m package of laminated limestone prevails. The dolostonelimestone shift is marked by an abrupt drop in 13 C values of
about 1 and up to 4 in 18 O values. Up section, 13 C gradually increases towards positive values, remaining homogeneous
around 0 or rising to around +4 (Fig. 2; Santos et al., 2000;
Alvarenga et al., 2007; Vieira et al., 2007; Kuchenbecker, 2011;
Lima, 2011).
In places, the thin basal pale dolostone is absent, most notably in
the classic Sete Lagoas Formation outcrops, between Belo Horizonte
to the south and Sete Lagoas to the north. Here, the Sete Lagoas
Formation sits atop paraconglomerate and rhythmite of the Carrancas Formation, generally with a sharp contact (Figs. 2a and 5f).
The lower Sete Lagoas Formation is characterized by deep-water
limestone bearing spectacular calcite crystal fan beds interpreted
as aragonite pseudomorphs. An excellent example is found in the
5 m wall of the Sambra Quarry (Inhamas Site), 50 km to the northwest of Belo Horizonte (Peryt et al., 1990; Hoppe et al., 2002; Vieira
et al., 2007; Babinski et al., 2007). Up section, storm-wave and tide
inuenced limestones predominate (Vieira et al., 2007). The 13 C
prole of the Sambra Quarry is identical to proles of the laminated
limestones that overlie the basal cap dolostone in other parts of the
So Francisco basin, with negative 13 C around 5 at the base,
quickly rising upwards, within 20 m, to values around 0 (Kaufman
et al., 2001; Vieira et al., 2007; Babinski et al., 2007). Aragonite
pseudomorphs were also described by Kuchenbecker (2011) in the
Arcos-Pains region, on the southern So Francisco craton (Fig. 2b),
in the same stratigraphic position.
The top of the rst shallowing-upwards cycle is characterized by
dissolution features, tepees, mud cracks, dolomitization and other
facies changes, as well as subtle variations in regional dip (Martins
and Lemos, 2007). Isotopically, this surface is marked by a positive
shift in 13 C values from around +4 below to +8 above, while in
other places this increase is less abrupt, with a gradual rise towards
more positive values (Fig. 2). This surface is also recognizable in
seismic proles (Zaln and Romeiro-Silva, 2007), and is, therefore,
an important unconformity in the middle of the Sete Lagoas Formation, which provides a chronostratigraphic tie that can be traced
across the whole So Francisco Basin (Fig. 2).
Above this unconformity, the upper Sete Lagoas Formation, up
to 160 m thick, is characterized by deep-water grey micrite, pelitic
rhythmite, and black, bituminous, crystalline limestone, locally
with dark Gymnosolenida columnar stromatolites (Marchese, 1974;
Vieira et al., 2007). Upwards, wave-related structures indicate the
progressive shallowing of the platform. The upper Sete Lagoas Formation generally displays 13 C values around +8 near the base,
rising upwards to values >10 (Fig. 2). The upper Sete Lagoas Formation is superseded by deep-water pelites of the Serra de Santa
Helena Formation, marking a ooding of the basin and the beginning of a new regressive cycle (Dardenne, 2000).

87 Sr/86 Sr values of 0.70740.7076 are characteristic of both the


lower and upper Sete Lagoas Formation everywhere (Fig. 2; Chang
et al., 1993; Kawashita, 1998; Misi and Veizer, 1998; Kaufman
et al., 2001; Babinski et al., 2007; Misi et al., 2007; Alvarenga et al.,
2007; Kuchenbecker, 2011; Lima, 2011). Black, oolitic limestones
of the Lagoa do Jacar Formation, in the upper Bambu Group, also
yield exceptionally high 13 C (up to +15) and exactly the same
87 Sr/86 Sr values (Kawashita, 1998; Misi and Veizer, 1998; Kaufman
et al., 2001; Santos et al., 2004; Misi et al., 2007; Lima, 2011). This
lack of secular variation in 87 Sr/86 Sr suggests the Bambu Group
was deposited over a relatively short time span (Misi et al., 2007).
Various attempts to directly date the deposition of the Sete
Lagoas Formation carbonates using the Pb-Pb whole rock method
have been made; most, however, have produced spurious or conicting results (Babinski et al., 1999). Most of the Pb-Pb ages,
around 520 Ma, are younger than the peak of tectonic activity
in the surrounding fold belts (630570 Ma), whose compressional deformation affects the Bambu Group at the cratonic
margins. Therefore these ages cannot represent the depositional
age of the carbonates. Other data sets suggest contamination
with Pb-bearing uids coming from the Archaean-Paleoproterozoic
basement (related to the genesis of Pb-Zn deposits within the upper
Sete Lagoas Formation, e.g. Misi et al., 2005), with regressions yielding apparent ages varying between 2.5 and 1.9 Ga or intercepting
the Pb growth curve at about 2.1 Ga (Babinski et al., 1999; Santos
et al., 2004).
Babinski et al. (1999) and DAgrella-Filho et al. (2000) proposed that a large scale uid percolation event reset the Pb isotope
compositions of the Bambu basin and caused widespread remagnetization throughout the So Francisco craton at the end of the
Brasiliano Orogeny. As a consequence, interpretation of Pb isotope
data and Pb-Pb ages for the Sete Lagoas Formation carbonates is
not straightforward. Nevertheless, Babinski and Kaufman (2003)
and Babinski et al. (2007) presented a Pb-Pb whole rock isochron
with an age of 740 22 Ma from samples of the crystal-fan facies
in the Sambra Quarry. This data is regarded by most authors as
the best evidence for a mid-Cryogenian (post-Sturtian) age of the
Sete Lagoas Formation. Recently, however, Rodrigues (2008) recovered Ediacaran-aged zircons from a foliated siltstone near the base
of the upper Sete Lagoas Formation, only 20 km to the southeast
of the Sambra Quarry. Analyses of 63 zircons (concordant within
10%) yielded a simple U-Pb provenance pattern with a dominant
peak at 650 Ma. A younger population of ve grains yielded a maximum depositional age for the upper Sete Lagoas Formation of
610 Ma. Consequently, the upper Sete Lagoas Formation is younger
than 610 Ma. If one considers the 740 22 Ma Pb-Pb isochron on
the Sambra Quarry to represent the age of deposition of these
carbonates, the middle Sete Lagoas unconformity would have to
represent a hiatus of 110 Ma. As previously noted, however, identical 87 Sr/86 Sr values below and above the unconformity argue
against a long-lived hiatus, particularly when one considers the
substantial long-term increase in 87 Sr/86 Sr through the Neoproterozoic (Halverson et al., 2007a). It seems more likely that the
identical values imply a short-lived break in deposition between
the lithologically similar lower and upper Sete Lagoas formations.
Furthermore, the abrupt break in 13 C values associated with this
surface is not observed everywhere in the So Francisco basin. In
places, a pattern of sharply rising 13 C is observed instead (e.g.
at Serra do Ramalho - Bahia, Misi et al., 2007; Fig. 2d), suggesting a combination of rapidly changing seawater composition or
condensed section, but inconsistent with a >100 m.y. depositional
hiatus.
Sr is a more signicant component of carbonates than Pb, with
primary concentrations of several thousand ppm, and is not prone
to redox changes as is U. Least altered 87 Sr/86 Sr values are identical
for carbonates with the same 13 C between Sambra and Tatiana

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858

quarries, just about 5 km apart, despite the fact that the latter
are heavily recrystalized and yielded a younger overlapping Pb-Pb
isochron of 681 50 Ma (Babinski et al., 2007). Hence, Sr isotopic
signatures clearly withstood the Pb isotopic resetting event proposed by Babinski et al. (1999) and DAgrella-Filho et al. (2000).
2.5. Fold belts
A number of diamictite-bearing units in the fold belts that surround the So Francisco craton are thought to record the precursor
rift to passive margin sequences, which were inverted and metamorphosed during the Brasiliano Orogeny, around 630570 Ma.
These sequences are generally thicker and more complex than the
diamictite-bearing cratonic units (Fig. 3); nevertheless, they are
broadly considered as chronostratigraphic equivalents that reect
a deep marine continuation of the cratonic platform sedimentation, mainly by down-slope gravity ow deposition (e.g. Uhlein
et al., 1999, 2004; Sial et al., 2009; Pedrosa-Soares et al., 2011). The
Sergipano fold belt is the only belt that contains a thick carbonate
succession; the others comprise mainly siliciclastic sediments with
no cap carbonate sequence. The younger detrital zircon U-Pb age
populations of most of these units yielded very similar Tonian ages
(8501000 Ma; Fig. 3; Pedrosa-Soares et al., 2000; Rodrigues et al.,
2010; Caxito et al., 2011; Amorim Dias, 2011; Babinski et al., 2012).
The only exceptions are the Rio Verde Formation schists, which
pass vertically downwards to the Cubato Formation diamictites of
the Braslia fold belt, and the Palestina Formation in the Sergipano
fold belt, both of which yielded 650640 Ma zircons (Fig. 3b and
d; Oliveira et al., 2005, 2006, 2010; Rodrigues et al., 2010; Amorim
Dias, 2011). As the onset of compressional deformation in both fold
belts is estimated at about 630 Ma (Pimentel et al., 1999; Dardenne,
2000; Valeriano et al., 2004; Oliveira et al., 2006, 2010), the deposition of these two units is tightly constrained to 640630 Ma,
which suggests a correlation with the global 635 Ma Marinoan
(end-Cryogenian) glaciation (Hoffmann et al., 2004; Condon et al.,
2005).
3. New data
3.1. Materials and methods
We present lithostratigraphic and isotopic data for a previously
unstudied section of the Sete Lagoas Formation in the north central
part of the So Francisco basin, near the town of Correntina, Bahia
(Figs. 2 and 6). Here, Archean to Paleoproterozoic gneiss and granites of the So Francisco craton basement (Cordani et al., 1979) are
directly overlain by the Sete Lagoas Formation carbonates (Fig. 6).
The section was sampled on a SN transect on the unpaved road
between Jaborandi and So Manoel, where the basement is exposed
(Fig. 6). Downcutting through the Paleogene peneplane surface
(King, 1956) by the Arrojado River provides access to a 127 m-thick
section of the Sete Lagoas Formation (Fig. 6).
The Correntina section begins with a two-meter thick massive to nely laminated pink dolostone (Fig. 7a), which grades
upward into a reddish to purple limestone rhythmite composed of
centimeter-scale undulating layers of calcilutite intercalated with
marly siltstone (Fig. 7b). The most common sedimentary features
are plane-parallel and undulating laminations, wavy and linsen
cross-laminations, marly mud drapes, cross-bedding in the marl
layers, and, locally, mud cracks, suggesting deposition in a tidal at
environment. This laminated limestone shows no facies change for
123 m; the top of the section is a black, plane-parallel laminated
calcarenite. Although no diamictite was observed between the Sete
Lagoas Formation and the basement, thin, meter-scale diamictite
layers at the base of the Sete Lagoas Formation have been described

45

in the Coribe region, 40 km to the south (Conceico Filho et al.,


2003). In lithostratigraphic terms, the Correntina section is very
similar to Sete Lagoas Formation sections fringing the northern
Braslia fold belt (Alvarenga et al., 2007), to the west, and in the
Serra do Ramalho area, to the east (Misi and Veizer, 1998; Misi
et al., 2007). It is unlike the classical Sete Lagoas Formation outcrops
between Belo Horizonte and Sete Lagoas that are characterized by
deep marine crystal-fan facies (Fig. 2).
In addition to the new section, we analyzed sixty-ve cobblesized carbonate clasts from four separate diamictite occurrences
on the So Francisco craton and in the surrounding Aracua and Rio
Preto fold belts (Fig. 2). Although this approach is not new, isotopic
data on carbonate clasts from these diamictites have not been previously pursued. As carbonate clasts are a common feature of most
diamictite occurrences (Fig. 5ce), they preserve information about
the source areas of these units, and therefore could yield signicant
provenance data and chemostratigraphic constraints.
3.2. Sampling and analytical procedures
Forty-seven samples of the Correntina section were collected
and slabbed perpendicular to lamination; thin sections were made
for the most representative samples. The petrographic analysis
aided in the choice of fresh and homogenous siliciclastic-free subdomains, lacking any recrystallization features; these domains
were micro-drilled to produce the powders from which aliquots
were used in the chemical and isotopic analysis. The same
methodology was used for the carbonate clast analysis. Sixty-ve
hand-sized specimens were collected and only the central part of
the clasts, free of any veins and matrix, were slabbed and microdrilled.
Carbon and oxygen isotope ratios were measured on a MicroMass Isoprime DI MulticarbTM mass spectrometer at the Stable
Isotopes Laboratory at GEOTOPUQAM in Montral, Canada. Circa
1 mg of each powdered sample was weighted in glass vials and
reacted individually with H3 PO4 after heating to 90 C for 1 h.
The released CO2 was collected through a coldnger and analyzed
using an in-house reference gas. Samples were calibrated to VPDB
(Vienna Pee Dee Belemnite) using standards NBS-18 and UQ6.
Errors are about 0.05 (1) for both 13 C and 18 O. 10% of the
samples were analyzed in duplicate to ensure data reproducibility.
Another aliquot of the powdered samples was taken for Sr
concentration and isotopic analysis. Approximately 15 mg of 40
selected samples was weighed in 1.5 ml centrifuge tubes and
leached sequentially twice in 1.0 ml 0.2 M ammonium acetate, for
half hour, in order to remove loosely bound Sr cations, including

radiogenic 87 Sr derived from 87 Rb decay (Montanez


et al., 1996;
Bailey et al., 2000; Halverson et al., 2007a). Samples were then
washed three times using ultrapure H2 O, and then reacted for 2 h in
1.0 ml 0.5 M acetic acid. After centrifuging, approximately 0.9 ml of
dissolved sample was recovered and placed in a Teon beaker.
This procedure recovers only the calcite fraction of the sample.
Samples were then dried down and dissolved in 1.0 ml 3 N HNO3 ,
from which a 0.5 ml aliquot was separated and diluted with ultrapure water to about 0.3 N HNO3 for concentration analysis. Sr was
separated from the remaining half of the sample via standard chromatographic techniques using EIChroM Sr-spec resin and eluted
with ultrapure H2 O. This procedure was repeated twice to ensure
that no Rb remained in the samples.
Sr concentrations were determined by Atomic Absorption on
a Perkin Elmer AAnalyst 100 spectrometer using N2 O and C2 H2
ames at the Trace Element Analysis Laboratory (TEAL), McGill University, Montral, and referenced against internal standards. Total
procedural blanks are negligible.
Sr isotope ratios were analyzed by thermal ionization mass spectrometry (TIMS) on a ThermoScientic Triton Plus multicollector

46

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858

Fig. 6. Simplied geological map and section of the Correntina region, BA. For location see Fig. 2. The map is in part from Souza et al. (2004).

mass spectrometer operating in static mode at the Radiogenic Isotope Geochemistry Laboratory at GEOTOPUQAM. All data were
corrected for internal mass bias using 86 Sr/88 Sr = 0.1194. Repeated
measurements of the NBS SRM 987 standard yielded a long-term
average of 0.710245 0.000007 (n = 7).
3.3. Results
3.3.1. Correntina section
The Correntina section yielded a coupled 13 C and 18 O pattern
(Table 1; Fig. 8). The basal pink dolostone preserves a drop in 13 C
values from 3.8 to 4.2 through the rst 1.5 m, accompanied by
a drop in 18 O values from 4.5 to 5.1. The red laminated limestone that sits immediately above it yielded 13 C values averaging

5 and 18 O around 11 for about 60 m up section. Above 60 m,


the 13 C values shift abruptly to around 0.5, accompanied by a
shift in 18 O values to 7. This change in isotopic values does
not correspond to a facies change or any obvious depositional hiatus. The values remain nearly constant through the remainder of
the section up to a massive black calcarenite where the only positive 13 C value is observed (+0.9). The section is capped by a
laminated black calcarenite with a 13 C value of 1.7.
Overall, the 13 C and 18 O isotopic prole of the Sete Lagoas
Formation in the Correntina region is similar to proles obtained
by Alvarenga et al. (2007) in the Formosa region to the west,
and fairly similar to other proles across the So Francisco basin
(Fig. 2bd; Misi et al., 2007; Kuchenbecker, 2011). Hence, the
Correntina section bridges the central part of the basin and the

Fig. 7. Sete Lagoas Formation carbonates at Correntina, Bahia: (a) hand sample of the pink, thinly laminated cap dolostone at the base of the section (0.5 meters); (b)
plane-parallel and undulose laminated limestone and marl intercalations; width of hammers head is about 10 cm.

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858

47

Table 1
13 C, 18 O and Sr isotope data from the Sete Lagoas Formation at Correntina, Bahia. 13 C and 18 O values in PDB (normalized to Pee Dee Belemnite).
Sample

Height to sea level (m)

COR0
COR0.5
COR1
COR1.5
COR2
COR7
COR8
COR9
COR10
COR11
COR12
COR13
COR20
COR20.5
COR21
COR21.5
COR22
COR22.5
COR28
COR36
COR37
COR46
COR47
COR50
COR51
COR52
COR53
COR54
COR55
COR60
COR61
COR62
COR63
COR64
COR65
COR66
COR67
COR68
COR69
COR70
COR71
COR72
COR75
COR97
COR105
COR126
COR127

525
525.5
526
526.5
527
532
533
534
535
536
537
538
545
545.5
546
546.5
547
547.5
553
561
562
571
572
575
576
577
578
579
580
585
586
587
588
589
590
591
592
593
594
595
596
597
600
622
630
651
652

13 C
3.77
3.89
3.94
4.17
5.75
5.60
5.67
5.43
5.26
4.95
4.75
5.31
5.67
5.66
5.70
6.05
5.72
5.61
5.07
5.02
5.59
4.72
4.43
5.21
5.18
5.04
4.91
4.85
4.28
0.30
0.35
0.37
0.36
0.54
0.52
0.57
0.59
0.58
0.59
0.78
0.31
0.37
0.97
0.88
0.79
0.99
1.72

sections fringing the Brasiliano fold belt margins (Fig. 2). However,
it differs signicantly from sections in the Sete Lagoas region
(on the southeastern margin of the So Francisco craton) insofar
as these lack the basal pink dolostone layer which preserves a
drop of 13 C values and coupled decline in 18 O (Fig. 2a). The rise
from negative values around 5 towards values around 0 in
the lower Sete Lagoas Formation usually occurs within 210 m.
Therefore, the isotopic break observed between 55 and 60 m in
the Correntina section probably correlates with this abrupt rise,
which is recorded by the crystal-fan facies on the Sete Lagoas and
Arcos-Pains regions (Fig. 2a and b; Vieira et al., 2007; Babinski
et al., 2007; Kuchenbecker, 2011). Hence, the Correntina section
likely represents the lower Sete Lagoas Formation.
Samples of the Correntina section above the 60 m mark have
Sr concentrations ([Sr]) between 250300 ppm, while a limestone
sample at 12 m have 173 ppm Sr and a sample from the cap dolostone is less than 50 ppm. Most of the samples have 87 Sr/86 Sr
ratios around 0.7080 (Table 1). The least radiogenic 87 Sr/86 Sr ratios,
around 0.7076, were obtained on two samples of the black calcarenite at the top of the section, which have the highest [Sr],
above 1000 ppm. These results are broadly consistent with previously reported 87 Sr/86 Sr for the Sete Lagoas Formation (Table 4;

18 O
4.48
4.60
4.80
5.10
10.79
10.86
11.02
10.91
10.76
10.48
10.43
12.59
11.04
10.98
10.98
11.66
10.91
10.77
13.34
7.68
12.23
11.05
10.70
11.06
10.91
10.78
10.62
10.67
8.18
7.30
7.34
7.22
7.26
7.24
7.21
7.25
7.23
7.12
7.14
7.14
7.39
7.28
6.95
6.77
7.21
6.65
6.46

Sr (ppm)

87

Sr/86 Sr (2)

<50

173.11

327.74
243.61

0.708021 (19)

246.79
256.13

0.708146 (11)

284.69

312.01

296.87
308.42
319.24
2171.76
1584.38

0.708005 (14)
0.708176 (8)
0.708499 (21)
0.707598 (5)
0.707567 (5)

Chang et al., 1993; Kawashita, 1998; Misi and Veizer, 1998; Babinski
et al., 2007; Alvarenga et al., 2007; Misi et al., 2007; Kuchenbecker,
2011; Lima, 2011). We tentatively consider a minimum cut-off [Sr]
of 1000 ppm as indicating the least altered 87 Sr/86 Sr to best approximate the original seawater ratio of about 0.7076.
3.3.2. Carbonate clast data
The wide range in 13 C values and consistently increasing
87 Sr/86 Sr signatures that characterizes Neoproterozoic seawater (Halverson et al., 2010; Halverson and Shields-Zhou, 2011)
make carbonate clast data (Table 2; Fig. 9) a promising tool for
provenance and chronological constraints (cf. Hork and Evans,
2011). Unfortunately, Sr concentrations are invariably lowered by
post-depositional alteration (Brand and Veizer, 1980; Banner and
Hanson, 1990; Halverson et al., 2007a), which hampers the recovery of primary 87 Sr/86 Sr ratios on carbonate clasts. Most of the
clasts analyzed yielded [Sr] < 50 ppm, with some preserving [Sr] of
around 100 ppm. The radiogenic signature of the 87 Sr/86 Sr ratios
(0.7130.728) clearly indicates that some samples have been heavily altered with respect to Sr. Despite the attempt to collect only
unweathered and homogeneous clasts, the fact that they represent
eroded fragments from the glacial substrate might explain their

48

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858

Fig. 8. 13 C and 18 O isotopic proles of the Correntina section. The insets magnify the basal 10 m of section.

Fig. 9. 13 C-18 O diagram of carbonate clasts compared with possible source areas in east central Brazil. See Table 2 for data. Possible source elds compiled from data from
the Paleoproterozoic Gandarela and Fecho do Funil formations (Sial et al., 2000; Bekker et al., 2003; Maheshwari et al., 2010); Mesoproterozoic Rio Pardo Grande Formation
and Parano Group (Santos et al., 2000, 2004; Alvarenga et al., 2007) and the Lapa Formation (Azmy et al., 2006), of purported late Mesoproterozoic age (Azmy et al., 2008).

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858


Table 2
13 C, 18 O and Sr isotope data from carbonate clasts, arranged by crescent 13 C for
each diamictite occurrence. 13 C and 18 O values in PDB (normalized to Pee Dee
Belemnite).
18 O

Sr (ppm)

87

Carrancas Formation
4.05
CAR7
4.23
CAR4
CAR6
4.13
CAR3
4.01
CAR2
4.00
CAR5
3.76
CAR1
3.74

11.23
6.83
8.03
6.56
6.72
6.53
6.32

155.14
<50
77.62
<50
<50
<50
<50

0.713020 (18)

Jequita Formation
JEQ14
JEQ03
JEQ18
JEQ19
JEQ12
JEQ24
JEQ17
JEQ15
JEQ01
JEQ02
JEQ13
JEQ09
JEQ20
JEQ04
JEQ21
JEQ05
JEQ27
JEQ07
JEQ06
JEQ11
JEQ23
JEQ25
JEQ29
JEQ08
JEQ30
JEQ22
JEQ26
JEQ28
JEQ16

10.17
9.83
13.88
10.80
13.31
11.48
11.08
11.09
9.76
12.75
14.35
12.13
11.24
10.12
11.18
12.22
10.61
9.68
12.03
11.54
9.85
10.16
10.20
8.69
10.22
9.80
9.66
6.86
7.06

Sample

13 C

7.70
4.73
4.06
3.83
3.15
2.95
2.37
2.33
1.78
1.60
1.46
1.44
0.86
0.84
0.81
0.81
0.79
0.61
0.59
0.45
0.33
0.29
0.05
0.00
0.01
0.55
0.64
1.91
2.12

Serra do Catuni Formation


SC09
6.75
SC07
6.32
SC04
6.31
SC08
6.21
5.94
SC13
5.85
SC06
5.21
SC11
4.89
SC05
4.69
SC15
3.86
SC03
3.80
SC18
3.53
SC12
2.85
SC10
1.55
SC01
2.41
SC17
2.60
SC02
7.60
SC19

13.05
10.08
12.33
12.78
11.61
13.11
11.99
12.84
15.16
15.59
15.37
12.32
8.52
14.85
9.31
7.68
15.36

Canabravinha Formation
CAN10
4.47
3.30
CAN6
3.09
CAN13
2.84
CAN8
2.81
CAN2
2.49
CAN9
2.48
CAN14
1.74
CAN11
1.44
CAN5
1.41
CAN1
0.11
CAN4
0.00
CAN7
3.01
J0.5 (matrix)

12.74
13.09
12.93
10.97
12.07
12.07
12.32
10.64
11.80
10.83
10.69
10.64
12.91

Sr/ 86 Sr ( 2)

49

poor retention of primary Sr isotope compositions and concentrations.


3.3.2.1. Canabravinha formation. At Canabravinha Creek, near
Monte Alegre dos Cardosos, BA (Figs. 2 and 4), 0.51 m thick layers
of metadiamictite are intercalated with lithic quartzite containing
graded bedding, climbing ripples and ripped-up shale clasts, and
with sericite-carbonate phyllite layers devoid of clasts. These rocks
are metamorphosed to lower greenschist facies. We collected 12
angular cobble-sized gray to black metalimestone clasts (Fig. 5e)
from the metadiamictite layers, which yielded a range of 13 C values from 4.5 to 0. 18 O values, on the other hand, exhibit a
more limited range of 13.1 to 10.6 (Table 2; Fig. 9).
Sample J0.5 is from a metamarl lens separating metadiamictite
and quartzite layers. This sample yielded a 13 C value of 3.0 and
a 18 O of 12.9 (Table 2). We believe that this is a detrital signature and is approximately equal to the metadiamictite matrix
values (which were otherwise not analyzed). No cap carbonate
sequence has been documented in the Rio Preto fold belt.

<50

<50

<50

<50
<50
<50
<50
<50

113.69

0.728088 (30)

<50
<50
<50

<50

99.18
98.09
<50
<50

0.729834 (8)
0.727384 (71)

3.3.2.2. Serra do Catuni Formation. Carbonate clasts from the Serra


do Catuni Formation of the Macabas Group, in the Aracua fold
belt, were collected in an inactive quarry at Couto de Magalhes,
22 km north of Diamantina, Minas Gerais (Figs. 2, 4b and 5d). As for
the Canabravinha Formation, rocks of the Serra do Catuni Formation
are also metamorphosed to lower greenschist facies. We collected
17 clasts which are mostly angular cobble- to boulder-sized, white
and light grey massive metalimestone (Fig. 5d). Sample SC03, a dark
metalimestone, is the only exception. These clasts yielded highly
variable 13 C, from 6.8 to +2.6 (Fig. 9). Sample SC19 yielded
the highest 13 C value, at +7.6. 18 O values are also quite variable,
between 15.6 and 8.5.
Stable isotope data on carbonate clasts from the Serra do Catuni
Formation were also reported by Martins (2006). Samples of 14
carbonate clasts were collected from metadiamictite layers outcropping in the Macabas and Onca riverbeds and yielded a wide
variation in 13 C values, from 7 to 0, with associated 18 O values ranging from 13.9 to 8.8. These results fall within the
range of our own data (Fig. 9).
3.3.2.3. Jequita Formation. An abandoned quarry at km 66 of the
BR-365 road, 36 km northeast of Jequita, provides one of the best
exposures of the Jequita Formation diamictite (Figs. 4b and 5c).
Here, the massive diamictite contains varied carbonate fragments,
as well as gneiss, granite, quartzite and metapelite clasts. The clasts
are angular to subrounded, up to boulder size, oating in a grey
pelitic matrix. Carbonate clasts comprise grey and black massive
and laminated limestone, stromatolites, and dolostone containing
grains of detrital quartz (Fig. 5c). A total of 29 carbonate clasts were
analyzed and those yielded 13 C values from 4.7 to +2.1, with
associated 18 O from 14 to 6.8, with the exception of a single
dark calcarenite clast (JEQ14) which yielded a very low 13 C value
of 7.7. Although similar, these values show less variation than
the Serra do Catuni Formation data (Table 2; Fig. 9).
Santos et al. (2004) previously reported 13 C between 2.1
and 0.6, with associated 18 O from 11.0 to 12.9, for ve
carbonate clasts of the Jequita Formation (Fig. 9), and 13 C from
3.1 to 2.1 and 18 O from 9.8 to 11.8 on four samples
of the matrix of the diamictite.
3.3.2.4. Carrancas Formation. We collected 7 carbonate clasts from
the Carrancas Formation paraconglomerate at its type outcrop, at
km 30 of the MG-424 road. Unlike the other diamictite occurrences,
carbonate clasts are mostly pale pink, rounded, cobble-sized inty
dolostone, indistinguishable from the thin dolostone found at the
base of the Sete Lagoas Formation elsewhere. Large cobbles and

50

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858

Table 3
Disponible stable isotope data from the thin cap dolostone unit at the base of the Sete Lagoas Formation. 13 C and 18 O values in PDB (normalized to Pee Dee Belemnite).
The top sample in each column is from the rst occurence of limestone above the cap dolostone, highlighting the 13 C and 18 O negative jump in the dolostone-limestone
transition.
Alvarenga et al. (2007)

Kuchenbecker (2011)

This study

Sample

Height (m)

13 C

18 O

Sample

Height (m)

13 C

18 O

Sample

Height (m)

13 C

18 O

Bz18E
Bz18D
Bz18C
Bz18B

7
6.5
3
1

5.7
4.7
4.4
3.2

9.1
5.3
5.4
5.6

M9
M8
M7
M6
M5
M4
M3

9
8
7
6
5
4
3

5.38
4.44
4.37
4.23
4.38
4.05
3.72

12.99
8.37
8.56
11.59
11.27
6.18
6.49

COR2
COR1.5
COR1
COR0.5
COR0

2
1.5
1
0.5
0

5.75
4.17
3.94
3.89
3.77

10.79
5.10
4.80
4.60
4.48

boulders of grey limestone are found near the top of the layer, close
to the sharp contact with the Sete Lagoas Formation (Fig. 5f). These
clasts are very lithologically similar to the overlying limestone of
the Sete Lagoas Formation. We collected two samples of these limestone clasts (samples CAR6 and CAR7). The other ve samples are
from dolostone clasts (CAR1 to CAR5).
The ve dolostone clasts yielded a restricted range of 13 C from
3.7 to 4.2, and 18 O from 6.3 to 6.8. These values are
different from any of the other analyzed diamictite clasts, but virtually identical to the basal Sete Lagoas dolostone values (Fig. 9).
The samples of the two limestone clasts (CAR6 and CAR7) yielded

13 C of 4.1 and 18 O of 8.0 and 11.2, respectively (Fig. 9),


which are nearly identical to the composition of the nearby lower
Sete Lagoas Formation limestone and also to the paraconglomerate matrix in the same section (Fig. 10; Vieira et al., 2007). These
distinct results reinforce the primary nature of the dolostone clast
data, as the 18 O values are much higher than the surrounding
matrix (Fig. 10).
3.3.3. Interpretation of the carbonate clast data
The stable isotope data suggests that most of the clasts came
from the erosion of a pre-glacial, carbonate platform with negative

Table 4
Sete Lagoas Formation 13 C, 18 O and least radiogenic 87 Sr/86 Sr data. Data from the Lagoa do Jacar Formation (upper Bambu Group) and from the Salitre Formation, Una
Group in Bahia are added for comparison.
Sample

13 C

18 O

Sr (ppm)

Mn/Sr

87

COR126
COR127
BB 16H
BB 16N
M40
M45
M46
M48
TMG16
TMG17
TMG18
TMG19
LU.10.4
LU.10.5
LU.10.6
Bz38
Bz37
Bz36
3.5
SR AM 111
MF 7-C

0.99
1.72
0.08
0.2
1.5
1.4
1.6
1.9
3.8
3.7
3.1
2.2
2.8
1.8
0.6
6.3
2.7
3.6

1.4

6.65
6.46
7.54
7.57
7.9
8.4
7.8
6.7
7.8
8.2
7.3
7.9
10.7
10.3
8.6
5
6
5.5

2171.76
1584.38
2328
643
1618
1817
1957
3304

1990
1616
3064
1733
1590
2050

0
0.2

0.02
0.13
0.08
0.01
0.01

0.707598
0.707567
0.70768
0.7078
0.707683
0.707622
0.707648
0.707493
0.7074
0.7075
0.7076
0.7081
0.7079
0.7081
0.7082
0.70748
0.70758
0.70745
0.70745
0.70755
0.70739

This study
This study
Lima (2011)
Lima (2011)
Kuchenbecker (2011)
Kuchenbecker (2011)
Kuchenbecker (2011)
Kuchenbecker (2011)
Vieira et al. (2007), Babinski et al. (2007)
Vieira et al. (2007), Babinski et al. (2007)
Vieira et al. (2007), Babinski et al. (2007)
Vieira et al. (2007), Babinski et al. (2007)
Vieira et al. (2007), Babinski et al. (2007)
Vieira et al. (2007), Babinski et al. (2007)
Vieira et al. (2007), Babinski et al. (2007)
Alvarenga et al. (2007)
Alvarenga et al. (2007)
Alvarenga et al. (2007)
Powis et al. (2001) in Misi et al. (2007)
Misi and Veizer (1998)
Kawashita (1998)

Sr/86 Sr

Reference

Lagoa do Jacar Formation


BDI 31A

BDI 31B

BDI 31C

BDI 31D
CA-IE-14

14
SR AM 1

3H

7.1

1260
954
1189
625
3874
1430
4050

0.04
0.03
0.02
0.02
0.01
0.01

0.70749
0.70756
0.70744
0.70757
0.70743
0.70767
0.70738

Lima (2011)
Lima (2011)
Lima (2011)
Lima (2011)
Powis et al. (2001) in Misi et al. (2007)
Misi and Veizer (1998)
Kawashita (1998)

Salitre Formation, Una Group


8.3
VML8
JC54
8.1
8M03
8
0.6
IR-AM-11
1.1
VML 7A
0.1
VML 6a
0.3
VML 6B
0.6
VML 6C
1.1
VML 7B
VML3
0.3

6.4
5.9
5.2
6.3
6.5
7.2
7
7.5
6.5
6.2

461
792
374
2384
1275
896
790
963
1429
1717

0.03
0.02
0.05
0.03
0.01
0.01
0.01
0.09
0.01
0.02

0.70769
0.70745
0.70759
0.70752
0.70765
0.70782
0.7078
0.70789
0.70765
0.7078

Misi and Veizer (1998)


Misi and Veizer (1998)
Misi and Veizer (1998)
Misi and Veizer (1998)
Misi and Veizer (1998)
Misi and Veizer (1998)
Misi and Veizer (1998)
Misi and Veizer (1998)
Misi and Veizer (1998)
Misi and Veizer (1998)

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858

Fig. 10. 13 C-18 O diagram of carbonate clast data from the Carrancas Formation
as compared to the thin cap dolostone layer found at the base of the Sete Lagoas
Formation (Table 3).

13 C. This result is hardly surprising, given that Neoproterozoic


glaciations are preceded worldwide by negative 13 C anomalies (Halverson et al., 2010; Halverson and Shields-Zhou, 2011).
Carbonate clasts from the end-Cryogenian (635 Ma) Ghaub
diamictite on Namibia, for example, display very similar results
(Halverson et al., 2005; Hoffman, 2011).
The isotopic data contradicts the informal assumption that carbonate clasts of diamictites in east central Brazil must all have
come from purported Mesoproterozoic successions containing carbonate sequences (Espinhaco Supergroup and Parano Group;
Fig. 9). The carbonates of these Mesoproterozoic successions bear
a monotonous 13 C signature between +0.1 and +2.7, with
87 Sr/86 Sr ranging from 0.7063 to 0.7068 (Santos et al., 2000, 2004;
Alvarenga et al., 2007; Fig. 9), and therefore are not potential
sources of negative 13 C clasts, although some of the positive 13 C
clasts could have come from these older units. An important question then arises: where, in east central Brazil, can a pre-glacial
carbonate platform with negative 13 C be found? This is a challenging question given that most of the sedimentary sequences
preceding diamictites in the So Francisco craton and surrounding
fold belts are predominantly siliciclastic. The only obvious exception is the Sergipano Fold Belt, where Sial et al. (2010) reported
negative 13 C values in carbonates of the Jacoca and Acau formations, which precede the end-Cryogenian Palestina Formation
diamictite (Fig. 3d).
The Domingas Formation in the Aracua fold belt is a stromatolitic dolostone unit that precedes the diamictite-bearing units
of the Macabas Group (not to be confused with the cratonic
Jequita Formation). Santos et al. (2004) presented a 10-sample
isotopic prole of the Domingas Formation, in which 13 C varied
from +0.7 at the base to 1.6 at the top. Although low resolution, this prole clearly demonstrates gradually lower 13 C values
towards the top, reminiscent of the worldwide Islay or Trezona
pre-glacial anomalies (Halverson et al., 2010). It could, therefore,
represent a prelude to the Jequita glaciation. Hence, Jequita Formation clasts, as proposed by Santos et al. (2004), could in part
be derived from the Domingas Formation. However, the Domingas
Formation is currently considered to belong to the Mesoproterozoic
Espinhaco Supergroup (Pedrosa-Soares et al., 2011) rather than the
basal Macabas Group, as previously proposed by Noce et al. (1997).

51

This controversy and the possibility that the Domingas Formation


was an important source of the clasts in the Jequita diamictites
highlights the need for additional eld and analytical data on the
carbonates in the Aracua fold belt.
The only other nearby negative 13 C carbonate unit older than
the Jequita Formation is the Lapa Formation of the Vazante Group,
in the Braslia fold belt (Figs. 1 and 9). The Lapa Formation is
composed mainly of laminated argillaceous dolostones, sitting
above a thin diamictite layer (DII, Azmy et al., 2006). Its isotopic
prole starts with strongly negative 13 C, around 8, quickly
rises to values around 0 and then remains nearly constant for
about 400 m (Azmy et al., 2006). Another negative 13 C anomaly
is found near the top of the succession, above a black shale horizon. The isotopic prole of the Lapa Formation (Misi et al., 2007;
Azmy et al., 2008) is similar to the pre-glacial Islay or Trezona
anomalies. Although the 87 Sr/86 Sr value of 0.7068 from the leastaltered dolomite sample is quite similar to mid-Cryogenian values
worldwide (and distinct from end-Cryogenian values), Azmy et al.
(2008) assigned a late Mesoproterozoic age for the Lapa Formation
(10001100 Ma), based on two Re-Os isochrons of 993 46 and
1100 77 Ma obtained from black shale samples associated with
the DII diamictite at the base of the Lapa Formation.
Another potential source of isotopically negative carbonates
is Paleoproterozoic strata on the So Francisco craton. Some
sequences preserve negative 13 C signatures (Fig. 9), most notably
in the Quadriltero Ferrfero region on the southeastern tip of the
So Francisco craton, south of Belo Horizonte (Figs. 1 and 2; Sial
et al., 2000; Bekker et al., 2003). Here, the 2.4 Ga Gandarela Formation yields 13 C values spanning a narrow range of 1.5 to +0.5,
while the overlying Fecho do Funil Formation yields positive values
around +6.5, which have been correlated to the Paleoproterozoic
Lomagundi event (2.22.1 Ga; Sial et al., 2000; Maheshwari et al.,
2010).
A plot of the carbonate clast data versus possible source areas
suggests that the clasts could not have come solely from the erosion
of Paleo- and Mesoproterozoic sources (Fig. 9). While the Lapa Formation may account for some of the clasts, it seems likely that the
main sources have been eroded, are covered, or not yet discovered.
The Carrancas Formation contains pale pink and inty dolostone clasts with a very narrow range of 13 C and 18 O values,
distinct from the other diamictite occurrences studied (Fig. 9). The
dolostone clasts have coupled 13 C and 18 O values that are isotopically indistinguishable from samples from the basal Sete Lagoas
cap dolostone, but unlike any other carbonate data from the So
Francisco craton (Table 3; Fig. 10). As previously described, the
Carrancas Formation consists of lenses of paraconglomerate locally
superseded by, or interleaved within, rhythmites. The most parsimonious interpretation of this data is that the Carrancas Formation
sampled the Sete Lagoas Formation itself, and thus post-dates
the Jequita Formation. Given that the Carrancas Formation is in
turn overlain by typical facies of the Sete Lagoas Formation, the
inescapable conclusion is that it is an intraformational, lateral
equivalent of the lower Sete Lagoas Formation. Given the association with rhythmites, we suggest that the Carrancas Formation
formed by the shedding of cap dolostone debris from a paleohigh
or escarpment in the basin into a deeper water environment otherwise characterized by limestone rhythmites. Analogous facies
relationships occur near shelf margin sections of the basal Ediacaran Maieberg Formation in Namibia (Hoffman and Halverson,
2008). This interpretation is consistent with the lack of evidence
for glacial deposition in the Carrancas Formation (Vieira et al.,
2007; Tuller et al., 2009). It may also account for why the cap dolostone, while widespread across most of the basin (Misi et al., 2007;
Alvarenga et al., 2007; Sial et al., 2009; Kuchenbecker, 2011; Lima,
2011; Misi et al., 2011; this study) is absent from the Sete Lagoas
type area.

52

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858

4. Discussion and correlation of the Sete Lagoas Formation


with other cap carbonate successions
Cap carbonate sequences occur above two Cryogenian glaciations the Sturtian (mid-Cryogenian) and Marinoan (endCryogenian; Kennedy et al., 1998; Hoffman et al., 1998) both
of which appear to have been global in extent (Sohl et al., 1998;
Macdonald et al., 2010a). Despite certain shared features, the cap
carbonate sequences of the Sturtian and Marinoan glaciations are
distinct from one another (Kennedy et al., 1998; Halverson et al.,
2005). The post-Marinoan cap carbonate contains unique sedimentary structures, occurs more widely, and is more complete,
preserving a distinctive basal transgressive unit the cap dolostone. This dolostone is typically 330 m thick (Hoffman et al.,
2007) and contains a suite of unique sedimentary structures that
include normal and reverse-graded peloids, plumb stromatolites
and giant wave ripples (Nogueira et al., 2003; Allen and Hoffman,
2005; Shields, 2005; Hoffman et al., 2007; Hoffman, 2011). 13 C
values in the cap dolostone characteristically range between 2
and 4, typically decreasing upwards (Halverson et al., 2005;
Hoffman et al., 2007). Whereas most Neoproterozoic dolomites
have relatively enriched 18 O values due to fractionation associated
with dolomitization (Halverson et al., 2007b), the best-preserved
cap dolostones have 18 O values between 5 and 7 (Shields,
2005; Hoffman et al., 2007). The cap dolostone is generally superseded by limestone that records a 12 negative shift in both 13 C
and 18 O (Halverson et al., 2005; Hoffman and Halverson, 2011).
This lower limestone often contains aragonite crystal fan pseudomorphs (e.g., Hayhook Fm. in NW Canada, Maieberg Fm. in NW
Namibia, Mount Doreen Fm. in Central Australia; Halverson et al.,
2005; Hoffman et al., 2007), implying deposition in a highly calcium
carbonate-saturated watermass. The age of the glacial-cap carbonate boundary is bracketed by a 635.5 1.2 Ma U-Pb TIMS age on
zircons from an ash fall tuff within the Ghaub diamictite in central
Namibia (Hoffmann et al., 2004) and a 635.2 0.6 Ma age on a tuff
in the lower Doushantuo Formation, south China (Condon et al.,
2005). The combination of the unique lithology and geochemistry
of the post-Marinoan cap carbonate, its widespread occurrence, and
the tight radiometric age constraints underpin the selection of its
base as the Global Stratotype Section and Point (GSSP) for the base
of the Ediacaran Period (Knoll et al., 2004).
Post-Sturtian cap carbonates, on the other hand, are usually
base-truncated, exhibiting a sharp basal rise in 13 C values from
4 up to +5. Lithologically, they are dark, organic- or sulphiderich limestones or mixed limestone-dolomite, lacking any of the
distinctive post-Marinoan cap carbonate sedimentary features, but
in places containing microbial roll-up structures (Halverson et al.,
2005; Macdonald et al., 2010b). The age and global character of
the Sturtian glaciation is at present the subject of much debate
(Hoffman et al., 1996; Frimmel et al., 1996; Brasier et al., 2000;
Lund et al., 2003; Fanning and Link, 2004; Halverson et al., 2005;
Kendall et al., 2006; Babinski et al., 2007; Macdonald et al., 2010a).
The Sturtian and Marinoan glaciations are both preceded and
followed by negative isotope anomalies, and the less severe Gaskiers (Ediacaran) glaciation appears to overlap the Shuram (or
Wonoka) 13 C anomaly the most extreme in the Neoproterozoic (Knoll et al., 1986; Hayes et al., 1999; Halverson et al., 2005).
In the case of the Sturtian and Marinoan events, 13 C dropped to
as low as 6 (Halverson, 2006; Prave et al., 2009) in shallow
water carbonates deposited prior to onset of glaciation: the Islay
and Trezona anomalies, respectively. On their own, these anomalies
cannot be used to distinguish between the Sturtian and Marinoan glaciations. However, whereas seawater 87 Sr/86 Sr ratios were
0.7066 during the Islay anomaly, they were 0.7072 during the
Trezona anomaly (Halverson and Shields-Zhou, 2011). Therefore,
the combination of primary carbon and strontium isotope data is

a robust tool to distinguish between Sturtian and Marinoan glacial


deposits where well preserved, immediately pre-glacial carbonates
occur.
The strontium isotope ratio of 0.70720.7073 at the onset of the
Trezona anomaly represented the highest seawater 87 Sr/86 Sr yet
recorded in Earths Precambrian history. Despite early predictions
that 87 Sr/86 Sr should fall precipitously during the snowball glaciation (due to a shutdown in the continental weathering ux of Sr;
Jacobsen and Kaufman, 1999), 87 Sr/86 Sr was virtually identical at
the time that aragonite crystal fans were deposited during the late
stages of post-glacial transgression (James et al., 2001; Halverson
et al., 2007a). However, 87 Sr/86 Sr appears to have subsequently
risen dramatically to values >0.7078 within a few million years of
the end of glaciation (Melezhik et al., 2009; Halverson et al., 2010),
likely due to a pulse of accelerated silicate weathering (Higgins and
Schrag, 2003).
The base of the Sete Lagoas Formation is marked, in most places,
by a distinctive thin pale and inty cap dolostone unit (010 m),
which records a drop in the 13 C values from 3.2 to 4.5, with
associated 18 O generally between 4.5 and 6.5. The lack of
this thin cap dolostone unit in the Sete Lagoas Formation type-area
is likely due at least in part to resedimentation, as suggested by
the dolostone clasts recovered from the underlying Carrancas Formation. Above the thin cap dolostone, a thick limestone succession
including crystal-fan facies interpreted as aragonite pseudomorphs
record a rapid rise in 13 C from 5 to 0, reecting recovery of
carbon cycle equilibrium in the post-glacial ocean.
The thin cap dolostone and the crystal-fan facies, interpreted
as seaoor precipitates, are two characteristic features of basal
Ediacaran (635 Ma) cap carbonate successions worldwide, and
were never described in sequences unambiguously related to older
(Sturtian) post-glacial successions. Indeed, the deposition of a thin
dolostone layer in the beginning of the Ediacaran Period was
regarded by Shields (2005) as possibly the only global depositional
event throughout the whole stratigraphic record. The crystal-fan
facies, although not unique to cap carbonates, does occur in basal
Ediacaran carbonates around the world, indicating a global CaCO3
supersaturation in the post-Marinoan ocean.
The very high 13 C values found in the upper half of the Sete
Lagoas Formation and also on the Lagoa do Jacar Formation
are higher than any other reported values in the Neoproterozoic
Era (Halverson et al., 2010; Halverson and Shields-Zhou, 2011).
Notably, these unusually high 13 C occur in black, fetid and sometimes oolitic limestones of both the Sete Lagoas and Lagoa do Jacar
formations (Alvarenga et al., 2007; Misi et al., 2007; and references
therein), which were likely deposited in a lagoonal or otherwise
restricted setting, perhaps due to the progressive development of
the Braslia and Aracua forelands. Several earlymiddle Ediacaran
sections exhibit sharp peaks in 13 C and unusual variability in shallow platform carbonates (Halverson and Shields-Zhou, 2011) that
are presumably related to local rather than global variation in seawater 13 C.
The least altered 87 Sr/86 Sr values obtained from throughout the
Sete Lagoas Formation consistently cluster around 0.70740.7076.
These values are characteristic of basal Ediacaran cap carbonates globally and distinct from the post-Sturtian values of
0.70680.7069 (Fig. 11; Jacobsen and Kaufman, 1999; Shields,
1999; Walter et al., 2000; Halverson et al., 2007a, 2010). Babinski
et al. (2007) reported a rise of 87 Sr/86 Sr from high [Sr] limestones of
the Sambra Quarry, from 0.7074 near the base to 0.7081 at the top
of the exposure, accompanying the rise of 13 C from 5 to 0.
Similarly, 87 Sr/86 Sr values rise from 0.7079 to 0.7082 in the Tatiana
Quarry, 5 km to the east (Babinski et al., 2007). This peak appears
to be a global feature of the basal Ediacaran Period (Melezhik et al.,
2009; Halverson et al., 2010), perhaps resulting from a pulse of elevated silicate weathering during the ultra-greenhouse world that

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858

Fig. 11. 87 Sr/86 Sr-13 C diagram showing available data from the Sete Lagoas Formation, compared to worldwide Sturtian and Marinoan carbonate data. Data from
the Lagoa do Jacar Formation of the upper Bambu Group and from the Salitre Formation of the Una Group is added for comparison. See Table 4 for data sources.
The Sturtian and Marinoan elds are drawn from the databases of Halverson et al.
(2007a, 2010), Halverson and Shields-Zhou (2011), and references therein.

followed after the Marinoan icehouse conditions (Shields, 2005;


Halverson et al., 2007a; Le Hir et al., 2009).
Fig. 11 shows a plot of 87 Sr/86 Sr versus 13 C data for the
Sete Lagoas Formation (Table 4), in comparison to the Sturtian
and Marinoan carbonates database from Halverson et al. (2007a),
Halverson and Shields-Zhou (2011), and references therein, which
includes well studied sections where the relative ages of the glacial
sequences are established. Almost all of the Sete Lagoas Formation
data fall within the Marinoan eld, and none within the Sturtian
eld, lending additional support to a lower Ediacaran age assignment for the Sete Lagoas Formation.
Comparison of the Sete Lagoas Formation with similar but better
geochronologically constrained sequences in Western Gondwana
also leads to the same interpretation. For example, in the West

53

Congo fold belt, which is the African counterpart of the Aracua


orogen that separated during the opening of the South Atlantic
Ocean (e.g. Pedrosa-Soares et al., 2008), a thick siliclastic-carbonate
sequence (West Congolian Group) contains two diamictite levels which are considered to be equivalent to the Sturtian and
Marinoan glacials: the Lower and Upper Mixtite formations, respectively (Fig. 12; Cahen, 1978; Tack et al., 2001; Frimmel et al., 2006;
Poidevin, 2007; Tait et al., 2011). The Lower Mixtite Formation is
superseded by the Haut Shiloango Group and the Upper Mixtite
Formation, by the Schisto-Calcaire Subgroup, which bears lithostratigraphic characteristics similar to the Bambu Group. An arkose
sample in the upper part of the Haut Shiloango Group yielded midto end-Cryogenian zircons (U-Pb SHRIMP II), which led Frimmel
et al. (2006) to assign it a minimum depositional age of 650 Ma.
This data reinforces the correlation of the Upper Mixtite and its
Schisto-Calcaire cap carbonate to the Marinoan glaciation.
The 13 C pattern of Sete Lagoas Formation is similar to the prole of the Schisto-Calcaire cap carbonate, and very different from
the underlying Haut Shiloango prole (Fig. 12). While the basal Ediacaran cap carbonate sequence of the Schisto-Calcaire Subgroup
displays an upwards increase of 13 C, from negative values at the
bottom towards positive values at the top, the Haut Shiloango
Group yields positive values between +3.2 and +8 (Fig. 12;
Frimmel et al., 2006). The most striking similarity of the SchistoCalcaire Subgroup with the Bambu Group is, however, the constant
87 Sr/86 Sr signature of 0.7075 (Fig. 12; Frimmel et al., 2006; Poidevin,
2007). The underlying Haut Shiloango Group displays lower values
(0.70680.7072; Frimmel et al., 2006; Poidevin, 2007) that span the
typical mid-Cryogenian range (Halverson and Shields-Zhou, 2011).
Based on this cumulative evidence, we favor an Ediacaran age and
chemostratigraphic correlation between the Bambu Group and the
Schisto-Calcaire Subgroup (Fig. 12), which were, after all, deposited
on the opposite sides of the same narrow, Red-Sea type Neoproterozoic ocean (a branch of the Adamastor Ocean; Pedrosa-Soares et al.,
2008; see inset in Fig. 12).
Similar reasoning can be applied to the Palestina Formation and
overlying Olhos Dgua Formation cap carbonate in the Sergipano
fold belt (Figs. 1 and 3d). The Palestina Formation contains 650
Ma detrital zircons (Oliveira et al., 2005, 2006, 2010), and the onset

Fig. 12. Chemostratigraphic correlation of lower Ediacaran successions on the So Francisco and Congo cratons. Inset map shows the approximate position of the stratigraphic
sections on a paelogeographic reconstruction of the So Francisco (SFC)Congo craton, during the lower Ediacaran. Note change in scale from columns a and c to column b.
See text for data sources. Dots in the upper Schisto-Calcaire Subgroup represent dispersion of 13 C data (Frimmel et al., 2006).

54

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858

of collisional tectonics in the Sergipano fold belt at 630 Ma favors


a late Cryogenian (Marinoan) age for this unit. The chemostratigraphy of the Olhos Dgua Formation compares with post-glacial
carbonate sequences across the CongoSo Francisco craton. In this
regard, the 13 C prole of the Olhos Dgua Formation is very similar to the Sete Lagoas Formation, including an abrupt increase to
values around +8 in the middle part of both formations (Fig. 12;
Sial et al., 2010). The 87 Sr/86 Sr values of both units also overlap,
between 0.7074 and 0.7081, reinforcing their correlation (Fig. 12).
The underlying Jequita and Palestina formations are therefore
broadly contemporaneous and Marinoan in age (Fig. 12).
The Bambu Group in the So Francisco basin is broadly considered as an equivalent to the Una Group in the Irec Basin,
eastern Bahia (Dardenne, 1979; Misi and Veizer, 1998). Here, the
Bebedouro Formation diamictites are overlain by the Salitre Formation cap carbonates that display nearly identical 13 C and 87 Sr/86 Sr
(0.70750.7077) proles as the Sete Lagoas Formation (Fig. 11; Misi
and Veizer, 1998; Guimares et al., 2011). Based on Sr and S isotopic
data of the Salitre Formation, Misi and Veizer (1998) attributed an
end-Cryogenian to early Ediacaran age for the Salitre Formation. We
concur and suggest that the Salitre Formation is equivalent to the
Sete Lagoas Formation, just as the underlying Bebedouro Formation
is correlative with the Jequita Formation.
Finally, in the Paraguay fold belt, on the southwestern margin
of the Amazon craton, the Araras Group also displays a upwards
increasing 13 C prole and 87 Sr/86 Sr signatures of 0.70740.7080
(Alvarenga et al., 2004, 2008; Nogueira et al., 2007) similar to the
lower Sete Lagoas Formation. Furthermore, the base of the Araras
Group is a dolostone (the 2032 m Mirassol Doeste Formation)
containing giant wave ripples and overlain by crystal fan facies
and is generally considered to be a typical basal Ediacaran cap carbonate, overlying the Marinoan Puga Formation glacials (Nogueira
et al., 2003; Allen and Hoffman, 2005; Alvarenga et al., 2008). By
extension, the Araras Group and Puga Formation correlate with the
Sete Lagoas and Jequita formations, respectively.

5. Conclusions
Based on available radiometric, lithostratigraphic, and
chemostratigraphic data, including new data presented here,
we suggest that the entire Sete Lagoas Formation is Ediacaran
in age. Hence, the lower part of this unit comprises the cap
carbonate sequence to the end-Cryogenian (Marinoan) glaciation,
which is represented by the Jequita Formation and equivalent
units elsewhere on the So Francisco craton (e.g. the Bebedouro
Formation) and in the neighboring fold belts (e.g. the Palestina
Formation of the Sergipano fold belt). This conclusion is based on
multiple data sets and lines of reasoning. First, the Sete Lagoas
Formation displays lithostratigraphic features that are similar
to basal Ediacaran cap carbonates worldwide, including a thin
basal pale pink cap dolostone unit and calcite crystal-fan facies
interpreted as aragonite pseudomorphs. Second, the 13 C and 18 O
proles of the Sete Lagoas Formation are nearly identical to those
of other basal Ediacaran cap carbonates, in particular through the
cap dolostone unit where 13 C drops from 3.7 to 5, accompanied by 18 O values around 5. Exceptionally, at the southern
tip of the So Francisco craton, the usual cap dolostone and its
associated isotopic pattern are not observed, most likely because
the basal Sete Lagoas Formation was eroded and resedimented
over the Sete Lagoas paleohigh. Here, the isotopic signature is
preserved in dolostone clasts within the Carrancas Formation,
which is an intraformational, lateral equivalent of the basal Sete
Lagoas Formation, and therefore belongs to the post-glacial cap
carbonate sequence. Third, 87 Sr/86 Sr signatures in the Sete Lagoas
Formation (0.70740.7076) are identical to those in other basal

Ediacaran cap carbonates worldwide. A sharp increase in 87 Sr/86 Sr


values from 0.7074 to 0.7081 in the crystal-fan facies at the base
of the Sete Lagoas Formation has been reported by Babinski et al.
(2007) and is also consistent with the global basal Ediacaran
87 Sr/86 Sr record (Halverson et al., 2010; Halverson and ShieldsZhou, 2011). Fourth, a chemostratigraphic correlation with the
Schisto-Calcaire Subgroup on the West Congo fold belt and with
the Olhos Dgua Formation on the Sergipano fold belt reinforces
a basal Ediacaran age for the Sete Lagoas Formation. All of these
units show similar 13 C and 87 Sr/86 Sr values. Both the Olhos
Dgua and Schisto-Calcaire cap carbonates are constrained to be
younger than 650 Ma, based on detrital zircon data (Oliveira et al.,
2005, 2006, 2010; Frimmel et al., 2006), and the Schisto-Calcaire
Subgroup was deposited on the opposite side of the same narrow,
Red-sea type Ediacaran ocean, with respect to the Sete Lagoas
Formation. And nally, 610 Ma old zircons have been recovered from the middle of the Sete Lagoas Formation (Rodrigues,
2008). Although there is an unconformity below the point where
these zircons were collected, it represents a short-lived hiatus,
as demonstrated by the lack of secular variation of the 87 Sr/86 Sr
signature, which is identical above and below the unconformity
(0.70740.7076).
The interpretation of the Sete Lagoas Formation as a basal Ediacaran cap carbonate and, therefore, the assignment of a Marinoan
age for the preceding glaciation recorded by the Jequita Formation, has major implications for the Neoproterozoic stratigraphic
framework and tectonic evolution of the So Francisco craton and
surrounding fold belts. Most of the Neoproterozoic sedimentary
record on the craton (i.e. the Bambu Group, including the Jequita
Formation and correlated diamictites) must be end-Cryogenian to
Ediacaran in age. Widespread continental rifting represented by
mac dikes and anorogenic granites at around 8501000 Ma suggest, however, that in the surrounding fold belts, older, Tonian to
Cryogenian sequences are probably represented, like in the African
counterpart of the Brasiliano orogenic system (Tack et al., 2001;
Pedrosa-Soares et al., 2008, 2011; Babinski et al., 2012). The difculty in assigning an absolute age for these sequences derives
from the lack of intercalated datable volcanic sequences. Detrital
zircon U-Pb data has thus far proven insufcient to distinguish
the younger sequences, since most stratigraphic units display very
similar youngest age peaks around 8501000 Ma, and no volumetrically signicant sources of younger Neoproterozoic zircons have
been found within the craton to date (Fig. 3; Buchwaldt et al., 1999;
Pedrosa-Soares et al., 2000; Rodrigues, 2008; Figueiredo et al.,
2009; Amorim Dias, 2011; Caxito et al., 2011; Babinski et al., 2012).
Chemostratigraphic data can be used as an alternative tool to rene
the chronostratigraphy of these sequences by correlations across
the So Francisco craton and globally.
Until robust radiometric ages are obtained from the Bambu
Group, the best test of the chronological framework proposed here
is the extent to which it simplies interpretations of the Neoproterozoic geological history of the So Francisco paleocontinent. If
the Jequita Formation records the Marinoan glaciation at 635 Ma,
then for 215 m.y. after the end of the Tonian rifting event around
850 Ma, the So Francisco paleocontinent must have been continuously eroded, acting as a higher, sediment source area for the
surrounding rift and passive margin basins. At around 630 Ma, the
onset of collisional tectonics between the Paran block to the west
and the So Francisco craton to the east (e.g. Pimentel et al., 1999;
Valeriano et al., 2004) initiated lithospheric exure in the foreland of the Braslia orogen, generating the accommodation space
for the Bambu Group to be deposited. Hence, only near the end
of the Brasiliano cycle did the So Francisco craton become a site
of widespread Neoproterozoic sedimentation. This relatively simple geological history is remarkably consistent with the cratonic
stratigraphic record.

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858

Acknowledgements
This work is the result of a sustained research interest on the
Bambu Group, supported mostly by the Universidade Federal de
Minas Gerais and Fundaco de Apoio Pesquisa de Minas Gerais,
Brazil. Prof. Lcio Fraga of the Universidade Federal dos Vales do
Jequitinhonha e Mucuri (UFVJM), Brazil, aided in the collection of
carbonate clasts near Diamantina (MG). The elemental and isotopic
analyses were supported by a Graduate Research Traineeship of FAC
at McGill University and GEOTOP, Montral, Canada. We thank Jean
Francois Hlie (UQAM) and Bill Minarik (McGill) for assistance in
acquisition of stable isotope and Sr element analysis, respectively.
The manuscript was greatly improved after suggestions by A.N. Sial
and an anonymous reviewer.

References
Alkmim, F.F., 2004. O que faz de um crton um crton? O Crton do So Francisco
e as revelaces almeidianas ao delimit-lo. In: Mantesso-Neto V., Bartorelli A.,
Carneiro C.D.R., Brito-Neves B.B. (orgs.) Geologia do Continente Sul-Americano:
evoluco da obra de Fernando Flvio Marques de Almeida, Beca, So Paulo, pp.
1735.
Alkmim, F.F., Martins-Neto, M.A., 2001. A bacia intracratnica do So Francisco:
Arcabouco Estrutural e cenrios evolutivos. In: Pinto, C.P., Martins-Neto, M.
(Eds.). A Bacia do So Francisco geologia e recursos naturais. SBG, Belo Horizonte,
pp. 930.
Allen, P.A., Hoffman, P.F., 2005. Extreme winds and waves in the aftermath of a
Neoproterozoic glaciation. Nature 433, 123127.
Almeida, F.F.M., 1977. O Crton do So Francisco. Revista Brasileira de Geocincias
7, 349364.
Alvarenga, C.J.S., Della Giustina, M.E.S., Silva, M.G.C., Santos, R.V., Gioia, S.M.C.,
Guimares, E.M., Dardenne, M.A., Sial, A.N., Ferreira, V.P., 2007. Variaces dos
istopos de C e Sr em carbonatos pr e ps-glaciaco Jequita (Esturtiano) na
regio de Bezerra-Formosa, Gois. Revista Brasileira de Geocincias 37, 147155.
Alvarenga, C.J.S., Santos, R.V., Dantas, E.L., 2004. C-O-Sr isotopic stratigraphy of cap
carbonates overlying Marinoan-age glacial diamictites in the Paraguay Belt,
Brazil. Precambrian Research 131, 121.
Alvarenga, C.J.S., Dardenne, M.A., Santos, R.V., Brod, E.R., Gioia, S.M.C.L., Sial, A.N.,
Dantas, E.L., Ferreira, V.P., 2008. Isotope stratigraphy of Neoproterozoic cap
carbonates in the Araras Group, Brazil. Gondwana Research 13, 469479.
Amorim Dias, P.H., 2011. Estratigraa e tectnica da Faixa Braslia na regio de Ibi,
Minas Gerais: caracterizaco e estudo de provenincia sedimentar dos sedimentos dos grupos Canastra e Ibi,com base em estudos isotpicos U-Pb e Sm-Nd.
Masters dissertation, Instituto de Geocincias, Universidade Federal de Minas
Gerais, Brazil.
Azmy, K., Kaufman, A.J., Misi, A., Oliveira, T.F., 2006. Isotope stratigraphy of the
Lapa Formation, So Francisco Basin, Brazil: implications for Late Neoproterozoic
glacial events in South America. Precambrian Research 149, 231248.
Azmy, K., Kendall, B., Creaser, R.A., Heaman, L., Oliveira, T.F., 2008. Global correlation
of the Vazante Group, So Francisco Basin, Brazil: ReOs and UPb radiometric
age constraints. Precambrian Research 164, 160172.
Babinski, M., Van Schmus, W.R., Chemale, F., 1999. PbPb dating and Pb isotope
geochemistry of Neoproterozoic carbonate rocks from the So Francisco basin,
Brazil: implications for the mobility of Pb isotopes during tectonism and metamorphism. Chemical Geology 160, 175199.
Babinski, M., Kaufman, A.J., 2003. First direct dating of a Neoproterozoic postglaciogenic cap carbonate. IV South American symposium on isotope geology.
Short Papers 1, 321323.
Babinski, M., Vieira, L.C., Trindade, R.I.F., 2007. Direct dating of the Sete Lagoas
cap carbonate (Bambu Group, Brazil) and implications for the Neoproterozoic
glacial events. Terra Nova 19, 401406.
Babinski, M., Pedrosa-Soares, A.C., Trindade, R.I.F., Martins, M., Noce, C.M., Liu, D.,
2012. Neoproterozoic glacial deposits from the Aracua orogen, Brazil: Age,
provenance and correlations with the So Francisco craton and West Congo
belt. Gondwana Research 21 (23), 451465.
Banner, J.L., Hanson, G.N., 1990. Calculation of simultaneous isotopic and trace element variations during waterrock interaction with application to carbonate
diagenesis. Geochimica et Cosmochimica Acta 55, 28832894.
Bailey, T., McArthur, J., Prince, H., Thirlwall, M., 2000. Dissolution methods for strontium isotope stratigraphy: whole rock analysis. Chemical Geology 167, 313319.
Bekker, A., Sial, A.N., Karhu, J.A., Ferreira, V.P., Noce, C.M., Kaufman, A.J., Romano,
A.W., Pimentel, M.M., 2003. Chemostratigraphy of carbonates from the Minas
Supergroup, Quadriltero Ferryfero (Iron Quadrangle), Brazil: A stratigraphic
record of early proterozoic atmospheric, biogeochemical and climactic change.
American Journal of Science 303, 865904.
Brand, U., Veizer, J., 1980. Chemical diagenesis of a multi-component carbonate
system. 1. Trace elements. Journal of Sedimentary Petrology 50, 12191236.
Brasier, M., McCarron, G., Tucker, R., Leather, J., Allen, P., Shields, G., 2000. New U-Pb
zircon dates for the Neoproterozoic Ghubrah glaciation and for the top of the
Huqf Supergroup, Oman. Geology 28, 175178.

55

Braun, O.P.G., 1968. Contribuico a estratigraa do Bambu. In: Congresso Brasileiro


de Geologia, 22, Belo Horizonte, 1968. Anais. Belo Horizonte: SBG, pp. 154166.
Buchwaldt, R., Toulkeridis, T., Babinski, M., Noce, C.M., Martins-Neto, M.A., Hercos,
C.M., 1999. Age determination and age related provenance analysis of the Proterozoic glaciation event in central eastern Brazil. In: South American Simp. on
Isotope Geology, 2, Crdoba, Argentina, Abstracts, pp. 387390.
Cahen, L., 1978. La stratigraphie et la tectonique du super-group Ouest-Congolien
dans les zones mdianes et externes de lorogne ouest-congolien (Pan-africain)
au bas Zaire et dans les regions limitrophes. Annales du Muse Royal de lAfrique
Centrale, Tervuren, Belgium, Science Geology 83, 150.
Castro, P.T.A., Dardenne, M.A., 2000. The sedimentology, stratigraphy and tectonic
context of the So Francisco Supergroup at the southwestern domain of the So
Francisco craton, Brazil. Revista Brasileira de Geocincias 30 (3), 439441.
Caxito, F.A., Dantas, E.L., Stevenson, R.K., Uhlein, A., 2011. Geochemical, detrital zircon (U-Pb) and Sm-Nd isotopic insights into the sedimentary provenance of the
Rio Preto Fold Belt, Northeastern Brazil. In: Schmitt, R.S., Trouw, R., Carvalho,
I.S., Collins, A. (Eds.). Gondwana 14, Abstracts, Universidade Federal do Rio de
Janeiro, Rio de Janeiro, p. 31.
Chang, H.K., Miranda, F.P., Magalhes, L., Alkmim, F.F., 1988. Consideraces sobre a
evoluco tectnica da bacia do So Francisco. In: Congresso Brasileiro de Geologia, 35, Belm, 1988. Anais, SBG, 5, pp. 20762090.
Chang, H.K., Kawashita, K., Alkmim, F.F., Moreira, M.Z., 1993. Consideraces sobre a
estratigraa isotpica do Grupo Bambu. In: SBG/SGM, II Simpsio sobre o Crton
do So Francisco, Salvador, Anais, pp. 195196.
Chaves, M.L.S., Guimares, J.T., Andrade, K.W., 2010. Litofcies glaciomarinhas na
Formaco Jequita: possveis implicaces na redistribuico de diamantes a oeste
da Serra do Espinhaco (MG). Revista Brasileira de Geocincias 40 (4), 516526.
Chemale, F., Alkmim, F.F., Endo, I., 1993. Late Proterozoic tectonism in the interior
of the So Francisco craton. In: Findlay, R.H., et al. (Eds.), Gondwana EightAssembly, Evolution and Dispersal. Balkema, pp. 2942.
Conceico Filho, V.M., Monteiro, M.D., Rangel, P.A., Garrido, I.A.A., 2003. Bacia do So
Francisco entre Santa Maria da Vitria e Iui, Bahia: geologia e potencialidade
econmica Salvador: CBPM, 2003. 76p.: il., mapa.(Srie Arquivos Abertos, 18).
Condon, D., Zhu, M., Bowring, S.A., Wang, W., Yang, A., Jin, Y., 2005. U-Pb ages from
the Neoproterozoic Doushantuo Formation, China. Science 308, 9598.
Cordani, U.G., Inda, H.A.V., Kawashita, K., 1979. O embasamento do Grupo Bambu na
regio de Correntina, Bacia do So Francisco, estado da Bahia. In: SBG, Simpsio
sobre a Geologia do Crton do So Francisco e suas Faixas Marginais, I, Salvador,
Resumos.
Costa, M.T., Branco, J.R., 1961. Roteiro para a excurso Belo Horizonte-Braslia. 14
Congresso Brasileiro de Geologia. UFMG, Inst. Pesq. Radioat., Publ. 15, 25p., Belo
Horizonte.
Cukrov, N., Alvarenga, C.J.S., Uhlein, A., 2005. Litofcies da glaciaco neoproterozica
nas porces sul do crton do So Francisco: Exemplos de Jequita (MG) e
Cristalina (GO). Revista Brasileira de Geocincias 35 (1), 6976.
DAgrella-Filho, M.S., Babinski, M., Trindade, R.I.F., Van Schmus, W.R., Ernesto, M.,
2000. Simultaneous remagnetization and UPb isotope resetting in Neoproterozoic carbonates of the So Francisco craton, Brazil. Precambrian Research
99, 179196.
Dardenne, M.A., 1978. Sntese sobre a estratigraa do Grupo Bambu no Brasil Central. In: Congresso Brasileiro de Geologia, 30, Recife, 1978. Anais., SBG, 2, pp.
597610.
Dardenne, M.A., 1978. Zonaco tectnica na borda ocidental do Crton do So Francisco. In: Congresso Brasileiro de Geologia, 30, Recife, 1978. Anais., SBG, 2, pp.
299308.
Dardenne, M.A., 1979. Les minralisations de plumb, zinc, uor du Protrozoique
Suprieur dans le Bresil Central. Thse Doctorat dEtat, Univ. Paris VI, Paris (FR),
251 p.
Dardenne, M.A., 2000. The Braslia Fold Belt. In: Cordani, U.G., Milani, E.J., Thomaz, A.,
Filho & Campos, D.A. 2000. Tectonic evolution of South America. Rio de Janeiro,
SBG, pp. 231263.
Egydio-Silva, M., Karmann, I., Trompette, R.R., 1989. Litoestratigraa do Supergrupo
Espinhaco e Grupo Bambu no noroeste do estado da Bahia. Revista Brasileira de
Geocincias 19 (2), 101112.
Fairchild, I.J., 1993. Balmy shores and icy wastes: the paradox of carbonates associated with glacial deposits in Neoproterozoic times. Sedimentology Review 1,
116.
Fanning, C.M., Link, P.K., 2004. U-Pb SHRIMP ages of Neoproterozoic (Sturtian)
glaciogenic Pocatello Formation, southeastern, Idaho. Geology 32, 881884.
Figueiredo, F.T., Almeida, R.P., Tohver, E., Babinski, M., Liu, D., Fanning, C.M., 2009.
Neoproterozoic glacial dynamics revealed by provenance of diamictites of the
Bebedouro Formation, So Francisco craton, Central Eastern Brazil. Terra Nova
21, 375385.
Frimmel, H.E., Kltzli, U.S., Siegfried, P.R., 1996. New Pb-Pb single zircon age constraints on the timing of Neoproterozoic glaciation and continental break-up in
Namibia. Journal of Geology 104, 459469.
Frimmel, H.E., Tack, L., Basei, M.S., Nutman, A.P., Boven, A., 2006. Provenance and
chemostratigraphy of the Neoproterozoic West Congolian Group in the Democratic Republic of Congo. Journal of African Earth Sciences 46, 221239.
Gaucher, C., Sial, A.N., Halverson, G.P., Frimmel, H.E., 2009. The Neoproterozoic and
Cambrian: a time of upheavals, extremes, and innovations. In: Gaucher, C., Sial,
A.N., Halverson, G.P., Frimmel, H.E. (Eds.), Neoproterozoic-Cambrian Tectonics,
Global Change and Evolution: A Focus on Southwestern Gondwana. Developments in Precambrian Geology, vol. 16. Elsevier, pp. 311.
Guimares, J.T., Misi, A., Pedreira, A.J., Dominguez, J.M.L., 2011. The Bebedouro Formation, Una Group, Bahia (Brazil). In: Arnaud, E., Halverson, G.P., Shields-Zhou,

56

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858

G. (Eds.), The Geological Record of Neoproterozoic Glaciations. Memoir 36. Geological Society of London, London, pp. 503508.
Gravenor, C.P., Monteiro, R., 1983. Ice-thrust features in the Proterozoic Macabas
Group, Jequita area, Minas Gerais, Brazil. Journal of Geology 91, 113116.
Halverson, G.P., 2006. A Neoproterozoic chronology. In: Xiao, S., Kaufman, A.J. (Eds.),
Neoproterozoic Geobiology and Paleobiology. Springer, pp. 231271.
Halverson, G.P., Shields-Zhou, G., 2011. Chemostratigraphy and the Neoproterozoic
glaciations. In: Arnaud, E., Halverson, G.P., Shields-Zhou, G. (Eds.), The Geological
Record of Neoproterozoic Glaciations. Memoir 36. Geological Society of London,
London, pp. 541546.
Halverson, G.P., Hoffman, P.F., Schrag, D.P., Maloof, A.C., Rice, A.H., 2005. Towards a
Neoproterozoic composite carbon-isotope record. Geological SocIety of American Bulletin 117, 11811207.
Halverson, G.P., Duds, F.O., Maloof, A.C., Bowring, S.A., 2007a. Evolution of the
87
Sr/86 Sr composition of Neoproterozoic seawater. Palaeogeography Palaeoclimatology Palaeoecology 256, 103129.
Halverson, G.P., Maloof, A.C., Schrag, D.P., Duds, F.O., Hurtgen, M.T., 2007b. Stratigraphy and geochemistry of a ca 800 Ma carbon isotope interval in northeastern
Svalbard. Chemical Geology 237, 2345.
Halverson, G.P., Wade, B.P., Hurtgen, M.T., Barovich, K.M., 2010. Neoproterozoic
chemostratigraphy. Precambrian Research 182, 337350.
Hayes, J.M., Strauss, H., Kaufman, A.J., 1999. The abundance of 13C in marine organic
matter and isotopic fractionation in the global biogeochemical cycle of carbon
during the past 800 Ma. Chemical Geology 161, 103126.
Hettich, M., 1977. A glaciaco proterozica no centro-norte de Minas Gerais. Revista
Brasileira de Geocincias 7, 87101.
Hettich, M., Karfunkel, J., 1978. Um esker, um varvito e seixos estriados no Grupo
Macabas, norte de Minas Gerais. Revista Escola de Minas de Ouro Preto 34, 58.
Higgins, J.A., Schrag, D.P., 2003. Aftermath of a snowball Earth. Geochemistry Geophysics Geosystem 43, 1, doi:10.1029/2002GC000403.
Hoffman, P.F., 2011. Strange bedfellows: glacial diamictite and cap carbonate from the Marinoan (635 ma) glaciation in Namibia. Sedimentology 58,
57119.
Hoffman, P.F., Halverson, G.P., 2008. Otavi Group of the western Northern Platform,
the Eastern Kaoko Zone and the western Northern Margin Zone. In: Miller, R.M.
(Ed.), The Geology of Namibia. Geological Survey of Namibia, Windhoek, pp.
69136.
Hoffman, P.F., Halverson, G.P., 2011. Neoproterozoic glacial record in the Mackenzie Mountains, northwest Canadian Cordillera. In: Arnaud, E., Halverson, G.P.,
Shields-Zhou, G.A. (Eds.), The Geological Record of Neoproterozoic Glaciations,
Memoir 36. Geological Society of London, pp. 397411.
Hoffman, P.F., Schrag, D.P., 2002. The snowball Earth hypothesis: testing the limits
of global change. Terra Nova 14, 129155.
Hoffman, P.F., Hawkins, D.P., Isachsen, C.E., Bowring, S.A., 1996, Precise U-Pb zircon
ages for early Damaran magmatism in the Summas Mountains and Welwitschia
Inlier, northern Damara belt, Namibia: Communications of the Geological Survey
of Namibia, vol. 11, pp. 4752.
Hoffman, P.F., Kaufman, A.J., Halverson, G.P., Schrag, D.P., 1998. A neoproterozoic
snowball Earth. Science 281, 13421346.
Hoffman, P.F., Halverson, G.P., Domack, E.W., Husson, J.M., Higgins, J.A., Schrag, D.P.,
2007. Are basal Ediacaran (635 Ma) post-glacial cap dolostones diachronous.
Earth Planetary Science Letter 258, 114131.
Hoffmann, K.H., Condon, D.J., Bowring, S.A., Crowley, J.L., 2004. A U-Pb zircon date
from the Neoproterozoic Ghaub Formation, Namibia: Constraints on Marinoan
glaciation. Geology 32, 817820.
Hoppe, A.H., Karfunkel, J., Noce, C.M., 2002. Stio Inhama, MG: Camadas aragonticas pr-cambrianas. In: Schobbenhaus, C., Campos, D.A., Queiroz, E.T., Winge,
M., Berbert-Born, M.L.C. (Eds.), Stios Geolgicos e Paleontolgicos do Brasil.
Braslia:DNPM, 2002. SIGEP 88, pp. 175180.
Hork, J.M., Evans, J.A., 2011. Early Neoproterozic limestones from the Gwna Group,
Anglesey. Geological Magazine 148, 7888.
Isotta, C.A.L., Rocha-Campos, A.C., Yoshida, R., 1969. Striated pavement of the Upper
Precambrian glaciation in Brazil. Nature 222, 466468.
Jacobsen, S.B., Kaufman, A.J., 1999. The Sr, C, and O isotopic evolution of Neoproterozoic seawater. Chemical Geology 161, 3757.
James, N., Narbonne, G., Kyser, T., 2001. Late Neoproterozoic cap carbonates:
Mackenzie Mountains, northwestern Canada: precipitation and global glaciation. Canadian Journal of Earth Sciences 38, 12291262.
Karfunkel, J., Hoppe, A., 1988. Late Precambrian glaciation in central-eastern Brazil:
Synthesis and model. Palaeogeography Palaeoclimatology Palaeoecology 65,
121.
Karfunkel, J., Hoppe, A., Noce, C.M., 2002. Serra da gua Fria e vizinhancas, MG: vestgios de glaciaco neoproterozica. In: Schobbenhaus, C., Campos, D.A., Queiroz,
E.T., Winge, M., Berbert-Born, M. (Eds.), Stios Geolgicos e Paleontolgicos do
Brasil, 1. DNPM, Braslia, Brazil, pp. 165173.
Kaufman, A.J., Varni, M.A., Misi, A., Brito-Neves, B.B., 2001. Anomalous d34S signatures in trace sulfate from a potential cap carbonate in the Neoproterozoic
Bambu Group, Brazil. In: Misi, A., Teixeira, J.B.G. (organizers), Proterozoic
Sediment-Hosted Base Metal Deposits of Western Gondwana, I Field Workshop,
Belo Horizonte and Paracatu (Minas Gerais), Brazil, pp. 6265.
Kawashita, K., 1998. Rochas carbonticas neoproterozicas da Amrica do Sul:
idades e inferncias quimioestratigrcas. Full Professor Thesis, University of
So Paulo, Brazil.
Kendall, B., Creaser, R.A., Selby, D., 2006. Re-Os geochronology of postglacial black
shales in Australia: Constraints on the timing of Sturtian glaciation. Geology
34 (9), 729732.

Kennedy, M.J., Runnegar, B., Prave, A.R., Hoffmann, K.-H., Arthur, M.A., 1998. Two or
four Neoproterozoic glaciations? Geology 26, 10591063.
King, L.C., 1956. A geomorfologia do Brasil Oriental. Revista Brasileira de Geograa
2, 147265.
Kirschvink, J.L., 1992. Late Proterozoic low-latitude global glaciation: the snowball
earth. In: Schopf, J.W., Klein, C. (Eds.), The Proterozoic Biosphere. Cambridge
University Press, Cambridge, pp. 5152.
Knoll, A.H., Walter, M.R., 1992. Latest Proterozoic stratigraphy and Earth history.
Nature 356, 673677.
Knoll, A.H., Hayes, J.M., Kaufman, A.J., Swett, K., Lambert, I.B., 1986. Secular variation
in carbon isotope ratios from upper Proterozoic successions of Svalbard and East
Greenland. Nature 321, 832837.
Knoll, A.H., Walter, M.R., Narbonne, G., Christie-Blick, N., 2004. A new period for the
geologic time-scale. Science 305, 621622.
Kuchenbecker, M., 2011. Quimioestratigraa e provenincia sedimentar da porco
basal do Grupo Bambu em Arcos (MG). Masters Dissertation, IGCUniversidade
Federal de Minas Gerais, Belo Horizonte, Brazil, 91 p.
Le Hir, G., Donnadieu, Y., Goddris, Y., Pierrehumbert, R.T., Halverson, G.P., Macouin,
M., Ndlec, A., Ramstein, G., 2009. The snowball Earth aftermath: exploring the
limits of continental weathering processes. Earth and Planetary Science Letters
277, 453463.
Lima, O.N.B., 2011. Estratigraa isotpica e evoluco sedimentar do Grupo Bambu
na borda ocidental do Crton do So Francisco: implicaco tectnica e paleoambiental. PhD Thesis, Instituto de Geocincias, Universidade de Braslia, 114
p.
Lund, K., Aleinikoff, J.N., Evans, K.V., Fanning, C.M., 2003. SHRIMP U-Pb geochronology of Neoproterozoic Windermere Supergroup, central Idaho: Implications
for rifting of western Laurentia and synchroneity of Sturtian glacial deposits.
Geological Society of America Bulletin 115, 349372.
Macdonald, F.A., Schmitz, M.D., Crowley, J.L., Roots, C.F., Jones, D.S., Maloof, A.C.,
Strauss, J.V., Cohen, P.A., Johnston, D.T., Schrag, D.P., 2010a. Calibrating the Cryogenian. Science 327, 12411243.
Macdonald, F.A., Strauss, J.V., Rose, C.V., Duds, F.O., Schrag, D.P., 2010b. Stratigraphy
of the Port Nolloth Group of Namibia and South Africa and implications for the
age of the Neoproterozoic iron formations. American Journal of Science 310,
862888.
Maheshwari, A., Sial, A.N., Gaucher, C., Bossi, J., Bekker, A., Ferreira, V.P., Romano,
A.W., 2010. Global nature of the Paleoproterozoic Lomagundi carbon isotope
excursion: A review of occurrences in Brazil, India, and Uruguay. Precambrian
Research 182 (4), 274299.
Marchese, H.C., 1974. Estromatlitos Gymnosolenides en el lado oriental de Minas
Gerais, Brasil. Revista Brasileira de Geocincias 4 (4), 257272.
Martins, M. S., 2006. Geologia dos diamantes e carbonados aluvionares da bacia do
Rio Macabas, MG. PhD thesis, Instituto de Geocincias, Universidade Federal
de Minas Gerais, Belo Horizonte.
Martins, M.S., Karfunkel, J., Noce, C.M., Babinski, M., Pedrosa-Soares, A.C., Sial, A.N.,
Liu, D., 2008. A seqncia prglacial do Grupo Macabas na rea-tipo e o registro
da abertura do rifte Aracua. Revista Brasileira de Geocincias 38, 768779.
Martins, M., Lemos, V.B., 2007. Anlise estratigrca das sequncias neoproterozicas da Bacia do So Francisco? Revista Brasileira de Geocincias 37
(4-suplemento), 156167.
Martins-Neto, M.A., Alkmim, F.F., 2001. Estratigraa e evoluco tectnica das bacias
neoproterozicas do paleocontinente So Francisco e suas margens: registro da
quebra de Rodnia e colagem de Gondwana. In: Pinto C.P., Martins-Neto M.A.
(Eds.), Bacia do So Francisco: Geologia e Recursos Naturais, SBG/MG, pp. 930.
Martins-Neto, M.A., Pedrosa-Soares, A.C., Lima, S.A.A., 2001. Tectono-sedimentary
evolution of sedimentary basins from Late Paleoproterozoic to Late Neoproterozoic in the So Francisco craton and Aracua fold belt, eastern Brazil. Sedimentary
Geology 141/142, 343370.
Melezhik, V., Gorokhov, I., Kuznetsov, A., Fallick, A., 2001. Chemostratigraphy of
Neoproterozoic carbonates: implications for blind dating. Terra Nova 13, 111.
Melezhik, V.A., Pokrovsky, B.G., Fallick, A.E., Kuznetsov, A.B., Bujakaite, M.I., 2009.
Constraints on the 87Sr/86Sr of Late Ediacaran seawater: insights from high-Sr
limestones. Journal of Geological Society (London) 166, 183191.
Misi, A., Veizer, J., 1998. Neoproterozoic carbonate sequences of the Una Group, Irece
Basin, Brasil: chemostratigraphy, age and correlations. Precambrian Research
89, 87100.
Misi, A., Iyer, S.S., Coelho, C.E.S., Tassinari, C.C.G., Franca-Rocha, W.J.S., Cunha, I.A.,
Gomes, A.S.R., Oliveira, T.F., Teixeira, J.B.G., Mnaco Filho, V., 2005. Sediment
hosted leadzinc deposits of the Neoproterozoic Bambu Group and correlative
sequences, So Francisco craton, Brazil: A review and a possible metallogenic
evolution model. Ore Geology Reviews 26, 263304.
Misi, A., Kaufman, A.J., Veizer, J., Powis, K., Azmy, K., Boggiani, P.C., Gaucher,
C., Teixeira, J.B., Sanches, A.L., Iyer, S.S., 2007. Chemostratigraphic correlation of Neoproterozoic successions in South America. Chemical Geology 237,
143167.
Misi, A., Kaufman, A.J., Azmy, K., Dardenne, M.A., Sial, A.N., Oliveira, T.F., 2011.
Neoproterozoic successions of the So Francisco craton, Brazil: the Bambu,
Una, Vazante and Vaza Barris/Miaba groups and their glaciogenic deposits. In:
Arnaud, E., Halverson, G.P., Shields-Zhou, G. (Eds.), The Geological Record of
Neoproterozoic Glaciations. Memoir 36. Geological Society of London, London,
pp. 509522.

I., Banner, J., Osleger, D., Borg, L., Bosserman, P., 1996. Integrated Sr isoMontanez,
tope variations and sea-level history of Middle and Upper Cambrian platform
carbonates: implications for the evolution of Cambrian seawater 87Sr/86. Sr.
Geology 24, 917920.

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858


Moraes, L.J., Guimares, D., 1931. The diamond-bearing region of Northern Minas
Gerais, Brazil. Economic Geology 26, 502530.
Noce, C.M., Pedrosa-Soares, A.C., Grossi-Sad, J.H., Baars, F.J., Guimares, M.V.,
Mouro, M.A.A., Oliveira, M.J.R., Roque, N.C., 1997. Nova Subdiviso Estratigrca
Regional do Grupo Macabas na Faixa Aracua: O Registro de uma Bacia Neoproterozica. Boletim do Ncleo Minas Gerais-Sociedade Brasileira de Geologia
14, 2931.
Nogueira, A.C.R., Riccomini, C., Sial, A.N., Moura, C.A.V., Fairchild, T.R., 2003. Softsediment deformation at the base of the Neoproterozoic Puga cap carbonate
(southwestern Amazon craton, Brazil): Conrmation of rapid icehouse to greenhouse transition in snowball Earth. Geology 31, 613616.
Nogueira, A.C.R., Riccomini, C., Sial, A.N., Moura, C.A.V., Trindade, R.I.F., Fairchild,
T.R., 2007. C and Sr isotope uctuations and paleoceanographic changes in the
Late Neoproterozoic Araras carbonate platform, southern Amazon craton, Brazil.
Chemical Geology 237, 168190.
Oliveira, E.P., Carvalho, M.J., Nascimento, R.S., Arajo, M.N.C., Dantas, D., Basilici, G.,
Bueno, J.F., McNaughton, N., 2005. Evidence from detrital zircon geochronology
and whole-rock SmNd isotopes for off-craton provenance of clastic metasedimentary units of the Sergipano belt, NE Brazil. In: X Simpsio Nacional de
Estudos Tectnicos, Curitiba, Boletim de Resumos Expandidos, pp. 308311.
Oliveira, E.P., Toteu, S.F., Arajo, M.N.C., Carvalho, M.J., Nascimento, R.S., Bueno, J.F.,
McNaughton, N., Basilici, G., 2006. Geologic correlation between the Neoproterozoic Sergipano belt (NE Brazil) and the Yaound schist belt (Cameroon, Africa).
Journal of African Earth Sciences 44, 470478.
Oliveira, E.P., Windley, B.F., Arajo, M.N.C., 2010. The Neoproterozoic Sergipano orogenic belt, NE Brazil: A complete plate tectonic cycle in western Gondwana.
Precambrian Research 181, 6484.
Pedrosa-Soares, A.C., Cordani, U., Nutman, A., 2000. Constraining the age of Neoproterozoic glaciation in eastern Brazil: First UPb (SHRIMP) data from detrital
zircons. Revista Brasileira de Geocincias 30, 5861.
Pedrosa-Soares, A.C., Alkmim, F.F., Tack, L., Noce, C.M., Babinski, M., Silva, L.C.,
Martins-Neto, M.A., 2008. Similarities and differences between the Brazilian
and African counterparts of the Neoproterozoic Aracua-West Congo orogen. In:
Pankhurst, R.J., Trouw, R.A.J., Brito Neves, B.B., De Wit, M.J. (Eds.), West Gondwana: Pre-Cenozoic Correlations Across the South Atlantic Region. Geological
Society, London, Special Publications 294, pp. 153172.
Pedrosa-Soares, A.C., Babinski, M., Noce, C., Martins, M.S., Queiroga, G. & Vilela, F.,
2011. The Neoproterozoic Macabas Group (Aracua orogen, SE Brazil) with
emphasis on the diamictite formations. In: Arnaud, E., Halverson, G., Shields, G.
(Eds.), The Geological Record of Neoproterozoic Glaciations, Geological Society
of London, Memoir 36, pp. 523534.
Pereira, L., Dardenne, M.A., Rosire, C.A., Pedrosa-Soares, A.C., 1994. Evoluco
Geolgica dos Grupos Canastra e Ibi na regio entre Coromandel e Guarda-Mor,
MG. Geonomos 2 (1), 2232.
Peryt, T.M., Hoppe, A., Bechstdt, T., Kster, J., Pierre, C., Richter, D.K., 1990. Late
Proterozoic aragonitic cement crusts, Bambu Group, Minas Gerais, Brazil. Sedimentology 37, 279286.
Pug, R., Schll, W.U., 1975. Proterozoic glaciation in Eastern Brazil: A review. Geologische Rundschau 64, 287299.
Pimentel, M.M., Fuck, R.A., Botelho, N.F., 1999. Granites and the geodynamic history
of the Braslia Belt, Central Brazil: a review. Lithos 46, 463483.
Poidevin, J.L., 2007. Stratigraphie isotopique du strontium et datation des formations carbonates et glaciogniques noprotrozoiques du Nord et de lOuest
du craton du Congo. Comptes Rendus Geoscience 339, 259273.
Prave, A.R., Fallick, A.E., Thomas, C.W., Graham, C.M., 2009. A composite C-isotope
prole for the Neoproterozoic of Scotland and Ireland. Journal of Geological
Society (London) 166, 845857.
Queiroga, G.N., Pedrosa-Soares, A.C., Noce, C.M., Alkmim, F.F., Pimentel, M.M., Dan
C., Suita, M.T.F., Prichard, H., 2007. Age of the
tas, E., Martins, M., Castaneda,
Ribeiro da Folha ophiolite, Aracua Orogen: The U-Pb zircon dating of a plagiogranite. Geonomos 15, 6165.
Ribeiro, J.H., Tuller, M.P., Pinho, J.M.M., Signorelli, N., Fboli, W.L., 2008. A fcies
diamictito da Formaco Carrancas, Grupo Bambu, na regio sudoeste da bacia
do So Francisco, Minas Gerais. In: Congresso Brasileiro de Geologia, 44, CuritibaPR, Anais, p. 913.
Rocha-Campos, A.C., Hasui, Y., 1981. Tillites of the Macabas Group (Proterozoic) in
central Minas Gerais and southern Bahia, Brazil. In: Hambrey, M.J., Harland, W.B.
(Eds.), Earthss pre-Pleistocene Glacial Record. Cambridge University Press, pp.
933939.
Rocha-Campos, A.C., Young, G.M., Santos, P.R., 1996. Re-examination of a striated pavement near Jequita, MG: implications for proterozoic stratigraphy and
glacial geology. Anais da Academia Brasileira de Cincias 68 (4), 593.
Rocha-Campos, A.C., Brito Neves, B.B., Babinski, M., Santos, P.R., Romano, A.W., 2007.
Laminito Moema: unidade Neoproterozica de provvel origem glaciognica,
no centro-leste do estado de Minas Gerais. In: Simpsio de Geologia
do Sudeste, 10, Diamantina, SBG-MG, Programaco e Livro de Resumos,
p. 90.
Rocha-Campos, A.C., Brito Neves, B.B., Babinski, M., Santos, P.R., Oliveira, S.M.B.,
Romano, A.W., 2007b. Moema laminites: a newly recognized Neoproterozoic
(?) glacigenic unit, So Francisco Basin, Brazil. In: Arnaud, E., Halverson, G.P.,
Shields-Zhou, G. (Eds.), The Geological Record of Neoproterozoic Glaciations.
Memoir 36. Geological Society of London, London, pp. 535540.
Rodrigues, J.B., 2008. Provenincia de sedimentos dos grupos Canastra, Ibi, Vazante
e BambuUm estudo de zirces detrticos e Idades Modelo Sm-Nd. Phd Thesis,
Universidade de Braslia, Brazil, 128 p.

57

Rodrigues, J.B., Pimentel, M.M., Dardenne, M.A., Armstrong, R., 2010. Age, provenance and tectonic setting of the Canastra and Ibi Groups (Braslia Belt, Brazil):
Implications for the age of a Neoproterozoic glacial event in central Brazil. Journal of South American Earth Sciences 29, 512521.
Romano, A.W., 2007. Par de Minas- SE.23-Z-C-IV, escala 1:100.000: nota explicativa, UFMG/CPRM, Belo Horizonte, 65 p.
Romano, A.W., Knauer, L.G., 2003. Evidncias da glaciaco neoproterozica na base
do Grupo Bamburegio de Onca do PitanguiMinas Gerais. In: Simpsio de
Geologia de Minas Gerais, 12, Ouro Preto, CD-ROM.
Santos, R.V., Alvarenga, C.J.S., Dardenne, M.A., Sial, A.N., Ferreira, V.P., 2000.
Carbon and oxygen isotope proles across Meso-Neoproterozoic limestones
from central Brazil: Bambu and Parano groups. Precambrian Research 104,
107122.
Santos, R.V., Alvarenga, C.J.S., Babinski, M., Ramos, M.L., Cukrov, N., Fonseca,
M.A., Sial, A.N., Dardenne, M.A., Noce, C.M., 2004. Carbon isotopes of
MesoproterozoicNeoproterozoic sequences from Southern So Francisco craton and Aracua Belt, Brazil: Paleographic implications. Journal of South
American Earth Sciences 18, 2739.
Shields, G., 1999. Working towards a new stratigraphic calibration scheme
for the NeoproterozoicCambrian. Eclogae Geologicae Helvitiae 92,
221233.
Shields, G.A., 2005. Neoproterozoic cap carbonates: a critical appraisal of existing
models and the plumeworld hypothesis. Terra Nova 17, 299310.
Sial, A.N., Ferreira, V.P., Almeida, A.R., Romano, A.W., Parente, C.V., Costa, M.L.,
Santos, V.H., 2000. Carbon isotope uctuations in Precambrian carbonate
sequences of several localities in Brazil. Anais da Academia Brasileira de Cincias
72 (4).
Sial, A.N., Dardenne, M.A., Misi, A., Pedreira, A.J., Gaucher, C., Ferreira, V.P., Silva
Filho, M.A., Uhlein, A., Pedrosa-Soares, A.C., Santos, R.V., Egydio-Silva, M., Babinski, M., Alvarenga, C.J.S., Fairchild, T.R., Pimentel, M.M., 2009. The So Francisco
Palaeocontinent. In: Gaucher, C., Sial, A.N., Halverson, G.P., Frimmel, H.E. (Eds):
Neoproterozoic-Cambrian Tectonics, Global Change and Evolution: a focus on
southwestern Gondwana. Developments in Precambrian Geology, 16, Elsevier,
pp. 3169.
Sial, A.N., Gaucher, C., Silva-Filho, M.A., Ferreira, V.P., Pimentel, M.M., Lacerda, L.D.,
Silva-Filho, E.V., Cezario, W., 2010. C-, Sr isotope and Hg chemostratigraphy
of Neoproterozoic cap carbonates of the Sergipano Belt, Northeastern Brazil.
Precambrian Research 182, 351372.
Sohl, L.E., Christie-Blick, N., Kent, D.V., 1998. Paleomagnetic polarity reversals in
Marinoan (ca. 600 Ma) glacial deposits of Australia: implications for the duration
of low-latitude glaciations in Neoproterozoic time. Geological Society of America
Bulletin 111, 11201139.
Souza, J.D., Kosin, M., Heineck, C.A., Lacerda Filho, J.V., Teixeira, L.R., Valente, C.R.,
Guimares, J.T., Bento, R.V., Borges, V.P., Santos, R.A., Leite, C.A., Neves, J.P.,
Oliveira, I.W.V., Carvalho, L.M., Pereira, L.H.M., Paes, V.J.C., 2004. Folha SD.23Braslia. In: Schobbenhaus, C., Goncalves, J.H., Santos, J.O.S., Abram, M.B., Leo
Neto, R., Matos, G.M.M., Vidotti, R.M., Ramos, M.A.B., Jesus, J.D.A. (eds). Carta
geolgica do Brasil ao milionsimo, Sistema de Informaces GeogrcasSIG,
Programa Geologia do Brasil, CPRM, Braslia. CD-ROM.
Tack, L., Wingate, M.T.D., Ligeois, J-P., Fernandez-Alonso, M., Deblond, A.,
2001. Early Neoproterozoic magmatism (1000910 Ma) of the Zadinian and
Mayumbian Groups (Bas-Congo): onset of Rodinia rifting at the western edge of
the Congo craton. Precambrian Research 110, 277306.
Tait, J., Delpomdor, F., Prat, A., Tack, L., Straathof, G., Nkula, V.K., 2011. Neoproterozoic sequences of the West Congo and Lindi/Ubangi Supergroups in the Congo
craton, Central Africa. In: Arnaud, E., Halverson, G.P., Shields-Zhou, G. (Eds.), The
Geological Record of Neoproterozoic Glaciations. Memoir 36. Geological Society
of London, London, pp. 185194.
Trompette, R.R., 1994. Geology of Western Gondwana (2000-500 Ma). Pan-AfricanBrasiliano aggregation of South America and Africa. A.A. Balkema, Rotterdam.
Tuller, M.P., Ribeiro, J.H., Signorelli, N., Feboli, W.L., Pinho, J.M.M., 2009. Projeto
Sete Lagoas-Abaet, Estado de Minas Gerais: texto explicativo, Belo Horizonte:
CPRM-BH.
Uhlein, A., Trompette, R.R., Alvarenga, C.J.S., 1999. Neoproterozoic glacial and gravitational sedimentation on a continental rifted margin: the Jequita-Macabas
sequence (Minas Gerais, Brazil). Journal of South American Earth Sciences 12,
435451.
Uhlein, A., Alvarenga, C.J.S., Trompette, R.R., Dupont, H.S.J.B., Egydio-Silva, M.,
Cukrov, N., Lima, O.N.B., 2004. Glaciaco neoproterozica sobre o crton do
So Francisco e faixas dobradas adjacentes. In: Mantesso-Neto V., Bartorelli A.,
Carneiro C.D.R., Brito-Neves B.B. (orgs.) Geologia do Continente Sul-Americano:
evoluco da obra de Fernando Flvio Marques de Almeida, Beca, So Paulo, pp.
539553.
Uhlein, A., Alvarenga, C.J.S., Dardenne, M.A., Trompette, R.R., 2011. The glaciogenic
Jequita Formation, southeastern Brazil. In: Arnaud, E., Halverson, G.P., ShieldsZhou, G. (Eds.), The Geological Record of Neoproterozoic Glaciations. Memoir
36. Geological Society of London, London, pp. 5166.
Valeriano, C.M., Simes, L.S.A., Teixeira, W., Heilbron, M., 2000. Southern Brasilia
belt (SE Brazil): tectonic discontinuities, KAr data and evolution during
the Neoproterozoic Brasiliano orogeny. Revista Brasileira de Geocincias 30,
195199.
Valeriano, C.M., Machado, N., Simonetti, A., Valadares, C.S., Seer, H.J., Simes, L.S.A.,
2004. UPb geochronology of the southern Braslia belt (SE-Brazil): sedimentary provenance, Neoproterozoic orogeny and assembly of West Gondwana.
Precambrian Research 130, 2755.

58

F.d.A. Caxito et al. / Precambrian Research 200203 (2012) 3858

Vieira, L.C., Trindade, R.I.F., Nogueira, A.C.R., Ader, M., 2007. Identication of a
Sturtian cap carbonate in the Neoproterozoic Sete Lagoas carbonate platform,
Bambu Group, Brazil. Comptes Rendus Geoscience 339, 240258.
Viveiros, J.F.M., Walde, D., 1976. Geologia da Serra do Cabral, Minas Gerais,
Brasil. Mnsterische Forschungshefte Geologie und Palaeontologie 38/39,
1527.

Walter, M.R., Veevers, J.J., Calver, C.R., Gorjan, P., Hill, A.C., 2000. Dating the 840544
Ma Neoproterozoic interval by isotopes of strontium, carbon, and sulfur in seawater, and some interpretive models. Precambrian Research 100, 371433.
Zaln, P.V., Romeiro-Silva, P.C., 2007. Proposta de mudanca signicativa na coluna
estratigrca da Bacia do So Francisco. In: Simpsio de Geologia do Sudeste,
10, Diamantina, SBG-MG, Programaco e Livro de Resumos, p. 96.

You might also like